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

Noncommutative Ergodic Optimization

Aidan Young1 University 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 (X,G,U)(X,G,U), and a real-valued continuous function fC(X)f\in C(X), what values can fdμ\int f\mathrm{d}\mu take when μ\mu is an invariant Borel probability measure on XX, 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 𝔄\mathfrak{A}, let Aut(𝔄)\operatorname{Aut}(\mathfrak{A}) denote the family of all *-automorphisms of 𝔄\mathfrak{A}. We endow Aut(𝔄)\operatorname{Aut}(\mathfrak{A}) with the point-norm topology, i.e. the topology induced by the pseudometrics

(Φ,Ψ)\displaystyle(\Phi,\Psi) Φ(a)Ψ(a)\displaystyle\mapsto\left\|\Phi(a)-\Psi(a)\right\| (a𝔄).\displaystyle(a\in\mathfrak{A}).

This topology makes Aut(𝔄)\operatorname{Aut}(\mathfrak{A}) a topological group [Blackadar, II.5.5.4].

We define a C*-dynamical system to be a triple (𝔄,G,Θ)(\mathfrak{A},G,\Theta) consisting of a unital C*-algebra 𝔄\mathfrak{A}, a topological group GG (called the phase group), and a point-continuous left group action Θ:GAut(𝔄)\Theta:G\to\operatorname{Aut}(\mathfrak{A}).

Notation 1.1.

Let (𝔄,G,Θ)(\mathfrak{A},G,\Theta) be a C*-dynamical system, and let FGF\subseteq G be a nonempty finite subset. We define AvgF:𝔄𝔄\operatorname{Avg}_{F}:\mathfrak{A}\to\mathfrak{A} by

AvgFx:=1|F|gFΘga.\operatorname{Avg}_{F}x:=\frac{1}{|F|}\sum_{g\in F}\Theta_{g}a.

Denote by 𝒮\mathcal{S} the family of all states on 𝔄\mathfrak{A} endowed with the weak*-topology, and by 𝒯\mathcal{T} the subfamily of all tracial states on 𝔄\mathfrak{A}. A state ϕ\phi on 𝔄\mathfrak{A} is called Θ\Theta-invariant (or simply invariant if the action Θ\Theta is understood in context) if ϕ=ϕΘg\phi=\phi\circ\Theta_{g} for all gGg\in G. Denote by 𝒮G𝒮\mathcal{S}^{G}\subseteq\mathcal{S} the family of all Θ\Theta-invariant states on 𝔄\mathfrak{A}, and by 𝒯G𝒯\mathcal{T}^{G}\subseteq\mathcal{T} the family of all Θ\Theta-invariant tracial states on 𝔄\mathfrak{A}. The set 𝒮G\mathcal{S}^{G} (resp. 𝒯G\mathcal{T}^{G}) is weak*-compact in 𝒮\mathcal{S} (resp. in 𝒯\mathcal{T}). Unless otherwise stated, whenever we deal with subspaces of 𝒮\mathcal{S}, we consider these subspaces equipped with the weak*-topology.

We will assume for the remainder of this section that (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is a C*-dynamical system such that 𝔄\mathfrak{A} is separable, and also that 𝒮G\mathcal{S}^{G}\neq\emptyset. This framework will include every system of the form (C(Y),G,Θ)(C(Y),G,\Theta), where YY is a compact metrizable topological space, the group GG is countable, discrete, and amenable, and Θ\Theta is of the form Θg:ffUg\Theta_{g}:f\mapsto f\circ U_{g} for all gGg\in G, where U:GYU:G\curvearrowright Y is a right action of GG on YY 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 Θ\Theta-invariance of certain states later.

Lemma 1.2.

Let (𝔄,G,Θ)(\mathfrak{A},G,\Theta) be a C*-dynamical system, and let GG be an amenable group. If (ϕk)k=1(\phi_{k})_{k=1}^{\infty} is a sequence in 𝒮\mathcal{S}, and 𝐅=(Fk)k=1\mathbf{F}=(F_{k})_{k=1}^{\infty} is a right Følner sequence for GG, then any weak*-limit point of the sequence (ϕkAvgFk)k=1\left(\phi_{k}\circ\operatorname{Avg}_{F_{k}}\right)_{k=1}^{\infty} is Θ\Theta-invariant. In particular, if KK is a nonempty, Θ\Theta-invariant, weak*-compact, convex subset of 𝒮\mathcal{S}, then K𝒮GK\cap\mathcal{S}^{G}\neq\emptyset.

Proof.

Let (ϕk)k=1(\phi_{k})_{k=1}^{\infty} be a sequence of states, and fix g0G,x𝔄g_{0}\in G,x\in\mathfrak{A}. Then

|ϕk(AvgFkΘg0x)ϕk(AvgFkx)|\displaystyle\left|\phi_{k}\left(\operatorname{Avg}_{F_{k}}\Theta_{g_{0}}x\right)-\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right)\right|
=\displaystyle= |1|Fk|[ϕk(gFkg0Θgx)ϕk(gFkΘgx)]|\displaystyle\left|\frac{1}{|F_{k}|}\left[\phi_{k}\left(\sum_{g\in F_{k}g_{0}}\Theta_{g}x\right)-\phi_{k}\left(\sum_{g\in F_{k}}\Theta_{g}x\right)\right]\right|
\displaystyle\leq |1|Fk|ϕk(gFkg0FkΘgx)|+|1|Fk|ϕk(gFkFkg0Θgx)|\displaystyle\left|\frac{1}{|F_{k}|}\phi_{k}\left(\sum_{g\in F_{k}g_{0}\setminus F_{k}}\Theta_{g}x\right)\right|+\left|\frac{1}{|F_{k}|}\phi_{k}\left(\sum_{g\in F_{k}\setminus F_{k}g_{0}}\Theta_{g}x\right)\right|
\displaystyle\leq |Fkg0ΔFk||Fk|x\displaystyle\frac{|F_{k}g_{0}\Delta F_{k}|}{|F_{k}|}\|x\|
k\displaystyle\stackrel{{\scriptstyle k\to\infty}}{{\to}} 0.\displaystyle 0.

Therefore, if k1<k2<k_{1}<k_{2}<\cdots is such that ψ=limϕkAvgFk\psi=\lim_{\ell\to\infty}\phi_{k_{\ell}}\circ\operatorname{Avg}_{F_{k_{\ell}}} exists, then

|ψ(Θgx)ψ(x)|lim sup|Fkg0ΔFk||Fk|x=0.\left|\psi(\Theta_{g}x)-\psi(x)\right|\leq\limsup_{\ell\to\infty}\frac{\left|F_{k_{\ell}}g_{0}\Delta F_{k_{\ell}}\right|}{\left|F_{k_{\ell}}\right|}\|x\|=0.

Finally, let KK be a nonempty, Θ\Theta-invariant, weak*-compact, convex subset of 𝒮\mathcal{S}. Let ϕ\phi be any state in KK, and consider the sequence (ϕAvgFk)k=1\left(\phi\circ\operatorname{Avg}_{F_{k}}\right)_{k=1}^{\infty}. By the convexity and Θ\Theta-invariance of KK, every term of this sequence is an element of KK, and since KK is compact, there exists a subsequence of this sequence which converges in KK. As has already been shown, that limit must be an element of 𝒮G\mathcal{S}^{G}. ∎

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 𝔄\mathfrak{A} be separable is not necessary for this proof of Lemma 1.2.

Corollary 1.4.

If 𝒯\mathcal{T}\neq\emptyset, and GG is amenable, then 𝒯G\mathcal{T}^{G}\neq\emptyset.

Proof.

Apply Lemma 1.2 to the case where K=𝒯K=\mathcal{T}. ∎

Definition 1.5.

We denote by \mathfrak{R} the real Banach space of all self-adjoint elements of 𝔄\mathfrak{A}, and denote by \mathfrak{R}^{\natural} the space of all real self-adjoint bounded linear functionals on 𝔄\mathfrak{A}.

Definition 1.6.

Let VV be a locally convex topological real vector space, and let KK be a compact subset of VV which is contained in a hyperplane that does not contain the origin. We call KK a simplex if the positive cone P={ck:c0,kK}P=\left\{ck:c\in\mathbb{R}_{\geq 0},k\in K\right\} defines a lattice ordering on PP={p1p2:p1,p2P}VP-P=\{p_{1}-p_{2}:p_{1},p_{2}\in P\}\subseteq V with respect to the partial order abbaPa\leq b\iff b-a\in P.

Remark 1.7.

In Definition 1.6, the assumption that KK 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 𝒮\mathcal{S} (and by extension 𝒮G,𝒯,𝒯G\mathcal{S}^{G},\mathcal{T},\mathcal{T}^{G}) lives in the real hyperplane {ϕ:ϕ(1)=1}\left\{\phi\in\mathfrak{R}^{\natural}:\phi(1)=1\right\} defined by the evaluation at 11.

We begin with the following lemma.

Lemma 1.8.
  1. (i)

    The spaces 𝒮,𝒮G,𝒯,𝒯G\mathcal{S},\mathcal{S}^{G},\mathcal{T},\mathcal{T}^{G} are compact and metrizable.

  2. (ii)

    If 𝒯\mathcal{T}\neq\emptyset, then the space 𝒯G\mathcal{T}^{G} is a simplex.

Before proving this lemma, we need to introduce some terminology. Let ϕ,ψ\phi,\psi be two positive linear functionals on a unital C*-algebra 𝔄\mathfrak{A}. We say that the two positive functionals are orthogonal, notated ϕψ\phi\perp\psi, if they satisfy either of the following two equivalent conditions:

  1. (a)

    ϕ+ψ=ϕ+ψ\|\phi+\psi\|=\|\phi\|+\|\psi\|.

  2. (b)

    For every ε>0\varepsilon>0 exists positive z𝔄z\in\mathfrak{A} of norm 1\leq 1 such that ϕ(1z)<ε,ψ(z)<ε\phi(1-z)<\varepsilon,\psi(z)<\varepsilon.

It is well-know that these conditions are equivalent [Pedersen, Lemma 3.2.3]. For every ϕ\phi\in\mathfrak{R}^{\natural}, there exist unique positive linear functionals ϕ+,ϕ\phi^{+},\phi^{-} such that ϕ=ϕ+ϕ\phi=\phi^{+}-\phi^{-}, and ϕ+ϕ\phi^{+}\perp\phi^{-}, called the Jordan decomposition of ϕ\phi [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 𝔄\mathfrak{A} be a unital C*-algebra, and ϕ\phi\in\mathfrak{R}^{\natural}. Suppose that ϕ(xy)=ϕ(yx)\phi(xy)=\phi(yx) for all x,y𝔄x,y\in\mathfrak{A}. Then ϕ±(xy)=ϕ±(yx)\phi^{\pm}(xy)=\phi^{\pm}(yx) for all x,y𝔄x,y\in\mathfrak{A}.

Proof.

Let 𝒰(𝔄)\mathcal{U}(\mathfrak{A}) denote the group of unitary elements in 𝔄\mathfrak{A}. For a unitary element u𝒰(𝔄)u\in\mathcal{U}(\mathfrak{A}), let AduAut(𝔄)\operatorname{Ad}_{u}\in\operatorname{Aut}(\mathfrak{A}) denote the inner automorphism

Adux=uxu.\operatorname{Ad}_{u}x=uxu^{*}.

Let ψ𝔄\psi\in\mathfrak{A}^{\prime}. We claim that ψ\psi is tracial if and only if ψAdu=ψ\psi\circ\operatorname{Ad}_{u}=\psi for all unitaries u𝒰(𝔄)u\in\mathcal{U}(\mathfrak{A}).

Let u𝒰(𝔄)u\in\mathcal{U}(\mathfrak{A}) be unitary, and x𝔄x\in\mathfrak{A} an arbitrary element. Then

ϕ(ux)\displaystyle\phi(ux) =ψ(u(xu)u)\displaystyle=\psi\left(u(xu)u^{*}\right)
=ψ(Adu(xu)).\displaystyle=\psi(\operatorname{Ad}_{u}(xu)).

So ψ(ux)=ψ(xu)\psi(ux)=\psi(xu) if and only if ψ(Adu(xu))=ψ(xu)\psi(\operatorname{Ad}_{u}(xu))=\psi(xu).

In one direction, suppose that ψ=ψAdu\psi=\psi\circ\operatorname{Ad}_{u} for all u𝒰(𝔄).u\in\mathcal{U}(\mathfrak{A}). Fix x,y𝔄x,y\in\mathfrak{A}. Then we can write y=j=14cjujy=\sum_{j=1}^{4}c_{j}u_{j} for some c1,,c4c_{1},\ldots,c_{4}\in\mathbb{C} and unitaries u1,,u4𝒰(𝔄)u_{1},\ldots,u_{4}\in\mathcal{U}(\mathfrak{A}) unitary. Then

ψ(xy)\displaystyle\psi(xy) =ψ(xj=14cjuj)\displaystyle=\psi\left(x\sum_{j=1}^{4}c_{j}u_{j}\right)
=j=14cjψ(xuj)\displaystyle=\sum_{j=1}^{4}c_{j}\psi(xu_{j})
=j=14cjψ(Aduj(xuj))\displaystyle=\sum_{j=1}^{4}c_{j}\psi(\operatorname{Ad}_{u_{j}}(xu_{j}))
=j=14cjψ(ujx)\displaystyle=\sum_{j=1}^{4}c_{j}\psi(u_{j}x)
=ψ((j=14cjuj)x)\displaystyle=\psi\left(\left(\sum_{j=1}^{4}c_{j}u_{j}\right)x\right)
=ψ(yx).\displaystyle=\psi(yx).

Thus ψ\psi is tracial.

In the other direction, suppose there exists u𝒰(𝔄)u\in\mathcal{U}(\mathfrak{A}) such that ψAduψ\psi\circ\operatorname{Ad}_{u}\neq\psi. Let y𝔄y\in\mathfrak{A} such that ψ(y)ψ(Aduy)\psi(y)\neq\psi(\operatorname{Ad}_{u}y), and let x=yux=yu^{*}. Then

ψ(xu)\displaystyle\psi(xu) =ψ(y)\displaystyle=\psi(y)
ψ(Aduy)\displaystyle\neq\psi(\operatorname{Ad}_{u}y)
=ψ(uyu)\displaystyle=\psi\left(uyu^{*}\right)
=ψ(ux).\displaystyle=\psi(ux).

Therefore ψ\psi is not tracial.

Now, if ϕ\phi\in\mathfrak{R}^{\natural} is tracial, then ϕAdu=ϕ\phi\circ\operatorname{Ad}_{u}=\phi for all u𝒰(𝔄)u\in\mathcal{U}(\mathfrak{A}). Then ϕ=ϕAdu=(ϕ+Adu)(ϕAdu)\phi=\phi\circ\operatorname{Ad}_{u}=\left(\phi^{+}\circ\operatorname{Ad}_{u}\right)-\left(\phi^{-}\circ\operatorname{Ad}_{u}\right). But ϕ±Adu=ϕ±\left\|\phi^{\pm}\circ\operatorname{Ad}_{u}\right\|=\left\|\phi^{\pm}\right\|, so it follows that ϕ=ϕ+Adu+ϕAdu\left\|\phi\right\|=\left\|\phi^{+}\circ\operatorname{Ad}_{u}\right\|+\left\|\phi^{-}\circ\operatorname{Ad}_{u}\right\|. Therefore ϕ=(ϕ+Adu)(ϕAdu)\phi=\left(\phi^{+}\circ\operatorname{Ad}_{u}\right)-\left(\phi^{-}\circ\operatorname{Ad}_{u}\right) is an orthogonal decomposition of ϕ\phi, and so it is the Jordan decomposition. This means that ϕ±=ϕ±Adu\phi^{\pm}=\phi^{\pm}\circ\operatorname{Ad}_{u}. Since this is true for all u𝒰(𝔄)u\in\mathcal{U}(\mathfrak{A}), it follows that ϕ±\phi^{\pm} are tracial. ∎

Proof of Lemma 1.8.
  1. (i)

    This all follows because 𝒮\mathcal{S} is a weak*-closed real subspace of the unit ball in the continuous dual of the separable Banach space \mathfrak{R}, and the spaces 𝒮G,𝒯,𝒯G\mathcal{S}^{G},\mathcal{T},\mathcal{T}^{G} are all closed subspaces of 𝒮\mathcal{S}.

  2. (ii)

    It is a standard fact that if 𝒯\mathcal{T}\neq\emptyset, then 𝒯\mathcal{T} is a simplex [Blackadar, II.6.8.11]. Let

    CG={cϕ:c0,ϕ𝒯G}C^{G}=\left\{c\phi:c\in\mathbb{R}_{\geq 0},\phi\in\mathcal{T}^{G}\right\}

    be the positive cone of 𝒯G\mathcal{T}^{G}, and let \mathfrak{R}^{\natural} denote the (real) space of all bounded self-adjoint tracial linear functionals on 𝔄\mathfrak{A}. Let EGE^{G} denote the (real) space of all bounded self-adjoint Θ\Theta-invariant linear functionals on 𝔄\mathfrak{A}. We already know that 𝒯\mathcal{T} lives in a hyperplane of \mathfrak{R}^{\natural} defined by the evaluation functional ϕϕ(1)\phi\mapsto\phi(1). It will therefore suffice to show that EG=CGCGE^{G}=C^{G}-C^{G}, and that EGE^{G} is a sub-lattice of \mathfrak{R}^{\natural}.

    Let ϕ+,ϕ0\phi^{+},\phi^{-}\geq 0 be positive functionals on 𝔄\mathfrak{A} such that ϕ=ϕ+ϕ\phi=\phi^{+}-\phi^{-} is tracial, and ϕ+ϕ\phi^{+}\perp\phi^{-}. By Lemma 1.9, we know that ϕ+,ϕ\phi^{+},\phi^{-} are tracial. We claim that if ϕEG\phi\in E^{G}, then ϕ+,ϕCG\phi^{+},\phi^{-}\in C^{G}. To prove this, let gGg\in G, and consider that ϕ+Θg,ϕΘg\phi^{+}\circ\Theta_{g},\phi^{-}\circ\Theta_{g} are both positive linear functionals such that ϕ=(ϕ+Θg)(ϕΘg)\phi=\left(\phi^{+}\circ\Theta_{g}\right)-\left(\phi^{-}\circ\Theta_{g}\right).

    We claim that (ϕ+Θg)(ϕΘg)\left(\phi^{+}\circ\Theta_{g}\right)\perp\left(\phi^{-}\circ\Theta_{g}\right). Fix ε>0\varepsilon>0. We know that there exists z𝔄z\in\mathfrak{A} such that z1,0z\|z\|\leq 1,0\leq z, and such that ϕ+(1z)<ε,ϕ(z)<ε\phi^{+}\left(1-z\right)<\varepsilon,\phi^{-}(z)<\varepsilon. Then Θg1(z)\Theta_{g^{-1}}(z) is a positive element of norm 1\leq 1 such that

    ϕ+(Θg(Θg1(1z)))\displaystyle\phi^{+}\left(\Theta_{g}\left(\Theta_{g^{-1}}(1-z)\right)\right) =ϕ+(1z)\displaystyle=\phi^{+}(1-z) <ε,\displaystyle<\varepsilon,
    ϕ(Θg(Θg1(z)))\displaystyle\phi^{-}\left(\Theta_{g}\left(\Theta_{g^{-1}}(z)\right)\right) =ϕ(z)\displaystyle=\phi^{-}(z) <ε.\displaystyle<\varepsilon.

