219

Characterization of Extremal KMS States on Groupoid C

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Characterization of Extremal KMS States on Groupoid C

Characterization of

Extremal KMS States on

Groupoid C*-Algebras

Rafael Pereira Lima

Dissertação apresentada

ao Instituto de Matemática e Estatística

da

Universidade de São Paulo

para

obtenção do título

de

Mestre em Ciências

Programa: Mestrado em Matemática Aplicada

Orientador: Prof. Dr. Rodrigo Bissacot

Durante o desenvolvimento deste trabalho o autor recebeu auxílio �nanceiro do CNPq

São Paulo, julho de 2019

Page 2: Characterization of Extremal KMS States on Groupoid C

Characterization of

Extremal KMS States on

Groupoid C*-Algebras

Esta dissertação contém as correções e alterações

sugeridas pela Comissão Julgadora durante a defesa

realizada por Rafael Pereira Lima em 1/7/2019.

O original encontra-se disponível no Instituto de

Matemática e Estatística da Universidade de São Paulo.

Comissão Julgadora:

• Prof. Dr. Cristian Ortiz (presidente) - IME-USP

• Prof. Dr. Alcides Buss - UFSC

• Prof. Dr. Alexandre Baraviera - UFRGS

Page 3: Characterization of Extremal KMS States on Groupoid C

Agradecimentos

Consegui escrever esta dissertação por causa da ajuda de algumas pessoas. Gostaria de

agradecer aos meus pais, minha irmã e minha madrinha pelo incentivo que sempre tive.

Sou grato pela ótima orientação do professor Rodrigo Bissacot, que me preparou para

a pesquisa e se preocupa com a carreira dos alunos, além de ser um exemplo para mim.

Aos colegas do grupo pela ajuda durante o desenvolvimento da dissertação, principalmente

Lucas, João, Rodrigo e Thiago, que revisaram o trabalho durante vários seminários. Em

particular, gostaria de agradecer ao Lucas, por me recomendar para o professor Rodrigo.

Agradeço ao professor Severino Toscano pela ajuda durante o mestrado. Especialmente,

sou muito grato pela ajuda e apoio do professor Paulo Cordaro, o que foi fundamental para

eu conseguir terminar o mestrado.

1

Page 4: Characterization of Extremal KMS States on Groupoid C

Resumo

Nesta dissertação de mestrado, estudamos um teorema de Neshveyev [17] que descreve todos

os estados KMS em uma C*-álgebra de um grupóide étale localmente compacto Hausdor�

satisfazendo o segundo axioma de enumerabilidade. Depois estudamos um resultado provado

por Thomsen [26] que caracteriza os estados KMS extremais nessa C*-álgebra para um

grupóide de Renault-Deaconu.

Palavras-chave: C*-álgebras, estados KMS, medidas conformes, grupóides.

2

Page 5: Characterization of Extremal KMS States on Groupoid C

Abstract

In this master's thesis we study a theorem due to Neshveyev [17] which describes all KMS

states on the groupoid C*-algebra for a locally compact Hausdor� second countable étale

groupoid. Then we study a result due to Thomsen [26] which characterizes the extremal

KMS states on this C*-algebra for a Renault-Deaconu groupoid.

Keyworkds: C*-algebras, KMS states, conformal measures, groupoids.

3

Page 6: Characterization of Extremal KMS States on Groupoid C

Contents

1 Introduction 6

2 Measure Theory 12

2.1 Radon-Nikodym Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.2 Pushforward Measure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2.3 Purely Atomic and Non-Atomic Measures . . . . . . . . . . . . . . . . . . . 17

