1 Introduction
In noncommutative geometry, given a metric space X, one defines certain
$\mathrm {C}^{*}$
-algebras of operators on a Hilbert space with the goal of coding certain aspects of the geometry of X in
$\mathrm {C}^{*}$
-algebraic terms. When interested in the large scale geometric properties of X, that is, in its coarse geometry, a well-known
$\mathrm {C}^{*}$
-algebra is to be considered: the uniform Roe algebra of X. This
$\mathrm {C}^{*}$
-algebra was introduced by Roe to study the index theory of elliptic operators on noncompact manifolds [Reference RoeRoe88, Reference RoeRoe93]. The interest in these algebras was then boosted due to their connection with the coarse Baum-Connes conjecture [Reference YuYu00]. More recently, these
$\mathrm {C}^{*}$
-algebras entered the realm of mathematical physics and researchers interested in topological insulators have been using them as observable algebras in order to describe topological phases. We refer the reader to [Reference KubotaKub17, Reference Ewert and MeyerEM19, Reference JonesJon21, Reference Ludewig and ThiangLT21, Reference BourneBou22] for the rapidly growing literature about uniform Roe algebras in mathematical physics.
The goal of this article is to look at uniform Roe algebras under yet another point of view motivated by mathematical physics: We study KMS states on uniform Roe algebras. Named after mathematical physicists Kubo, Martin, and Schwinger, KMS states are states defined on any C*-algebra A admitting a flow, that is, a strongly continuous one-parameter group
$\{\sigma _t\}_{t\in {\mathbb R}}$
of automorphisms, thought of as the time development of observables of an idealized infinite system of particles. Among the many equivalent definitions of such states, we adopt the one that requires our state
$\varphi $
to satisfy the relation

for every a in A and every analytic element b in A. This condition has been noted by Kubo, Martin, and Schwinger in the late 1950’s, as being satisfied by the grand canonical ensembles in the Gibbs equilibrium formalism for finite systems. Observing that this condition in fact characterizes the Gibbs states, Haag, Hugenholtz, and Winnink later proposed this as a criterion for equilibrium.
The parameter
$\beta $
appearing above is the same parameter weighing the average energy and the entropy in the expression for the free energy in the variational deduction of Gibbs states, and it is often thought of as the reciprocal of the temperature. While our abstract treatment of KMS states will not really involve the physical meaning of
$\beta $
, it is crucial to realize that the existence and uniqueness of KMS states depend in a very fundamental way on
$\beta $
, so much so that we shall refer to states satisfying the above condition as
$(\sigma ,\beta )$
-KMS states, following the the modern literature standards.
Crucially, among the most interesting features of KMS states is the abrupt change in behavior as
$\beta $
crosses certain thresholds. In classical infinite particle systems, a sudden change with temperature is often referred to as a phase transition, which is what one observes when a gas liquefies when cooled down or when a magnet spontaneously loses its magnetization when heated beyond a critical temperature. Thus, if for example there is a unique
$(\sigma ,\beta )$
-KMS state for every
$\beta $
greater than some fixed
$\beta _0$
, while there are many
$(\sigma ,\beta _0)$
-KMS states, one says that a phase transition has happened at the critical value
$\beta _0$
.
It is well known that uniform Roe C*-algebras may be described as the reduced groupoid C*-algebra of a principal, ample, étale groupoid [Reference Skandalis, Tu and YuSTY02]. Moreover, in his 1980 thesis [Reference RenaultRen80, Proposition II.5.4], Renault described a method to study KMS states on groupoid C*-algebras in terms of quasi-invariant measures satisfying a certain Radon-Nikodym condition. Even though we do not directly employ Renault’s result here, much of what we do here may be interpreted as studying such quasi-invariant measures.
Before giving a detailed description of this article and our main findings, we start with some basic definitions.
1.1 Coarse geometry and uniform Roe algebras
A map
$h\colon (X,d)\to (Y,\partial )$
between metric spaces is called coarse if for all
$r>0$
, there is
$s>0$
such that

With coarse maps being the morphisms of interest, local properties of the metric spaces are irrelevant in coarse geometry and one usually restricts themselves to discrete spaces. In fact, for our goals, we will assume the metric spaces to be uniformly locally finite (abbreviated as u.l.f.), that is, they have the property that for each
$r>0$
their balls of radius r are uniformly bounded in size by a finite quantity.Footnote
1
Given a set X,
$\ell _2(X)$
denotes the Hilbert space of square-summable maps
$X\to \mathbb {C}$
and
$(\delta _x)_{x\in X}$
denotes its canonical orthonormal basis. The space of bounded operators on
$\ell _2(X)$
is denoted by
$\mathcal {B}(\ell _2(X))$
and
$\mathcal {K}(\ell _2(X))$
denotes its ideal of compact operators.
Definition 1.1 Let
$(X,d)$
be a u.l.f. metric space. The propagation of an operator
$a\in \mathcal {B}(\ell _2(X))$
is defined by

The
$^{*}$
-algebra of all operators with finite propagation, denoted by
$\mathrm {C}^{*}_u[X]$
, is the algebraic uniform Roe algebra of
$(X,d)$
. The norm closure of
$\mathrm {C}^{*}_u[X]$
, denoted by
$\mathrm {C}^{*}_u(X)$
, is the uniform Roe algebra of
$(X,d)$
.
Uniform Roe algebras code coarse geometric properties of X in terms of
$\mathrm {C}^{*}$
-algebraic properties. For instance, it is known that X has Yu’s property A if and only if
$\mathrm {C}^{*}_u(X)$
is nuclear [Reference Brown and OzawaBO08, Theorem 5.5.7]. Also, it has been recently shown that this construction is rigid in the sense that if the
$\mathrm {C}^{*}$
-algebras
$\mathrm {C}^{*}_u(X)$
and
$\mathrm {C}^{*}_u(Y)$
are isomorphic, then X and Y must be coarsely equivalent [Reference Baudier, Braga, Farah, Khukhro, Vignati and WillettBBF+22b, Theorem 1.2].
1.2 Flows and KMS states on uniform Roe algebras
Given a
$\mathrm {C}^{*}$
-algebra A, an action
$\sigma \colon \mathbb {R}\curvearrowright A$
is a flow if it is strongly continuousFootnote
2
and
$\sigma _t\colon A\to A$
is an isomorphism for all
$t\in \mathbb {R}$
.
Quantum mechanical systems in thermal equilibrium can be described by their so called KMS states. The number
$\beta $
in the definition below should be interpreted as the inverse of the temperature of the system.
Definition 1.2 Let A be a
$\mathrm {C}^{*}$
-algebra and
$\sigma $
be a flow on A. For
$\beta \in \mathbb {R}$
, we say that a state
$\varphi $
on A is a
$(\sigma ,\beta )$
-KMS state if

for all
$a\in A$
and all analytic
$b\in A$
.Footnote
3
In order to study KMS states on uniform Roe algebras, one must first identify natural flows in them. We now introduce such flows. Given a set X and a map
$h\colon {X \to \mathbb {R}}$
, we denote by
$\bar h$
the X-by-X diagonal matrix of reals such that its
$(x,x)$
-entry is
$h(x)$
for all
$x\in X$
and all other entries are zero. Notice that
$\bar h$
canonically induces a bounded operator on
$\ell _2(X)$
if and only if h is bounded.
Definition 1.3 Let X be a u.l.f. metric space and
$h\colon X\to \mathbb {R}$
be a coarse map. We denote by
$\sigma _{h}$
the flow on
$\mathrm {C}^{*}_u(X)$
given by,

for all
$t\in \mathbb {R} $
and all
$a\in \mathrm {C}^{*}_u(X)$
.
Notice that the hypothesis on
$h\colon X\to \mathbb {R}$
being coarse is important so that
$\sigma _h$
is indeed a flow. Indeed, the action
$\sigma _h$
is strongly continuous if and only if h is coarse (see Proposition 2.1). All flows on uniform Roe algebras considered in this article will be of the form above for some appropriate
$h\colon X\to \mathbb {R}$
. In order to have any hope of understanding the KMS states for those flows, we must first understand the analytic elements of
$\mathrm {C}^{*}_u(X)$
or, more precisely, a
$^{*}$
-subalgebra of analytic operators of
$\mathrm {C}^{*}_u(X)$
which is dense in it. We have:
Proposition 1.4 Let X be a u.l.f. metric space and
$h\colon X\to \mathbb {R}$
be a map.
-
1. If h is bounded, then every element of
$\mathrm {C}^{*}_u(X)$ is analytic for
$\sigma _h$
-
2. If h is coarse, then every element of
$\mathrm {C}^{*}_u[X]$ is analytic for
$\sigma _h$ .
The reader may wonder how strong is the restriction of only working with flows of the form above. As we show in Proposition 2.2, if
$\sigma \colon \mathbb {R}\curvearrowright \mathrm {C}^{*}_u(X) $
is an arbitrary flow which leaves the Cartan masa
$\ell _\infty (X)$
invariant, that is,
$\sigma _t(\ell _\infty (X))\subseteq \ell _\infty (X)$
for all
$t\in \mathbb {R}$
, then there is a coarse map
$h\colon X\to \mathbb {R}$
such that
$\sigma =\sigma _h$
.Footnote
4
This corroborates to our claim that such flows form a very natural and general class of flows on those algebras.
1.3 Main results
It is often common in the study of KMS states on a given
$\mathrm {C}^{*}$
-algebra A that there is some “natural”
$\mathrm {C}^{*}$
-subalgebra
$B\subseteq A$
and a conditional expectation
$E\colon A\to B$
such that the KMS states
$\varphi \colon A\to \mathbb {C}$
factor through E. We show that this is also the case in our setting with the “natural”
$\mathrm {C}^{*}$
-subalgebra through which the KMS states factor being the
$\mathrm {C}^{*}$
-algebra of all bounded maps
$X\to \mathbb {C}$
, denoted by
$\ell _\infty (X)$
. Precisely, throughout these notes, we identify
$\ell _\infty (X)$
with the
$\mathrm {C}^{*}$
-algebra of diagonal operators on
$\ell _2(X)$
in the usual way: Given
$a=(a_x)_{x\in X}\in \ell _\infty (X)$
and
$\xi =(\xi _x)_{x\in X}\in \ell _2(X)$
, we let

Given
$A\subseteq X$
,
$\chi _A\in \ell _\infty (X)$
denotes the canonical orthogonal projection
$\ell _2(X)\to \ell _2(A)$
.
We show the following:
Theorem 1.5 Let X be a u.l.f. metric space,
$h\colon X\to \mathbb {R}$
be a coarse map, and
$\beta \in \mathbb {R}$
. If
$\varphi $
is a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(X)$
, then
$\varphi =\varphi \circ E$
, where
$E\colon \mathrm {C}^{*}_u(X)\to \ell _\infty (X)$
is the canonical conditional expectation (see Figure 1).

Figure 1 KMS states on
$\mathrm {C}^{*}_u(X)$
factor through
$\ell _\infty (X)$
, see Section 2.2 for the precise definition of E.
Theorem 1.5 is an extremely powerful tool in our study of KMS states on uniform Roe algebras and most of our results deeply depend on it. For instance, it allows us to understand the case of a flow given by a bounded map
$h\colon X\to \mathbb {R}$
in terms of amenability: for h bounded,
$\mathrm {C}^{*}_u(X)$
has a
$(\sigma _h,\beta )$
-KMS states if and only if X is amenable (see Theorem 2.7). Moreover, Theorem 1.5 allows us to reduce the study of KMS states on uniform Roe algebras to two cases (see Proposition 4.1):
-
(I) strongly continuous KMS states, and
-
(II) KMS states which vanish on the the ideal of compact operators.
The strongly continuous case is the simplest one and the next result summarizes what happens:
Theorem 1.6 Let X be a u.l.f. metric space,
$h\colon X\to \mathbb {R}$
be a coarse map, and
$\beta \geq 0$
. There are strongly continuous
$(\sigma _{h},\beta )$
-KMS states on
$\mathrm {C}^{*}_u(X)$
if and only if

Moreover, a function
$\varphi \colon \mathrm {C}^{*}_u(X)\to \mathbb {C}$
is a strongly continuous
$(\sigma _{ h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(X)$
if and only if

for all
$a=[a_{x,y}]\in \mathrm {C}^{*}_u(X)$
. In particular, whenever they exist, strongly continuous
$(\sigma _h,\beta )$
-KMS states are unique.
In other words, the strongly continuous KMS states are exactly the Gibbs states provided
$e^{-\beta \bar h}$
is trace class (see [Reference Bratteli and RobinsonBR97, Section 6.2.2]). This is of course no big surprise since the strongly continuous states on any operator algebra containing the compacts correspond precisely with the strongly continuous states defined on the whole
$\mathcal {B}(\ell _2(X))$
.
With the strongly continuous case being well understood, we then proceed to study the much more interesting case of KMS states which vanish on the compact operators. This property allows us to factor those states through the uniform Roe corona of X.
Definition 1.7 [Reference Braga, Farah and VignatiBFV21, Definition 1.2]
Let X be a u.l.f. metric space. The uniform Roe corona of X is the
$\mathrm {C}^{*}$
-algebra given by

We denote by
$\pi =\pi _X\colon \mathrm {C}^{*}_u(X)\to \mathrm {Q}^{*}_u(X)$
the canonical quotient map.
A state
$\varphi $
on
$\mathrm {C}^{*}_u(X)$
which vanishes on
$\mathcal {K}(\ell _2(X))$
gives rise to a well-defined state
$\psi $
on
$\mathrm {Q}^{*}_u(X)$
determined by

Moreover, given a coarse map
$h\colon X\to \mathbb {R}$
, the flow
$\sigma _h$
canonically induces a flow on the corona
$\mathrm {Q}^{*}_u(X)$
. Precisely, as
$\sigma _{h}$
leaves
$\mathcal {K}(\ell _2(X))$
invariant, that is,

we obtain a flow
$\sigma _{h}^\infty $
on
$\mathrm {Q}^{*}_u(X)$
by letting

In other words,
$\sigma _{h}^\infty $
is a flow on
$\mathrm {Q}^{*}_u(X)$
which makes the following diagram commute.