    Therefore (ϕ+Θg)(ϕΘg)\left(\phi^{+}\circ\Theta_{g}\right)-\left(\phi^{-}\circ\Theta_{g}\right) is a Jordan decomposition of ϕ\phi, and since the Jordan decomposition is unique, it follows that ϕ+=ϕ+Θg,ϕ=ϕΘg\phi^{+}=\phi^{+}\circ\Theta_{g},\phi^{-}=\phi^{-}\circ\Theta_{g}, i.e. that ϕ+,ϕCG\phi^{+},\phi^{-}\in C^{G}. This means that EG=CGCGE^{G}=C^{G}-C^{G}.

    We now want to show that EG=CGCGE^{G}=C^{G}-C^{G} is a sublattice of EE, i.e. that it is closed under the lattice operations. Let ϕ,ψEG\phi,\psi\in E^{G}. For this calculation, we draw on the identities listed in [PositiveOperators, Theorem 1.3]. Then

    ϕψ\displaystyle\phi\lor\psi =(((ϕψ)+ψ)(0+ψ))\displaystyle=\left(\left((\phi-\psi)+\psi\right)\lor(0+\psi)\right)
    =((ϕψ)0)+ψ\displaystyle=\left((\phi-\psi)\lor 0\right)+\psi
    =(ϕψ)++ψ,\displaystyle=(\phi-\psi)^{+}+\psi,
    ϕψ\displaystyle\phi\land\psi =((ϕψ)+ψ)(0+ψ)\displaystyle=\left((\phi-\psi)+\psi\right)\land(0+\psi)
    =((ϕψ)0)+ψ\displaystyle=\left((\phi-\psi)\land 0\right)+\psi
    =(((ϕψ))0)+ψ\displaystyle=-\left((-(\phi-\psi))\lor 0\right)+\psi
    =(ψϕ)++ψ.\displaystyle=-(\psi-\phi)^{+}+\psi.

    Therefore, if EGE^{G} is a real linear space and is closed under the operations ϕϕ+,ϕϕ\phi\mapsto\phi^{+},\phi\mapsto\phi^{-}, then it is also closed under the lattice operations. Thus EGE^{G} is a sublattice of \mathfrak{R}^{\natural}.

    Hence, the subset 𝒯G\mathcal{T}^{G} 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 S1,S2S_{1},S_{2} be convex spaces. We call a map T:S1S2T:S_{1}\to S_{2} an affine map if for every v,wS1;t[0,1]v,w\in S_{1};\;t\in[0,1], we have

T(tv+(1t)w)=tT(v)+(1t)T(w).T(tv+(1-t)w)=tT(v)+(1-t)T(w).

In the case where S2S_{2}\subseteq\mathbb{R}, we call TT an affine functional.

Definition 1.11.

Let KK be a convex subset of a locally convex real topological vector space VV.

  1. (a)

    A point kKk\in K is called an extreme point of KK if for every pair of points k1,k2Kk_{1},k_{2}\in K and parameter t[0,1]t\in[0,1] such that k=tk1+(1t)k2k=tk_{1}+(1-t)k_{2}, either k1=k2k_{1}=k_{2} or t{0,1}t\in\{0,1\}. In other words, we call kk extreme if there is no nontrivial way of expressing kk as a convex combination of elements of KK.

  2. (b)

    The set of all extreme points of KK is denoted eK\partial_{e}K.

  3. (c)

    A subset FF of KK is called a face if for every pair k1,k2K,t(0,1)k_{1},k_{2}\in K,t\in(0,1) such that tk1+(1t)k2Ftk_{1}+(1-t)k_{2}\in F, we have that k1,k2Fk_{1},k_{2}\in F.

  4. (d)

    A face FF of KK is called an exposed face of KK if there exists a continuous affine functional :K\ell:K\to\mathbb{R} such that (x)=0\ell(x)=0 for all xFx\in F, and (y)<0\ell(y)<0 for all yKFy\in K\setminus F.

  5. (e)

    A point kKk\in K is called an exposed point of KK if {k}\{k\} is an exposed face of KK.

  6. (f)

    Given a subset \mathcal{E} of KK, the closed convex hull of \mathcal{E} is written as co¯()\overline{\operatorname{co}}(\mathcal{E}).

We now introduce the basic concepts in our treatment of ergodic optimization.

Definition 1.12.

Let xx\in\mathfrak{R} be a self-adjoint element, and let K𝒮GK\subseteq\mathcal{S}^{G} be a compact convex subset of 𝒮G\mathcal{S}^{G}. Define a value m(x|K)m\left(x|K\right) by

m(x|K):=supψKψ(x).m\left(x|K\right):=\sup_{\psi\in K}\psi(x).

We say a state ϕK\phi\in K is (x|K)(x|K)-maximizing if ϕ(x)=m(x|K)\phi(x)=m(x|K). Let Kmax(x)KK_{\mathrm{max}}(x)\subseteq K denote the set of all (x|K)(x|K)-maximizing states. A state ϕK\phi\in K is called uniquely (x|K)(x|K)-maximizing if Kmax(x)={ϕ}K_{\mathrm{max}}(x)=\{\phi\}.

Remark 1.13.

We note here a trivial inequality: If K1K2K_{1}\subseteq K_{2} are compact convex subsets of 𝒮G\mathcal{S}^{G}, then m(x|K1)m(x|K2)m\left(x|K_{1}\right)\leq m\left(x|K_{2}\right), and in particular, we will always have m(x|K1)m(x|𝒮G)m\left(x|K_{1}\right)\leq m\left(x|\mathcal{S}^{G}\right).

We will single out one type of compact convex subset of 𝒮G\mathcal{S}^{G} which will prove important later. Given a subset A𝔄A\subseteq\mathfrak{A}, set

Ann(A):={ϕ𝒮G:Akerϕ}.\operatorname{Ann}(A):=\left\{\phi\in\mathcal{S}^{G}:A\subseteq\ker\phi\right\}.

When 𝔄\mathfrak{I}\subseteq\mathfrak{A} is a Θ\Theta-invariant closed ideal of 𝔄\mathfrak{A}, we have a bijective correspondence between the states in Ann()\operatorname{Ann}(\mathfrak{I}) and the states on 𝔄/\mathfrak{A}/\mathfrak{I} invariant under the action induced by Θ\Theta. We will be referring to this set again in Sections 2 and 3, when values of the form m(a|Ann(A))m\left(a|\operatorname{Ann}(A)\right) come up in reference to certain ergodic averages. We observe that Ann({0})=𝒮G\operatorname{Ann}(\{0\})=\mathcal{S}^{G}, and that AB𝔄Ann(A)Ann(B)A\subseteq B\subseteq\mathfrak{A}\Rightarrow\operatorname{Ann}(A)\supseteq\operatorname{Ann}(B). There is also no a priori guarantee that Ann(A)\operatorname{Ann}(A)\neq\emptyset, since for example Ann({1})=\operatorname{Ann}(\{1\})=\emptyset. However, Proposition 1.14 gives sufficient conditions for Ann(A)\operatorname{Ann}(A) to be nonempty.

Proposition 1.14.

Let A𝔄A\subseteq\mathfrak{A} be such that ΘgAA\Theta_{g}A\subseteq A for all gGg\in G. Suppose there exists a state on 𝔄\mathfrak{A} which vanishes on AA. Then Ann(A)\operatorname{Ann}(A)\neq\emptyset. In particular, if 𝔄\mathfrak{I}\subsetneq\mathfrak{A} is a proper closed two-sided ideal of 𝔄\mathfrak{A} for which Θg=\Theta_{g}\mathfrak{I}=\mathfrak{I} for all gGg\in G, then Ann()\operatorname{Ann}(\mathfrak{I})\neq\emptyset.

Proof.

Let K𝒮K\subseteq\mathcal{S} denote the family of all (not necessarily invariant) states on 𝔄\mathfrak{A} which vanish on AA. Then if ϕK\phi\in K and aAa\in A, then ΘgaA\Theta_{g}a\in A, so ϕΘg\phi\circ\Theta_{g} vanishes on AA. Therefore ΘgKK\Theta_{g}K\subseteq K for all gGg\in G. It follows from Lemma 1.2 that K𝒮G=Ann(A)K\cap\mathcal{S}^{G}=\operatorname{Ann}(A)\neq\emptyset.

Suppose 𝔄\mathfrak{I}\subsetneq\mathfrak{A} is a proper closed two-sided ideal of 𝔄\mathfrak{A} for which Θg=\Theta_{g}\mathfrak{I}=\mathfrak{I} for all gGg\in G, and let π:𝔄𝔄/\pi:\mathfrak{A}\twoheadrightarrow\mathfrak{A}/\mathfrak{I} be the canonical quotient map. Let Θ~:GAut(𝔄/)\tilde{\Theta}:G\to\operatorname{Aut}(\mathfrak{A}/\mathfrak{I}) be the induced action of GG on 𝔄/\mathfrak{A}/\mathfrak{I} by Θ~g(a+)=Θga+\tilde{\Theta}_{g}(a+\mathfrak{I})=\Theta_{g}a+\mathfrak{I}. Let ψ\psi be a Θ~\tilde{\Theta}-invariant state on 𝔄/\mathfrak{A}/\mathfrak{I}. Then ψπ\psi\circ\pi is a Θ\Theta-invariant state on 𝔄\mathfrak{A} which vanishes on \mathfrak{I}, i.e. ψπAnn()\psi\circ\pi\in\operatorname{Ann}(\mathfrak{I}). ∎

Proposition 1.15.

Let K𝒮GK\subseteq\mathcal{S}^{G} be a nonempty compact convex subset of 𝒮G\mathcal{S}^{G}, and let xx\in\mathfrak{R}. Then Kmax(x)K_{\mathrm{max}}(x) is a nonempty, compact, exposed face of KK.

Proof.

To see that Kmax(x)K_{\mathrm{max}}(x) is nonempty, for each nn\in\mathbb{N}, let ϕnK\phi_{n}\in K such that ϕn(x)m(x|K)1n\phi_{n}(x)\geq m(x|K)-\frac{1}{n}. Then since KK is compact, the sequence (ϕn)n=1(\phi_{n})_{n=1}^{\infty} has a convergent subsequence. Let ϕ\phi be the limit of a convergent subsequence of (ϕn)n=1(\phi_{n})_{n=1}^{\infty}. Then ϕ\phi is (x|K)(x|K)-maximizing.

To see that Kmax(x)K_{\mathrm{max}}(x) is compact, consider that

Kmax(x)={ϕK:ϕ(x)=m(x|K)},K_{\mathrm{max}}(x)=\left\{\phi\in K:\phi(x)=m(x|K)\right\},

which is a closed subset of KK. As for being an exposed face, consider the continuous affine functional :K\ell:K\to\mathbb{R} given by

(ϕ)=ϕ(x)m(x|K).\ell(\phi)=\phi(x)-m(x|K).

Then the functional \ell exposes Kmax(x|K)K_{\mathrm{max}}(x|K), since it is nonpositive on all of KK and vanishes exactly on Kmax(x)K_{\mathrm{max}}(x). ∎

The following result describes the ways in which some ergodic optimizations interact with equivariant *-homomorphisms of C*-dynamical systems.

Theorem 1.16.

Let (𝔄,G,Θ),(𝔄~,G,Θ~)\left(\mathfrak{A},G,\Theta\right),\left(\tilde{\mathfrak{A}},G,\tilde{\Theta}\right) be two C*-dynamical systems, and let π:𝔄𝔄~\pi:\mathfrak{A}\to\tilde{\mathfrak{A}} be a surjective *-homomorphism such that

Θ~gπ\displaystyle\tilde{\Theta}_{g}\circ\pi =πΘg\displaystyle=\pi\circ\Theta_{g} (gG).\displaystyle(\forall g\in G).

Let 𝒮~G\tilde{\mathcal{S}}^{G} denote the space of Θ~\tilde{\Theta}-invariant states on Θ~\tilde{\Theta}. Then m(π(a)|𝒮~G)=m(a|Ann(kerπ))m\left(\pi(a)|\tilde{\mathcal{S}}^{G}\right)=m\left(a|\operatorname{Ann}(\ker\pi)\right).

Proof.

Let 𝒮~G\tilde{\mathcal{S}}^{G} denote the space of Θ~\tilde{\Theta}-invariant states on 𝔄~\tilde{\mathfrak{A}}. We claim that there is a natural bijective correspondence between 𝒮~G\tilde{\mathcal{S}}^{G} and Ann(kerπ)\operatorname{Ann}(\ker\pi). If ϕ\phi is a Θ~\tilde{\Theta}-invariant state on 𝔄~\tilde{\mathfrak{A}}, then we can pull it back to a Θ\Theta-invariant state ϕ0\phi_{0} on 𝔄\mathfrak{A} by

ϕ0=ϕπ.\phi_{0}=\phi\circ\pi.

This ϕ0\phi_{0} obviously vanishes on kerπ\ker\pi, and is Θ\Theta-invariant by virtue of the equivariance property of π\pi. Conversely, if we start with a Θ\Theta-invariant state ψ\psi on 𝔄\mathfrak{A} that vanishes on kerπ\ker\pi, then we can push it to a Θ~\tilde{\Theta}-invariant state ψ~\tilde{\psi} on 𝔄~\tilde{\mathfrak{A}} by

ψ~π=ψ.\tilde{\psi}\circ\pi=\psi.

We claim now that

m(a|Ann(kerπ))=m(π(a)|𝒮~G).m\left(a|\operatorname{Ann}(\ker\pi)\right)=m\left(\pi(a)|\tilde{\mathcal{S}}^{G}\right).

Let ϕ\phi be a (π(a)|𝒮~G)\left(\pi(a)|\tilde{\mathcal{S}}^{G}\right)-maximizing state on 𝔄~\tilde{\mathfrak{A}}. Then ϕπAnn(kerπ)\phi\circ\pi\in\operatorname{Ann}(\ker\pi), so

m(π(a)|𝒮~G)=ϕ(π(a))m(a|Ann(kerπ)).m\left(\pi(a)|\tilde{\mathcal{S}}^{G}\right)=\phi(\pi(a))\leq m\left(a|\operatorname{Ann}(\ker\pi)\right).

On the other hand, if ψAnn(kerπ)\psi\in\operatorname{Ann}(\ker\pi) is (a|Ann(kerπ))\left(a|\operatorname{Ann}(\ker\pi)\right)-maximizing, then let ψ~\tilde{\psi} be such that ψ~π=ψ\tilde{\psi}\circ\pi=\psi. Then ψ~𝒮~G\tilde{\psi}\in\tilde{\mathcal{S}}^{G}, so

m(a|Ann(kerπ))=ψ(a)=ψ~(π(a))m(a|𝒮~G).m\left(a|\operatorname{Ann}(\ker\pi)\right)=\psi(a)=\tilde{\psi}(\pi(a))\leq m\left(a|\tilde{\mathcal{S}}^{G}\right).

The assumption in Theorem 1.16 that π\pi 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 \mathbb{Z} 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 𝒮~G\tilde{\mathcal{S}}^{G} and certain compact convex subsets of Ann(kerπ)\operatorname{Ann}(\ker\pi). For example under the same hypotheses, if 𝒯\mathcal{T}\neq\emptyset, then the proof could be modified in a simple manner to establish that m(π(a)|𝒯~G)=m(a|Ann(kerπ)𝒯G)m\left(\pi(a)|\tilde{\mathcal{T}}^{G}\right)=m\left(a|\operatorname{Ann}(\ker\pi)\cap\mathcal{T}^{G}\right), where 𝒯~G\tilde{\mathcal{T}}^{G} denotes the Θ~\tilde{\Theta}-invariant tracial states on 𝔄~\tilde{\mathfrak{A}}. 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 KK be a compact metrizable simplex. Then every closed face of KK is exposed.

Proof.

See [Davies, Theorem 7.4]. ∎

The theorem we are building to in this section is as follows.

Theorem 1.18.

Let K𝒮GK\subseteq\mathcal{S}^{G} be a compact simplex. Then the closed faces of KK are exactly the sets of the form Kmax(x)K_{\mathrm{max}}(x) for some xx\in\mathfrak{R}.

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 K𝒮GK\subseteq\mathcal{S}^{G} be a compact simplex, and let :K\ell:K\to\mathbb{R} be a continuous affine functional. Then there exists a continuous linear functional ~:span¯(K)\tilde{\ell}:\overline{\operatorname{span}}_{\mathbb{R}}(K)\to\mathbb{R} such that ~|K=\tilde{\ell}|_{K}=\ell.

To prove this theorem, we break it up into several parts, attaining the extension ~\tilde{\ell} as the final step of a few subsequent extensions of \ell.

Lemma 1.20.

Let K𝒮GK\subseteq\mathcal{S}^{G} be a compact metrizable simplex, and let :K\ell:K\to\mathbb{R} be a continuous affine functional. Let P={cϕ:c0,ϕK}P=\left\{c\phi:c\in\mathbb{R}_{\geq 0},\phi\in K\right\}. Then there exists a continuous functional 1:P\ell_{1}:P\to\mathbb{R} satisfying the following conditions for all f1,f2P;c0f_{1},f_{2}\in P;c\in\mathbb{R}_{\geq 0}:

  1. (a)

    1(cf1)=c1(f1)\ell_{1}(cf_{1})=c\ell_{1}(f_{1}),

  2. (b)

    1(f1+f2)=1(f1)+2(f2)\ell_{1}(f_{1}+f_{2})=\ell_{1}(f_{1})+\ell_{2}(f_{2}),

  3. (c)

    1|K=\ell_{1}|_{K}=\ell.

Proof.

Note that every nonzero element of PP can be expressed uniquely as cϕc\phi for some c0{0},ϕKc\in\mathbb{R}_{\geq 0}\setminus\{0\},\phi\in K. As such we define

1(cϕ)={c(ϕ)c>00c=0\ell_{1}(c\phi)=\begin{cases}c\ell(\phi)&c>0\\ 0&c=0\end{cases}

It is immediately clear that this 1\ell_{1} satisfies conditions (a) and (c), leaving only (b) to check.