2.4 Measures on Locally Compact Spaces . . . . . . . . . . . . . . . . . . . . . . 18

2.5 µ-Measurable Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

2.6 Vector-Valued Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3 Groupoids 41

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

3.2 Topological Groupoids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

3.3 Groupoid C*-Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

4 Renault's Disintegration Theorem 74

4.1 Haar Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

4.2 Borel Hilbert Bundles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

4.3 Renault's Disintegration Theorem . . . . . . . . . . . . . . . . . . . . . . . . 92

4

Page 7: Characterization of Extremal KMS States on Groupoid C

5 Neshveyev's Theorems 95

5.1 KMS States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

5.2 First Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

5.3 Second Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

6 Renault-Deaconu Groupoid 159

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159

6.2 Full orbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171

6.3 Conformal Measures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175

6.4 KMS States on the Renault-Deaconu Groupoid . . . . . . . . . . . . . . . . 201

7 Concluding Remarks 214

5

Page 8: Characterization of Extremal KMS States on Groupoid C

Chapter 1

Introduction

The purpose of this thesis is to �nd all KMS states on groupoid C*-algebras when the

groupoid satis�es certain topological conditions. This result was proved by Neshveyev in

[17]. Later we study a theorem due to Thomsen [26] which applies Neshveyev's theorem to

a Renault-Deaconu groupoid to characterize its extremal KMS states.

Groupoids are a generalization of groups where not every pair of elements can be multiplied

but each element has an inverse. This structure can be seen as a collection of arrows attached

to points on a plane, as shown in Figure 1.1. Such arrows can be composed if the end (called

range) of the �rst arrow is the source of the second. The inverse is obtained by reversing the

direction of the arrow and each point is identi�ed with an element of the groupoid assuming

its corresponding vector is the null vector.

s(g)

r(g)

s(h)

r(h)

gh

(a)

r(g)

s(g) = r(h)

s(h)g

h

gh

(b)

s(g)

r(g)

s(g)

r(g)

s(g) = g−1g

r(g) = gg−1

g g−1

(c)

Figure 1.1: Groupoids can be seen as arrows on a plane. s(g) and r(g) denote the sourceand range of g. (a) g and h are not composable, since s(g) 6= r(h); (b) The composition ofg and h is gh; (c) g−1 is the inverse of g. Note that g−1g = s(g) and gg−1 = r(g).

6

Page 9: Characterization of Extremal KMS States on Groupoid C

Given a groupoid G, G(2) is the set of composable elements. It consists of all pairs of

elements in G which can be multiplied. G(0) ⊂ G is the set of units. G is endowed with the

multiplication (also called composition) and inversion operations. r, s : G → G(0) are the

range and source maps. Later we will de�ne formally the notion of groupoids. The results

on groupoids in this thesis can be found in Rodrigo Frausino's thesis [9]. In fact, this thesis

can be seen as a sequel of his work because he also describes groupoid C*-algebras and the

Renault-Deaconu groupoid. In addition, many results here are based on his work.

Under certain conditions, we can equip the groupoid with a topology in such a way that

r, s are local homeomorphisms and the sets Gxx = s−1(x)∩ r−1(x) are discrete and countable

groups, and we assume this topology satis�es other conditions. In this case, we can equip

the space of continuous and compactly supported functions on G, denoted by Cc(G), with an

involution and a convolution which is not the pointwise multiplication. Then Cc(G) becomes

a ∗-algebra, not necessarily commutative.

In order to de�ne the groupoid C*-algebra C*(G), we equip Cc(G) with a norm which

depends on the ∗-representations of Cc(G). Then C*(G) is de�ned as the completion of

Cc(G) with respect to this norm.

Let c be a continuous R-valued 1-cocycle, that is, a continuous function c : G → R such

that c(g1g2) = c(g1) + c(g2) for (g1, g2) ∈ G(2). Then we �x a dynamics on C*(G) de�ned by

τt(f)(g) = eitc(g)f(g) for every f ∈ Cc(G), g ∈ G and t ∈ R. For f ∈ Cc(G), we can extend

the de�nition of τ to complex parameters, that is, τz(f) is well-de�ned. Given β ∈ R, we say

that a state ϕ on C*(G) is a KMS state if ϕ(f1τiβ(f2)) = ϕ(f2f1) for every f1, f2 ∈ Cc(G).