We show that the study of
$(\sigma _h,\beta )$
-KMS states on
$\mathrm {C}^{*}_u(X)$
which vanish on the ideal of compact operators completely reduces to the study of
$(\sigma _h^\infty ,\beta )$
-KMS states on
$\mathrm {Q}^{*}_u(X)$
in a canonical way. Precisely:
Proposition 1.8 Let X be a u.l.f. metric space,
$h\colon X\to \mathbb {R}$
be a coarse map, and
$\beta \in \mathbb {R}$
. A state
$\psi $
on
$\mathrm {Q}^{*}_u(X)$
is a
$(\sigma _{h}^\infty ,\beta )$
-KMS state if and only if
$\varphi =\psi \circ \pi $
is a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(X)$
. Moreover, the assignment

is an affine isomorphism from the set of all
$(\sigma _h^\infty ,\beta )$
-KMS states on
$\mathrm {Q}^{*}_u(X)$
to the set of all
$(\sigma _h,\beta )$
-KMS states on
$\mathrm {C}^{*}_u(X)$
which vanish on
$\mathcal {K}(\ell _2(X))$
.
Guided by Proposition 1.8, we then focus on KMS states on the corona algebra
$\mathrm {Q}^{*}_u(X)$
. For that, we show some general results about KMS states on arbitrary
$\mathrm {C}^{*}$
-algebras with respect to arbitrary flows (see Section 3 for details). In a nutshell, we show that the extreme KMS states on an arbitrary
$\mathrm {C}^{*}$
-algebra A are influenced by the center of A, denoted by
$\mathcal {Z}(A)$
, and its
$\mathrm {C}^{*}$
-subalgebras. Returning to our coarse setting, this brings up a seemingly unexpected link between KMS states on uniform Roe algebras and the Higson corona of metric spaces. More precisely, given a u.l.f. metric space X, we denote its Higson compactification by
$hX$
and its Higson corona by
$\nu X=h X\backslash X$
.Footnote
5
The space of continuous functions on the Higson compactification,
$C(hX)$
, is canonically seen as a
$\mathrm {C}^{*}$
-subalgebra of
$\ell _\infty (X)$
, which in turn allow us to canonically identify the continuous functions on its corona,
$C(\nu X)$
, with a
$\mathrm {C}^{*}$
-subalgebra of
$\mathrm {Q}^{*}_u(X)$
. Under this identifications, it has been recently shown that

(see [Reference Baudier, Braga, Farah, Vignati and WillettBBF+22a, Proposition 3.6]).
This link between KMS states and the Higson corona is essential in the analysis of KMS states which vanish on the compacts. Precisely, the next result summarizes our findings on the topic.
Theorem 1.9 Let X be a u.l.f. metric space,
$h\colon X\to \mathbb {R}$
be a coarse map, and
$\beta \in \mathbb {R}$
.
-
1. For any extreme
$ (\sigma _{h}^\infty ,\beta )$ -KMS state
$\psi $ on
$\mathrm {Q}^{*}_u(X)$ , there is
$x\in \nu X$ such that
$$\begin{align*}\psi(a)=a(x) \ \text{ for all }\ a\in C(\nu X).\end{align*}$$
-
2. Suppose there is a
$(\sigma _{h}^\infty ,\beta )$ -KMS state on
$\mathrm {Q}^{*}_u(X)$ whose restriction to
$C(\nu X)$ is faithful. Then, for any
$x\in \nu X$ , there is an extreme
$(\sigma _{h}^\infty ,\beta )$ -KMS state
$\psi $ on
$\mathrm {Q}^{*}_u(X)$ such that
$$\begin{align*}\psi(a)=a(x) \ \text{ for all }\ a\in C(\nu X).\end{align*}$$
Our methods give us a strong control on the support of KMS states on
$\mathrm {C}^{*}_u(X)$
. In order to state this control, a definition is in place.
Definition 1.10 Let X be a u.l.f. metric space,
$x\in \nu X$
, and
$\varphi $
be a state on
$\mathrm {C}^{*}_u(X)$
. We say that
$\varphi $
is supported on x if for all neighborhoods
$U\subseteq h X$
of x, we have
$\varphi (\chi _{U\cap X})=1$
.
Theorem 1.11 Let X be a u.l.f. metric space,
$h\colon X\to \mathbb {R}$
be a coarse map, and
$\beta \in \mathbb {R}$
. The following holds:
-
1. Any extreme
$(\sigma _h,\beta )$ -KMS state on
$\mathrm {C}^{*}_u(X)$ which vanishes on the compacts is supported at some element of
$\nu X$ .
-
2. If there is a
$(\sigma _h,\beta )$ -KMS state on
$\mathrm {C}^{*}_u(X)$ which vanishes on the compacts and such that its induced state on
$\mathrm {Q}^{*}_u(X)$ is faithful on
$C(\nu X)$ , then for every
$x\in \nu X$ there is a
$(\sigma _h,\beta )$ -KMS state on
$\mathrm {C}^{*}_u(X)$ supported on x.
In fact, both Theorems 1.9 and 1.11 have versions that hold with
$C(\nu X)$
being substituted by arbitrary unital
$\mathrm {C}^{*}$
-subalgebras of
$C(\nu X)$
(see Theorems 4.4 and 4.6).
In Section 4.1, we show that the Higson corona of any infinite u.l.f. metric space contains
$2^{2^{\aleph _0}}$
elements (see Theorem 4.14). This result has been first obtained in [Reference KeeslingKee94, Theorem 3], but we chose to present an alternative and self-contained proof here for the readers convenience. As a consequence of this result, Theorem 1.9 and Proposition 1.8 above imply that if there is a
$(\sigma _{h}^\infty ,\beta )$
-KMS state on
$\mathrm {Q}^{*}_u(X)$
whose restriction to
$C(\nu X)$
is faithful, then there are
$2^{2^{\aleph _0}}$
extreme KMS states in both
$\mathrm {Q}^{*}_u(X)$
and
$\mathrm {C}^{*}_u(X)$
(see Corollary 4.16).
1.4 Applications
Our methods can be applied to specific metric spaces. Notice that Theorem 1.6 implies that if the balls of X have polynomial growth, then
$\mathrm {C}^{*}_u(X)$
will have
$(\sigma _{h},\beta )$
-KMS states for any
$\beta>0$
and any “reasonable”
$h\colon X\to \mathbb {R}$
. Indeed, suppose h is such that there is
$L>0$
and
$x_0\in X$
for which

Suppose now p is a polynomial controling the growth of the balls of X, that is, every ball in X centered at
$x_0$
of radius r has at most
$p(r)$
elements. Then, the series
$\sum _{x\in X}e^{-\beta h(x)}$
converges to a finite number for any
$\beta>0$
. Therefore, in order to find examples with interesting phase transition, it is advisable to look for metric spaces with large growth. This makes the n-branching trees natural spaces to apply our theory to.
We point out that, due to the technical aspects of Theorems 1.9 and 1.11, the result below is not a mere corollary of the results above and a deeper analysis of Higson coronas as well as of the weak
$^{*}$
-limit of their strongly continuous KMS states is needed. The study of invariant means on semigroups developed by Chou in [Reference ChouCho69] is also essential for the precise computation of the cardinality of extreme KMS states presented below.
Given
$n\in \mathbb {N}$
, let
$T_n$
denote the n-branching tree, that is,
$T_n=\{\emptyset \}\cup \bigcup _{k=1}^\infty \{1,\ldots , n\}^k$
and we endow
$T_n$
with its canonical graph distance (see Section 5 for details). The branches of
$T_n$
are denoted by
$[T_n]$
, that is,
$[T_n]=\{1,\ldots , n\}^{\mathbb {N}}$
. Given
$\bar x=(x_i)_{i=1}^\infty \in [T_n]$
, we let
$\bar x | k=(x_1,\ldots , x_k)\in T_n$
and
$\bar x| k^\smallfrown T_n$
denotes the words in
$T_n$
which start with
$\bar x| k$
.
Theorem 1.12 Given
$n\in \mathbb {N}$
, let
$T_n$
denote the n-branching tree endowed with its graph distance d and let
$\emptyset $
denote its root. Let
$h\colon T_n\to \mathbb {R}$
be given by
$h(x)=d(x,\emptyset )$
for all
${x\in T_n}$
. Then, there is a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(T_n)$
if and only if
$\beta \geq \log (n)$
. Moreover,
-
1. For
$\beta> \log (n)$ , there is a unique
$(\sigma _{h},\beta )$ -KMS state on
$\mathrm {C}^{*}_u(T_n)$ and this state is strongly continuous.
-
2. For
$\beta =\log (n)$ , the
$(\sigma _{h},\beta )$ -KMS states on
$\mathrm {C}^{*}_u(T_n)$ all vanish on
$\mathcal {K}(\ell _2(T_n))$ . Moreover, for all
$\bar x\in [T_n]$ , there are
$2^{2^{\aleph _0}}$ extreme
$(\sigma _{h},\beta )$ -KMS states
$\varphi $ on
$\mathrm {C}^{*}_u(T_n)$ such that
$$\begin{align*}\varphi(\chi_{\bar x|k^\smallfrown T_n})=1 \ \text{ for all }\ k\in\mathbb{N}.\end{align*}$$
Conversely, any extreme
$(\sigma _{h},\beta )$ -KMS state on
$\mathrm {C}^{*}_u(T_n)$ satisfies the above for an appropriate
$\bar x\in [T_n]$ .
For inverse temperature
$\beta>\log (n)$
, we actually have a precise formula for its unique KMS state (see Theorem 5.4).
Finally, in Section 5.4, we discuss a somewhat unusual phenomenon known as chaotic convergence of KMS states. In order to explain what this means, consider a flow
$\sigma $
on a C*-algebra A admitting a unique KMS state at inverse temperature
$\beta $
, say
$\varphi _\beta $
, for every
$\beta $
in an interval of the form
$(\beta _0,\beta _0+\varepsilon )$
, so that it makes sense to ask whether or not the limit

exists (here the limit should be taken with respect to the weak
$^{*}$
topology). The most commonly observed behavior (see [Reference van Enter and RuszelvER07, Reference Chazottes and HochmanCH10, Reference Coronel and Rivera-LetelierCRL15, Reference Bissacot, Garibaldi and ThieullenBGT18]) is when this limit exists, even when
$\beta _0$
is critical, that is, even when there are multiple
$(\sigma ,\beta _0 )$
-KMS states.
By chaotic convergence of KMS states it is meant a situation where the above fails in the sense that there are different sequences
$\beta _n$
converging to
$\beta _0$
from above for which the corresponding limit states differ. This chaotic behavior has been observed for ground states [Reference Bissacot, Garibaldi and ThieullenBGT18], that is, regarding the limit as
$\beta \to \infty $
, but we are not aware of too many situations where this phenomenon happens at finite temperatures.
As detailed in Theorem 5.15 below, we analyze this question for
$\mathrm {C}^{*}_u(T_n)$
as
$\beta $
approaches
$\log (n)$
from above, showing that such chaotic behavior is indeed present.
2 Basics on KMS states on uniform Roe algebras
In this section, we start our study of KMS states of uniform Roe algebras and prove several general properties which will be essential throughout these notes. We also present some simple examples by studying the KMS states of the simplest coarse space:
$\{n^2\mid n\in \mathbb {N}\}$
. We start this section introducing some notation which was left out from Section 1.
Given a set X and
$x,y\in X$
, we let
$e_{x,y}\in \mathcal {B}(\ell _2(X))$
be the rank 1 partial isometry sending
$\delta _y$
to
$\delta _x$
. If
$A\subseteq X$
, we let

where the letters
$\mathrm {SOT}$
above mean that the sum converges with respect to the strong operator topology. In other words,
$\chi _A$
is the canonical orthogonal projection
$\ell _2(X)\to \ell _2(A)$
. Under the identification of
$\ell _\infty (X)$
with the
$\mathrm {C}^{*}$
-subalgebra of
$\mathrm {C}^{*}_u(X)$
consisting of the diagonal operators, we have that
$\chi _A\in \ell _\infty (X)$
for all
$A\subseteq X$
. The
$\mathrm {C}^{*}$
-algebra of functions
$X\to \mathbb {C}$
which vanish at infinity is identified with the compact operators in
$\ell _\infty (X)$
, that is,

The following description of operators in
$\mathrm {C}^{*}_u[X]$
will be very useful for our goals: Firstly, recall that a partial bijection of X is a bijection
$f\colon A\to B$
between subsets A and B of X. If moreover

then we say that f is a partial translation. Given any partial translation
$f\colon A\subseteq X\to B\subseteq X$
, we define an operator
$v_f$
on
$\ell _2(X)$
by letting

So, each
$v_f$
is a partial isometry and the algebraic uniform Roe algebra is linearly spanned by products of elements in
$\ell _\infty (X)$
by those partial isometries. Precisely, we have

(see [Reference Špakula and WillettŠW17, Lemma 2.4] for details).
2.1 Flows and analytic elements
Our very first result shows that the actions
$\sigma _h\colon \mathbb {R}\curvearrowright \mathrm {C}^{*}_u(X)$
are indeed flows if and only if h is coarse.
Proposition 2.1 Let X be a u.l.f. metric space and
$h\colon X\to \mathbb {R}$
be a map. Then h is coarse if and only if the action
$\sigma _h$
given by Definition 1.3 is strongly continuous, that is,

is continuous for all
$a\in \mathrm {C}^{*}_u(X)$
.
Proof Suppose first that h is coarse. Since
$\mathrm {C}^{*}_u[X]$
is dense in
$\mathrm {C}^{*}_u(X)$
, it is enough to show that the map in (2.2) is continuous for each
$a\in \mathrm {C}^{*}_u[X]$
. Moreover, since
$\mathrm {C}^{*}_u[X]$
is spanned by the subset of all
$av_f$
, for
$a\in \ell _\infty (X)$
and
$f\colon A\subseteq X\to B\subseteq X$
a partial translation, it is enough to notice that (2.2) holds for all such elements
$av_f$
. Fix such a and f. Then, as f is a partial bijection, we have that

Since f is a partial translation and h is coarse, we have

Therefore, it follows from (2.3) and the intermediate value theorem that

is continuous.
Suppose now that the action
$\sigma _h$
is strongly continuous. Suppose towards a contradiction that h is not coarse. Then there is
$r>0$
, and sequences
$(x_i)_i$
and
$(y_i)_i$
in X such that
$\lim _i|h(x_i)-h(y_i)|=\infty $
and
$d(x_i,y_i)\leq r$
for all
$i\in \mathbb {N}$
. As X is u.l.f., those sequences cannot be bounded, so, by going to a subsequence if necessary, we assume that
$(x_i)_i$
and
$(y_i)_i$
are sequences of distinct points of X. We can then define a map

and this map is a partial translation. So,
$v_f\in \mathrm {C}^{*}_u[X]$
and, since
$\sigma _h$
is strongly continuous, we have that