Now, suppose that f1=c1ϕ1,f2=c2ϕ2f_{1}=c_{1}\phi_{1},f_{2}=c_{2}\phi_{2} for some ϕ1,ϕ2K;c1,c20\phi_{1},\phi_{2}\in K;c_{1},c_{2}\in\mathbb{R}_{\geq 0}. Consider first the case where at least one of c1,c2c_{1},c_{2} are nonzero. Then

f1+f2\displaystyle f_{1}+f_{2} =c1ϕ1+c2ϕ2\displaystyle=c_{1}\phi_{1}+c_{2}\phi_{2}
=(c1+c2)(c1c1+c2ϕ1+c2c1+c2ϕ2)\displaystyle=(c_{1}+c_{2})\left(\frac{c_{1}}{c_{1}+c_{2}}\phi_{1}+\frac{c_{2}}{c_{1}+c_{2}}\phi_{2}\right)
1(f1+f2)\displaystyle\Rightarrow\ell_{1}(f_{1}+f_{2}) =1(c1ϕ1+c2ϕ2)\displaystyle=\ell_{1}\left(c_{1}\phi_{1}+c_{2}\phi_{2}\right)
=(c1+c2)(c1c1+c2ϕ1+c2c1+c2ϕ2)\displaystyle=(c_{1}+c_{2})\ell\left(\frac{c_{1}}{c_{1}+c_{2}}\phi_{1}+\frac{c_{2}}{c_{1}+c_{2}}\phi_{2}\right)
[because \ell is affine] =(c1+c2)(c1c1+c2(ϕ1)+c2c1+c2(ϕ2))\displaystyle=(c_{1}+c_{2})\left(\frac{c_{1}}{c_{1}+c_{2}}\ell(\phi_{1})+\frac{c_{2}}{c_{1}+c_{2}}\ell(\phi_{2})\right)
=c1(ϕ1)+c2(ϕ2)\displaystyle=c_{1}\ell(\phi_{1})+c_{2}\ell(\phi_{2})
=1(c1ϕ1)+1(c2ϕ2)\displaystyle=\ell_{1}(c_{1}\phi_{1})+\ell_{1}(c_{2}\phi_{2})
=1(f1)+1(f2).\displaystyle=\ell_{1}(f_{1})+\ell_{1}(f_{2}).

In the event that c1=c2=0c_{1}=c_{2}=0, then the additivity property attains trivially.

It remains now to show that 1\ell_{1} is continuous. We will check continuity at nonzero points in PP, and then at 0P0\in P. First, consider the case where cϕP{0}c\phi\in P\setminus\{0\}, and c0,ϕKc\in\mathbb{R}_{\geq 0},\phi\in K. Suppose that (cnϕn)n(c_{n}\phi_{n})_{n} is a sequence in PP converging in the weak*-topology to cϕc\phi. We claim that cncc_{n}\to c in \mathbb{R}, and ϕnϕ\phi_{n}\to\phi in the weak*-topology.

We first observe that (cnϕn)(1)=cn(c_{n}\phi_{n})(1)=c_{n}, so (cn)n(c_{n})_{n} converges in 0\mathbb{R}_{\geq 0} to cc, meaning in particular that for sufficiently large nn, we have that cn[c2,3c2]c_{n}\in\left[\frac{c}{2},\frac{3c}{2}\right]. Now, if λ:\lambda:\mathfrak{R}\to\mathbb{R} is a norm-continuous linear functional, then

λ(ϕn)\displaystyle\lambda(\phi_{n}) =1cnλ(cnϕn)\displaystyle=\frac{1}{c_{n}}\lambda(c_{n}\phi_{n})
1cλ(cϕ)\displaystyle\to\frac{1}{c}\lambda(c\phi)
=λ(ϕ).\displaystyle=\lambda(\phi).

Therefore cnc,ϕnϕc_{n}\to c,\phi_{n}\to\phi. Thus we can compute

|1(cϕ)1(cnϕn)|\displaystyle\left|\ell_{1}(c\phi)-\ell_{1}(c_{n}\phi_{n})\right| |1(cϕ)1(cnϕ)|+|1(cnϕ)1(cnϕn)|\displaystyle\leq\left|\ell_{1}(c\phi)-\ell_{1}(c_{n}\phi)\right|+\left|\ell_{1}(c_{n}\phi)-\ell_{1}(c_{n}\phi_{n})\right|
=|ccn||(ϕ)|+|cn||(ϕ)(ϕn)|\displaystyle=|c-c_{n}|\cdot\left|\ell(\phi)\right|+|c_{n}|\cdot\left|\ell(\phi)-\ell(\phi_{n})\right|
|ccn|(supϕK|(ϕ)|)+3c2|(ϕ)(ϕn)|\displaystyle\leq|c-c_{n}|\left(\sup_{\phi\in K}|\ell(\phi)|\right)+\frac{3c}{2}\left|\ell(\phi)-\ell(\phi_{n})\right|
n0,\displaystyle\stackrel{{\scriptstyle n\to\infty}}{{\to}}0,

where supϕK|(ϕ)|\sup_{\phi\in K}|\ell(\phi)| must be finite because KK is weak*-compact, and |(ϕ)(ϕn)|n0|\ell(\phi)-\ell(\phi_{n})|\stackrel{{\scriptstyle n\to\infty}}{{\to}}0 because \ell is weak*-continuous.

Now, suppose that (cnϕn)n=1(c_{n}\phi_{n})_{n=1}^{\infty} converges to 0. Then again we have that cn0c_{n}\to 0 by the same argument used above (i.e. cn=(cnϕn)(1)c_{n}=(c_{n}\phi_{n})(1)). Therefore

|1(cnϕn)|=|cn||(ϕn)||cn|(supϕK|(ϕ)|)n0.\left|\ell_{1}(c_{n}\phi_{n})\right|=|c_{n}|\cdot\left|\ell(\phi_{n})\right|\leq|c_{n}|\left(\sup_{\phi\in K}|\ell(\phi)|\right)\stackrel{{\scriptstyle n\to\infty}}{{\to}}0.

We can thus conclude that 1\ell_{1} is weak*-continuous. ∎

Lemma 1.21.

Let 1,P\ell_{1},P be as in Lemma 1.20, and let V=PPV=P-P. Then there exists a continuous linear functional ~:V\tilde{\ell}:V\to\mathbb{R} such that ~|P=1\tilde{\ell}|_{P}=\ell_{1}.

Proof.

Define ~:V\tilde{\ell}:V\to\mathbb{R} by

~(v)=1(v+)1(v),\tilde{\ell}(v)=\ell_{1}\left(v^{+}\right)-\ell_{1}\left(v^{-}\right),

where v+,vv^{+},v^{-} are meant in the sense of the lattice structure VV possesses by virtue of KK being a simplex.

Our first claim is that if f,gPf,g\in P such that v=fgv=f-g, then ~(v)=1(f)1(g)\tilde{\ell}(v)=\ell_{1}(f)-\ell_{1}(g). To see this, we observe that f+v=g+v+Pf+v^{-}=g+v^{+}\in P. Therefore

1(f+v)\displaystyle\ell_{1}\left(f+v^{-}\right) =1(g+v+)\displaystyle=\ell_{1}\left(g+v^{+}\right)
=1(f)+1(v)\displaystyle=\ell_{1}(f)+\ell_{1}\left(v^{-}\right) =1(g)+1(v+)\displaystyle=\ell_{1}(g)+\ell_{1}\left(v^{+}\right)
1(f)1(g)\displaystyle\Rightarrow\ell_{1}(f)-\ell_{1}(g) =1(v+)1(v)\displaystyle=\ell_{1}\left(v^{+}\right)-\ell_{1}\left(v^{-}\right)
=~(v).\displaystyle=\tilde{\ell}(v).

This makes linearity fairly straightforward to check. First, to confirm additivity, let v,wVv,w\in V. Then v+w=(v++w+)(v+w)v+w=\left(v^{+}+w^{+}\right)-\left(v^{-}+w^{-}\right), where v++w+,v+wPv^{+}+w^{+},v^{-}+w^{-}\in P. Thus

~(v+w)\displaystyle\tilde{\ell}(v+w) =1(v++w+)1(v+w)\displaystyle=\ell_{1}\left(v^{+}+w^{+}\right)-\ell_{1}\left(v^{-}+w^{-}\right)
=1(v+)+1(w+)1(v)1(w)\displaystyle=\ell_{1}\left(v^{+}\right)+\ell_{1}\left(w^{+}\right)-\ell_{1}\left(v^{-}\right)-\ell_{1}\left(w^{-}\right)
=1(v+)1(v)+1(w+)1(w)\displaystyle=\ell_{1}\left(v^{+}\right)-\ell_{1}\left(v^{-}\right)+\ell_{1}\left(w^{+}\right)-\ell_{1}\left(w^{-}\right)
=~(v)+~(w).\displaystyle=\tilde{\ell}(v)+\tilde{\ell}(w).

To check homogeneity, let cc\in\mathbb{R}. If c0c\geq 0, then cv+,cvPcv^{+},cv^{-}\in P, and cv+cv=cvcv^{+}-cv^{-}=cv; on the other hand, if c0c\leq 0, then cv,cv+P-cv^{-},-cv^{+}\in P, and cv=cv+cv+cv=-cv^{-}+cv^{+}. In both cases, homogeneity is straightforward to show. This proves that ~\tilde{\ell} is linear.

It is also quick to show that ~|P=1\tilde{\ell}|_{P}=\ell_{1}, since if vPv\in P, then v=v+v=v^{+}, so ~(v)=1(v+)0=1(v)\tilde{\ell}(v)=\ell_{1}\left(v^{+}\right)-0=\ell_{1}(v).

It remains now to show that ~\tilde{\ell} is continuous. By [RudinFunctional, Theorem 1.18], it will suffice to show that ker~\ker\tilde{\ell} is weak*-closed. To prove the kernel is closed, let (vn)n=1(v_{n})_{n=1}^{\infty} be a sequence in ker~\ker\tilde{\ell} converging in the weak*-topology to vVv\in V. By the Uniform Boundedness Principle, it follows that supnvn<\sup_{n}\|v_{n}\|<\infty. By rescaling, we can assume without loss of generality that vn1\|v_{n}\|\leq 1 for all nn\in\mathbb{N}, and since the unit ball BVB\subseteq V is weak*-closed by Banach-Alaoglu, we can infer that v1\|v\|\leq 1.

Since the unit ball BB is weak*-compact, it follows that the sequences (vn+)n=1,(vn)n=1\left(v_{n}^{+}\right)_{n=1}^{\infty},\left(v_{n}^{-}\right)_{n=1}^{\infty} have convergent subsequences. Let (nj)j=1(n_{j})_{j=1}^{\infty} be a subsequence along which vnj+m1P,vnjm2Pv_{n_{j}}^{+}\to m_{1}\in P,v_{n_{j}}^{-}\to m_{2}\in P. Then if xx\in\mathfrak{R}, then

v(x)\displaystyle v(x) =limnvn(x)\displaystyle=\lim_{n\to\infty}v_{n}(x)
=limn(vn+(x)vn(x))\displaystyle=\lim_{n\to\infty}\left(v_{n}^{+}(x)-v_{n}^{-}(x)\right)
=limj(vnj+(x)vnj(x))\displaystyle=\lim_{j\to\infty}\left(v_{n_{j}}^{+}(x)-v_{n_{j}}^{-}(x)\right)
=(limjvnj+(x))(limjvnj(x))\displaystyle=\left(\lim_{j\to\infty}v_{n_{j}}^{+}(x)\right)-\left(\lim_{j\to\infty}v_{n_{j}}^{-}(x)\right)
=m1(x)m2(x).\displaystyle=m_{1}(x)-m_{2}(x).

Therefore v=m1m2v=m_{1}-m_{2}, so

~(v)\displaystyle\tilde{\ell}(v) =~(m1)~(m2)\displaystyle=\tilde{\ell}(m_{1})-\tilde{\ell}(m_{2})
=(limj~(vnj+))(limj~(vnj))\displaystyle=\left(\lim_{j\to\infty}\tilde{\ell}\left(v_{n_{j}}^{+}\right)\right)-\left(\lim_{j\to\infty}\tilde{\ell}\left(v_{n_{j}}^{-}\right)\right)
=limj(~(vnj+)~(vnj))\displaystyle=\lim_{j\to\infty}\left(\tilde{\ell}\left(v_{n_{j}}^{+}\right)-\tilde{\ell}\left(v_{n_{j}}^{-}\right)\right)
=limj~(vnj)\displaystyle=\lim_{j\to\infty}\tilde{\ell}\left(v_{n_{j}}\right)
=limj0\displaystyle=\lim_{j\to\infty}0
=0.\displaystyle=0.

Therefore, we can conclude that ~\tilde{\ell} is weak*-continuous. ∎

Proof of Theorem 1.19.

This follows from Lemmas 1.20 and 1.21. ∎

Proof of Theorem 1.18.

Let FKF\subseteq K be a closed face of KK. By Lemma 1.17, the face FF is exposed, so let :K\ell:K\to\mathbb{R} be a weak*-continuous affine functional such that

(k)\displaystyle\ell(k) =0\displaystyle=0 (kF),\displaystyle(\forall k\in F),
(k)\displaystyle\ell(k) <0\displaystyle<0 (kKF).\displaystyle(\forall k\in K\setminus F).

Set

V={c1ϕ1c2ϕ2:c1,c20;ϕ1,ϕ2K},V=\left\{c_{1}\phi_{1}-c_{2}\phi_{2}:c_{1},c_{2}\in\mathbb{R}_{\geq 0};\phi_{1},\phi_{2}\in K\right\},

and let ~:V\tilde{\ell}:V\to\mathbb{R} be a continuous linear extension of \ell to VV whose existence is promised by Theorem 1.19. We can then extend ~:V\tilde{\ell}:V\to\mathbb{R} to a weak*-continuous linear functional :\ell^{\prime}:\mathfrak{R}^{\natural}\to\mathbb{R} [PositiveOperators, Theorem 3.6]. There thus exists some xx\in\mathbb{R} such that (ϕ)=ϕ(x)\ell^{\prime}(\phi)=\phi(x) for all ϕ\phi\in\mathfrak{R}^{\natural} [Baggett, Theorem 5.2]. In particular, we have (v)=v(x)\ell^{\prime}(v)=v(x) for all vVv\in V. Therefore F=Kmax(x)F=K_{\mathrm{max}}(x).

The converse is contained in Proposition 1.15. ∎

In particular, we can recover the following corollary.

Corollary 1.22.

If ϕeK\phi\in\partial_{e}K, then there exists xx\in\mathfrak{R} such that ϕ\phi is uniquely (x|K)(x|K)-maximizing, i.e. such that {ϕ}=Kmax(x)\{\phi\}=K_{\mathrm{max}}(x).

Proof.

The singleton {ϕ}\{\phi\} 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 xx-maximizing states simpliciter, but of a state that is maximizing relative to a compact convex subset KK of 𝒮G\mathcal{S}^{G}, especially a compact simplex KK. 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 Θ\Theta-invariant.

Since Theorem 1.18 applies in cases where KK is a simplex, we will conclude this section by describing some situations where 𝒮G\mathcal{S}^{G} is a compact metrizable simplex.

For each ϕ𝒮G\phi\in\mathcal{S}^{G}, let πϕ:𝔄(ϕ)\pi_{\phi}:\mathfrak{A}\to\mathscr{B}(\mathscr{H}_{\phi}) be the GNS representation corresponding to ϕ\phi. Define a unitary representation uϕ:G𝕌(ϕ)u_{\phi}:G\to\mathbb{U}(\mathscr{H}_{\phi}) of GG by

uϕ(g)πϕ(a)=πϕ(Θg1(a)),u_{\phi}(g)\pi_{\phi}(a)=\pi_{\phi}\left(\Theta_{g^{-1}}(a)\right),

extending this from πϕ(𝔄)\pi_{\phi}(\mathfrak{A}) to ϕ\mathscr{H}_{\phi}. Set

Eϕ={vϕ:uϕ(v)=v for all gG}.E_{\phi}=\left\{v\in\mathscr{H}_{\phi}:u_{\phi}(v)=v\textrm{ for all }g\in G\right\}.

Let Pϕ:ϕEϕP_{\phi}:\mathscr{H}_{\phi}\twoheadrightarrow E_{\phi} be the orthogonal projection (in the functional-analytic sense) of ϕ\mathscr{H}_{\phi} onto EϕE_{\phi}. We call the C*-dynamical system (𝔄,G,Θ)(\mathfrak{A},G,\Theta) a GG-abelian system if for every ϕ𝒮G\phi\in\mathcal{S}^{G}, the family of operators {Pϕπϕ(a)Pϕ(ϕ):a𝔄}\left\{P_{\phi}\pi_{\phi}(a)P_{\phi}\in\mathscr{B}(\mathscr{H}_{\phi}):a\in\mathfrak{A}\right\} is mutually commutative.

We record here a handful of germane facts about GG-abelian systems.

Proposition 1.23.

If (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is GG-abelian, then 𝒮G\mathcal{S}^{G} is a simplex.

Proof.

See [Sakai, Theorem 3.1.14]. ∎

Definition 1.24.

We call a system (𝔄,G,Θ)(\mathfrak{A},G,\Theta) asymptotically abelian if there exists a sequence (gn)n=1(g_{n})_{n=1}^{\infty} in GG such that

[Θgna,b]n0\left[\Theta_{g_{n}}a,b\right]\stackrel{{\scriptstyle n\to\infty}}{{\to}}0

for all a,b𝔄a,b\in\mathfrak{A}, where [,]\left[\cdot,\cdot\right] is the Lie bracket [x,y]=xyyx[x,y]=xy-yx on 𝔄\mathfrak{A}.

Proposition 1.25.

If (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is asymptotically abelian, then it is also GG-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 (𝔐,ρ)(\mathfrak{M},\rho) consisting of a von Neumann algebra 𝔐\mathfrak{M} and a faithful tracial normal state ρ\rho on 𝔐\mathfrak{M}. An automorphism of a W*-probability space (𝔐,ρ)(\mathfrak{M},\rho) is a *-automorphism T:𝔐𝔐T:\mathfrak{M}\to\mathfrak{M} such that ρT=ρ\rho\circ T=\rho, i.e. an automorphism of 𝔐\mathfrak{M} which preserves ρ\rho. A W*-dynamical system is a quadruple (𝔐,ρ,G,Ξ)(\mathfrak{M},\rho,G,\Xi), where (𝔐,ρ)(\mathfrak{M},\rho) is a W*-probability space, and Ξ:GAut(𝔐,ρ)\Xi:G\to\operatorname{Aut}(\mathfrak{M},\rho) is a left action of a discrete topological group GG (called the phase group) on 𝔐\mathfrak{M} by ρ\rho-preserving automorphisms of 𝔐\mathfrak{M}, i.e. such that ρ(Ξgx)=ρ(x)\rho(\Xi_{g}x)=\rho(x) for all gG,x𝔐g\in G,x\in\mathfrak{M}. Importantly, if (𝔐,ρ,G,Ξ)(\mathfrak{M},\rho,G,\Xi) is a W*-dynamical system, then (𝔐,G,Ξ)(\mathfrak{M},G,\Xi) 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 GG 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 2(𝔐,ρ)\mathcal{L}^{2}(\mathfrak{M},\rho) to be the Hilbert space defined by completing 𝔐\mathfrak{M} with respect to the inner product x,yρ=ρ(yx)\left<x,y\right>_{\rho}=\rho\left(y^{*}x\right), i.e. the Hilbert space associated with the faithful GNS representation of 𝔐\mathfrak{M} induced by ρ\rho.

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 (𝔐,ρ,G,Ξ)(\mathfrak{M},\rho,G,\Xi) be a W*-dynamical system. A C*-model of (𝔐,ρ,G,Ξ)(\mathfrak{M},\rho,G,\Xi) is a quadruple (𝔄,G,Θ;ι)(\mathfrak{A},G,\Theta;\iota) consisting of a C*-dynamical system (𝔄,G,Θ)(\mathfrak{A},G,\Theta) and a *-homomorphism ι:𝔄𝔐\iota:\mathfrak{A}\to\mathfrak{M} such that

  1. (a)

    ι(𝔄)\iota(\mathfrak{A}) is dense in the weak operator topology of 𝔐\mathfrak{M},

  2. (b)

    Ξg(ι(𝔄))=ι(𝔄)\Xi_{g}\left(\iota(\mathfrak{A})\right)=\iota(\mathfrak{A}) for all gGg\in G, and

  3. (c)

    Ξgι=ιΘg\Xi_{g}\circ\iota=\iota\circ\Theta_{g} for all gGg\in G.

We call the C*-model (𝔄,G,Θ;ι)(\mathfrak{A},G,\Theta;\iota) faithful if ι\iota is also injective.

We remark that we can turn any C*-model into a faithful C*-model through a quotienting process. If ι\iota was not injective, then we could instead consider ι~:𝔄/kerι𝔐\tilde{\iota}:\mathfrak{A}/\ker\iota\hookrightarrow\mathfrak{M}. In the case where 𝔄\mathfrak{A} 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 𝔄=C(X)\mathfrak{A}=C(X), where XX is a compact metrizable topological space, and let 𝔐=L(X,μ)\mathfrak{M}=L^{\infty}(X,\mu) for some Borel probability measure μ\mu. Let ι:C(X)L(X,μ)\iota:C(X)\to L^{\infty}(X,\mu) be the (not necessarily injective) map that maps a continuous function on XX to its equivalence class in L(X,μ)L^{\infty}(X,\mu). It can be seen that fkerιf\in\ker\iota if and only if the open set {xX:f(x)0}\left\{x\in X:f(x)\neq 0\right\} is of measure 0, or equivalently if f|supp(μ)=0f|_{\operatorname{supp}(\mu)}=0, and in particular that ι\iota is injective if and only if μ\mu is strictly positive (i.e. μ\mu assigns positive measure to all nonempty open sets). As such, we can identify C(X)/kerιC(X)/\ker\iota with C(supp(μ))C(\operatorname{supp}(\mu)). Let Y=supp(μ)Y=\operatorname{supp}(\mu) denote the support of μ\mu on XX, and let π:C(X)C(Y)\pi:C(X)\twoheadrightarrow C(Y) be the quotient map (which corresponds to a restriction from XX to YY, i.e. πf=f|Y\pi f=f|_{Y}). Then algebraically, we have a commutative diagram

C(X){C(X)}C(Y){C(Y)}L(X,μ){L^{\infty}(X,\mu)}π\scriptstyle{\pi}ι\scriptstyle{\iota}ι~\scriptstyle{\tilde{\iota}}

So in the commutative case, we can make ι:C(X)L(X,μ)\iota:C(X)\to L^{\infty}(X,\mu) injective by looking at ι~:C(Y)L(Y,μ)L(X,μ)\tilde{\iota}:C(Y)\to L^{\infty}(Y,\mu)\cong L^{\infty}(X,\mu), i.e. by using the support YY to model (Y,μ)(X,μ)(Y,\mu)\cong(X,\mu).

Importantly, so long as 2(𝔐,ρ)\mathcal{L}^{2}(\mathfrak{M},\rho) is separable, any W*-dynamical system (𝔐,ρ,G,Ξ)(\mathfrak{M},\rho,G,\Xi) will admit a faithful separable C*-model. To construct such a C*-model, it suffices to take some separable C*-subalgebra 𝔅𝔐\mathfrak{B}\subseteq\mathfrak{M} which is dense in 𝔐\mathfrak{M} with respect to the weak operator topology, then let 𝔄\mathfrak{A} be the norm-closure of the span of gG(Ξg𝔅)\bigcup_{g\in G}\left(\Xi_{g}\mathfrak{B}\right). We then define Θg=Ξg|𝔄\Theta_{g}=\Xi_{g}|_{\mathfrak{A}} and let ι:𝔄𝔐\iota:\mathfrak{A}\hookrightarrow\mathfrak{M} be the inclusion map.

One last important concept in this section and the next will be unique ergodicity. A C*-dynamical system (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is called uniquely ergodic if 𝒮G\mathcal{S}^{G} 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 (𝔄,,Θ)(\mathfrak{A},\mathbb{Z},\Theta), let a𝔄a\in\mathfrak{A} be a positive element. We define the gauge of aa to be

Γ(a):=limk1kj=0k1Θja.\Gamma(a):=\lim_{k\to\infty}\frac{1}{k}\left\|\sum_{j=0}^{k-1}\Theta_{j}a\right\|.

To prove this limit exists, it suffices to observe that the sequence (j=0k1Θja)k=1\left(\left\|\sum_{j=0}^{k-1}\Theta_{j}a\right\|\right)_{k=1}^{\infty} is subadditive, since

j=0k+1Θja\displaystyle\left\|\sum_{j=0}^{k+\ell-1}\Theta_{j}a\right\| j=0k1Θja+j=kk+1Θja\displaystyle\leq\left\|\sum_{j=0}^{k-1}\Theta_{j}a\right\|+\left\|\sum_{j=k}^{k+\ell-1}\Theta_{j}a\right\|
=j=0k1Θja+Θkj=01Θja\displaystyle=\left\|\sum_{j=0}^{k-1}\Theta_{j}a\right\|+\left\|\Theta_{k}\sum_{j=0}^{\ell-1}\Theta_{j}a\right\|
=j=0k1Θja+j=01Θja.\displaystyle=\left\|\sum_{j=0}^{k-1}\Theta_{j}a\right\|+\left\|\sum_{j=0}^{\ell-1}\Theta_{j}a\right\|.

Therefore, by the Subadditivity Lemma, the sequence (1kj=0k1Θja)k=1\left(\frac{1}{k}\left\|\sum_{j=0}^{k-1}\Theta_{j}a\right\|\right)_{k=1}^{\infty} converges, and we have the equality

limk1kj=0k1Θja=infk1kj=0k1Θja.\lim_{k\to\infty}\frac{1}{k}\left\|\sum_{j=0}^{k-1}\Theta_{j}a\right\|=\inf_{k\in\mathbb{N}}\frac{1}{k}\left\|\sum_{j=0}^{k-1}\Theta_{j}a\right\|.

We have the following characterization of Γ\Gamma in the language of ergodic optimization.

Theorem 2.4.

Let (𝔄,,Θ)(\mathfrak{A},\mathbb{Z},\Theta) be a C*-dynamical system. Then if a𝔄a\in\mathfrak{A} is a positive element, then Γ(a)=m(a|𝒮G)\Gamma(a)=m\left(a|\mathcal{S}^{G}\right).

Proof.

For each kk\in\mathbb{N}, choose a state σk\sigma_{k} on 𝔄\mathfrak{A} such that

σk(1kj=0k1Θja)=1kj=0k1Θja.\sigma_{k}\left(\frac{1}{k}\sum_{j=0}^{k-1}\Theta_{j}a\right)=\left\|\frac{1}{k}\sum_{j=0}^{k-1}\Theta_{j}a\right\|.

Let ωk=1kj=0k1σkΘj\omega_{k}=\frac{1}{k}\sum_{j=0}^{k-1}\sigma_{k}\circ\Theta_{j}, so

ωk(x)\displaystyle\omega_{k}(x) =1kj=0k1σk(Θjx)\displaystyle=\frac{1}{k}\sum_{j=0}^{k-1}\sigma_{k}\left(\Theta_{j}x\right)
=σk(1kj=0k1Θjx),\displaystyle=\sigma_{k}\left(\frac{1}{k}\sum_{j=0}^{k-1}\Theta_{j}x\right),
ωk(a)\displaystyle\omega_{k}(a) =σk(1kj=0k1Θja)\displaystyle=\sigma_{k}\left(\frac{1}{k}\sum_{j=0}^{k-1}\Theta_{j}a\right)
=1kj=0k1Θja.\displaystyle=\left\|\frac{1}{k}\sum_{j=0}^{k-1}\Theta_{j}a\right\|.

Let ω𝒮\omega\in\mathcal{S} be a weak*-limit point of (ωk:k)\left(\omega_{k}:k\in\mathbb{N}\right), and let k1<k2<k_{1}<k_{2}<\cdots be a subsequence such that ωknnω\omega_{k_{n}}\stackrel{{\scriptstyle n\to\infty}}{{\to}}\omega in the weak*-topology. By Lemma 1.2, we know that ω\omega is Θ\Theta-invariant. Therefore ω(a)=Γ(a)\omega(a)=\Gamma(a), and ω\omega is a Θ\Theta-invariant state on 𝔄\mathfrak{A}, so

Γ(a)=ω(a)m(a|𝒮).\Gamma(a)=\omega(a)\leq m\left(a|\mathcal{S}^{\mathbb{Z}}\right).

Now, we prove the opposite inequality. Let ϕ𝒮\phi\in\mathcal{S}^{\mathbb{Z}}. Then

ϕ(a)\displaystyle\phi(a) =ϕ(Avgka)\displaystyle=\phi\left(\operatorname{Avg}_{k}a\right)
Avgka\displaystyle\leq\left\|\operatorname{Avg}_{k}a\right\|
=1kj=0k1Θja\displaystyle=\frac{1}{k}\left\|\sum_{j=0}^{k-1}\Theta_{j}a\right\|
=1kj=0k1Θja\displaystyle=\frac{1}{k}\left\|\sum_{j=0}^{k-1}\Theta_{j}a\right\| (k)\displaystyle(\forall k\in\mathbb{N})
ϕ(a)\displaystyle\Rightarrow\phi(a) infk1kj=0k1Θja\displaystyle\leq\inf_{k\in\mathbb{N}}\frac{1}{k}\left\|\sum_{j=0}^{k-1}\Theta_{j}a\right\|
=Γ(a)\displaystyle=\Gamma(a)
supψ𝒮ψ(a)\displaystyle\Rightarrow\sup_{\psi\in\mathcal{S}^{\mathbb{Z}}}\psi(a) Γ(a).\displaystyle\leq\Gamma(a).

Therefore

m(a|𝒮)=supψ𝒮ψ(a)Γ(a).m\left(a|\mathcal{S}^{\mathbb{Z}}\right)=\sup_{\psi\in\mathcal{S}^{\mathbb{Z}}}\psi(a)\leq\Gamma(a).

This establishes the identity. ∎

Corollary 2.5.

Let (𝔐,ρ,,Ξ)(\mathfrak{M},\rho,\mathbb{Z},\Xi) be a W*-dynamical system, and let (𝔄,,Θ;ι)(\mathfrak{A},\mathbb{Z},\Theta;\iota) be a C*-model of (𝔐,ρ,,Ξ)(\mathfrak{M},\rho,\mathbb{Z},\Xi). If a𝔄a\in\mathfrak{A} is a positive element, then

Γ(ι(a))=m(a|Ann(kerι)).\Gamma(\iota(a))=m\left(a|\operatorname{Ann}(\ker\iota)\right).
Proof.

Write 𝔄~=ι(𝔄)𝔐\tilde{\mathfrak{A}}=\iota(\mathfrak{A})\subseteq\mathfrak{M}, and let Θ~:Aut(𝔄~)\tilde{\Theta}:\mathbb{Z}\to\operatorname{Aut}\left(\tilde{\mathfrak{A}}\right) be the action Θ~n=Ξn|𝔄~\tilde{\Theta}_{n}=\Xi_{n}|_{\tilde{\mathfrak{A}}} obtained by restricting Ξ\Xi to 𝔄~\tilde{\mathfrak{A}}. Write 𝒮~\tilde{\mathcal{S}}^{\mathbb{Z}} for the space of Θ~\tilde{\Theta}-invariant states on 𝔄~\tilde{\mathfrak{A}}.

We can write Γ𝔐(ι(a))=Γ𝔄~(ι(a))\Gamma_{\mathfrak{M}}(\iota(a))=\Gamma_{\tilde{\mathfrak{A}}}(\iota(a)). By Theorem 2.4, we know that Γ𝔄~(ι(a))=m(ι(a)|𝒮~)\Gamma_{\tilde{\mathfrak{A}}}(\iota(a))=m\left(\iota(a)|\tilde{\mathcal{S}}^{\mathbb{Z}}\right), and by Theorem 1.16, we know that m(ι(a)|𝒮~)=m(a|Ann(kerι))m\left(\iota(a)|\tilde{\mathcal{S}}^{\mathbb{Z}}\right)=m\left(a|\operatorname{Ann}(\ker\iota)\right). ∎

Remark 2.6.

Corollary 2.5 can be regarded as an operator-algebraic extension of Lemma 2.3 from [Assani-Young]. The assumption that (𝔄,G,Θ;ι)(\mathfrak{A},G,\Theta;\iota) is faithful can be understood as analogous to the assumption of strict positivity in that paper.

This Γ\Gamma value provides an alternative characterization of unique ergodicity, at least under some additional Choquet-theoretic hypotheses.

Theorem 2.7.

Let (𝔐,ρ,,Ξ)(\mathfrak{M},\rho,\mathbb{Z},\Xi) be a W*-dynamical system, and let (𝔄,,Θ;ι)(\mathfrak{A},\mathbb{Z},\Theta;\iota) be a faithful C*-model of (𝔐,ρ,,Ξ)(\mathfrak{M},\rho,\mathbb{Z},\Xi). Then the following conditions are related by the implications (i)\iff(ii)\Rightarrow(iii).

  1. (i)

    The C*-dynamical system (𝔄,,Θ)(\mathfrak{A},\mathbb{Z},\Theta) is uniquely ergodic.

  2. (ii)

    The C*-dynamical system (𝔄,,Θ)(\mathfrak{A},\mathbb{Z},\Theta) is strictly ergodic.

  3. (iii)

    Γ(ι(a))=ρ(ι(a))\Gamma(\iota(a))=\rho(\iota(a)) for all positive a𝔄a\in\mathfrak{A}.

Further, if 𝒮\mathcal{S}^{\mathbb{Z}} is a simplex, then (iii)\Rightarrow(i).

Proof.

(i)\Rightarrow(ii) Suppose that (𝔄,,Θ)(\mathfrak{A},\mathbb{Z},\Theta) is uniquely ergodic. Then ρι\rho\circ\iota is an invariant state on 𝔄\mathfrak{A}, so it follows that ρι\rho\circ\iota is the unique invariant state on 𝔄\mathfrak{A}. But ρι\rho\circ\iota is also a faithful state on 𝔄\mathfrak{A}, so it follows that (𝔄,,Θ)(\mathfrak{A},\mathbb{Z},\Theta) is strictly ergodic.

(ii)\Rightarrow(i) Trivial.

(i)\Rightarrow(iii) Suppose that (𝔄,,Θ)(\mathfrak{A},\mathbb{Z},\Theta) is uniquely ergodic, and let a𝔄a\in\mathfrak{A} be positive. Let ϕ\phi be a 𝒮\mathcal{S}^{\mathbb{Z}}-maximizing state for aa. Then ϕ=ρι\phi=\rho\circ\iota, since both ϕ\phi and ρι\rho\circ\iota are invariant states on 𝔄\mathfrak{A}, and (𝔄,,Θ)(\mathfrak{A},\mathbb{Z},\Theta) is uniquely ergodic. Thus ϕ=ρι\phi=\rho\circ\iota, so Γ(ι(a))=ϕ(a)=ρ(ι(a))\Gamma(\iota(a))=\phi(a)=\rho(\iota(a)).

(iii)\Rightarrow(i) Suppose that 𝒮\mathcal{S}^{\mathbb{Z}} is a simplex, but that (𝔄,,Θ)(\mathfrak{A},\mathbb{Z},\Theta) is not uniquely ergodic. By the Krein-Milman Theorem, there exists two distinct extreme points of 𝒮\mathcal{S}^{\mathbb{Z}}, and in particular there exists an extreme point ϕ𝒮\phi\in\mathcal{S}^{\mathbb{Z}} of 𝒮\mathcal{S}^{\mathbb{Z}} distinct from ρι\rho\circ\iota. Then by Corollary 1.22, there exists a𝔄a\in\mathfrak{A} self-adjoint such that {ϕ}=𝒮max(a)\{\phi\}=\mathcal{S}_{\mathrm{max}}^{\mathbb{Z}}(a). We can assume that aa is positive, since otherwise we could replace aa with a+ra+r for a sufficiently large positive real number r>0r>0, and 𝒮max(a)=𝒮max(a+r)\mathcal{S}_{\mathrm{max}}^{\mathbb{Z}}(a)=\mathcal{S}_{\mathrm{max}}^{\mathbb{Z}}(a+r). Then Γ(ι(a))=ϕ(a)\Gamma(\iota(a))=\phi(a). But by the assumption that ϕ\phi is uniquely (a|𝒮)\left(a|\mathcal{S}^{\mathbb{Z}}\right)-maximizing, it follows that ρ(ι(a))<ϕ(a)\rho(\iota(a))<\phi(a). Therefore Γ(ι(a))ρ(ι(a))\Gamma(\iota(a))\neq\rho(\iota(a)), meaning that (iii) does not attain. Thus ¬\neg(i)¬\Rightarrow\neg(iii). ∎

3. Unique ergodicity and gauge: the amenable setting

For the duration of this section, we assume that (𝔐,ρ,G,Ξ)(\mathfrak{M},\rho,G,\Xi) is a W*-dynamical system with 2(𝔐,ρ)\mathcal{L}^{2}(\mathfrak{M},\rho) separable. Assume further that (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is a C*-dynamical system such that 𝔄\mathfrak{A} is separable, and that GG is amenable. It follows from Corollary 1.4 that 𝒮G\mathcal{S}^{G}\neq\emptyset.

In this section, we expand upon some of the ideas presented in Section 2, generalizing from the case of actions of \mathbb{Z} to actions of a countable discrete amenable group GG. 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 GG is amenable.

We define the weak topology on a C*-algebra 𝔄\mathfrak{A} to be the topology generated by the states on 𝔄\mathfrak{A}, i.e.

x\displaystyle x ψ(x)\displaystyle\mapsto\psi(x) (ψ𝒮).\displaystyle(\psi\in\mathcal{S}).