Fix
$\delta>0$
such that

Notice now that

Hence, picking
$i\in \mathbb {N}$
large enough so that

we obtain that
$\|\sigma _t(v_f)-v_f\|\geq 2$
; contradiction.
We now show that our choice of only dealing with flows of the form
$\sigma _h$
for some coarse map
$h\colon X\to \mathbb {R}$
does not represent a big restriction in a sense.
Proposition 2.2 Let X be a u.l.f. metric space and let
$\sigma \colon \mathbb {R}\curvearrowright \mathrm {C}^{*}_u(X)$
be a flow leaving
$\ell _\infty (X)$
invariant, that is,
$\sigma (\ell _\infty (X))\subseteq \ell _\infty (X)$
for all
$t\in \mathbb {R}$
. Then, there is a coarse map
$h\colon X\to \mathbb {R}$
such that
$\sigma =\sigma _h$
.
Proof We first notice that the condition of
$\sigma \colon \mathbb {R}\curvearrowright \mathrm {C}^{*}_u(X)$
leaving
$\ell _\infty (X)$
invariant implies that
$\sigma _t$
is the identity on
$\ell _\infty (X)$
for all
$t\in \mathbb {R}$
. Indeed, as
$\sigma _0$
is by hypothesis the identity on
$\mathrm {C}^{*}_u(X)$
, we have that
$\sigma _0(e_{x,x})=e_{x,x}$
for all
$x\in X$
. As
$\sigma _t$
is an isomorphism for all
$t\in \mathbb {R}$
,
$\sigma _t(e_{x,x})$
must be a projection for all
$t\in \mathbb {R}$
and all
$x\in X$
. Therefore, since
$t\in \mathbb {R}\mapsto \sigma _t(e_{x,x})\in \ell _\infty (X)$
is continuous, this shows that
$\sigma _t(e_{x,x})=e_{x,x}$
for all
$t\in \mathbb {R}$
and all
$x\in X$
. Hence,
$\sigma _t$
must be the identity on
$c_0(X)$
for all
$t\in \mathbb {R}$
. As isomorphisms of uniform Roe algebras are strongly continuous [Reference Špakula and WillettŠW13, Lemma 3.1], this shows that each
$\sigma _t$
is the identity on
$\ell _\infty (X)$
are desired.
Fix
$x\in X$
. For each
$\xi \in \ell _2(X)$
, let
$r_\xi $
be the rank one operator given by

For each
$t\in \mathbb {R}$
, define an operator
$u_t$
on
$\ell _2(X)$
by letting

Claim 2.3 We have

In particular,
$u_t\in \ell _\infty (X)$
for all
$t\in \mathbb {R}$
.
Proof First notice that

Hence, by the arbitrariness of a above, this implies that

for all
$\xi \in \ell _2(X)$
, all
$t\in \mathbb {R}$
, and all
$a\in \mathrm {C}^{*}_u(X)$
.
For the last claim, notice that, as each
$\sigma _t$
is the identity on
$\ell _\infty (X)$
, the previous paragraph implies that each
$u_t$
commutes with the elements of
$\ell _\infty (X)$
. As
$\ell _\infty (X)$
is a maximal abelian subalgebra of
$\mathrm {C}^{*}_u(X)$
, this gives that
$u_t\in \ell _\infty (X)$
for all
$t\in \mathbb {R}$
.
Claim 2.4 The family
$(u_t)_t$
is a one-parameter unitary group, that is,
$t\in \mathbb {R}\mapsto u_t\xi \in \ell _2(X)$
is continuous for all
$\xi \in \ell _2(X)$
,
$u_{t+s}=u_tu_s$
for all
$t,s\in \mathbb {R}$
, and each
$u_t$
is a unitary,
Proof First notice that, as
$t\in \mathbb {R}\mapsto \sigma _t(r_\xi )\in \mathrm {C}^{*}_u(X)$
is continuous,
$t\in \mathbb {R}\mapsto u_t\xi \in \ell _2(X)$
is also continuous for all
$\xi \in \ell _2(X)$
. Also, using (2.4), we have

for all
$\xi \in \ell _2(X)$
and all
$t,s\in \mathbb {R}$
. Finally, as each
$u_t$
is an element of
$\ell _\infty (X)$
with norm at most one satisfying
$u_{t}u_{-t}=1$
, this also shows that
$u_t$
is a unitary.
By Claims 2.3 and 2.4, there is map
$h\colon X\to \mathbb {R}$
such that

Therefore, by Claim 2.3, we have that
$\sigma =\sigma _h$
. By Proposition 2.1, it follows that h must be coarse.
In order to study the KMS states on uniform Roe algebras which are given by the flows defined above, it is essential to understand the analytic elements of this flow. This is precisely the content of Proposition 1.4.
Proof of Proposition 1.4
(1) If h is bounded,
$\bar h$
is a bounded operator on
$\ell _2(X)$
. Therefore, the analyticity of
$e^z$
gives that

is analytic for all
$a\in \mathrm {C}^{*}_u(X)$
.
(2) Since
$\mathrm {C}^{*}_u[X]$
is spanned by the subset of all
$av_f$
, for
$a\in \ell _\infty (X)$
and
$f\colon A\subseteq X\to B\subseteq X$
a partial translation, it is enough to show that each such
$av_f$
is analytic. Fix such a and
$f\colon A\subseteq X\to B\subseteq X$
, and let
$g\colon X\to \mathbb {R}$
be given by

A simple computation gives that

for all
$t\in \mathbb {R}$
. As
$d(f(x),x)\leq r$
for all
$x\in A$
, g is bounded. Then, the analyticity of
$e^z$
implies the that

is analytic; so,
$av_f$
is analytic.
2.2 Factoring KMS-states through
$\ell _\infty (X)$
It is often common in the study of KMS states on a given
$\mathrm {C}^{*}$
-algebra A that there is some “natural”
$\mathrm {C}^{*}$
-subalgebra
$B\subseteq A$
and a conditional expectation
$E\colon A\to B$
such that the KMS states
$\varphi \colon A\to \mathbb {C}$
factor through E; precisely,
$\varphi =\varphi \circ E$
, so the diagram below commutes.

We now show that this also happens with KMS state on uniform Roe algebras.
Recall,
$\ell _\infty (X)$
is a Cartan masa of
$\mathrm {C}^{*}_u(X)$
and the conditional expectation
$E\colon \mathrm {C}^{*}_u(X)\to \ell _\infty (X)$
is simply deleting the matrix entries of the operators on
$\mathrm {C}^{*}_u(X)$
which are not in the main diagonal. Precisely, the canonical conditional expectation
$E\colon \mathrm {C}^{*}_u(X)\to \ell _\infty (X)$
is defined as follows:

for all
$a=[a_{x,y}]\in \mathrm {C}^{*}_u(X)$
and all
$x,y\in X$
.
Proof of Theorem 1.5
As
$\mathrm {C}^{*}_u[X]$
is dense in
$\mathrm {C}^{*}_u(X)$
, it is enough to show that
$\varphi (a)=\varphi (E(a))$
for all
$a\in \mathrm {C}^{*}_u[X]$
. Moreover, as
$\mathrm {C}^{*}_u[X]$
is the span of all
$av_f$
, where
$ a\in \ell _\infty (X)$
and f is a partial translation on X, it is enough to show that
$\varphi (av_f)=0$
for all
$a\in \ell _\infty (X)$
and all partial translations
$f\colon A\subseteq X\to B\subseteq X$
such that
$f(x)\neq x$
for all
$x\in A$
; fix a and f as such.
Let
$r=\sup _{x\in A}d(x,f(x))$
; as f is a partial translation, r is finite. As X is u.l.f., there is a partition

such that each
$A_i$
is
$2r$
-separated, that is,
$d(x,y)>2r$
for all
$i\in \{1,\ldots , n\}$
and all distinct
$x,y\in A_i$
. Therefore,

for all
$i\in \{1,\ldots , n\}$
and all distinct
$x,y\in A_i$
; in particular,
$x\neq f(y)$
. Moreover, as
$f(x)\neq x$
for all
$x\in A$
, this shows that

for all
$i\in \{1,\ldots , n\}$
.
For each
$i\in \{1,\ldots , n\}$
, let
$f_i=f\restriction A_i$
. So, (2.5) implies that
$\chi _{A_i}v_{f_i}=0$
for all
$i\in \{1,\ldots , n\}$
. Therefore, since

we conclude that

Since
$v_f=v_{f_1}+\ldots +v_{f_n}$
, this finishes the proof.
As KMS states on uniform Roe algebras factor through the canonical conditional expectation
$E\colon \mathrm {C}^{*}_u(X)\to \ell _\infty (X)$
, it will be very useful to have a condition on when a state
$\varphi $
on
$\mathrm {C}^{*}_u(X)$
satisfies the KMS condition which depends only on operators on
$\ell _\infty (X)$
. We first introduce some notation which will be used in the next proof. Given
$a=(a_y)_y\in \ell _{\infty }(X)$
and a partial bijection
$f\colon A\subseteq X\to B\subseteq X$
, we let
$a_{\circ f}\in \ell _\infty (X)$
be the operator given by

for all
$x\in X$
.Footnote
6
The relevance of this notation is due to the fact that

Theorem 2.5 Let X be a u.l.f. metric space,
$h\colon X\to \mathbb {R}$
be coarse, and
$\beta \in \mathbb {R}$
. Suppose
$\varphi $
is a state on
$\ell _\infty (X)$
. Then
$\varphi $
satisfies

for all partial translations
$f\colon A\to f(A)$
on X if and only if
$\varphi \circ E$
is a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(X)$
; where
$E\colon \mathrm {C}^{*}_u(X)\to \ell _\infty (X)$
is the canonical conditional expectation.
Proof Suppose first that
$\varphi $
is a
$(\sigma _{h},\beta )$
-KMS on
$\mathrm {C}^{*}_u(X)$
. Let
$f\colon A\to f(A)$
be a partial translation on X. Then,
$\chi _{f(A)}=\chi _{f(A)}v_fv_f^{*}$
. As

the KMS condition gives that

Suppose now that
$\varphi $
satisfies (2.6). First, notice that as
$\ell _\infty (X)$
is linearly generated by the characteristic functions on X, this implies that

for all partial translations f on X and all
$c\in \ell _\infty (\mathrm {Im}(f))$
. By abuse of notation, we extend
$\varphi $
to the whole
$\mathrm {C}^{*}_u(X)$
and still denote it by
$\varphi $
, that is,
$\varphi =\varphi \circ E$
. In order to show that
$\varphi \circ E$
is a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(X)$
, it is enough to show the KMS condition for elements of the form
$av_f$
, where
$a\in \ell _\infty (X)$
and f is a partial translation of X.
Fix
$a,b\in \ell _\infty (X)$
and partial translations f and g on X. Let

and notice that
$g\restriction f(A)=(f\restriction A)^{-1}$
. We can then write

Notice that the last two terms in the right handside of the equality above are in the kernel of the conditional expectation E. Therefore,

For this reason, it is enough to check the KMS condition for partial translations of X which are inverse of each other. For now on, assume that
$g=f^{-1}$
.
Let us now show the KMS condition holds. Firstly, notice that

Then, letting
$c=ba_{\circ f^{-1}}$
, we have that
$c\in \ell _\infty (\mathrm {Im}(f))$
and

Therefore, (2.7) gives that

This shows that
$\varphi $
is a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(X)$
.
2.3 Amenable spaces
A priori, our flows of interest
$\sigma _{h}$
are given by any coarse map
$h\colon X\to \mathbb {R}$
(see Proposition 1.4). Therefore, being automatically coarse, bounded maps form a natural class of maps to produce flows in uniform Roe algebras. However, as we show in this subsection, the existence of KMS states for such flows reduces to the amenability of the metric space, equivalently, to the uniform Roe algebra having a positive unital trace (see [Reference RoeRoe03, Theorem 4.6]). Recall:
Definition 2.6 A u.l.f. metric space X is amenable if there is a nonzero finitely additive measure
$\mu \colon \mathcal {P}(X)\to [0,\infty )$
which is invariant, that is,
$\mu (A)=\mu (B)$
for all
$A,B\subseteq X$
such that there is a partial translation
$f\colon A\to B$
. We call such measure an invariant mean.
Theorem 2.7 Let X be a u.l.f. metric space,
$h\colon X\to \mathbb {R}$
be a bounded map, and
$\beta \in \mathbb {R}$
. Then,
$\mathrm {C}^{*}_u(X)$
has a
$(\sigma _{h},\beta )$
-KMS state if and only if X is amenable.
Before proving Theorem 2.7, we isolate a straightforward lemma which highlights the relation between the trace and the KMS condition when the KMS state is given by elements in the
$\mathrm {C}^{*}$
-algebra.
Lemma 2.8 Let A be a
$C^{*}$
-algebra and
$u\in A$
be invertible. Consider the following assignments:
-
1. For each functional
$\tau $ on A, let
$\varphi _{\tau ,u}$ be the functional given by
$\varphi _{\tau ,u}(a)=\tau (au)$ for all
$a\in A$ .
-
2. For each functional
$\varphi $ on A, let
$\tau _{\varphi ,u}$ be the functional given by
$\tau _{\varphi ,u}(a)=\varphi (au^{-1})$ for all
$a\in A$ .
The assignment
$ \tau \mapsto \varphi _\tau $
defines a bijection between the functionals
$\tau $
on A such that
$\tau (ab)=\tau (ba)$
and the functionals
$\varphi $
on A such that
$ \varphi (ab)=\varphi (buau^{-1}) $
for all
$a,b\in A$
; the inverse of this assignment is
$\varphi \mapsto \tau _{\varphi ,u}$
with the appropriate domain/codomain.
Proof of Theorem 2.7
We start recalling a well-known fact about uniform Roe algebras: a u.l.f. metric space has a positive unital trace if and only if it is amenable [Reference RoeRoe03, Theorem 4.6]. In fact, if
$\mu $
is a nontrivial invariant mean on X, say
$\mu (X)=1$
, and
$E\colon \mathrm {C}^{*}_u(X)\to \ell _\infty (X)$
is the canonical conditional expectation, then

defines a positive unital trace on
$\mathrm {C}^{*}_u(X)$
. On the other hand, if
$\tau $
is a positive unital trace on
$\mathrm {C}^{*}_u(X)$
, then

defines an invariant mean on X.
Suppose then that X is amenable and that
$\tau $
is the trace on
$\mathrm {C}^{*}_u(X)$
given by a nontrivial invariant mean
$\mu $
on X as above. By Lemma 2.8,
$\varphi _{\tau , e^{\beta \bar h}}$
satisfies the
$(\sigma _{h},\beta )$
-KMS condition. Moreover, using the formula of
$\tau $
, we have that

Therefore,
$\varphi $
is positive and, as
$t=\sup _{x\in X} |h(x)|<\infty $
, we have that