In other words, the weak topology is the topology in which a net (xi)i(x_{i})_{i\in\mathscr{I}} converges to xx if and only if (ψ(xi))i(\psi(x_{i}))_{i\in\mathscr{I}} converges to ψ(x)\psi(x) for every state ψ\psi on 𝔄\mathfrak{A}. We say the net (xi)i(x_{i})_{i\in\mathscr{I}} converges weakly to xx if it converges in the weak topology.

Theorem 3.1.

Let (𝔄,G,Θ)(\mathfrak{A},G,\Theta) be a C*-dynamical system. Then the following conditions are equivalent.

  1. (i)

    (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is uniquely ergodic.

  2. (ii)

    There exists a right Følner sequences (Fk)k=1(F_{k})_{k=1}^{\infty} for GG and a linear functional ϕ:𝔄\phi:\mathfrak{A}\to\mathbb{C} such that for all x𝔄x\in\mathfrak{A}, the sequence (AvgFkx)k=1\left(\operatorname{Avg}_{F_{k}}x\right)_{k=1}^{\infty} converges in norm to ϕ(x)11\phi(x)1\in\mathbb{C}1.

  3. (iii)

    There exists a left Følner sequences (Fk)k=1(F_{k})_{k=1}^{\infty} for GG and a linear functional ϕ:𝔄\phi:\mathfrak{A}\to\mathbb{C} such that for all x𝔄x\in\mathfrak{A}, the sequence (AvgFkx)k=1\left(\operatorname{Avg}_{F_{k}}x\right)_{k=1}^{\infty} converges weakly to ϕ(x)11\phi(x)1\in\mathbb{C}1.

  4. (iv)

    There exists a state ϕ\phi on 𝔄\mathfrak{A} such that for every right Følner sequence (Fk)k=1(F_{k})_{k=1}^{\infty} for GG, the sequence (AvgFkx)k=1\left(\operatorname{Avg}_{F_{k}}x\right)_{k=1}^{\infty} converges in norm to ϕ(x)11\phi(x)1\in\mathbb{C}1.

  5. (v)

    There exists a state ϕ\phi on 𝔄\mathfrak{A} such that for every left Følner sequence (Fk)k=1(F_{k})_{k=1}^{\infty} for GG, the sequence (AvgFkx)k=1\left(\operatorname{Avg}_{F_{k}}x\right)_{k=1}^{\infty} converges weakly to ϕ(x)11\phi(x)1\in\mathbb{C}1.

Proof.

Assume throughout that any x𝔄x\in\mathfrak{A} is nonzero.

(ii)\Rightarrow(iii) Follows from the existence of two-sided Følner sequence.

(iv)\Rightarrow(v) Follows from the existence of two-sided Følner sequence.

(iv)\Rightarrow(ii) Trivial.

(v)\Rightarrow(iii) Trivial.

(iii)\Rightarrow(i) Suppose that AvgFkxϕ(x)11\operatorname{Avg}_{F_{k}}x\to\phi(x)1\in\mathbb{C}1 weakly for all x𝔄x\in\mathfrak{A}. We claim that ϕ\phi is the unique invariant state of (𝔄,G,Θ)(\mathfrak{A},G,\Theta). First, we demonstrate that ϕ\phi is Θ\Theta-invariant. Fix g0Gg_{0}\in G, and fix ε>0\varepsilon>0. Choose K1,K2,K3K_{1},K_{2},K_{3}\in\mathbb{N} such that

k\displaystyle k K1\displaystyle\geq K_{1} |ϕ(ϕ(x)1)ϕ(AvgFkx)|\displaystyle\Rightarrow\left|\phi(\phi(x)1)-\phi\left(\operatorname{Avg}_{F_{k}}x\right)\right| <ε3,\displaystyle<\frac{\varepsilon}{3},
k\displaystyle k K2\displaystyle\geq K_{2} |ϕ(Θg0ϕ(x)1)ϕ(Θg0AvgFkx)|\displaystyle\Rightarrow\left|\phi(\Theta_{g_{0}}\phi(x)1)-\phi(\Theta_{g_{0}}\operatorname{Avg}_{F_{k}}x)\right| <ε3,\displaystyle<\frac{\varepsilon}{3},
k\displaystyle k K3\displaystyle\geq K_{3} |g0FkΔFk||Fk|\displaystyle\Rightarrow\frac{|g_{0}F_{k}\Delta F_{k}|}{|F_{k}|} <ε3x.\displaystyle<\frac{\varepsilon}{3\|x\|}.

The K1,K2K_{1},K_{2} exist because we know that in the weak topology, the functionals ϕ,ϕΘg0\phi,\phi\circ\Theta_{g_{0}} are both continuous, and K3K_{3} exists by the amenability of GG. Let K=max{K1,K2,K3}K=\max\{K_{1},K_{2},K_{3}\}. Then if kKk\geq K, then

|ϕ(Θg0x)ϕ(x)|\displaystyle\left|\phi(\Theta_{g_{0}}x)-\phi(x)\right|\leq |ϕ(Θg0x)ϕ(Θg0AvgFkx)|\displaystyle\left|\phi(\Theta_{g_{0}}x)-\phi(\Theta_{g_{0}}\operatorname{Avg}_{F_{k}}x)\right|
+|ϕ(Θg0AvgFkx)ϕ(AvgFkx)|+|ϕ(AvgFkx)ϕ(x)|\displaystyle+\left|\phi(\Theta_{g_{0}}\operatorname{Avg}_{F_{k}}x)-\phi(\operatorname{Avg}_{F_{k}}x)\right|+\left|\phi(\operatorname{Avg}_{F_{k}}x)-\phi(x)\right|
\displaystyle\leq ε3+|ϕ(Θg0AvgFkx)ϕ(AvgFkx)|+ε3\displaystyle\frac{\varepsilon}{3}+\left|\phi(\Theta_{g_{0}}\operatorname{Avg}_{F_{k}}x)-\phi(\operatorname{Avg}_{F_{k}}x)\right|+\frac{\varepsilon}{3}
=\displaystyle= 2ε3+|ϕ(Θg0AvgFkx)ϕ(AvgFkx)|\displaystyle\frac{2\varepsilon}{3}+\left|\phi(\Theta_{g_{0}}\operatorname{Avg}_{F_{k}}x)-\phi(\operatorname{Avg}_{F_{k}}x)\right|
=\displaystyle= 2ε3+|ϕ(1|Fk|(gFkΘg0gx)(1|Fk|gFkΘgx))|\displaystyle\frac{2\varepsilon}{3}+\left|\phi\left(\frac{1}{|F_{k}|}\left(\sum_{g\in F_{k}}\Theta_{g_{0}g}x\right)-\left(\frac{1}{|F_{k}|}\sum_{g\in F_{k}}\Theta_{g}x\right)\right)\right|
=\displaystyle= 2ε3+|ϕ(1|Fk|(gg0FkΘgx)(1|Fk|gFkΘgx))|\displaystyle\frac{2\varepsilon}{3}+\left|\phi\left(\frac{1}{|F_{k}|}\left(\sum_{g\in g_{0}F_{k}}\Theta_{g}x\right)-\left(\frac{1}{|F_{k}|}\sum_{g\in F_{k}}\Theta_{g}x\right)\right)\right|
=\displaystyle= 2ε3+|ϕ(1|Fk|(gg0FkFkΘgx)(1|Fk|gFkg0FkΘgx))|\displaystyle\frac{2\varepsilon}{3}+\left|\phi\left(\frac{1}{|F_{k}|}\left(\sum_{g\in g_{0}F_{k}\setminus F_{k}}\Theta_{g}x\right)-\left(\frac{1}{|F_{k}|}\sum_{g\in F_{k}\setminus g_{0}F_{k}}\Theta_{g}x\right)\right)\right|
\displaystyle\leq 2ε3+|ϕ(1|Fk|gg0FkFkΘgx)|+|ϕ(1|Fk|gFkg0FkΘgx)|\displaystyle\frac{2\varepsilon}{3}+\left|\phi\left(\frac{1}{|F_{k}|}\sum_{g\in g_{0}F_{k}\setminus F_{k}}\Theta_{g}x\right)\right|+\left|\phi\left(\frac{1}{|F_{k}|}\sum_{g\in F_{k}\setminus g_{0}F_{k}}\Theta_{g}x\right)\right|
<\displaystyle< 2ε3+|g0FkΔFk||Fk|x\displaystyle\frac{2\varepsilon}{3}+\frac{|g_{0}F_{k}\Delta F_{k}|}{|F_{k}|}\|x\|
=\displaystyle= ε.\displaystyle\varepsilon.

Therefore ϕ\phi is Θ\Theta-invariant. To see that it is positive, it suffices to observe that x0AvgFkx0x\geq 0\Rightarrow\operatorname{Avg}_{F_{k}}x\geq 0, meaning that ϕ(x)=limkϕ(AvgFkx)0\phi(x)=\lim_{k\to\infty}\phi(\operatorname{Avg}_{F_{k}}x)\geq 0. To see that ϕ(1)=1\phi(1)=1, we just observe that AvgFk1=1\operatorname{Avg}_{F_{k}}1=1 for all kk\in\mathbb{N}.

Now we show that ϕ\phi is the unique Θ\Theta-invariant state. Let ψ\psi be any invariant state. Then

ψ(x)\displaystyle\psi(x) =ψ(AvgFkx)\displaystyle=\psi(\operatorname{Avg}_{F_{k}}x)
kψ(ϕ(x)1)\displaystyle\stackrel{{\scriptstyle k\to\infty}}{{\to}}\psi(\phi(x)1)
=ϕ(x)ψ(1)\displaystyle=\phi(x)\psi(1)
=ϕ(x).\displaystyle=\phi(x).

Therefore ψ=ϕ\psi=\phi, and so (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is uniquely ergodic.

(i)\Rightarrow(iv) Fix a right Følner sequence (Fk)k=1(F_{k})_{k=1}^{\infty}, and assume for contradiction that (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is uniquely ergodic with Θ\Theta-invariant state ϕ\phi, but that there exists x𝔄x\in\mathfrak{A} such that (AvgFkx)k=1\left(\operatorname{Avg}_{F_{k}}x\right)_{k=1}^{\infty} does not converge in norm to a scalar, and in particular does not converge in norm to ϕ(x)1\phi(x)1. Since we can decompose xx into its real and imaginary parts, we can assume that x𝔄sax\in\mathfrak{A}_{\mathrm{sa}}. Fix ε0>0\varepsilon_{0}>0 for which there exists an infinite sequence k1<k2<k_{1}<k_{2}<\cdots such that AvgFknxϕ(x)1ε0\left\|\operatorname{Avg}_{F_{k_{n}}}x-\phi(x)1\right\|\geq\varepsilon_{0}. Then for each nn\in\mathbb{N} exists a state ψn\psi_{n} on 𝔄\mathfrak{A} such that |ψn(AvgFknxϕ(x)1)|=AvgFknxϕ(x)1\left|\psi_{n}\left(\operatorname{Avg}_{F_{k_{n}}}x-\phi(x)1\right)\right|=\left\|\operatorname{Avg}_{F_{k_{n}}}x-\phi(x)1\right\|.

Set

ωn=ψnAvgFkn,\omega_{n}=\psi_{n}\circ\operatorname{Avg}_{F_{k_{n}}},

so ωn(xϕ(x)1)=ψn(AvgFknxϕ(x)1)\omega_{n}(x-\phi(x)1)=\psi_{n}\left(\operatorname{Avg}_{F_{k_{n}}}x-\phi(x)1\right). Then (ωn)n=1(\omega_{n})_{n=1}^{\infty} has a subsequence, call it (ωnj)j=1(\omega_{n_{j}})_{j=1}^{\infty} which converges in the weak*-topology to some ω\omega. This ω\omega is also a state on 𝔄\mathfrak{A}, and by Lemma 1.2, we know ω\omega is Θ\Theta-invariant. But ωϕ\omega\neq\phi, since

|ω(x)ϕ(x)|\displaystyle\left|\omega(x)-\phi(x)\right| =limj|ωnj(x)ϕ(x)|\displaystyle=\lim_{j\to\infty}\left|\omega_{n_{j}}(x)-\phi(x)\right|
=limj|ωnj(xϕ(x)1)|\displaystyle=\lim_{j\to\infty}\left|\omega_{n_{j}}(x-\phi(x)1)\right|
=limj|ψnj(AvgFknjxϕ(x)1)|\displaystyle=\lim_{j\to\infty}\left|\psi_{n_{j}}(\operatorname{Avg}_{F_{k_{n_{j}}}}x-\phi(x)1)\right|
=limjAvgFknjxϕ(x)1\displaystyle=\lim_{j\to\infty}\left\|\operatorname{Avg}_{F_{k_{n_{j}}}}x-\phi(x)1\right\|
ε0.\displaystyle\geq\varepsilon_{0}.

This contradicts (𝔄,G,Θ)(\mathfrak{A},G,\Theta) 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 𝐅=(Fk)k=1\mathbf{F}=(F_{k})_{k=1}^{\infty} be a right Følner sequence.

  1. (a)

    Let (𝔄,G,Θ)(\mathfrak{A},G,\Theta) be a C*-dynamical system, and let 𝐅=(Fk)k=1\mathbf{F}=(F_{k})_{k=1}^{\infty} be a right Følner sequence for GG. Then if a𝔄a\in\mathfrak{A} is a positive element, then the sequence (1|Fk|gFkΘga)k=1\left(\left\|\frac{1}{|F_{k}|}\sum_{g\in F_{k}}\Theta_{g}a\right\|\right)_{k=1}^{\infty} converges to m(a|𝒮G)m\left(a|\mathcal{S}^{G}\right).

  2. (b)

    Let (𝔄,G,Θ;ι)(\mathfrak{A},G,\Theta;\iota) be a faithful C*-model of (𝔐,ρ,G,Ξ)(\mathfrak{M},\rho,G,\Xi). Then the following conditions are related by the implications (i)\iff(ii)\Rightarrow(iii).

    1. (i)

      The C*-dynamical system (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is uniquely ergodic.

    2. (ii)

      The C*-dynamical system (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is strictly ergodic.

    3. (iii)

      Γ(ι(a))=ρ(ι(a))\Gamma(\iota(a))=\rho(\iota(a)) for all positive a𝔄a\in\mathfrak{A}.

    Further, if 𝒮G\mathcal{S}^{G} is a simplex, then (iii)\Rightarrow(i).

Theorem 3.4.

Let (𝔄,G,Θ)(\mathfrak{A},G,\Theta) be a C*-dynamical system, and let 𝐅=(Fk)k=1\mathbf{F}=(F_{k})_{k=1}^{\infty} be a right Følner sequence for GG. Then if a𝔄a\in\mathfrak{A} is a positive element, then the sequence (1|Fk|gFkΘga)k=1\left(\left\|\frac{1}{|F_{k}|}\sum_{g\in F_{k}}\Theta_{g}a\right\|\right)_{k=1}^{\infty} converges to m(a|𝒮G)m\left(a|\mathcal{S}^{G}\right).

Proof.

For each kk\in\mathbb{N}, choose a state σk\sigma_{k} on 𝔄\mathfrak{A} such that

σk(1|Fk|gFkΘga)=1|Fk|gFkΘga.\sigma_{k}\left(\frac{1}{|F_{k}|}\sum_{g\in F_{k}}\Theta_{g}a\right)=\left\|\frac{1}{|F_{k}|}\sum_{g\in F_{k}}\Theta_{g}a\right\|.

Let ωk=1|Fk|gFkσkΘg\omega_{k}=\frac{1}{|F_{k}|}\sum_{g\in F_{k}}\sigma_{k}\circ\Theta_{g}, so

ωk(x)\displaystyle\omega_{k}(x) =1|Fk|gFkσk(Θgx)\displaystyle=\frac{1}{|F_{k}|}\sum_{g\in F_{k}}\sigma_{k}(\Theta_{g}x)
=σk(1|Fk|gFkΘgx),\displaystyle=\sigma_{k}\left(\frac{1}{|F_{k}|}\sum_{g\in F_{k}}\Theta_{g}x\right),
ωk(a)\displaystyle\omega_{k}(a) =σk(1|Fk|gFkΘga)\displaystyle=\sigma_{k}\left(\frac{1}{|F_{k}|}\sum_{g\in F_{k}}\Theta_{g}a\right)
=1|Fk|gFkΘga.\displaystyle=\left\|\frac{1}{|F_{k}|}\sum_{g\in F_{k}}\Theta_{g}a\right\|.

This means that in order to show that (1|Fk|gGΘga)k=1\left(\left\|\frac{1}{|F_{k}|}\sum_{g\in G}\Theta_{g}a\right\|\right)_{k=1}^{\infty} converges to m(a|𝒮G)m\left(a|\mathcal{S}^{G}\right), it suffices to show that ωk(a)km(a|𝒮G)\omega_{k}(a)\stackrel{{\scriptstyle k\to\infty}}{{\to}}m\left(a|\mathcal{S}^{G}\right). So for the remainder of this proof, we are going to be looking instead at the sequence (ωk)k=1(\omega_{k})_{k=1}^{\infty}.

Let k1<k2<k_{1}<k_{2}<\cdots be some sequence such that (ωkn)n=1\left(\omega_{k_{n}}\right)_{n=1}^{\infty} converges in the weak*-topology to some ω\omega. It follows from Lemma 1.2 that ω\omega is Θ\Theta-invariant. To see that (ωk(a)k=1=(1|Fk|gGΘgι(a))k=1(\omega_{k}(a)_{k=1}^{\infty}=\left(\left\|\frac{1}{|F_{k}|}\sum_{g\in G}\Theta_{g}\iota(a)\right\|\right)_{k=1}^{\infty} converges to m(a|𝒮G)m\left(a|\mathcal{S}^{G}\right), it will suffice to show that every limit point ω\omega of (ωk:k)\left(\omega_{k}:k\in\mathbb{N}\right) satisfies

ω𝒮maxG(a).\omega\in\mathcal{S}_{\mathrm{max}}^{G}(a).

This follows because if there existed a subsequence k1<k2<k_{1}<k_{2}<\cdots of (ωk)k=1(\omega_{k})_{k=1}^{\infty} such that ωkn(a)nzm(a|𝒮G)\omega_{k_{n}}(a)\stackrel{{\scriptstyle n\to\infty}}{{\to}}z\neq m\left(a|\mathcal{S}^{G}\right), then by compactness, that subsequence (ωkn:n)\left(\omega_{k_{n}}:n\in\mathbb{N}\right) would have some subsequence converging to some ω\omega^{\prime} for which ω(a)=zm(a|𝒮G)\omega^{\prime}(a)=z\neq m\left(a|\mathcal{S}^{G}\right), meaning in particular that ω𝒮maxG(a)\omega^{\prime}\not\in\mathcal{S}_{\mathrm{max}}^{G}(a).