Therefore, normalizing
$\varphi $
, we obtain a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(X)$
.
Suppose now that
$\varphi $
is a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(X)$
. By Lemma 2.8,
$\tau _{\varphi ,e^{\beta \bar h}}$
satisfies the trace condition, that is,
$\tau _{\varphi ,e^{\beta \bar h}}(ab)=\tau _{\varphi ,e^{\beta \bar h}}(ba)$
for all
$a,b\in \mathrm {C}^{*}_u(X)$
. As
$\varphi $
is positive and factors through the canonical conditional expectation
$\mathrm {C}^{*}_u(X)\to \ell _\infty (X)$
(Theorem 1.5),
$\tau _{\varphi , e^{\beta \bar h}}$
is also positive. Finally, it follows form our definition of t that

So, normalizing
$\tau $
, we obtain a positive unital trace on X.
As Theorem 2.7 completely takes care of bounded maps, we can now restrict our analyses to unbounded coarse maps
$h\colon X\to \mathbb {R}$
.
2.4 Strongly continuous KMS states
This section deals with strongly continuous KMS states. As we shall see below, those states are the easiest to get and, whenever they exist, they are unique (Theorem 1.6). We also show that the set of
$\beta $
’s for which a strongly continuous KMS state exists must be either of the form
$(t,\infty )$
or
$[t,\infty )$
, for some
$t\geq 0$
(Corollary 2.10 for a precise statement).
Proposition 2.9 Let X be a u.l.f. metric space,
$h\colon X\to [0,\infty )$
be an unbounded coarse map, and
$\beta <0$
. If
$\varphi $
is a
$(\sigma _{h},\beta )$
-KMS state, then
$\varphi (e_{x,x})=0$
for all
$x\in X$
. In particular, there are no strongly continuous
$(\sigma _{ h},\beta )$
-KMS states on
$\mathrm {C}^{*}_u(X)$
.
Proof Fix
$x\in X$
. As h is unbounded, there is a sequence
$(x_n)_n$
in X such that
$\lim _n h(x_n)=\infty $
. Then, if
$\varphi $
is a
$(\sigma _{ h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(X)$
, we have

As
$(\varphi (e_{x_n,x_n}))_n$
is bounded and
$\beta <0$
, we conclude that
$\varphi (e_{x,x})=0$
by letting n go to infinity.
Proof of Theorem 1.6
Suppose
$\varphi $
is a strongly continuous
$(\sigma _{ h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(X)$
. Fix
$x_0\in X$
(this can be thought of as the “center” of X). Since all maps
$(f_x\colon \{x_0\}\to \{x\})_{x\in X}$
are partial translations, the KMS condition gives us that

for all
$x\in X$
(see Theorem 2.5). As
$\varphi $
is strongly continuous,

So,
$\varphi (e_{x_0,x_0})\neq 0$
and

must be finite (as well as independent on
$x_0$
). The formula for
$\varphi $
in the statement of the theorem then follows immediately from the strong continuity of
$\varphi $
.
Suppose now
$Z(\beta )$
is finite and
$\varphi $
is given as in the statement of the theorem. Clearly,
$\varphi $
is a strongly continuous state on
$\mathrm {C}^{*}_u(X)$
. Moreover, if
$f\colon A\to B$
is a partial translation of X, then, by the formula of
$\varphi $
, we have

So, by Theorem 2.5,
$\varphi $
is a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(X)$
.
The following is a straightforward consequence of Proposition 2.9 and Theorem 1.6.
Corollary 2.10 Let X be a u.l.f. metric space and
$h\colon X\to [0,\infty )$
be an unbounded coarse map. The subset of all
$\beta \in \mathbb {R}$
for which there are strongly continuous
$(\sigma _{ h},\beta )$
-KMS states on
$\mathrm {C}^{*}_u(X)$
is either of the form
$(t,\infty )$
or
$[t,\infty )$
for some
$t\geq 0$
.
Remark 2.11 Throughout these notes, we will see many examples for which the set of
$\beta $
’s admitting are strongly continuous
$(\sigma _{ h},\beta )$
-KMS states on
$\mathrm {C}^{*}_u(X)$
are of the form
$(\beta _0,\infty )$
for some
$\beta _0>0$
. This could give the impression this must always be the case, however, this is not so. For instance, let
$X=\{n\in \mathbb {N}\mid n\geq 3\}$
and let
$h(x)=\log (x\log ^2(x))$
for all
$x\in X$
(the restriction of
$x\geq 3$
is simply so that h is well defined). In this case,

and this series converges if and only if
$\beta \geq 1$
.
2.5 The simplest coarse space
Under the optics of coarse geometry, the simplest infinite metric space is the coarse disjoint union of singletons; that is, any metric space which is bijectively coarsely equivalent to

where
$X_0$
is endowed with the usual metric d on the natural numbers. In this subsection, we study KMS states on
$X_0$
. The simplicity of the geometry of
$X_0$
makes any map
$h\colon X_0\to Y$
, where Y is another metric space, be automatically coarse. Also, given any
$r>0$
, there is a finite
$F\subseteq X_0\times X_0$
such that

Therefore, it follows that

Proposition 2.12 Let
$(X_0,d)$
be the coarse disjoint union of singletons described above. If
$\varphi $
is a state on
$\ell _\infty (X_0)$
such that
$\varphi \restriction c_0(X_0)=0$
, then
$\varphi \circ E$
is a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(X_0)$
for all
$h\colon X_0\to \mathbb {R}$
and all
$\beta \in \mathbb {R}$
; where
$E\colon \mathrm {C}^{*}_u(X_0)\to \ell _\infty (X_0)$
denotes the canonical conditional expectation.
Proof Let
$f\colon A\subseteq X_0\to B\subseteq X_0$
be a partial translation. Then, there must be a partition
$A=A_1\sqcup A_2$
such that
$f(x)=x$
for all
$x\in A_1$
and
$|A_2|<\infty $
. As
${\varphi \restriction c_0(X_0)=0}$
, we have that

Similarly, we have

The result then follows from Theorem 2.5.
Remark 2.13 Here is a more conceptual way of obtaining Proposition 2.12: notice that since
$\mathrm {C}^{*}_u(X_0)=\ell _\infty (X_0)+\mathcal {K}(\ell _2(X_0))$
, we must have
$\mathrm {Q}^{*}_u(X_0)\cong \ell _\infty /c_0$
; so,
$\mathrm {Q}^{*}_u(X_0)$
is abelian. Moreover, as
$\sigma _h$
is the identity on
$\ell _\infty (X_0)$
, the flow
$\sigma _h^\infty $
induced by
$\sigma _h$
on
$\mathrm {Q}^{*}_u(X_0)$
is trivial (see Section 1 for the definition of
$\sigma _h^\infty $
). In particular, any state on
$\mathrm {Q}^{*}_u(X_0)$
is KMS for any
$\beta $
. The result is then a corollary of Proposition 1.8.
We now restrict our study of KMS states on
$X_0$
to a specific map h. This will allow us to find all KMS states on
$\mathrm {C}^{*}_u(X_0)$
for the corresponding flow. For the sake of generality, we first isolate a result which does not depend on X being the coarse disjoint union of singletons per se.
Corollary 2.14 Let d be any u.l.f. metric on
$\mathbb {N}$
for which the map
$h(x)=\log (x)$
is coarse and let
$\beta \in \mathbb {R}$
. If
$\varphi $
is a strongly continuous
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(\mathbb {N},d)$
, then
$\beta>1$
and

for all
$[a_{x,y}]\in \mathrm {C}^{*}_u(\mathbb {N},d)$
.
Proof This is a straightforward consequence of Theorem 1.6.
We can now describe the KMS states on
$X_0$
completely with
$h=\log $
. Precisely:
Corollary 2.15 Let
$X_0=\{n^2\mid n\in \mathbb {N}\}$
be the coarse disjoint union of singletons described above,
$\beta \in \mathbb {R}$
, and
$h\colon X\to \mathbb {R}$
be given by
$h(x)=\log (\sqrt {x})$
for all
$x\in X_0$
. The
$(\sigma _{h},\beta )$
-KMS states of
$\mathrm {C}^{*}_u(X_0)$
are precisely the following:
-
1. Any state on
$\mathrm {C}^{*}_u(X_0)$ which vanishes on
$c_0(X_0)$ ,
-
2. If
$\beta>1$ , then
$\mathrm {C}^{*}_u(X_0)$ has a unique strongly continuous
$(\sigma _{h},\beta )$ -KMS state and this state is given by
$$\begin{align*}\varphi([a_{x,y}])=\frac{1}{\sum_{n=1}^\infty \frac{1}{n^\beta}}\sum_{n=1}^\infty \frac{a_{n^2,n^2}}{n^\beta},\end{align*}$$
for all
$[a_{x,y}]\in \mathrm {C}^{*}_u(X_0)$ , and
-
3. for
$\beta>1$ , any convex combination of the states above.
3 Intermission
As seen in Theorem 1.6, strongly continuous KMS states on uniform Roe algebras are completely understood; so we are left to understand the strongly discontinuous case. In this section, before explicitly perusing this goal, we take a short break from uniform Roe algebras per se, and present some results about KMS states on arbitrary
$\mathrm {C}^{*}$
-algebras with respect to arbitrary flows. The technical results herein will be essential in the analysis to follow of KMS states on uniform Roe algebras which are strongly discontinuous.
We start by properly stating the settings of this section. But firstly, we recall some standard notation: if A is a
$\mathrm {C}^{*}$
-algebra, then
$\mathcal {Z}(A)$
denotes the center of A, that is,

Moreover, if K is a compact Hausdorff space, then
$C(K)$
denotes the
$\mathrm {C}^{*}$
-algebra of all continuous functions
$K\to \mathbb {C}$
.
Assumption 3.1 Throughout this section, we fix a unital
$\mathrm {C}^{*}$
-algebra A, a flow
$\sigma $
on A, and
$\beta \in \mathbb {R}$
. Moreover, we fix a unital
$\mathrm {C}^{*}$
-subalgebra
$C\subseteq A$
contained in
$\mathcal {Z}(A)$
, and identify C with
$ C(\Omega (C))$
via the Gelfand transform; here
$\Omega (C)$
denotes the spectrum of C.
Proposition 3.2 In the setting of Assumption 3.1: If
$\varphi $
is a
$(\sigma ,\beta )$
-KMS state on A and
$c\in A$
is a positive element in the center of A with
$\varphi (c)\neq 0$
, then the state
$\varphi _c$
on A defined by

is a
$(\sigma ,\beta )$
-KMS state on A.
Proof First notice that, as
$c\in \mathcal {Z}(A)$
, then
$ac$
is positive for all positive
$a\in A$
. Therefore,
$\varphi _c$
is indeed a state. Given
$a,b\in A$
, with b analytic, we have

So,
$\varphi _c$
is a
$(\sigma ,\beta )$
-KMS state on A.
Proposition 3.3 In the setting of Assumption 3.1: If
$\varphi $
is an extreme
$(\sigma ,\beta )$
-KMS state on A, then there is
$x\in \Omega (C)$
such that

In particular, letting

we have that
$\varphi \restriction J_x=0$
.
Proof By Riesz representation theorem, there is a probability measure
$\mu $
on
$\Omega (C)$
such that

Let
$K\subseteq \Omega (C)$
be the support of
$\mu $
. Let us show that K is a singleton. In order to prove this, suppose by contradiction that there are two distinct points
$x,y\in K$
. By Urysohn’s lemma, we can pick a positive
$k\in C(\Omega (C))$
with
$\|k\|\leq 1$
and such that
$k(x)=1$
and
$k(y)=0$
. Setting
$\ell =1-k$
, we have that both k and
$\ell $
are not identically zero on K, so both
$\varphi (k)$
and
$\varphi (\ell )$
are nonzero. By Proposition 3.2,
$\varphi _k$
and
$\varphi _\ell $
are
$(\sigma ,\beta )$
-KMS states on A, and it is clear that

where
$\lambda =\varphi (k)$
. Since
$\varphi _k\neq \varphi _\ell $
, this contradicts the assumption that
$\varphi $
is an extreme
$(\sigma ,\beta )$
-KMS state. So, K contains only one point, say
$K=\{x\}$
. Therefore,
$\mu $
must be the dirac measure on
$\{x\}$
, which gives that

The last claim follows straightforwardly from the above.
Definition 3.4 In the setting of Assumption 3.1:
-
1. We denote the set of all
$(\sigma ,\beta )$ -KMS states on A by
$K_\beta $ .
-
2. For each
$x\in \Omega (C)$ , let
$$\begin{align*}J_x=\{a\in C=C(\Omega(C))\mid a(x)=0\} \text{ and }K^x_\beta=\{\varphi\in K_\beta\mid \varphi\restriction J_x=0\}.\end{align*}$$
It is plainly clear that each
$K^x_\beta $
is a weak
$^{*}$
-closed convex subset of
$K_\beta $
.
Recall that if C is a convex subset in a vector space, and
$E\subseteq C$
is convex, then E is said to be an extreme subset of C if, for any pair of points x and y in C, such that
$\lambda x +(1-\lambda )y\in E$
, with
$0<\lambda <1$
, one has that both x and y lie in E. For example, if x is an extreme point of C, then
$\{x\}$
is an extreme subset of C.
Proposition 3.5 In the setting of Assumption 3.1: For all
$x\in \Omega (C)$
, one has that
$K^x_\beta $
is an extreme subset of
$K_\beta $
.
Proof Pick
$\varphi \in K^x_\beta $
and assume that

where
$\varphi _1,\varphi _2\in K_\beta $
and
$\lambda \in (0,1)$
. Denoting by
$\psi $
,
$\psi _1$
, and
$\psi _2$
the restrictions of
$\varphi $
,
$\varphi _1$
, and
$\varphi _2$
to C, respectively, it is apparent that

By Proposition 3.3,
$\psi $
is a character of
$C=C(\Omega (C))$
. Hence,
$\psi $
is an extreme point of the unit ball of the dual of C. This shows that
$\psi =\psi _1=\psi _2$
, which in turn implies that both
$\varphi _1$
and
$\varphi _2$
vanish on
$J_x$
. Therefore,
$\varphi _1,\varphi _2\in K^x_\beta $
as desired.
We can now present the main result of this section. In it,
$\mathrm {Ext}(K_\beta )$
(resp.
$\mathrm {Ext}(K_\beta ^x)$
) denotes the subset of all extreme elements of
$K_\beta $
(resp.
$\mathrm {Ext}(K_\beta ^x)$
).
Theorem 3.6 In the setting of Assumption 3.1: We have

Moreover, if there is a
$(\sigma ,\beta )$
-KMS state on A whose restriction to C is faithful, then
$K^x_\beta \neq \emptyset $
for all
$x\in \Omega (C)$
. In particular, if such KMS state exists, we have that

Proof By Proposition 3.3, every extreme point
$\varphi $
of
$K_\beta $
lies in some
$K^x_\beta $
and, in this case,
$\varphi $
is evidently an extreme point of
$K^x_\beta $
. Conversely, as each
$K^x_\beta $
is an extreme subset of
$K_\beta $
(Proposition 3.5), every extreme point of any
$K^x_\beta $
is an extreme point of
$K_\beta $
.
Suppose now that there is a
$(\sigma ,\beta )$
-KMS state
$\varphi $
on A whose restriction to C is faithful. Fix
$x\in \Omega (C)$
and let us show
$K^x_\beta \neq \emptyset $
. Let
$\mathcal {V}$
be the family of all open subsets of
$\Omega (C)$
which contain x and, for each
$V\in \mathcal {V}$
, let
$h_V\colon \Omega (C)\to [0,1]$
be a continuous function such that
$h_V(x)=1$
and
$h_V(y)=0$
for all
$y\not \in V$
. By the faithfulness of
$\varphi $
,
$\varphi (h_V)\neq 0$
for all
$V\in \mathcal {V}$
. Therefore, by Proposition 3.2, each
$\varphi _V=\varphi _{h_V}$
is a
$(\sigma ,\beta )$
-KMS state on A.
Consider
$\mathcal {V}$
as a directed set with the usual reverse containment order. By Banach-Alaoglu theorem,
$K_\beta $
is weak
$^{*}$
-compact. Hence, by passing to a subset if necessary, we can assume that
$(\varphi _V)_{V\in \mathcal {V}}$
converges to some
$\psi \in K_\beta $
in the weak
$^{*}$
-topology. As
$\psi $
is a limit of
$(\varphi _V)_{V\in \mathcal {V}}$
and as
$\lim _{V,\mathcal {V}}\|a h_V\|=0$
, for all
$a\in J_x$
, the state
$\psi $
must vanish on
$J_x$
. This shows that
$\varphi \in K^x_\beta $
and
$K^x_\beta $
cannot be empty as desired.
The last claim is a straightforward consequence of the above.
4 Factoring KMS states through the uniform Roe corona
In this section, we return to the setting of uniform Roe algebras and study strongly discontinuous KMS states (the strongly continuous case was already completely treated in Theorem 1.6). We start noticing that, in order to study such states, it is enough to study the KMS states which vanish on the ideal of compact operators. Precisely:
Proposition 4.1 Let X be a u.l.f. metric space,
$h\colon X\to \mathbb {R}$
be coarse, and
$\beta \in \mathbb {R}$
. Suppose
$\varphi $
is a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(X)$
and define a positive functional
$\psi $
on
$\mathrm {C}^{*}_u(X)$
by letting

where
$\mathcal {F}$
is the net of all finite subsets of X ordered by reverse inclusion. Then,
$\psi $
is well defined and
-
1.
$\psi $ is strongly continuous and satisfies the
$(\sigma _{h},\beta )$ -KMS condition, and
-
2.
$\varphi -\psi $ is a positive functional which satisfies the
$(\sigma _{h},\beta )$ -KMS condition and vanishes on
$\mathcal {K}(\ell _2(X))$ .
Proof The fact that
$\psi $
is well defined follows straightforwardly from the fact that
$\varphi $
is positive and factors through
$\ell _\infty (X)$
(Theorem 1.5). Positivity and strong continuity of
$\psi $
are then completely immediate. It is also immediate that
$\psi \leq \varphi $
, so
$\varphi -\psi $
is also positive. Since
$\psi \restriction \mathcal {K}(\ell _2(X))=\varphi \restriction \mathcal {K}(\ell _2(X))$
,
$\varphi -\psi $
vanishes on the compacts. We are only left to show that both
$\psi $
and
$\varphi -\psi $
satisfy the
$(\sigma ,\beta )$
-KMS condition. But this is an immediate consequence of Theorem 2.5 and the formula of
$\psi $
.
Theorem 1.6 and Proposition 4.1 show that, in order to understand the KMS states on uniform Roe algebras, we only need to focus of the states which vanish on the ideal of compact operators. For the remainder of this section, this will be our focus. Since the compacts form an ideal, we can factor those states through the quotient algebra. For that, recall that the uniform Roe corona of X is

(see Definition 1.7). If
$\varphi $
is a state on
$\mathrm {C}^{*}_u(X)$
which vanishes on
$\mathcal {K}(\ell _2(X))$
, then
$\varphi $
gives rise to a well-defined state
$\psi $
on
$\mathrm {Q}^{*}_u(X)$
determined by

Moreover, given a coarse map
$h\colon X\to \mathbb {R}$
, the flow
$\sigma _{h}$
induces a flow
$\sigma _{h}^\infty $
on
$\mathrm {Q}^{*}_u(X)$
by letting

(see Section 1.3 for more details).
Proposition 1.8 highlights the relations between
$\varphi $
and
$\psi $
, and
$\sigma _{h}$
and
$\sigma _{h}^\infty $
defined above.
Proof of Proposition 1.8.
Notice that if b is an analytic element in
$\mathrm {C}^{*}_u(X)$
for
$\sigma _{h}$
, then
$\pi (b)$
is analytic for
$\sigma _{h}^\infty $
and, moreover,

Therefore, the image of the set of all analytic elements in
$\mathrm {C}^{*}_u(X)$
under
$\pi $
forms a dense set of analytic elements in
$\mathrm {Q}^{*}_u(X)$
. Consequently, in order to check that a state
$\psi $
on
$\mathrm {Q}^{*}_u(X)$
is a
$(\sigma _{h}^\infty ,\beta )$
-KMS state, it suffices to prove that

for all
$a,b\in A$
with b analytic. Observing that the left-hand-side above coincides with
$(\psi \circ \pi )(a\sigma _{h,i\beta }(b))$
and that the right-hand-side equals
$(\psi \circ \pi )(ba)$
, the first statement of the proposition follows. The second statement in turn follows from the first one immediately.
Proposition 1.8 then reduces our problem to the one of understanding the KMS states on the uniform Roe corona
$\mathrm {Q}^{*}_u(X)$
. In view of Section 3, it will be useful to study the center
$\mathrm {Q}^{*}_u(X)$
as well as its
$\mathrm {C}^{*}$
-subalgebras. This brings up a seemingly unexpected link between KMS states and the Higson corona. Recall:
Definition 4.2 Let X be a u.l.f. metric space.
-
1. A bounded function
$f\colon X\to \mathbb {C}$ is a Higson function if for all
$\varepsilon>0$ and all
$R>0$ there is a finite
$F\subseteq X$ such that
$$\begin{align*}\forall x,y\in X\setminus F, \ d(x,y)<R\ \text{ implies }\ |f(x)-f(y)|<\varepsilon.\end{align*}$$
The set of all Higson functions on X forms a
$\mathrm {C}^{*}$ -subalgebra of
$\ell _\infty (X)$ which we denote by
$C_h(X)$ .
-
2. The spectrum of
$C_h(X)$ , denoted by
$hX$ , is called the Higson compactification of X. So, the Gelfand transform gives us the identification
$C(hX)=C_h(X)$ .
-
3. The boundary
$\nu X=hX\setminus X$ is called the Higson corona and we have the identification
$C(\nu X)=C_h(X)/c_0(X)$ .
Notice that, as
$C_h(X)\subseteq \ell _\infty (X)$
, we may canonically view
$C_h(X)/c_0(X)$
as a
$\mathrm {C}^{*}$
-subalgebra of
$\mathrm {Q}^{*}_u(X)$
; so, by the identification
$C(\nu X)=C_h(X)/c_0(X)$
, we have

It has been recently observed that the center of
$\mathrm {Q}^{*}_u(X)$
is precisely the Higson corona of X. Indeed, the following was proven in [Reference Baudier, Braga, Farah, Vignati and WillettBBF+22a, Proposition 3.6] as a consequence of [Reference Špakula and ZhangŠZ20, Theorem 3.3].
Proposition 4.3 Given a u.l.f. metric space X, we have that

We now apply our results of Section 3 to our coarse setting. In what follows, if C is a unital
$\mathrm {C}^{*}$
-algebra,
$\Omega (C)$
denotes the spectrum of C. So,
$\Omega (C)$
is a compact Hausdoff topological space and we use the identification
$C=C(\Omega (C))$
given by the Gelfand transform.
Theorem 4.4 Let X be a u.l.f. metric space,
$h\colon X\to \mathbb {R}$
be a coarse map, and
$\beta \in \mathbb {R}$
. Let C be a unital
$\mathrm {C}^{*}$
-subalgebra of
$C(\nu X)$
.
-
1. For any extreme
$ (\sigma _{h}^\infty ,\beta )$ -KMS state
$\psi $ on
$\mathrm {Q}^{*}_u(X)$ , there is
$x\in \Omega (C)$ such that
$$\begin{align*}\psi(a)=a(x) \ \text{ for all }\ a\in C=C(\Omega(C)).\end{align*}$$
-
2. Suppose there is a
$(\sigma _{h}^\infty ,\beta )$ -KMS state on
$\mathrm {Q}^{*}_u(X)$ whose restriction to C is faithful. Then, for any
$x\in \Omega (C)$ , there is an extreme
$(\sigma _{h}^\infty ,\beta )$ -KMS state
$\psi $ on
$\mathrm {Q}^{*}_u(X)$ such that
$$\begin{align*}\psi(a)=a(x) \ \text{ for all }\ a\in C=C(\Omega(C)).\end{align*}$$
Proof This is a mere corollary of Theorem 3.6.
We now obtain Theorem 1.11 by proving a more general version of it. For that, we first generalize Definition 1.10.
Definition 4.5 Let X be a u.l.f. metric space and
$\bar X$
be a compactification of X.
-
1. We call
$\bar X$ Higson compatible if
$$\begin{align*}f\restriction X\in C_h(X)\ \text{ for all }\ f\in C(\bar X).\end{align*}$$
-
2. If
$\bar X$ is Higson compatible and
$x\in \bar X$ , we say that a state
$\varphi $ on
$\mathrm {C}^{*}_u(X)$ is
$\bar X$ -supported on x if for all neighborhoods
$U\subseteq \bar X$ of x, we have
$\varphi (\chi _{U\cap X})=1$ .
Notice that if
$\bar X$
is a Higson compatible compactification of X, then
$C(\bar X)$
can be canonically identified with a
$\mathrm {C}^{*}$
-subalgebra of
$C_h(X)$
, which in turn allows us to identify
$C(\bar X)/c_0(X)$
with a
$\mathrm {C}^{*}$
-subalgebra of
$C(\nu X)\subseteq \mathrm {Q}^{*}_u(X)$
.Footnote
7
Theorem 4.6 Let X be a u.l.f. metric space,
$h\colon X\to \mathbb {R}$
be a coarse map, and
$\beta \in \mathbb {R}$
. Let
$\bar X$
be a Higson compatible compactification of X. The following holds:
-
1. Any extreme
$(\sigma _h,\beta )$ -KMS state on
$\mathrm {C}^{*}_u(X)$ which vanishes on the compacts is
$\bar X$ -supported at some element of
$\bar X$ .
-
2. If there is a
$(\sigma _h,\beta )$ -KMS state on
$\mathrm {C}^{*}_u(X)$ which vanishes on the compacts and such that its induced state on
$\mathrm {Q}^{*}_u(X)$ is faithful on
$C(\bar X)/c_0(X)$ , then for every
$x\in \bar X$ there is a
$(\sigma _h,\beta )$ -KMS state on
$\mathrm {C}^{*}_u(X)$ which is
$\bar X$ -supported on x.
4.1 The size of the Higson corona
We show that the Higson corona of an infinite u.l.f. metric space must always have
$2^{2^{\aleph _0}}$
many elements (Theorem 4.14). Together with the previous results in this section, this will give us a very strong control of the cardinality of KMS states on
$\mathrm {C}^{*}_u(X)$
.
In this subsection, we work a lot with partial bijections f of X and it will be useful to be able to write “
$f(A)$
” regardless of whether
$A\subseteq \mathrm {Dom}(f)$
. We then define: given any set X, a partial bijection
$f\colon \mathrm {Dom}(f)\to \mathrm {Im}(f)$
of X, and
$A\subseteq X$
, we let

Also, given partial bijections f and g of X, we let
$g\circ f$
be the partial bijection from
$ f^{-1}[\mathrm {Dom}(g)]$
to
$g[\mathrm {Im}(f)]$
defined by
$g\circ f(x)=g(f(x)) $
for all
$x\in f^{-1}[\mathrm {Dom}(g)]$
.
The following lemma is an easy exercise and we leave the details to the reader.
Lemma 4.7 Let f and g be partial bijections of X. Then

for all
$A,B\subseteq X$
.
Definition 4.8 Let X be a u.l.f. metric space. A subset
$A\subseteq X$
is thin if
$f[A]\cap A$
is finite for all partial translations f of X which do not fix points, that is, such that
$f(x)\neq x$
for all
$x\in \mathrm {Dom}(f)$
.
Lemma 4.9 Every infinite u.l.f. metric space contains an infinite thin subset.
Proof If
$(X,d)$
is infinite and u.l.f., then X is unbounded. Hence, we can inductively pick a sequence
$(x_i)_{i\in \mathbb {N}}$
in X such that

for all
$\ell>k$
in
$\mathbb {N}$
. The set
$A=\{x_i\mid i\in \mathbb {N}\}$
is clearly thin.Footnote
8
Proposition 4.10 Let X be a u.l.f. metric space,
$C\subseteq X$
be thin, and let
$C=A\sqcup B$
be a partition of C. If f and g are partial translations of X, then
$f[A]\cap g[B]$
is finite.
Proof By Lemma 4.7, it is enough to show that
$(g^{-1}\circ f)[A]\cap B$
is finite. As the composition of partial translations is still a partial translation, it is enough to show that
$ f[A]\cap B$
is finite for any partial translation f of X. Fix such f and, replacing A with
$A\cap \mathrm {Dom}(f)$
, we also assume that
$A\subseteq \mathrm {Dom}(f)$
. Let us show that
$f(A)\cap B$
is finite. Set