So let k1<k2<k_{1}<k_{2}<\cdots be some sequence such that (ωkn)n=1\left(\omega_{k_{n}}\right)_{n=1}^{\infty} converges in the weak*-topology to some ω\omega. As has already been remarked, we have that ω𝒮G\omega\in\mathcal{S}^{G}, so ω(a)m(a|𝒮G)\omega(a)\leq m\left(a|\mathcal{S}^{G}\right). We prove the opposite inequality. Let ϕ𝒮G\phi\in\mathcal{S}^{G}. Then

ϕ(a)\displaystyle\phi(a) =ϕ(1|Fkn|gFknΘga)\displaystyle=\phi\left(\frac{1}{\left|F_{k_{n}}\right|}\sum_{g\in F_{k_{n}}}\Theta_{g}a\right) (ϕ is Θ-invariant)\displaystyle\left(\textrm{$\phi$ is $\Theta$-invariant}\right)
1|Fkn|gFknΘga\displaystyle\leq\left\|\frac{1}{\left|F_{k_{n}}\right|}\sum_{g\in F_{k_{n}}}\Theta_{g}a\right\|
=ωkn(a)\displaystyle=\omega_{k_{n}}(a)
ϕ(a)\displaystyle\Rightarrow\phi(a) limnωkn(a)\displaystyle\leq\lim_{n\to\infty}\omega_{k_{n}}(a)
=ω(a).\displaystyle=\omega(a).

Therefore ω(a)supψ𝒮Gψ(a)=m(a|𝒮G)\omega(a)\geq\sup_{\psi\in\mathcal{S}^{G}}\psi(a)=m\left(a|\mathcal{S}^{G}\right). 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 (𝔄,G,Θ)(\mathfrak{A},G,\Theta) be a C*-dynamical system, and let 𝐅=(Fk)k=1\mathbf{F}=(F_{k})_{k=1}^{\infty} be a right Følner sequence for GG. Let ϕ𝒮G\phi\in\mathcal{S}^{G}. Then (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is uniquely ergodic if and only if for every positive element a𝔄a\in\mathfrak{A}, the sequence (1|Fk|gFkΘga)k=1\left(\left\|\frac{1}{|F_{k}|}\sum_{g\in F_{k}}\Theta_{g}a\right\|\right)_{k=1}^{\infty} converges to ϕ(a)\phi(a).

Proof.

(\Rightarrow) Suppose (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is uniquely ergodic. Then ϕ(a)=m(a|𝒮G)\phi(a)=m\left(a|\mathcal{S}^{G}\right) for all positive a𝔄a\in\mathfrak{A}, so by Theorem 3.4 1|Fk|gFkΘgakϕ(a)\left\|\frac{1}{|F_{k}|}\sum_{g\in F_{k}}\Theta_{g}a\right\|\stackrel{{\scriptstyle k\to\infty}}{{\to}}\phi(a).

(\Leftarrow) We’ll prove the contrapositive. Suppose (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is not uniquely ergodic. Then there exists an extreme point ψ\psi of 𝒮G\mathcal{S}^{G} different from ϕ\phi. By Corollary 1.22, there exists a𝔄a\in\mathfrak{A} self-adjoint such that {ϕ}=𝒮maxG(a)\{\phi\}=\mathcal{S}_{\mathrm{max}}^{G}(a). We can assume that aa is positive, replacing aa by a+ra+r for a sufficiently large positive real number r>0r>0 otherwise. Thus limk1|Fk|gFkΘga=ψ(a)>ϕ(a)\lim_{k\to\infty}\left\|\frac{1}{|F_{k}|}\sum_{g\in F_{k}}\Theta_{g}a\right\|=\psi(a)>\phi(a). ∎

Definition 3.7.

Given a C*-dynamical system (𝔄,G,Θ)(\mathfrak{A},G,\Theta), a positive element a𝔄a\in\mathfrak{A}, and a right Følner sequence 𝐅=(Fk)k=1\mathbf{F}=(F_{k})_{k=1}^{\infty} for GG, we define the gauge of aa to be the limit

Γ(x):=limk1|Fk|gFkΘgx.\Gamma(x):=\lim_{k\to\infty}\left\|\frac{1}{|F_{k}|}\sum_{g\in F_{k}}\Theta_{g}x\right\|.

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 (𝔄,G,Θ)(\mathfrak{A},G,\Theta), and independent of the right Følner sequence 𝐅=(Fk)k=1\mathbf{F}=(F_{k})_{k=1}^{\infty}. As such, even though the gauge as we have described it is computed using a right Følner sequence 𝐅=(Fk)k=1\mathbf{F}=(F_{k})_{k=1}^{\infty}, we do not need to include 𝐅\mathbf{F} in our notation for Γ\Gamma.

Corollary 3.8.

Let (𝔄,G,Θ),(𝔄~,G,Θ~)\left(\mathfrak{A},G,\Theta\right),\left(\tilde{\mathfrak{A}},G,\tilde{\Theta}\right) be two C*-dynamical systems, and let π:𝔄𝔄~\pi:\mathfrak{A}\to\tilde{\mathfrak{A}} be a *-homomorphism (not necessarily surjective) such that

Θ~gπ\displaystyle\tilde{\Theta}_{g}\circ\pi =πΘg\displaystyle=\pi\circ\Theta_{g} (gG).\displaystyle(\forall g\in G).

Let 𝒮~G\tilde{\mathcal{S}}^{G} denote the space of Θ~\tilde{\Theta}-invariant states on Θ~\tilde{\Theta}. Then m(π(a)|𝒮~G)=m(a|Ann(kerπ))m\left(\pi(a)|\tilde{\mathcal{S}}^{G}\right)=m\left(a|\operatorname{Ann}(\ker\pi)\right).

Proof.

Let 𝔅=π(𝔄)\mathfrak{B}=\pi(\mathfrak{A}), and let H:GAut(𝔅)H:G\to\operatorname{Aut}(\mathfrak{B}) be the action Hg=Θ~g|𝔅H_{g}=\tilde{\Theta}_{g}|_{\mathfrak{B}}. Let KK denote the space of all HH-invariant states on 𝔅\mathfrak{B}. Then

m(π(a)|𝒮~G)\displaystyle m\left(\pi(a)|\tilde{\mathcal{S}}^{G}\right) =Γ𝔄~(π(a))\displaystyle=\Gamma_{\tilde{\mathfrak{A}}}(\pi(a)) (Theorem 3.4)\displaystyle\left(\textrm{Theorem \ref{Gauge exists for C*-dynamical systems}}\right)
=Γ𝔅(π(a))\displaystyle=\Gamma_{\mathfrak{B}}(\pi(a))
=m(π(a)|K)\displaystyle=m\left(\pi(a)|K\right) (Theorem 3.4)\displaystyle\left(\textrm{Theorem \ref{Gauge exists for C*-dynamical systems}}\right)
=m(a|Ann(kerπ))\displaystyle=m\left(a|\operatorname{Ann}(\ker\pi)\right) (Theorem 1.16).\displaystyle\left(\textrm{Theorem \ref{Ergodic optimization through *-homomorphisms}}\right).

Corollary 3.9.

Let (𝔐,ρ,G,Ξ)(\mathfrak{M},\rho,G,\Xi) be a W*-dynamical system, and let (𝔄,G,Θ;ι)(\mathfrak{A},G,\Theta;\iota) be a C*-model of (𝔐,ρ,,Ξ)(\mathfrak{M},\rho,\mathbb{Z},\Xi). Then if a𝔄a\in\mathfrak{A} is a positive element, then

Γ(ι(a))=m(a|Ann(kerι)).\Gamma(\iota(a))=m\left(a|\operatorname{Ann}(\ker\iota)\right).
Proof.

Write 𝔄~=ι(𝔄)𝔐\tilde{\mathfrak{A}}=\iota(\mathfrak{A})\subseteq\mathfrak{M}, and let Θ~:GAut(𝔄~)\tilde{\Theta}:G\to\operatorname{Aut}\left(\tilde{\mathfrak{A}}\right) be the action Θ~g=Ξg|𝔄~\tilde{\Theta}_{g}=\Xi_{g}|_{\tilde{\mathfrak{A}}} obtained by restricting Ξ\Xi to 𝔄~\tilde{\mathfrak{A}}. Write 𝒮~G\tilde{\mathcal{S}}^{G} for the space of Θ~\tilde{\Theta}-invariant states on 𝔄~\tilde{\mathfrak{A}}.

We know Γ𝔐(ι(a))=Γ𝔄~(ι(a))\Gamma_{\mathfrak{M}}(\iota(a))=\Gamma_{\tilde{\mathfrak{A}}}(\iota(a)). By Theorem 3.4, we know that Γ𝔄~(ι(a))=m(ι(a)|𝒮~G)\Gamma_{\tilde{\mathfrak{A}}}(\iota(a))=m\left(\iota(a)|\tilde{\mathcal{S}}^{G}\right), and by Theorem 1.16, we know that

m(ι(a)|𝒮~G)=m(a|Ann(kerι)).m\left(\iota(a)|\tilde{\mathcal{S}}^{G}\right)=m\left(a|\operatorname{Ann}(\ker\iota)\right).

This brings us to our characterization of unique ergodicity with respect to the gauge for C*-models.

Theorem 3.10.

Let (𝔐,ρ,G,Ξ)(\mathfrak{M},\rho,G,\Xi) be a W*-dynamical system, and let (𝔄,G,Θ;ι)(\mathfrak{A},G,\Theta;\iota) be a faithful C*-model of (𝔐,ρ,G,Ξ)(\mathfrak{M},\rho,G,\Xi). Then the following conditions are related by the implications (i)\iff(ii)\Rightarrow(iii).

  1. (i)

    The C*-dynamical system (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is uniquely ergodic.

  2. (ii)

    The C*-dynamical system (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is strictly ergodic.

  3. (iii)

    Γ(a)=ρ(ι(a))\Gamma(a)=\rho(\iota(a)) for all positive a𝔄a\in\mathfrak{A}.

Further, if 𝒮G\mathcal{S}^{G} is a simplex, then (iii)\Rightarrow(i).

Proof.

(i)\Rightarrow(ii) Suppose that (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is uniquely ergodic. Then ρι\rho\circ\iota is an invariant state on 𝔄\mathfrak{A}, so it follows that ρι\rho\circ\iota is the unique invariant state. But ρι\rho\circ\iota is also faithful, so it follows that (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is strictly ergodic.

(ii)\Rightarrow(i) Trivial.

(i)\Rightarrow(iii) Suppose that (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is uniquely ergodic, and let a𝔄a\in\mathfrak{A} be positive. Let ϕ\phi be an (a|𝒮G)\left(a|\mathcal{S}^{G}\right)-maximizing state on 𝔄\mathfrak{A}. Then ϕ=ρι\phi=\rho\circ\iota, since both are invariant states and (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is uniquely ergodic. Then ϕ=ρι\phi=\rho\circ\iota, so Γ(a)=ϕ(a)=ρ(ι(a))\Gamma(a)=\phi(a)=\rho(\iota(a)).

(iii)\Rightarrow(i) Suppose that 𝒮G\mathcal{S}^{G} is a simplex, but that (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is not uniquely ergodic. Let ϕ𝒮G\phi\in\mathcal{S}^{G} be an extreme point of 𝒮G\mathcal{S}^{G} different from ρι\rho\circ\iota. Then by Corollary 1.22, there exists a𝔄a\in\mathfrak{A} self-adjoint such that {ϕ}=𝒮maxG(a)\{\phi\}=\mathcal{S}_{\mathrm{max}}^{G}(a). We can assume that aa is positive, since otherwise we could replace aa with a+ra+r for a sufficiently large positive real number r>0r>0, and 𝒮max(a)=𝒮max(a+r)\mathcal{S}_{\mathrm{max}}^{\mathbb{Z}}(a)=\mathcal{S}_{\mathrm{max}}^{\mathbb{Z}}(a+r). Then Γ(a)=ϕ(a)\Gamma(a)=\phi(a). But by the assumption that ϕ\phi is uniquely (a|𝒮G)\left(a|\mathcal{S}^{G}\right)-maximizing, it follows that ρ(ι(a))<ϕ(a)\rho(\iota(a))<\phi(a). Therefore Γ(a)ρ(ι(a))\Gamma(a)\neq\rho(\iota(a)), meaning that (iii) does not attain. Thus ¬\neg(i)¬\Rightarrow\neg(iii). ∎

4. A noncommutative Herman ergodic theorem

For the duration of this section, we assume that (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is a C*-dynamical system such that 𝔄\mathfrak{A} is separable, and that GG is amenable.

Let 𝐅=(Fk)k=1\mathbf{F}=(F_{k})_{k=1}^{\infty} be a right Følner sequence for GG. Write 𝒫𝐅(S)\mathscr{P}^{\mathbf{F}}(S) to denote the set of all limit points of sequences of the form (ϕkAvgFk)k=1\left(\phi_{k}\circ\operatorname{Avg}_{F_{k}}\right)_{k=1}^{\infty}, where ϕkS\phi_{k}\in S for all kk\in\mathbb{N}. Because 𝐅\mathbf{F} is right Følner, we know from Lemma 1.2 that if SS is nonempty, then 𝒫𝐅(S)\mathscr{P}^{\mathbf{F}}(S) will be a nonempty compact subset of 𝒮G\mathcal{S}^{G}. In particular, if S𝒮GS\supseteq\mathcal{S}^{G}, then 𝒫𝐅(S)=𝒮G\mathscr{P}^{\mathbf{F}}(S)=\mathcal{S}^{G} for any choice of 𝐅\mathbf{F}. Moreover, if SS is convex and Θ\Theta-invariant, then 𝒫𝐅(S)=S¯\mathscr{P}^{\mathbf{F}}(S)=\overline{S}.

Question 4.1.

Is 𝒫𝐅(S)\mathscr{P}^{\mathbf{F}}(S) dependent on 𝐅\mathbf{F} in general?

We now define two quantities.

Notation 4.2.

Let 𝐅\mathbf{F} be a right Følner sequence for GG, and SS a nonempty subset of 𝒮\mathcal{S}. Let xx\in\mathfrak{R}. Define

a¯𝐅,S(x)\displaystyle\overline{a}_{\mathbf{F},S}(x) :=sup{ψ(x):ψ𝒫𝐅(x)},\displaystyle:=\sup\left\{\psi(x):\psi\in\mathscr{P}^{\mathbf{F}}(x)\right\},
a¯𝐅,S(x)\displaystyle\underline{a}_{\mathbf{F},S}(x) :=inf{ψ(x):ψ𝒫𝐅(x)},\displaystyle:=\inf\left\{\psi(x):\psi\in\mathscr{P}^{\mathbf{F}}(x)\right\},
d¯𝐅,S(x)\displaystyle\overline{d}_{\mathbf{F},S}(x) :=limk(sup{ϕ(AvgFkx):ϕS}),\displaystyle:=\lim_{k\to\infty}\left(\sup\left\{\phi\left(\operatorname{Avg}_{F_{k}}x\right):\phi\in S\right\}\right),
d¯𝐅,S(x)\displaystyle\underline{d}_{\mathbf{F},S}(x) :=limk(inf{ϕ(AvgFkx):ϕS}).\displaystyle:=\lim_{k\to\infty}\left(\inf\left\{\phi\left(\operatorname{Avg}_{F_{k}}x\right):\phi\in S\right\}\right).

The values a¯𝐅,S,d¯𝐅,S\overline{a}_{\mathbf{F},S},\overline{d}_{\mathbf{F},S} can be compared to the α\alpha and δ\delta 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 d¯𝐅,S(x),d¯𝐅,S(x)\overline{d}_{\mathbf{F},S}(x),\underline{d}_{\mathbf{F},S}(x) are well-defined when S𝒮GS\subseteq\mathcal{S}^{G} is compact, convex, and Θ\Theta-invariant. Moreover, they satisfy

a¯𝐅,S(x)\displaystyle\overline{a}_{\mathbf{F},S}(x) =d¯𝐅,S(x),\displaystyle=\overline{d}_{\mathbf{F},S}(x), a¯𝐅,S(x)\displaystyle\underline{a}_{\mathbf{F},S}(x) =d¯𝐅,S(x).\displaystyle=\underline{d}_{\mathbf{F},S}(x).
Proof.

We’ll prove that a¯𝐅,S(x)=d¯𝐅,S(x)\overline{a}_{\mathbf{F},S}(x)=\overline{d}_{\mathbf{F},S}(x), as the proof that a¯𝐅,S(x)=d¯𝐅,S(x)\underline{a}_{\mathbf{F},S}(x)=\underline{d}_{\mathbf{F},S}(x) is very similar. We know a priori that 𝒫𝐅(S)=S¯\mathscr{P}^{\mathbf{F}}(S)=\overline{S}.

Let (ϕk)k=1(\phi_{k})_{k=1}^{\infty} be a sequence in SS such that for each kk\in\mathbb{N}, we have

sup{ϕ(AvgFkx):ϕS}1/kϕk(AvgFkx)sup{ϕ(AvgFkx):ϕS}.\sup\left\{\phi\left(\operatorname{Avg}_{F_{k}}x\right):\phi\in S\right\}-1/k\leq\phi_{k}(\operatorname{Avg}_{F_{k}}x)\leq\sup\left\{\phi\left(\operatorname{Avg}_{F_{k}}x\right):\phi\in S\right\}.

We know that any limit point of (ϕkAvgFk)k=1\left(\phi_{k}\circ\operatorname{Avg}_{F_{k}}\right)_{k=1}^{\infty} is in S¯\overline{S}. Let k1<k2<k_{1}<k_{2}<\cdots be chosen such that limϕk(AvgFkx)=lim supkϕk(AvgFkx)\lim_{\ell\to\infty}\phi_{k_{\ell}}\left(\operatorname{Avg}_{F_{k_{\ell}}}x\right)=\limsup_{k\to\infty}\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right). We can assume that (ϕkAvgFk)=1\left(\phi_{k_{\ell}}\circ\operatorname{Avg}_{F_{k_{\ell}}}\right)_{\ell=1}^{\infty} is weak*-convergent to a state ψS¯\psi\in\overline{S}, passing to a subsequence if necessary. Then

lim supkϕk(AvgFkx)=limϕk(AvgFkx)=ψ(x)a¯𝐅,S(x).\limsup_{k\to\infty}\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right)=\lim_{\ell\to\infty}\phi_{k_{\ell}}\left(\operatorname{Avg}_{F_{k_{\ell}}}x\right)=\psi(x)\leq\overline{a}_{\mathbf{F},S}(x).

Assume for contradiction that lim infkϕk(AvgFkx)<a¯𝐅,S(x)\liminf_{k\to\infty}\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right)<\overline{a}_{\mathbf{F},S}(x). Let ψ𝒫𝐅(S)\psi^{\prime}\in\mathscr{P}^{\mathbf{F}}(S) be such that ψ(x)>lim infkϕk(AvgFkx)\psi^{\prime}(x)>\liminf_{k\to\infty}\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right). Then

ψ(x)\displaystyle\psi^{\prime}(x) =ψ(AvgFkx)\displaystyle=\psi^{\prime}\left(\operatorname{Avg}_{F_{k}}x\right) ϕk(AvgFkx)1/k.\displaystyle\leq\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right)-1/k.