Then, as
$A\cap B=\emptyset $
, we have

Let
$f_1=f\restriction A_1$
. Then
$f_1$
has no fixed points and

Since C is thin,
$f_1[C]\cap C$
must be finite. So,
$f(A)\cap B$
is finite.
Given a u.l.f. metric space X, let
$\beta X$
denote the Stone–Čech compactification of X.Footnote
9
Since X is discrete,
$\beta X$
can be identified with the space of ultrafilters on X endowed with the Stone topology, that is, the topology generated by open sets of the form

for all
$A\subseteq X$
. Given
$A\subseteq X$
, we let
$\bar A$
denote the closure of A in
$\beta X$
and let
$\hat A=\bar A\setminus A$
. By the defining property of
$\beta X$
, any element in
$\ell _\infty (X)$
extends to one in
$C(\beta X)$
. This defines a canonical isomorphism between
$\ell _\infty (X)$
and
$C(\beta X)$
, and we identify those algebras under this isomorphism. We identify
$C(\hat X)$
with
$C(\beta X)/c_0(X)$
via Gelfand transform. Hence, under these identifications, we have

We now define invariant subsets of the Stone–Čech compactification. For that, recall that, by the defining property of
$\beta X$
, any partially defined map
$f\colon \mathrm {Dom}(f)\subseteq X\to \mathrm {Im}(f)\subseteq X$
can be continuously extended to a (necessarily surjective) map
$\overline {\mathrm {Dom}(f)}\to \overline {\mathrm {Im}(f)}$
. By abuse of notation, we still denote this map by f.
Definition 4.11 Let X be a u.l.f. metric space and
$A\subseteq \beta X$
. We say that A is invariant if
$f[A]\subseteq A$
for all partial translations f of X.
For the next lemma, notice that if
$L\subseteq \hat X$
is a clopen subset, then
$\chi _L\in C(\hat X)$
. Hence, it makes sense to wonder whether
$\chi _L$
can also be in
$C(\nu X)\subseteq C(\hat X)$
.
Lemma 4.12 Let X be a u.l.f. metric space and
$L\subseteq \hat X$
be an invariant clopen subset. Then
$\chi _L\in C(\nu X)$
.
Proof By Proposition 4.3, it is enough to notice that
$\chi _L$
is in the center of
$\mathrm {Q}^{*}_u(X)$
. Hence, since
$\mathrm {C}^{*}_u[X]$
is dense in
$\mathrm {C}^{*}_u(X)$
and spanned by
$av_f$
, where
$a\in \ell _\infty (X)$
and f is a partial translation of X, we only need to show that
$\chi _{L}$
commutes with
$w_f=\pi (v_f)$
for all partial translations f of X. Fix such partial translation f and let
$A=\mathrm {Dom}(f)$
and
$B=\mathrm {Im}(f)$
. Then,
$w_f= \chi _{\hat B} w_f\chi _{\hat A}$
and

notice that
$f[L]=\hat B\cap L$
. Indeed, since L is invariant and f is a partial translation,
$f[L]\subset \hat B\cap L$
. On the other hand, as
$f^{-1}$
is also a partial translation, we have
$f^{-1}[L]\subseteq L$
. Hence, as
$\hat B\cap L\subseteq f[f^{-1}[L] ]$
, we also have
$ \hat B\cap L\subseteq f[L]$
. We can then conclude from (4.1) that
$w_f\chi _L=\chi _Lw_f$
. As the partial translation f was arbitrary, we conclude that
$\chi _L\in C(\nu X)$
as desired.
Lemma 4.13 Let X be a u.l.f. metric space and
$C\subseteq X$
be thin. If
$\omega ,\omega '\in \hat C$
are distinct, then there are disjoint invariant open subsets
$U,V\subseteq \hat X$
such that
$\omega \in U$
and
$\omega '\in V$
.
Proof Since
$\omega ,\omega '\in \bar C$
, it follows that
$C\in \omega $
and
$C\in \omega '$
. As
$\omega \neq \omega '$
, there is
$D\subseteq X$
such that
$D\in \omega $
and
$D\not \in \omega '$
. Hence,

Therefore,
$\omega \in \hat A$
and
$\omega '\in \hat B$
. Let
$\mathcal {PT}$
denote the set of all partial translations of X and define

Clearly, U and V are open, invariant and contain
$\omega $
and
$\omega '$
, respectively. We only need to notice they are also disjoint. For that, notice that Proposition 4.10 implies that
$f[A]\cap g[B]$
is finite for all
$f,g\in \mathcal {PT}$
. But then
$\widehat {f[A]}\cap \widehat {g[B]}=\emptyset $
for all
$f,g\in \mathcal {PT}$
, which in turn implies that
$U\cap V=\emptyset $
.
Theorem 4.14 Let X be an infinite u.l.f. metric space. Then,
$\nu X$
has at least
$2^{2^{\aleph _0}}$
elements.
Proof Let
$p\colon \hat X\to \nu X$
be the continuous surjection such that the canonical identification of
$C(\nu X)$
with a
$\mathrm {C}^{*}$
-subalgebra of
$C(\hat X)$
is given by the map

Let
$C\subseteq X$
be an infinite thin subset given by Lemma 4.9. As
$\hat C$
is the set of all nonprincipal ultrafilters on C and C is countable, we have that
$|\hat C|=2^{2^{\aleph _0}}$
. Therefore, in order to obtain that
$\nu X$
has
$2^{2^{\aleph _0}}$
elements, it is enough to show that p is injective on
$\hat C$
.
Let
$\omega ,\omega '\in \hat C$
be distinct. By Lemma 4.13, there are disjoint invariant open subsets
$U,V\subseteq \hat X$
containing
$\omega $
and
$\omega '$
, respectively. As
$\beta X$
is extremely disconnected,
$\bar U$
is clopen in
$\hat X$
which implies that the characteristic function of
$\bar U$
,
$\chi _{\bar U}$
, is a continuous function in
$C(\hat X)$
. As
$\bar U$
is invariant, Lemma 4.12 shows that
$\chi _{\bar U}\in C(\nu X)$
. Therefore, since we clearly have
$\chi _{\bar U}(\omega )=1$
and
$\chi _{\bar U}(\omega ')=0$
, this shows that
$p(\omega )\neq p(\omega ')$
.
Remark 4.15 We would like to observe that Theorem 4.14 is only valid for metric u.l.f. spaces. Precisely, Higson coronas can be defined more generally for coarse spaces — for brevity, we do not define coarse spaces here, the reader can find the precise definition in [Reference RoeRoe03] or [Reference Baudier, Braga, Farah, Khukhro, Vignati and WillettBBF+22b, Section 5]. It is known that every perfectly normal compact Hausdorff space is homeomorphic to the Higson corona of some u.l.f. coarse space (see [Reference Banakh and ProtasovBP20, p. 2]). It is however not surprising that the Higson corona of nonmetrizable u.l.f. coarse spaces can be much smaller since there will be fewer Higson functions in this case. The proof of Theorem 4.14 cannot hold outside the metrizable world since thin sets may not exist. For instance, if
$\mathcal {E}_{\max }$
is the maximal u.l.f. coarse structure on an infinite set X (see [Reference Baudier, Braga, Farah, Khukhro, Vignati and WillettBBF+22b, Subsection 1.3] for the precise definition), then it is clear that
$(X,\mathcal {E}_{\max })$
has no infinite thin subsets.
Corollary 4.16 Let X be an infinite u.l.f. metric space,
$h\colon X\to \mathbb {R}$
be a coarse map, and
$\beta \in \mathbb {R}$
. If there is a
$(\sigma ^\infty _h,\beta )$
-KMS state on
$\mathrm {Q}^{*}_u(X)$
whose restriction to
$C(\nu X)$
is faithful, then there are
$2^{2^{\aleph _0}}$
extreme
$(\sigma ^\infty _h,\beta )$
-KMS states on
$\mathrm {Q}^{*}_u(X)$
. In particular, there are
$2^{2^{\aleph _0}}$
extreme
$(\sigma _h,\beta )$
-KMS states on
$\mathrm {C}^{*}_u(X)$
which vanish on
$\mathcal {K}(\ell _2(X))$
.
5 Applications: Branching trees
In this section, we apply the theory of KMS states on uniform Roe algebras developed above to n-branching trees. Recall that, as mentioned in the introduction, the choice for those spaces are, in a sense, very natural. Precisely, as explained in Section 1.4, as long as
$h\colon X\to \mathbb {R}$
is such that
$h(x)$
is bounded below by an affine map in terms of
$d(x,x_0)$
for a given
$x_0\in X$
, there will always be
$(\sigma _{h},\beta )$
-KMS states on
$\mathrm {C}^{*}_u(X)$
for all
$\beta>0$
as long as X has polynomial growth. Therefore, in order to find more interesting phase transitions, it is natural to look at metric spaces with exponential growth.
5.1 n-branching trees
Given a set
$\Gamma $
, we let
$\Gamma ^{<\infty }$
be the set of all finite words on
$\Gamma $
, including the empty word; which we denote by
$\emptyset $
. In other words, if
$\gamma \in \Gamma ^{<\infty }$
, then either
$\gamma =\emptyset $
or
$\gamma =(\gamma _1,\ldots , \gamma _n)$
for some
$n\in \mathbb {N}$
and some
$\gamma _1,\ldots , \gamma _n\in \Gamma $
. Given
$\gamma \in \Gamma ^{<\infty }$
, if
$\gamma =\emptyset $
, we say that the length of
$\gamma $
is
$0$
, if
$\gamma =(\gamma _1,\ldots , \gamma _n)$
, we say that the length of
$\gamma $
is n; either way, we denote the length of
$\gamma $
by
$|\gamma |$
and we write
$\gamma =(\gamma _1,\ldots , \gamma _{|\gamma |})$
(here it is understood that if
$|\gamma |=0$
, then
$\gamma =\emptyset $
). Given
$\gamma ,\gamma '\in \Gamma ^{<\infty }$
we denote the concatenation of
$\gamma $
and
$\gamma '$
by
$\gamma ^\smallfrown \gamma '$
, that is,

Definition 5.1 Let
$n\in \mathbb {N}$
and consider
$\Gamma =\{1,\ldots , n\}$
. We make
$\Gamma ^{<\infty }$
into a graph by saying that any two distinct elements
$\gamma ,\gamma '\in \Gamma ^{<\infty }$
are adjacent if there is
$k\in \Gamma $
such that either
$\gamma _1=\gamma _2^\smallfrown k$
or
$\gamma _2=\gamma _1^\smallfrown k$
. This defines a graph structure on
$\Gamma ^{<\infty }$
making it into a connected (undirected) graph. We can then see
$\Gamma ^{<\infty }$
as a metric space endowed with the shortest path distance. We call this metric space the n-branching tree and denote it by
$T_n$
.
For simplicity, we now isolate the setting of this subsection.
Assumption 5.2 Let
$n\in \mathbb {N}$
and let
$T_n$
be the n-branching tree endowed with the shortest path metric, denoted by d. Let
$h\colon T_n\to \mathbb {R}$
be the function given by
$h(x)=d(x,\emptyset )$
for all
$x\in T_n$
5.2 Strongly continuous KMS states on
$\mathrm {C}^{*}_u(T_n)$
We start with a simple lemma about states on
$\ell _\infty $
. In the next lemma,
$\ell _\infty =\ell _\infty (\mathbb {N})$
and
$c_0=c_0(\mathbb {N})$
.
Lemma 5.3 Let
$\varphi $
be a state on
$\ell _\infty $
. If
$\varphi \restriction c_0$
has norm 1, then
$\varphi $
is strongly continuous.
Proof For each j, denote by
$b_j=\varphi (\chi _{\{j\}})$
. Then, as
$\varphi \restriction c_0$
has norm 1,

Let a be a positive element in
$\ell _\infty $
with norm at most 1. Then, as
$\varphi $
is positive, we have that

So, upon taking the limit as
$k\to \infty $
, we get

Applying the same reasoning to
$1-a$
we deduce that

Hence,

Now, for an arbitrary
$a\in \ell _\infty $
, splitting a into its real and imaginary parts, and splitting each such parts into their positive and negative parts, the previous paragraph imply that
$\varphi ( a)= \sum _{j=1}^\infty a_jb_j$
, so the lemma follows.
The next result is a partial version of Theorem 1.12.
Theorem 5.4 In the setting of Assumption 5.2: Given
$\beta \in \mathbb {R}$
, there is a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(T_n)$
if and only if
$\beta \geq \log (n)$
. Moreover,
-
1. For
$\beta> \log (n)$ , there is a unique
$(\sigma _{h},\beta )$ -KMS state
$\varphi _\beta $ on
$\mathrm {C}^{*}_u(T_n)$ and
$\varphi _\beta $ is given by
$$\begin{align*}\varphi_\beta([a_{x,y}])=\sum_{y\in T_n} a_{y,y}\Big(e^{-\beta|y|}-ne^{-\beta (|y|+1)}\Big)\end{align*}$$
for all
$[a_{x,y}]\in \mathrm {C}^{*}_u(T_n)$ .
-
2. For
$\beta =\log (n)$ , the
$(\sigma _{h},\beta )$ -KMS states on
$\mathrm {C}^{*}_u(T_n)$ all vanish on
$\mathcal {K}(\ell _2(T_n))$ .
Proof Suppose
$\varphi $
is a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(T_n)$
. Notice that, for each
$y\in T_n$
, the map
$f\colon T_n\to T_n$
given by
$f(x)=x^{\smallfrown }y$
, for all
$x\in T_n$
, is a partial translation; indeed,
$d(x,f(x))=|y|$
for all
$x\in T_n$
. So, each
$v_f$
belongs to
$\mathrm {C}^{*}_u[T_n]$
. Then, for each
$y\in T_n$
, we have

For each
$y\in T_n$
, set

Hence, as
$\chi _{T_n}=v_f^{*}v_f$
and
$\chi _{T_n^\smallfrown y}=v_fv_f^{*}$
, we must have

for all
$y\in T_n$
; which implies

Since for each
$y\in T_n$
, we have

where “
$\bigsqcup $
” denotes disjoint union, (5.1) implies that

for all
$y\in T_n$
.
As
$\varphi $
is positive, each
$\varphi (e_{y,y})$
must be positive. So, (5.2) gives that

Solving for
$\beta $
, this implies
$\beta \geq \log (n)$
. Moreover, as (5.2) must hold regardless of
$\beta $
, this also shows that the
$(\sigma _{h},\log (n))$
-KMS states on
$\mathrm {C}^{*}_u(T_n)$
all vanish on
$c_0(T_n)$
. Since such states factors through
$\ell _\infty (T_n)$
(Theorem 1.5), (2) follows.
We must now show that if
$\beta \geq \log (n)$
, then
$(\sigma _{h},\beta )$
-KMS states exist. This will however be an immediate consequence of (1). Indeed, the set of all
$\beta $
’s for with
$(\sigma _{h},\beta )$
-KMS states exist is always a closed set (see [Reference Bratteli and RobinsonBR97, Proposition 5.3.25]).
We now show (1) holds. For this, suppose
$\beta>\log (n)$
and let us show that any given
$(\sigma _h,\beta )$
-KMS state
$\varphi $
must have the required form. Notice that
$\varphi \restriction \ell _\infty (T_n)$
is a state on
$\ell _\infty (T_n)$
. Moreover, the computations above show that

for all
$a=(a_y)_y\in c_0(T_n)$
. Hence, an easy computation gives

where
$\mathcal {F}$
is the net of all finite subsets of
$T_n$
ordered by reverse inclusion. Therefore, it follows that
$\|\varphi \restriction c_0(T_n)\|=1$
and, by Lemma 5.3,
$\varphi \restriction \ell _\infty $
is strongly continuous. This implies that (5.3) holds for all
$a=(a_y)_y\in \ell _\infty (T_n)$
. In order to notice that this holds for arbitrary elements of
$\mathrm {C}^{*}_u(T_n)$
, let
$E\colon \mathrm {C}^{*}_u(X)\to \ell _\infty (X)$
be the canonical conditional expectation and recall that, by Theorem 1.5, we have
$\varphi =\varphi \circ E$
. This proves the uniqueness part of (1).
We are left to notice that a
$\varphi $
given by the formula above is indeed a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(T_n)$
. This will be done by using Theorem 2.5.Footnote
10
So, let
$f\colon A\to f(A)$
be a partial translation on X. On one hand, we have that

On the other hand,

The change of variable
$y=f(x)$
give us

As
$\varphi =\varphi \circ E$
, Theorem 2.5 gives us that
$\varphi $
is a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(X)$
.
5.3 KMS states on
$\mathrm {C}^{*}_u(T_n)$
vanishing on compacts
In order to complete the proof of Theorem 1.12, we must further analyze the case
$\beta =\log (n)$
. According to Theorem 5.4, the KMS states for this inverse temperature will all vanish on the ideal of compact operators and we can then make use of the material of Section 4. Moreover, ideas in [Reference ChouCho69, Lemma 3] will also be extremely useful in order to compute to the precise cardinality of the set of extreme
$(\sigma _h,\beta )$
-KMS states on
$\mathrm {C}^{*}_u(T_n)$
.
5.3.1 Precise cardinality of the set of KMS states on
$\mathrm {C}^{*}_u(T_n)$
for
$\beta =\log (n)$
We start by setting up some notation. Given
$y\in T_n$
, consider the map

Then, by the defining property of
$\beta T_n$
,
$\tilde y$
can be extended to a continuous map
${\beta T_n\to \beta T_n}$
which, by abuse of notation, we still denote by
$\tilde y$
. Notice that

where the closures above are taken in
$\beta T_n$
(see Lemma [Reference ChouCho69, Lemma 2.1]). We call a subset
$A\subseteq \beta T_n$
right-invariant
Footnote
11
if

The following is a particular case of [Reference ChouCho69, Lemma 2 and Proposition 4.1], except that Chou prefers to work with left translations.Footnote 12
Lemma 5.5 Given
$n\in \mathbb {N}$
,
$\beta T_n$
contains at least
$2^{2^{\aleph _0}}$
nonempty, mutually disjoint, closed, invariant subsets.Footnote
13
Theorem 5.6 In the setting of Assumption 5.2: If
$\beta =\log (n)$
, then there are
$2^{2^{\aleph _0}}$
extreme
$(\sigma _h,\beta )$
-KMS states on
$\mathrm {C}^{*}_u(T_n)$
.
Before proving Theorem 5.6, let us isolate an easy lemma for further reference. Precisely, the next result is simply a more specialized version of Theorem 2.5.
Lemma 5.7 In the setting of Assumption 5.2: Suppose
$\varphi $
is a state on
$\ell _\infty (T_n)$
such that

for all
$y\in T_n$
and all
$A\subseteq T_n$
. Then,
$\varphi \circ E$
is a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(T_n)$
; where
$E\colon \mathrm {C}^{*}_u(T_n)\to \ell _\infty (T_n)$
is the canonical conditional expectation.
Proof Any partial translation f of
$T_n$
can be written as a disjoint union
$f=\bigcup _{i=1}^kf_i$
, where each
$f_i$
is a composition of partial isometries of the form

for some
$y\in T_n$
and
$A\subseteq T_n$
, and partial isometries of the form

for some
$y\in T_n$
and
$A\subseteq T_n$
. Therefore, by Theorem 2.5, it is enough to notice that (5.4) holds for partial isometries of the second kind. For that, fix
$y\in T_n$
and
$A\subseteq T_n$
, and let
$g\colon \tilde y( A)\to A$
be the partial translation given by
$g(\tilde y(x))=x$
for all
$x\in A$
. Then, since

our assumption on
$\varphi $
implies that

So, we are done.
Proof of Theorem 5.6.
We start establishing some convention. Firstly, recall that
$\ell _\infty (T_n)$
is canonically isomorphic to
$C(\beta T_n)$
. In order to keep track of notation, if
$a\in \ell _\infty (T_n)$
, we write
$\bar a$
to denote a as an element of
$C(\beta T_n)$
. Notice that, if
$a=\chi _A$
for some
$A\subseteq T_n$
, then

where the closure
$\bar A$
is taken in
$\beta T_n$
. Therefore, if
$\varphi $
is a state on
$\ell _\infty (T_n)$
, we can view it as a state on
$C(\beta T_n)$
, that is,
$\varphi $
is a Borel measure on
$\beta T_n$
and

With this in mind, we define the support of
$\varphi $
as the support of
$\varphi $
as a Borel measure on
$\beta T_n$
and denote it by
$\operatorname {\mathrm {supp}}(\varphi )\subseteq \beta T_n$
. Suppose now that
$\varphi $
is a state on
$\mathrm {C}^{*}_u(T_n)$
. Then,
$\varphi \restriction \ell _\infty (T_n)$
is a state on
$\ell _\infty (T_n)$
and, by abuse of notation, we write

We now start the proof. By Lemma 5.5, there is a family
$(L_j)_{j\in J}$
of nonempty, mutually disjoint, closed, invariant subsets of
$\beta T_n$
such that
$|J|=2^{2^{\aleph _0}}$
. Fix
$j\in J$
and, for simplicity, let
$L=L_j$
. Denote the subset of all
$(\sigma _h,\beta )$
-KMS states on
$\mathrm {C}^{*}_u(T_n)$
which vanish on the compacts by
$K_{\beta }$
and define

Clearly,
$K_{\beta }^L$
is convex and weak
$^{*}$
-compact. Let us show
$K_{\beta }^L$
is nonempty.
By Theorem 5.4,
$K_\beta \neq \emptyset $
. From now on, we fix
$\varphi \in K_{\beta }$
. As L is nonempty, fix also
$\omega \in L$
. We define a state
$\psi $
on
$\ell _\infty (T_n)$
as follows: for each
$a\in \ell _\infty (T_n)$
, let
$\tilde a\in \ell _\infty (T_n)$
be given by

We then let
$\psi $
be the state on
$\ell _\infty (T_n)$
given by

We extend
$\psi $
to the whole
$\mathrm {C}^{*}_u(T_n)$
in the usual way, that is, we let
$\psi =\psi \circ E$
where
$E\colon \mathrm {C}^{*}_u(T_n)\to \ell _\infty (T_n)$
is the canonical conditional expectation. Since it is immediate that
$\psi $
is indeed a state on
$\mathrm {C}^{*}_u(T_n)$
, we only need to show that
$\psi $
satisfies the required KMS condition and that
$\operatorname {\mathrm {supp}}(\psi )\subseteq L$
.
For the KMS conditions, let
$y\in T_n$
and
$A\subseteq T_n$
; so,
$\tilde y\restriction A\colon A\to \tilde y(A)$
is a partial translation on
$T_n$
. Notice that

for all
$x\in T_n$
. In order to understand
$ \chi _{ \tilde y(\bar A)}(\tilde x(\omega )) $
, notice that

and

Therefore, as
$\{x\in T_n\mid |x|<|y|\}$
is finite and as
$\varphi $
vanishes on compacts, letting

we have that
$\varphi (\chi _{\tilde y(B)})=\varphi (\chi _C)$
. By (5.5), we have
$\widetilde {\chi _{\tilde y(A)}}=\chi _C$
and our discussion gives

As
$h- h\circ \tilde y$
is bounded,
$( h- h\circ \tilde y)^\sim $
is well defined. Let
$(z_j)_j$
be a net of elements of
$T_n$
converging to
$\omega $
. Notice that

Therefore,

for all
$x\in T_n$
. By the definition of B, it is clear that
$\chi _B=\widetilde {\chi _A}$
. Therefore,

By (5.6) and (5.7), we conclude that

As
$y\in T_n$
and
$A\subseteq T_n$
were arbitrary, this shows that
$\psi $
is a
$(\sigma _h,\beta )$
-KMS state on
$\mathrm {C}^{*}_u(T_n)$
.
Let us notice that
$\operatorname {\mathrm {supp}}(\psi )\subseteq L$
. Suppose
$\omega '\not \in L$
. Then there is
$A\subseteq T_n $
such that
$\omega '\in \bar A$
and
$\bar A\cap L=\emptyset $
. As
$\omega \in L$
and L is invariant,
$\tilde x(\omega )\in L$
for all
$x\in T_n$
. Hence,

for all
$x\in T_n$
. Then, thinking of
$\psi $
as being defined on
$C(\beta T_n)$
as described above, we have that
$\psi (\chi _A)=0$
. This shows that
$\operatorname {\mathrm {supp}}(\psi )\subseteq L$
and we concluded our proof that
$K^L_\beta \neq \emptyset $
.
Since
$j\in J$
was arbitrary, we have that each
$K^{L_j}_\beta $
is convex, weak
$^{*}$
compact, and nonempty. Hence, Krein–Milman theorem implies that each of them contains extreme points. Since
$(L_j)_{j\in J}$
are disjoint, this implies that there are
$2^{2^{\aleph _0}}$
many extreme points and we are done.
5.3.2 Localization of KMS states on
$\mathrm {C}^{*}_u(T_n)$
for
$\beta =\log (n)$
.
We are left to notice that a version of Theorem 5.6 holds along every branch of
$T_n$
. For that, we must further analyze the Higson corona of
$T_n$
. More precisely, we must identify a
$\mathrm {C}^{*}$
-subalgebra of
$C(\nu T_n)$
which will help us to locate the KMS states on
$\mathrm {C}^{*}_u(T_n)$
for inverse temperature
$\beta =\log (n)$
better.
We first introduce some notation. Firstly, let
$[T_n]$
denote the branches of
$T_n$
, that is,

Given
$\bar x=(x_j)_{j=1}^\infty \in [T_n]$
and
$k\in \mathbb {N}$
, we let
$\bar x|k$
be the initial segment of
$\bar x$
with k letters, that is,

We now set

and endow
$\mathcal T_n$
with an appropriate topology. For that, we first extend the concatenation operation: for
$y\in T_n$
and
$\bar x\in [T_n]$
, we let

Given any
$y\in T_n$
, we let

that is,
$y^\smallfrown \mathcal T_n$
denotes the set of words, finite or not, which “start” with y. We define
$y^\smallfrown T_n$
and
$y^\smallfrown [T_n]$
analogously, that is,

We endow
$\mathcal T_n$
with the topology generated by

So,
$T_n$
is an open subset of
$\mathcal T_n$
and the inclusion

is a homeomorphic embedding with dense range. Moreover, it is easy to see that
$\mathcal T_n$
is a compact space. Hence,
$\mathcal T_n$
is a compactification of
$T_n$
.
As
$T_n$
is dense in
$\mathcal T_n$
, this allow us to see
$C(\mathcal T_n)$
as a
$\mathrm {C}^{*}$
-subalgebra of
$\ell _\infty (T_n)$
in a canonical way. Precisely, we identify
$C(\mathcal T_n)$
with the image of the following injective
$^{*}$
-homomorphism

Lemma 5.8 Let
$n\in \mathbb {N}$
and consider the n-branching tree
$T_n$
. Then:
-
1. For all
$y\in T_n$ , the projection
$\chi _{y^\smallfrown T_n}$ is a Higson function.
-
2. The Banach space
$$\begin{align*}C_n=\overline{\mathrm{span}}\{\chi_{y^\smallfrown T_n}\mid y\in T_n\}\end{align*}$$
is a
$\mathrm {C}^{*}$ -algebra contained in
$C_h(T_n)$ .
-
3. Under the identification of
$C(\mathcal T_n)$ with the
$\mathrm {C}^{*}$ -subalgebra of
$\ell _\infty (T_n)$ described above, we have
$C_n=C(\mathcal T_n)$ . In particular, the compactification
$\mathcal T_n$ is Higson compatible.
In particular, identifying
$C([T_n])=C(\mathcal T_n)/c_0(T_n)$
via Gelfand transfom, we have that
$C([T_n])\subseteq \mathrm {Q}^{*}_u(T_n)$
.
Proof (1) Fix
$y\in T_n$
. Let
$\varepsilon>0$
and
$R>0$
. Let

Then, if
$x,z\in T_n\setminus F$
and
$d(x,z)<R$
, we must have that either both x and z are in
$y^\smallfrown T_n$
, or neither of them are. In either case, we have

so
$\chi _{y^\smallfrown T_n}$
is a Higson function.
(2) It is evident that
$C_n$
is closed under the adjoint operator. So, we only need to show that
$C_n$
is also closed under product. If
$x,z\in T_n$
, we write
$x\leq z$
if
$|x|\leq |z|$
and
$x_i=z_i$
for all
$i\in \{1,\ldots , |x|\}$
. The fact that C is a
$\mathrm {C}^{*}$
-algebra follows from the straightforward fact that, for all
$x,z\in T_n$
, we have

So,
$C_n$
is closed under multiplication. The fact that
$C_n\subseteq C_h(T_n)$
follows from (1).
(3) We start noticing that

In particular, if
$\bar x\in [T_n]$
and
$a=\sum _{y\in T_n}a_y\chi _{y^\smallfrown T_n}$
is as above, the limit

exists. We can then define an
$^{*}$
-isomorphic embedding
$\Phi \colon C_n\to C(\mathcal T_n)$
by letting

It is straightforward to show that
$\Phi $
is indeed well-defined, that is,
$\Phi (a)$
is a continuous function on
$\mathcal T_n$
for all
$a\in C_n$
. Moreover, it is also clear
$\Phi $
is an injective
$^{*}$
-homomorphism and that