Let k1<k2<k_{1}^{\prime}<k_{2}^{\prime}<\cdots such that (ϕk(AvgFkx))=1\left(\phi_{k_{\ell}^{\prime}}\left(\operatorname{Avg}_{F_{k_{\ell}}^{\prime}}x\right)\right)_{\ell=1}^{\infty} converges to lim infkϕk(AvgFkx)\liminf_{k\to\infty}\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right). Then

ψ(x)\displaystyle\psi^{\prime}(x) limϕk(AvgFkx)\displaystyle\leq\lim_{\ell\to\infty}\phi_{k_{\ell}^{\prime}}\left(\operatorname{Avg}_{F_{k_{\ell}^{\prime}}}x\right) =lim infkϕk(AvgFkx)\displaystyle=\liminf_{k\to\infty}\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right) <ψ(x),\displaystyle<\psi(x),

a contradiction. Therefore we conclude that lim infkϕk(AvgFkx)a¯𝐅,S(x)\liminf_{k\to\infty}\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right)\geq\overline{a}_{\mathbf{F},S}(x). Thus

a¯𝐅,S(x)lim infkϕk(AvgFkx)lim supkϕk(AvgFkx)a¯𝐅,S(x).\overline{a}_{\mathbf{F},S}(x)\leq\liminf_{k\to\infty}\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right)\leq\limsup_{k\to\infty}\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right)\leq\overline{a}_{\mathbf{F},S}(x).

Thus we can conclude that d¯𝐅,S(x)\overline{d}_{\mathbf{F},S}(x) is well-defined and equal to d¯𝐅,S(x)\overline{d}_{\mathbf{F},S}(x). ∎

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 \mathbb{Z}. Moreover, it extends the result to C*-dynamical systems. Finally, it allows us to relate convergence in certain seminorms to relative ergodic optimizations.

Let (𝔄,G,Θ)(\mathfrak{A},G,\Theta) be a C*-dynamical system, where GG is an amenable group. Given a nonempty subset SS of 𝒮\mathcal{S}, define the seminorm S\|\cdot\|_{S} on 𝔄\mathfrak{A} by

xS:=supϕS|ψ(x)|.\|x\|_{S}:=\sup_{\phi\in S}\left|\psi(x)\right|.
Theorem 4.5.

Let 𝐅\mathbf{F} be a right Følner sequence for GG, and S𝒮S\subseteq\mathcal{S}. Let xx\in\mathfrak{R}, and λ\lambda\in\mathbb{R}. Then the following are equivalent.

  1. (i)

    {ψ(x):ψ𝒫𝐅(S)}={λ}\left\{\psi(x):\psi\in\mathscr{P}^{\mathbf{F}}(S)\right\}=\{\lambda\}.

  2. (ii)

    limkAvgFkxλS=0\lim_{k\to\infty}\left\|\operatorname{Avg}_{F_{k}}x-\lambda\right\|_{S}=0.

Proof.

(i)\Rightarrow(ii): We prove the contrapositive. Suppose there exists ε0>0\varepsilon_{0}>0 and k1<k2<k_{1}<k_{2}<\cdots such that

AvgFkxλS\displaystyle\left\|\operatorname{Avg}_{F_{k_{\ell}}}x-\lambda\right\|_{S} >ε0\displaystyle>\varepsilon_{0} ().\displaystyle(\forall\ell\in\mathbb{N}).

For each kk\in\mathbb{N}, choose ϕkS\phi_{k}\in S such that |ϕk(AvgFkxλ)|12AvgFkxλS\left|\phi_{k}\left(\operatorname{Avg}_{F_{k}}x-\lambda\right)\right|\geq\frac{1}{2}\left\|\operatorname{Avg}_{F_{k}}x-\lambda\right\|_{S}. Then in particular we know that

|ϕk(AvgFkxλ)|\displaystyle\left|\phi_{k_{\ell}}\left(\operatorname{Avg}_{F_{k_{\ell}}}x-\lambda\right)\right| >ε0/2\displaystyle>\varepsilon_{0}/2 ().\displaystyle(\forall\ell\in\mathbb{N}).

By the weak*-compactness of 𝒮\mathcal{S}, there must exist a weak*-convergent subsequence of (ϕkAvgk)=1\left(\phi_{k_{\ell}}\circ\operatorname{Avg}_{k_{\ell}}\right)_{\ell=1}^{\infty}. Assume without loss of generality that (ϕkAvgk)=1\left(\phi_{k_{\ell}}\circ\operatorname{Avg}_{k_{\ell}}\right)_{\ell=1}^{\infty} converges in the weak* topology, and write ψ=limϕkAvgk\psi=\lim_{\ell\to\infty}\phi_{k_{\ell}}\circ\operatorname{Avg}_{k_{\ell}}. Then

|ψ(xλ)|\displaystyle|\psi(x-\lambda)| =|limϕk(AvgFkxλ)|\displaystyle=\left|\lim_{\ell\to\infty}\phi_{k_{\ell}}\left(\operatorname{Avg}_{F_{k_{\ell}}}x-\lambda\right)\right|
=lim|ϕk(AvgFkxλ)|\displaystyle=\lim_{\ell\to\infty}\left|\phi_{k_{\ell}}\left(\operatorname{Avg}_{F_{k_{\ell}}}x-\lambda\right)\right|
ε0/2.\displaystyle\geq\varepsilon_{0}/2.

Therefore ψ(x)λ\psi(x)\neq\lambda, meaning that {ψ(x):ψ𝒫𝐅(S)}{λ}\left\{\psi(x):\psi\in\mathscr{P}^{\mathbf{F}}(S)\right\}\neq\{\lambda\}.

(ii)\Rightarrow(i): Suppose that limkAvgFkxλS=0\lim_{k\to\infty}\left\|\operatorname{Avg}_{F_{k}}x-\lambda\right\|_{S}=0. Let (ϕk)k=1(\phi_{k})_{k=1}^{\infty} be a sequence in SS, and let (ϕkAvgFk)=1\left(\phi_{k_{\ell}}\circ\operatorname{Avg}_{F_{k_{\ell}}}\right)_{\ell=1}^{\infty} be a weak*-convergent subsequence of (ϕkAvgFk)k=1\left(\phi_{k}\circ\operatorname{Avg}_{F_{k}}\right)_{k=1}^{\infty} with limit ψ\psi. Then

|ψ(xλ)|\displaystyle\left|\psi(x-\lambda)\right| =|limϕk(AvgFkxλ)|\displaystyle=\left|\lim_{\ell\to\infty}\phi_{k_{\ell}}\left(\operatorname{Avg}_{F_{k_{\ell}}}x-\lambda\right)\right|
=lim|ϕk(AvgFkxλ)|\displaystyle=\lim_{\ell\to\infty}\left|\phi_{k_{\ell}}\left(\operatorname{Avg}_{F_{k_{\ell}}}x-\lambda\right)\right|
lim supAvgFkxλS\displaystyle\leq\limsup_{\ell\to\infty}\left\|\operatorname{Avg}_{F_{k}}x-\lambda\right\|_{S}
=0.\displaystyle=0.

Therefore {ψ(x):ψ𝒫𝐅(S)}={λ}\left\{\psi(x):\psi\in\mathscr{P}^{\mathbf{F}}(S)\right\}=\{\lambda\}. ∎

Remark 4.6.

An alternate proof of Theorem 4.5 using nonstandard analysis is presented in Section 5.

Corollary 4.7.

Let 𝐅\mathbf{F} be a right Følner sequence for GG. Let xx\in\mathfrak{R}, and λ\lambda\in\mathbb{R}. Then the following are equivalent.

  1. (i)

    {ψ(x):ψ𝒮G}={λ}\left\{\psi(x):\psi\in\mathcal{S}^{G}\right\}=\{\lambda\}.

  2. (ii)

    limkAvgFkxλ=0\lim_{k\to\infty}\left\|\operatorname{Avg}_{F_{k}}x-\lambda\right\|=0.

Proof.

Apply Theorem 4.5 in the case where S=𝒮S=\mathcal{S}, implying that S=\|\cdot\|_{S}=\|\cdot\| and 𝒫𝐅(S)=𝒮G\mathscr{P}^{\mathbf{F}}(S)=\mathcal{S}^{G}. ∎

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 𝐅=(Fk)k=1\mathbf{F}=(F_{k})_{k=1}^{\infty} be a right Følner sequence for GG. A C*-dynamical system (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is uniquely ergodic if and only if (AvgFkx)k=1\left(\operatorname{Avg}_{F_{k}}x\right)_{k=1}^{\infty} converges in norm to an element of 1𝔄\mathbb{C}1\subseteq\mathfrak{A} for all x𝔄x\in\mathfrak{A}.

Proof.

()(\Rightarrow): By taking real and imaginary parts, we can reduce to the case where xx is self-adjoint. If (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is uniquely ergodic, then {ψ(x):ψ𝒮G}\left\{\psi(x):\psi\in\mathcal{S}^{G}\right\} is singleton, so by Corollary 4.7 the averages will converge to a scalar.

()(\Leftarrow): Conversely, if (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is not uniquely ergodic, then there exist two states ψ1,ψ2𝒮G\psi_{1},\psi_{2}\in\mathcal{S}^{G} for which there exists yy\in\mathfrak{R} such that ψ1(y)ψ2(y)\psi_{1}(y)\neq\psi_{2}(y), implying that {ψ(y):ψ𝒮G}\left\{\psi(y):\psi\in\mathcal{S}^{G}\right\} is not singleton. Corollary 4.7 then tells us that (AvgFkx)k=1\left(\operatorname{Avg}_{F_{k}}x\right)_{k=1}^{\infty} 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 (𝔄,G,Θ)(\mathfrak{A},G,\Theta) is a C*-dynamical system, and that 𝔘\mathfrak{U} is a universe that contains 𝔄,G,\mathfrak{A},G,\mathbb{C}. Assume that :𝔘𝔘*:\mathfrak{U}\mapsto\mathfrak{U}^{\prime} is a countably saturated universe embedding. We say that xx\in\prescript{*}{}{\mathbb{C}} is unlimited if |x|>n|x|>n for all nn\in\mathbb{N}, and limited otherwise. Let 𝕃=n{z:zn}\mathbb{L}=\bigcup_{n\in\mathbb{N}}\left\{z\in\prescript{*}{}{\mathbb{C}}:\|z\|\leq n\right\} denote the external ring of limited elements of \prescript{*}{}{\mathbb{C}}. For z,wz,w\in\prescript{*}{}{\mathbb{C}}, we write zwz\simeq w if |zw|<1/n|z-w|<1/n for all nn\in\mathbb{N}. This \simeq is an equivalence relation on \prescript{*}{}{\mathbb{C}}. We define the shadow sh:𝕃\operatorname{sh}:\mathbb{L}\twoheadrightarrow\mathbb{C} to be the \mathbb{C}-linear functional mapping z𝕃z\in\mathbb{L} to the unique (standard) complex number ww\in\mathbb{C} for which zwz\simeq w. The shadow is also order-preserving on 𝕃\mathbb{L}\cap\prescript{*}{}{\mathbb{R}}. Let :={K:k(Kk)}=\prescript{*}{}{\mathbb{N}}_{\infty}:=\left\{K\in\prescript{*}{}{\mathbb{N}}:\forall k\in\mathbb{N}\;(K\geq k)\right\}=\prescript{*}{}{\mathbb{N}}\setminus\mathbb{N} denote the unlimited hypernaturals.

We have the following nonstandard analogue of Lemma 1.2.

Lemma 5.1.

Let (𝔄,G,Θ)(\mathfrak{A},G,\Theta) be a C*-dynamical system, and let GG be an amenable group. Consider a sequence in (ϕk)k=1(\phi_{k})_{k=1}^{\infty} in 𝒮\mathcal{S}, and a right Følner sequence 𝐅=(Fk)k=1\mathbf{F}=(F_{k})_{k=1}^{\infty} for GG. Let KK\in\prescript{*}{}{\mathbb{N}}_{\infty} be an unlimited hypernatural, and define a state ω:𝔄\omega:\mathfrak{A}\to\mathbb{C} by

ω(x)=sh(ϕK(AvgFKx)).\omega(x)=\operatorname{sh}\left(\prescript{*}{}{\phi_{K}\left(\operatorname{Avg}_{F_{K}}x\right)}\right).

Then ω\omega is a well-defined Θ\Theta-invariant state, and is a limit point of the sequence (ϕkAvgFk)k=1\left(\phi_{k}\circ\operatorname{Avg}_{F_{k}}\right)_{k=1}^{\infty}.

Proof.

First, we take up the well-definedness of ω\omega. If x𝔄x\in\mathfrak{A}, then

k(|ϕk(AvgFkx)|x),\forall k\in\mathbb{N}\;\left(\left|\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right)\right|\leq\|x\|\right),

and so by the Transfer Principle

k(|ϕk(AvgFkx)|x).\forall k\in\prescript{*}{}{\mathbb{N}}\;\left(\left|\prescript{*}{}{\phi}_{k}\left(\operatorname{Avg}_{F_{k}}x\right)\right|\leq\|x\|\right).

In particular, it follows that |ϕK(AvgFKx)|x\left|\prescript{*}{}{\phi_{K}\left(\operatorname{Avg}_{F_{K}}x\right)}\right|\leq\|x\|, meaning that ϕK(AvgFKx)𝕃\prescript{*}{}{\phi_{K}\left(\operatorname{Avg}_{F_{K}}x\right)}\in\mathbb{L}. Thus ω(x)\omega(x) is well-defined. We can similarly prove that ω\omega is positive and unital by applying the Transfer Principle to the sentences

kx𝔄\displaystyle\forall k\in\mathbb{N}\;\forall x\in\mathfrak{A}\; (ϕk(AvgFk(xx))0),\displaystyle\left(\phi_{k}\left(\operatorname{Avg}_{F_{k}}\left(x^{*}x\right)\right)\geq 0\right),
k\displaystyle\forall k\in\mathbb{N}\; (ϕk(AvgFk1)=1).\displaystyle\left(\phi_{k}\left(\operatorname{Avg}_{F_{k}}1\right)=1\right).

To prove the Θ\Theta-invariance of ω\omega, we recall from a familiar argument (see proof of Lemma 1.2) that if g0G,x𝔄g_{0}\in G,x\in\mathfrak{A}, then

|ϕk(AvgFkΘg0x)ϕk(AvgFkx)||Fkg0ΔFk||Fk|xk0.\left|\phi_{k}\left(\operatorname{Avg}_{F_{k}}\Theta_{g_{0}}x\right)-\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right)\right|\leq\frac{|F_{k}g_{0}\Delta F_{k}|}{|F_{k}|}\left\|x\right\|\stackrel{{\scriptstyle k\to\infty}}{{\to}}0.

It follows from a classical result of nonstandard analysis [Goldblatt, Theorem 6.1.1] that |ϕK(AvgFKΘg0x)ϕK(AvgFKx)||FKg0ΔFK||FK|x0\left|\prescript{*}{}{\phi}_{K}\left(\operatorname{Avg}_{F_{K}}\Theta_{g_{0}}x\right)-\prescript{*}{}{\phi}_{K}\left(\operatorname{Avg}_{F_{K}}x\right)\right|\leq\frac{\left|F_{K}g_{0}\Delta F_{K}\right|}{\left|F_{K}\right|}\|x\|\simeq 0, meaning that ω(x)=ω(Θg0x)\omega(x)=\omega\left(\Theta_{g_{0}}x\right).

Finally, we argue that ω\omega is a limit point of (ϕkAvgFk)k=1\left(\phi_{k}\circ\operatorname{Avg}_{F_{k}}\right)_{k=1}^{\infty}. For n,,k0;x1,,x𝔄n,\ell,k_{0}\in\mathbb{N};x_{1},\ldots,x_{\ell}\in\mathfrak{A}, consider the sentence σx1,,x;n,k0\sigma_{x_{1},\ldots,x_{\ell};n,k_{0}} given by

k\displaystyle\exists k\in\mathbb{N}\; [(kk0)(min1j|ω(xj)ϕk(AvgFkxj)|<1/n)].\displaystyle\left[(k\geq k_{0})\land\left(\min_{1\leq j\leq\ell}\left|\omega(x_{j})-\phi_{k}\left(\operatorname{Avg}_{F_{k}}x_{j}\right)\right|<1/n\right)\right].

Then σx1,,x;n,k0\prescript{*}{}{\sigma}_{x_{1},\ldots,x_{\ell};n,k_{0}} is true for all n,,k0;x1,,x𝔄n,\ell,k_{0}\in\mathbb{N};x_{1},\ldots,x_{\ell}\in\mathfrak{A}, witnessed by KK. Therefore, it follows from the Transfer Principle that σx1,,x;n,k0\sigma_{x_{1},\ldots,x_{\ell};n,k_{0}} is true for all n,,k0;x1,,x𝔄n,\ell,k_{0}\in\mathbb{N};x_{1},\ldots,x_{\ell}\in\mathfrak{A}. We know that

{{ψ𝒮:min1j|ω(xj)ψ(xj)|<1/n}:n,;x1,,x𝔄}\left\{\left\{\psi\in\mathcal{S}:\min_{1\leq j\leq\ell}|\omega(x_{j})-\psi(x_{j})|<1/n\right\}:n,\ell\in\mathbb{N};x_{1},\ldots,x_{\ell}\in\mathfrak{A}\right\}

is a neighborhood basis for ω\omega in the weak* topology. Thus we have shown that ω\omega is a limit point of the sequence (ϕkAvgFk)k=1\left(\phi_{k}\circ\operatorname{Avg}_{F_{k}}\right)_{k=1}^{\infty}. ∎

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 (ϕk)k=1(\phi_{k})_{k=1}^{\infty} in 𝒮\mathcal{S}, every limit point of the sequence (ϕkAvgFk)k=1\left(\phi_{k}\circ\operatorname{Avg}_{F_{k}}\right)_{k=1}^{\infty} can be written as sh(ϕK(AvgFKx))\operatorname{sh}\left(\prescript{*}{}{\phi_{K}\left(\operatorname{Avg}_{F_{K}}x\right)}\right) for some KK\in\prescript{*}{}{\mathbb{N}}_{\infty}. To see this, choose k1<k2<k_{1}<k_{2}<\cdots such that ψ=limϕkAvgFk\psi=\lim_{\ell\to\infty}\phi_{k_{\ell}}\circ\operatorname{Avg}_{F_{k_{\ell}}} exists. Let 𝒩\mathcal{N} be a countable neighborhood basis for ψ\psi in the weak*-topology, and for each U𝒩,kU\in\mathcal{N},k\in\mathbb{N}, let SU,kS_{U,k} be the set

SU,k={k:(kk)(ϕkAvgFkU)}.S_{U,k}=\left\{k^{\prime}\in\mathbb{N}:\left(k^{\prime}\geq k\right)\land\left(\phi_{k^{\prime}}\circ\operatorname{Avg}_{F_{k^{\prime}}}\in U\right)\right\}.