We are left to notice that the
$\Phi $
is subjective. For that, we show that the image of

under
$\Phi $
is dense in
$C(\mathcal T_n)$
. Fix
$f\in C(\mathcal T_n)$
and
$\varepsilon>0$
. As f is continuous and
$[T_n]$
is compact, we can pick
$y_1,\ldots , y_k\in T_n$
such that

and

By (5.8), there is a finite set
$F\subseteq T_n$
such that

For simplicity, assume
$F\cap y_j^\smallfrown T_n=\emptyset $
for all
$j\in \{1,\ldots ,k\}$
and let
$a\in \ell _\infty (X)$
be given by

It is straightforward to check that

and that
$\|\Phi (a)-f\|\leq \varepsilon $
.
The next couple of results will focus more on KMS states on
$\mathrm {Q}^{*}_u(T_n)$
and will not be necessary for the main result of this section per se (Theorem 1.12). The reader interested only in Theorem 1.12 can safely skip to Lemma 5.12.
Definition 5.9 In the setting of Assumption 5.2: For each
$\beta>\log (n)$
, let
$\varphi _\beta $
be the
$(\sigma _{h},\beta )$
-KMS state in Theorem 5.4. If
$(\beta _k)_k\subseteq (\log (n),\infty )$
is a sequence converging to
$\log (n)$
and
$ \mathcal {U}$
is a nonprincipal ultrafilter on
$\mathbb {N}$
, then

is a
$(\sigma _{h},\log (n))$
-KMS state on
$\mathrm {C}^{*}_u(T_n)$
. We call any such KMS states a limiting KMS state. By Theorem 5.4, those states always vanish on
$\mathcal {K}(\ell _2(T_n))$
.
Corollary 5.10 In the setting of Assumption 5.2: Let
$\beta =\log (n)$
and
$\varphi $
be a limiting
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(T_n)$
. Let
$\psi $
be the
$(\sigma _{h}^\infty ,\beta )$
-KMS on
$\mathrm {Q}^{*}_u(T_n)$
such that
${\varphi =\psi \circ \pi }$
. Then, the restriction of
$\psi $
to
$C([T_n])$
is faithful.
Proof Let
$\mu $
be the probability measure on
$[T_n]$
given by Riesz representation theorem by restricting
$\psi $
to
$C([T_n])$
, that is,

Since
$\varphi $
is a limiting
$(\sigma _{h},\log (n))$
-KMS state, let
$(\beta _k)_k\subseteq (\log (n),\infty )$
be a sequence converging to
$\log (n)$
and
$\mathcal {U}$
be a nonprincipal ultrafilter such that

By the formula of each
$\varphi _{\beta _k}$
given by Theorem 5.4, it follows that

Hence, by the formula of
$\varphi $
, we have

This shows that
$\mu $
is the Bernoulli measure on
$[T_n]=\{1,\ldots ,n\}^{\mathbb {N}}$
. Since the support of the Bernoulli measure is the whole
$[T_n]$
, this shows that
$\varphi $
is faithful on
$C([T_n])$
. This completes the proof.
Corollary 5.11 In the setting of Assumption 5.2: If
$\beta =\log (n)$
, then for all
$\bar x\in [T_n]$
there is an extreme
$(\sigma ^\infty _h,\beta )$
-KMS state
$\psi $
on
$\mathrm {Q}^{*}_u(T_n)$
such that

Moreover, if
$\varphi $
is a limiting
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(T_n)$
, then the
$(\sigma _{h}^\infty ,\beta )$
-KMS state
$\psi $
on
$\mathrm {Q}^{*}_u(T_n)$
determined by
$\varphi =\psi \circ \pi $
is not extreme.
Proof The first assertion follows from Theorem 4.4 and Corollary 5.10. For the second assertion, notice that if
$\psi =\varphi \circ \pi $
were extreme, then there would be
$\bar x\in [T_n]$
such that
$\psi $
vanishes on the ideal

(see Proposition 3.3). However, it was shown in the proof of Corollary 5.10 that
$\psi $
is faithful on
$C([T_n])$
; contradiction.
We now return to the proof of Theorem 1.12. The following lemma is trivial and we isolate it for further reference.
Lemma 5.12 Let
$n\in \mathbb {N}$
and
$T_n$
be the n-branching tree. Given any
$\bar x,\bar y\in [T_n]$
there is an isometry
$f\colon T_n\to T_n$
such that
$f(\bar x|k)=\bar y|k$
.
Given a metric space X and an isometry
$f\colon X\to X$
, we let
$u_f\colon \ell _2(X)\to \ell _2(X)$
be the (linear) isometry determined by

Lemma 5.13 Let
$n\in \mathbb {N}$
and
$T_n$
be the n-branching tree. Let
$f\colon T_n\to T_n$
be an isometry and consider the (linear) isometry
$u_f\colon \ell _2(T_n)\to \ell _2(T_n)$
defined above. Then, the map

is an affine isometry of the set of
$(\sigma _h,\beta )$
-KMS states on
$\mathrm {C}^{*}_u(T_n)$
to itself.
Proof It is enough to notice that
$\varphi \circ \mathrm {Ad}(u_f)$
is a
$(\sigma _h,\beta )$
-KMS state on
$\mathrm {C}^{*}_u(T_n)$
given that the same holds for
$\varphi $
. Indeed, once this is done the result follows since this map will clearly be an affine isometry with inverse
$\varphi \to \varphi \circ \mathrm {Ad}(u_f^{*})$
.
Notice that

Indeed, any isometry of the tree
$T_n$
must satisfy
$f(\emptyset )=\emptyset $
. Therefore, for each
$x\in T_n$
, we have

Using this, an immediate computation gives us that

for all
$t\in \mathbb {R}$
, all
$a\in \mathrm {C}^{*}_u(T_n)$
, and all
$x,y\in T_n$
. In other words, the flow
$\sigma _h$
is invariant under
$\mathrm {Ad}(u_f)$
. This shows that
$\varphi \circ \mathrm {Ad}(u_f)$
must be a
$(\sigma _h,\beta )$
-KMS state on
$\mathrm {C}^{*}_u(T_n)$
given that
$\varphi $
is one (equivalently, this could also be shown with the help of Theorem 2.5).
Theorem 5.14 In the setting of Assumption 5.2: If
$\beta =\log (n)$
, then for each
$\bar x\in [T_n]$
there are
$2^{2^{\aleph _0}}$
extreme
$(\sigma _h,\beta )$
-KMS states
$\varphi $
on
$\mathrm {C}^{*}_u(T_n)$
such that

Proof Fix
$\beta =\log (n)$
. By Theorem 5.6, there are
$2^{2^{\aleph _0}}$
extreme
$(\sigma _h,\beta )$
-KMS states on
$\mathrm {C}^{*}_u(T_n)$
. By Lemma 5.8,
$\mathcal T_n$
is a Higson compatible compactification of
$T_n$
. Therefore, Theorem 4.6 implies that for any extreme
$(\sigma _h,\beta )$
-KMS state
$\varphi $
on
$\mathrm {C}^{*}_u(T_n)$
, there is
$\bar x\in [T_n]$
such that

Therefore, since
$|[T_n]|=2^{\aleph _0}$
, a pigeonhole argument implies that there is at least one
$\bar x\in [T_n]$
for which there are
$2^{2^{\aleph _0}}$
extreme
$(\sigma _h,\beta )$
-KMS states on
$\mathrm {C}^{*}_u(T_n)$
satisfying (5.9) for
$\bar x$
. Fix such
$\bar x\in [T_n]$
.
Let now
$\bar y\in [T_n]$
be arbitrary and let
$f\colon T_n\to T_n$
be an isometry such that

(Lemma 5.12). Clearly, we must have that

Hence,
$\mathrm {Ad}(u_f)(\chi _{\bar y|k^\smallfrown T_n})=\chi _{\bar x|k^\smallfrown T_n}$
for all
$k\in \mathbb {N}$
and, if
$\varphi $
satisfies (5.9) for
$\bar x$
, it follows that

The result then follows from Lemma 5.13.
Proof of Theorem 1.12.
Theorem 5.4 gives that there is a
$(\sigma _h,\beta )$
-KMS state on
$\mathrm {C}^{*}_u(T_n)$
if and only if
$\beta \geq \log (n)$
. Moreover, item (1) and the first claim of item (2) of Theorem 1.12 also follow from Theorem 5.4.
We are left to notice that the second and third claim of Theorem 1.12(2) hold. From now on, let
$\beta =\log (n)$
. By Lemma 5.8,
$\mathcal T_n$
is a Higson compatible compactification of
$T_n$
. Therefore, Theorem 4.6 implies that any extreme
$(\sigma _h,\beta )$
-KMS state
$\varphi $
on
$\mathrm {C}^{*}_u(T_n)$
must have the required form, that is, there must be
$\bar x\in [T_n]$
such that

Finally, the fact that for each
$\bar x\in [T_n]$
, there are
$2^{2^{\aleph _0}}$
extreme
$(\sigma _h,\beta )$
-KMS states on
$\mathrm {C}^{*}_u(T_n)$
satisfying the above is now simply Theorem 5.14.
5.4 Obtaining distinct KMS states on
$\mathrm {C}^{*}_u(T_n)$
for
$\beta =\log (n)$
We finish the article presenting a more concrete way of obtaining distinct KMS states for inverse temperature
$\beta =\log (n)$
. Precisely, if
$(\beta _n)_n$
is a sequence converging to
$\log (n)$
from the right and
$(\varphi _{\beta _n})_n$
is a sequence of states such that each
$\varphi _{\beta _n}$
is a
$(\sigma _h,\beta _n)_n$
-KMS state on
$\mathrm {C}^{*}_u(T_n)$
, then
$w^{*}\text {-}\lim _{n,\mathcal {U}}\varphi _{\beta _n}$
is a
$(\sigma _h, \log (n))$
-KMS state, where
$\mathcal {U}$
is an arbitrary nonprincipal ultrafilter on
$\mathbb {N}$
. The next theorem shows that, picking different sequences
$(\beta _n)$
as above, this procedure may give us distinct
$(\sigma _h, \log (n))$
-KMS states. As mentioned at the end of Section 1.4, this kind of behavior is unusual (see [Reference van Enter and RuszelvER07]) and known as chaotic behavior of chaotic convergence of KMS states.
Theorem 5.15 In the setting of Assumption 5.2: different sequences
$(\beta _n)_n$
converging to
$\log (n)$
may converge to distinct
$(\sigma _h,\log (n))$
-KMS states on
$\mathrm {C}^{*}_u(T_n)$
.
Proof Let
$\beta =\log (n)$
. For each
$\beta '>\log (n)$
, let
$\varphi _{\beta '} $
be the
$(\sigma _{h},\beta ')$
-KMS state on
$\mathrm {C}^{*}_u(T_n)$
given by Theorem 5.4(1), that is,

for all
$[a_{x,y}]\in \mathrm {C}^{*}_u(T_n)$
. Given any nonprincipal ultrafilter
$\mathcal {U}$
on
$\mathbb {N}$
and any sequence
$(\beta _n)_n$
converging to
$\beta $
from the right, we know that
$w^{*}\text {-}\lim _{n,\mathcal {U}}\varphi _{\beta _n}$
is a
$(\sigma _{h},\beta )$
-KMS state on
$\mathrm {C}^{*}_u(T_n)$
. Our strategy will be to construct different sequences
$(\beta _n)_n$
as above which give us different
$(\sigma _{h},\beta )$
-KMS states on
$\mathrm {C}^{*}_u(T_n)$
. For that, some manipulations with the formula of
$\varphi _{\beta '}$
will be useful. Firstly, given
$E\subseteq T_n$
and
$k\in \mathbb {N}$
, write

and notice that
$|E_k|\leq n^k$
. Then, given an arbitrary
$\beta '>\beta $
, we have

Applying the change of variables
$\tau =ne^{-{\beta '}}$
and letting
$a_k=|E_k|/n^k$
for each
$k\geq 0$
, we have that each
$a_k$
is in
$[0,1]$
and

Moreover,
$\beta '\to \log (n)$
from the right if and only if
$\tau \to 1$
from the left. At last, notice that if E is such that there are
$p<q\in \mathbb {N}$
with

then

This finishes the manipulations in the formula of
$\varphi _{\beta '}$
that we will need.
We now construct increasing sequences
$(\tau _k)_k$
and
$(\theta _k)_k$
converging to
$1$
, and sequences
$(p_k)_k$
and
$(q_k)_k$
of natural numbers by induction for which the following holds:
-
•
$p_k<q_k<p_{k+1}-1$ for all
$k\in \mathbb {N}$ ,
-
•
$\tau _k^{p_k}-\tau _k^{q_k+1}>1/2$ for all
$k\in \mathbb {N}$ , and
-
•
$\theta _k^{p_m}-\theta _k^{q_m+1}< 2^{-m-2}$ for all
$k,m\in \mathbb {N}$ .
This can be easily done as follows: let
$k\geq 2$
and suppose
$(\tau _m)_{m=1}^{k-1}$
,
$(\theta _m)_{m=1}^{k-1}$
,
$(p_m)_{m=1}^{k-1}$
, and
$(q_m)_{m=1}^{k-1}$
where chosen appropriately; step
$1$
of the induction can clearly be done. Step k of the induction goes as follows. Pick
$p_k>q_{k-1}+1$
such that
$\theta _m^{p_k}<2^{-k-2}$
for all
$m\leq k-1$
. Then pick
$\tau _k\in (\tau _{k-1},1)$
such that
$\tau _k^{p_k}>3/4$
and
$q_k>p_k$
with
$\tau _k^{q_k+1}<1/4$
. Choose now
$\theta _k\in (\theta _{k-1},1)$
with
$\theta _k^{p_m}-\theta _k^{q_m+1}<2^{-m-2}$
for all
$m\leq k$
. This finishes the induction.
We now use the sequences constructed in the previous paragraph to finish the proof. Precisely, we show that if
$\mathcal {U}$
is a nonprincipal ultrafilter on
$\mathbb {N}$
, then

For this, let
$E\subseteq T_n$
be given by

Then
$|E_k|=n^k$
if
$k\in \bigcup _{m=1}^\infty [p_m,q_m]$
and
$|E_k|=0$
otherwise. Hence, letting
${a_k=|E_k|/n^k}$
as above, we have that

Therefore, using (5.10) above, we have

for all
$k\in \mathbb {N}$
. On the other hand,

for all
$k\in \mathbb {N}$
. Therefore, we conclude that

This finishes the proof.
Acknowledgments
B. M. Braga would like to thank Alcides Buss and the Universidade Federal of Santa Catarina (UFSC) for an invitation which led him to meet Ruy Exel and, as a consequence, made this project possible. R. Exel would like to express his thanks to IMPA for funding a two week visit to Rio de Janeiro during which a large part of this project was developed.