Then {SU,k}U𝒩,k\left\{S_{U,k}\right\}_{U\in\mathcal{N},k\in\mathbb{N}} has the finite intersection property, and so by the countable saturation of our universe embedding, it follows that there exists KK\in\prescript{*}{}{\mathbb{N}} such that

KU𝒩,kSU,k,K\in\bigcap_{U\in\mathcal{N},k\in\mathbb{N}}\prescript{*}{}{S}_{U,k},

which is necessarily unlimited. Then for any x𝔄x\in\mathfrak{A}, we have that |ϕK(AvgFKx)ψ(x)|<1/n\left|\prescript{*}{}{\phi}_{K}\left(\operatorname{Avg}_{F_{K}}x\right)-\psi(x)\right|<1/n for all nn\in\mathbb{N}, so

sh(ϕK(AvgFKx))=ψ(x).\operatorname{sh}\left(\prescript{*}{}{\phi}_{K}\left(\operatorname{Avg}_{F_{K}}x\right)\right)=\psi(x).

This correspondence can be generalized in the following result.

Proposition 5.2.

Let Ω=(Ω,τ)\Omega=(\Omega,\tau) be a compact Hausdorff topological space, and let :𝔘𝔘*:\mathfrak{U}\to\mathfrak{U}^{\prime} be a countably saturated extension of a universe 𝔘\mathfrak{U} containing Ω\Omega and \mathbb{N}. Let \simeq be the equivalence relation on Ω\prescript{*}{}{\Omega} defined by

xy\displaystyle x\simeq y \displaystyle\iff Uτ(xUyU).\displaystyle\forall U\in\tau\;\left(x\in\prescript{*}{}{U}\leftrightarrow y\in\prescript{*}{}{U}\right).

Define a map sh:ΩΩ\operatorname{sh}:\prescript{*}{}{\Omega}\to\Omega that sends xΩx\in\prescript{*}{}{\Omega} to the unique yΩy\in\Omega such that xyx\simeq y, and let (xk)k=1(x_{k})_{k=1}^{\infty} be a sequence in Ω\Omega. Then the map sh\operatorname{sh} is well-defined.

Further, set

LS((xk)k=1)\displaystyle\operatorname{LS}\left((x_{k})_{k=1}^{\infty}\right)
=\displaystyle= {ωΩ:Uτk[(ωU)(k((kK)(xkU)))]}.\displaystyle\left\{\omega\in\Omega:\forall U\in\tau\;\forall k\in\mathbb{N}\;\left[\left(\omega\in U\right)\rightarrow\left(\exists k^{\prime}\in\mathbb{N}\;\left(\left(k^{\prime}\geq K\right)\land\left(x_{k^{\prime}}\in U\right)\right)\right)\right]\right\}.

Then

{sh(xK):K}LS((xk)k=1).\left\{\operatorname{sh}\left(\prescript{*}{}{x}_{K}\right):K\in\prescript{*}{}{\mathbb{N}}_{\infty}\right\}\subseteq\operatorname{LS}\left((x_{k})_{k=1}^{\infty}\right).

In addition, if * is κ\kappa-saturated for some uncountable cardinal κ>||\kappa>|\mathcal{B}|, where \mathcal{B} is some topological basis \mathcal{B} of τ\tau, then {sh(xK):K}=LS((xk)k=1).\left\{\operatorname{sh}\left(\prescript{*}{}{x}_{K}\right):K\in\prescript{*}{}{\mathbb{N}}_{\infty}\right\}=\operatorname{LS}\left((x_{k})_{k=1}^{\infty}\right).

Proof.

The fact that in a compact topological space, for every xΩx\in\prescript{*}{}{\Omega} exists some yΩy\in\Omega such that xyx\simeq y can be found in [AppliedNSA, Theorem 1.6 of Chapter 3]. As for the uniqueness, assume for contradiction that there existed y,zΩy,z\in\Omega such that xyz,yzx\simeq y\simeq z,y\neq z. Using the Hausdorff property, choose open neighborhoods U,VU,V such that yU,zV,UV=y\in U,z\in V,U\cap V=\emptyset. Then yUy\in U, so by the Transfer Principle yUy\in\prescript{*}{}{U}. Thus zUz\in\prescript{*}{}{U}, because yzy\simeq z. Therefore zUz\in U by the Transfer Principle, a contradiction. Thus sh:ΩΩ\operatorname{sh}:\prescript{*}{}{\Omega}\to\Omega is well-defined.

Let KK\in\prescript{*}{}{\mathbb{N}}_{\infty}, and consider y=sh(xK)y=\operatorname{sh}\left(\prescript{*}{}{x}_{K}\right). Let 𝒩y={U:yU}\mathcal{N}_{y}=\left\{U\in\mathcal{B}:y\in U\right\}, where \mathcal{B} is a topological basis for τ\tau, and consider for k,U𝒩yk\in\mathbb{N},U\in\mathcal{N}_{y} the sentence σU,k\sigma_{U,k} defined by

k[(kk)(xkU)].\exists k^{\prime}\in\mathbb{N}\;\left[\left(k^{\prime}\geq k\right)\land\left(x_{k}\in U\right)\right].

Then σk,U\prescript{*}{}{\sigma}_{k,U} is true for all k,U𝒩yk\in\mathbb{N},U\in\mathcal{N}_{y}, since xKU\prescript{*}{}{x}_{K}\in\prescript{*}{}{U} and KkK\geq k for all kk\in\mathbb{N}, so it follows that σk,U\sigma_{k,U} is true for all k,U𝒩yk\in\mathbb{N},U\in\mathcal{N}_{y}. Since 𝒩y\mathcal{N}_{y} forms a neighborhood basis for yy, it follows that yLS((xk)k=1)y\in\operatorname{LS}\left((x_{k})_{k=1}^{\infty}\right).

Now suppose that * is κ\kappa-saturated for some uncountable cardinal κ>||\kappa>|\mathcal{B}|, and let ωLS((xk)k=1)\omega\in\operatorname{LS}\left((x_{k})_{k=1}^{\infty}\right). Let 𝒩ω={U:ωU}\mathcal{N}_{\omega}=\left\{U\in\mathcal{B}:\omega\in U\right\}. For k,U𝒩ωk\in\mathbb{N},U\in\mathcal{N}_{\omega}, consider the set

Sk,U={k:(kk)(xkU)}.S_{k,U}=\left\{k^{\prime}\in\mathbb{N}:\left(k^{\prime}\geq k\right)\land\left(x_{k^{\prime}}\in U\right)\right\}.

Then {Sk,U:k,U𝒩ω}\left\{S_{k,U}:k\in\mathbb{N},U\in\mathcal{N}_{\omega}\right\} has the finite intersection property, and thus there exists Kk,U𝒩ωSk,UK\in\bigcap_{k\in\mathbb{N},U\in\mathcal{N}_{\omega}}\prescript{*}{}{S}_{k,U}. Thus xKU\prescript{*}{}{x}_{K}\in\prescript{*}{}{U} for all U𝒩ωU\in\mathcal{N}_{\omega}, and KK\in\prescript{*}{}{\mathbb{N}}_{\infty}. Thus ω=sh(xK)\omega=\operatorname{sh}\left(\prescript{*}{}{x}_{K}\right). ∎

Remark 5.3.

Our definitions of \simeq and sh\operatorname{sh} in the statement of Proposition 5.2 is consistent with our definition of \simeq on 𝕃\mathbb{L} in the following sense. We can write 𝕃=n{z:|z|n}\mathbb{L}=\bigcup_{n\in\mathbb{N}}\prescript{*}{}{\left\{z\in\mathbb{C}:|z|\leq n\right\}}. If x,y𝕃x,y\in\mathbb{L}, then there exists nn\in\mathbb{N} such that max{|x|,|y|}n\max\{|x|,|y|\}\leq n. Then x,y{z:|z|n}x,y\in\prescript{*}{}{\left\{z\in\mathbb{C}:|z|\leq n\right\}}. The set {z:|z|n}\left\{z\in\mathbb{C}:|z|\leq n\right\} is compact, and the definition of \simeq on that compact space in the sense of Proposition 5.2 will agree with our definition of \simeq on 𝕃\mathbb{L} 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.

Proof of Theorem 3.4 using nonstandard analysis.

For each kk\in\mathbb{N}, choose a state ϕk\phi_{k} on 𝔄\mathfrak{A} such that

ϕk(AvgFka)=AvgFka.\phi_{k}\left(\operatorname{Avg}_{F_{k}}a\right)=\left\|\operatorname{Avg}_{F_{k}}a\right\|.

Fix KK\in\prescript{*}{}{\mathbb{N}}_{\infty}, and let ω:𝔄\omega:\mathfrak{A}\to\mathbb{C} be the state

ω(x)=sh(ϕK(AvgFKx)).\omega(x)=\operatorname{sh}\left(\prescript{*}{}{\phi}_{K}\left(\operatorname{Avg}_{F_{K}}x\right)\right).

Lemma 5.1 tells us that ω\omega is Θ\Theta-invariant. We argue now that ω(a)=m(a|𝒮G)\omega(a)=m\left(a|\mathcal{S}^{G}\right). This follows because if ψ𝒮G\psi\in\mathcal{S}^{G}, then

ψ(a)=ψ(AvgFka)AvgFka=ϕk(AvgFka)\psi(a)=\psi\left(\operatorname{Avg}_{F_{k}}a\right)\leq\left\|\operatorname{Avg}_{F_{k}}a\right\|=\phi_{k}\left(\operatorname{Avg}_{F_{k}}a\right)

for all kk\in\mathbb{N}, and thus we can apply the Transfer Principle to the sentence k(ψ(a)ϕk(AvgFka))\forall k\in\mathbb{N}\;\left(\psi(a)\leq\phi_{k}\left(\operatorname{Avg}_{F_{k}}a\right)\right) to infer

ψ(a)ϕK(AvgFKa)ψ(a)ω(a).\psi(a)\leq\prescript{*}{}{\phi}_{K}\left(\operatorname{Avg}_{F_{K}}a\right)\Rightarrow\psi(a)\leq\omega(a).

Therefore, we’ve proven that AvgFKam(a|𝒮G)\prescript{*}{}{\left\|\operatorname{Avg}_{F_{K}}a\right\|}\simeq m\left(a|\mathcal{S}^{G}\right) for all KK\in\prescript{*}{}{\mathbb{N}}_{\infty}. Therefore by a classical result of nonstandard analysis [Goldblatt, Theorem 6.1.1], it follows that limkAvgFka=m(a|𝒮G)\lim_{k\to\infty}\left\|\operatorname{Avg}_{F_{k}}a\right\|=m\left(a|\mathcal{S}^{G}\right). ∎

Proof of Proposition 4.3 using nonstandard analysis.

We’ll prove that a¯𝐅,S(x)=d¯𝐅,S(x)\overline{a}_{\mathbf{F},S}(x)=\overline{d}_{\mathbf{F},S}(x), as the proof that a¯𝐅,S(x)=d¯𝐅,S(x)\underline{a}_{\mathbf{F},S}(x)=\underline{d}_{\mathbf{F},S}(x) is very similar. We know a priori that 𝒫𝐅(S)=S¯\mathscr{P}^{\mathbf{F}}(S)=\overline{S}.

Let (ϕk)k=1(\phi_{k})_{k=1}^{\infty} be a sequence in SS such that for each kk\in\mathbb{N}, we have

sup{ϕ(AvgFkx):ϕS}1/kϕk(AvgFkx)sup{ϕ(AvgFkx):ϕS}.\sup\left\{\phi\left(\operatorname{Avg}_{F_{k}}x\right):\phi\in S\right\}-1/k\leq\phi_{k}(\operatorname{Avg}_{F_{k}}x)\leq\sup\left\{\phi\left(\operatorname{Avg}_{F_{k}}x\right):\phi\in S\right\}.

Let KK\in\prescript{*}{}{\mathbb{N}}_{\infty}, and let ω:𝔄\omega:\mathfrak{A}\to\mathbb{C} be the state ω(y)=sh(ϕK(AvgFKy))\omega(y)=\operatorname{sh}\left(\prescript{*}{}{\phi}_{K}\left(\operatorname{Avg}_{F_{K}}y\right)\right). Then ω𝒫𝐅(S)\omega\in\mathscr{P}^{\mathbf{F}}(S), so ω(x)a¯𝐅,S(x)\omega(x)\leq\overline{a}_{\mathbf{F},S}(x).

To prove the opposite inequality, let ψ𝒫𝐅(S)=S¯\psi\in\mathscr{P}^{\mathbf{F}}(S)=\overline{S}. Then

ψ(x)\displaystyle\psi(x) =ψ(AvgFkx)\displaystyle=\psi\left(\operatorname{Avg}_{F_{k}}x\right)
sup{ϕ(AvgFkx):ϕS}\displaystyle\leq\sup\left\{\phi\left(\operatorname{Avg}_{F_{k}}x\right):\phi\in S\right\}
ϕk(AvgFkx)+1/k\displaystyle\leq\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right)+1/k (k).\displaystyle(\forall k\in\mathbb{N}).

Thus the sentence

k(ψ(x)ϕk(AvgFkx)+1/k)\forall k\in\mathbb{N}\;\left(\psi(x)\leq\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right)+1/k\right)

is true. Applying the Transfer Principle then tells us that ψ(x)ϕK(AvgFKx)+1/K\psi(x)\leq\prescript{*}{}{\phi}_{K}\left(\operatorname{Avg}_{F_{K}}x\right)+1/K, implying that ψ(x)ω(x)\psi(x)\leq\omega(x). Taking a supremum over ψS¯=𝒫𝐅(S)\psi\in\overline{S}=\mathscr{P}^{\mathbf{F}}(S) tells us that

a¯𝐅,S(x)ω(x).\overline{a}_{\mathbf{F},S}(x)\leq\omega(x).

Therefore ϕK(x)a¯𝐅,S(x)\prescript{*}{}{\phi}_{K}(x)\simeq\overline{a}_{\mathbf{F},S}(x) for all KK\in\prescript{*}{}{\mathbb{N}}. Thus, by a classical result of nonstandard analysis [Goldblatt, Theorem 6.1.1], it follows that limkAvgFka=a¯𝐅,S(x)\lim_{k\to\infty}\left\|\operatorname{Avg}_{F_{k}}a\right\|=\overline{a}_{\mathbf{F},S}(x). ∎

Proof of Theorem 4.5 using nonstandard analysis.

(i)\Rightarrow(ii): Suppose {ψ(x):ψ𝒫𝐅(S)}={λ}\left\{\psi(x):\psi\in\mathscr{P}^{\mathbf{F}}(S)\right\}=\{\lambda\}. For each kk\in\mathbb{N}, choose ϕkS\phi_{k}\in S such that |ϕk(AvgFkxλ)|12AvgFkxλS\left|\phi_{k}\left(\operatorname{Avg}_{F_{k}}x-\lambda\right)\right|\geq\frac{1}{2}\left\|\operatorname{Avg}_{F_{k}}x-\lambda\right\|_{S}. Fix KK\in\prescript{*}{}{\mathbb{N}}, and let ω:𝔄\omega:\mathfrak{A}\to\mathbb{C} be the state

ω(y)=sh(ϕK(AvgFKy)).\omega(y)=\operatorname{sh}\left(\prescript{*}{}{\phi}_{K}\left(\operatorname{Avg}_{F_{K}}y\right)\right).

Lemma 5.1 tells us that ω𝒫𝐅(S)\omega\in\mathscr{P}^{\mathbf{F}}(S). Thus ω(x)=λ\omega(x)=\lambda. Therefore |ϕK(AvgFKx)λ|0\left|\prescript{*}{}{\phi}_{K}\left(\operatorname{Avg}_{F_{K}}x\right)-\lambda\right|\simeq 0 for all KK\in\prescript{*}{}{\mathbb{N}}_{\infty}, meaning a classical result of nonstandard analysis [Goldblatt, Theorem 6.1.1] tells us that limk|ϕk(AvgFkx)λ|=0\lim_{k\to\infty}\left|\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right)-\lambda\right|=0. But because AvgFkxλS2|ϕk(AvgFkx)λ|\left\|\operatorname{Avg}_{F_{k}}x-\lambda\right\|_{S}\leq 2\left|\phi_{k}\left(\operatorname{Avg}_{F_{k}}x\right)-\lambda\right| for all kk\in\mathbb{N}, we can conclude that limkAvgFkxλS=0\lim_{k\to\infty}\left\|\operatorname{Avg}_{F_{k}}x-\lambda\right\|_{S}=0.

(ii)\Rightarrow(i): Suppose that limkAvgFkxλS=0\lim_{k\to\infty}\left\|\operatorname{Avg}_{F_{k}}x-\lambda\right\|_{S}=0. Let (ϕk)k=1(\phi_{k})_{k=1}^{\infty} be a sequence in SS, and let ω:𝔄\omega:\mathfrak{A}\to\mathbb{C} be the state

ω(y)=sh(ϕK(AvgFKy)).\omega(y)=\operatorname{sh}\left(\prescript{*}{}{\phi}_{K}\left(\operatorname{Avg}_{F_{K}}y\right)\right).

Then

|ω(xλ)||ϕK(AvgFKxλ)|AvgFKxλS0.\left|\omega(x-\lambda)\right|\simeq\left|\prescript{*}{}{\phi}_{K}\left(\operatorname{Avg}_{F_{K}}x-\lambda\right)\right|\leq\left\|\prescript{*}{}{\operatorname{Avg}}_{F_{K}}x-\lambda\right\|_{S}\simeq 0.

Therefore ω(x)=λ\omega(x)=\lambda. We can then take a supremum to get

sup(ϕk)k=1S,K|sh(ϕK(AvgFKx))λ|=0.\sup_{\left(\phi_{k}\right)_{k=1}^{\infty}\in S^{\mathbb{N}},K\in\prescript{*}{}{\mathbb{N}}_{\infty}}\left|\operatorname{sh}\left(\prescript{*}{}{\phi}_{K}\left(\operatorname{Avg}_{F_{K}}x\right)\right)-\lambda\right|=0.

But in light of Proposition 5.2, we know that

𝒫𝐅(S)={ysh(ϕK(AvgFKy)):(ϕk)k=1S,K},\mathscr{P}^{\mathbf{F}}(S)=\left\{y\mapsto\operatorname{sh}\left(\prescript{*}{}{\phi}_{K}\left(\operatorname{Avg}_{F_{K}}y\right)\right):(\phi_{k})_{k=1}^{\infty}\in S^{\mathbb{N}},K\in\prescript{*}{}{\mathbb{N}}_{\infty}\right\},

so this shows that ψ(x)=λ\psi(x)=\lambda for all ψ𝒫𝐅(S)\psi\in\mathscr{P}^{\mathbf{F}}(S). ∎

Acknowledgments

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.

References