KMS states characterizes the equilibrium states in quantum statistical mechanics. A

theorem due to Neshveyev describes every KMS state ϕ on C*(G) by an explicit formula.

In fact, there is a correspondence between ϕ and a pair (µ, {ϕx}x∈G(0)) satisfying some

conditions, such that µ is a probability measure on G(0) and each ϕx is a state on C*(Gxx).

An important step in the proof of this theorem is the Renault's disintegration theorem [15],

which will be used to obtain {ϕx}x∈G(0) and µ when a KMS state ϕ on C*(G) is given.

In the �nal part of the thesis, we de�ne the Renault-Deaconu groupoid and prove Thom-

7

Page 10: Characterization of Extremal KMS States on Groupoid C

sen's theorem.

Let X be a locally compact, second countable, locally Hausdor� space. Given σ : X → X

a local homeomorphism, the Renault-Deaconu groupoid is de�ned by

G = {(x, k, y) : k = n−m,σn(x) = σm(y)},

with composition (x, k, y)(y, l, z) = (x, k + l, z) and inversion (x, k, y)−1 = (y,−k, x).

Although the de�nition of G is abstract, it is useful to have an intuition about this

structure. Note that the sequence {σn(x)}n∈N can be seen as a trajectory starting at x.

Given y ∈ X, (x, k, y) ∈ G means that the trajectories of x and y eventually meet. k can be

interpreted as the delay of one trajectory with respect to the other. Figure 1.2 shows this

idea.

y

x · σ(y) ·

σ(x) σ2(y)

σ2(x) = σ3(y)

Figure 1.2: If (x, k, y) ∈ G then the trajectories {σl(x)}l∈N and {σl(y)}l∈N eventually meet.k can be seen as the delay of one trajectory with respect to the other. In this �gure, k = −1,since σ2(x) = σ3(y).

Given a continuous function F : X → R, we can de�ne a continuous R-valued 1-cocycle

cF by

cF (x, k, y) =n−1∑j=0

F (σj(x))−m−1∑j=0

F (σj(y))

for n,m ∈ N such that k = n − m and σn(x) = σm(y). In fact, there exists a bijection

between R-valued 1-cocycles on G and continuous real-valued functions on X. Then we

de�ne the dynamics on C*(G) by τt(f)(g) = eitcF (g)f(g). We want to describe the KMS

8

Page 11: Characterization of Extremal KMS States on Groupoid C

states on C*(G) with respect to this dynamics.

Since extremal KMS states are su�cient to describe all KMS states on a C*-algebra,

Thomsen's theorem characterizes only the extremal KMS states on the full C*-algebra of

this groupoid. In this case, we show that the probability measures corresponding to the

KMS states are eβF -conformal measures on X.

The orbit O(x) of x denotes the set of points y ∈ X such that (x, k, y) ∈ G for some k.

There is a bijection between orbits in X and the set of extremal atomic eβF -conformal prob-

ability measures on X. Thomsen's theorem divides extremal KMS-states ϕ corresponding

to measures m in three cases:

• when m is continuous;

• when m purely atomic and corresponds to a periodic orbit;

• when m purely atomic and corresponds to an aperiodic orbit.

In each case the theorem gives a formula for ϕ.

This thesis is structured in the following way:

Chapter 2: we recall some concepts of measure theory. This chapter is important to

understand the properties of the measures corresponding to KMS states on grupoid C*-

algebras. We also de�ne the integral of vector-valued functions on a Banach space.

Chapter 3: we de�ne groupoids and topological groupoids. Then we de�ne the groupoid

C*-algebra and prove some properties of this C*-algebra.

Chapter 4: we de�ne concepts necessary to understand Renault's disintegration theorem

and we state this theorem. However, we do not prove this result.

Chapter 5: we de�ne KMS states on arbitrary C*-algebras and prove some properties.

Then we prove two theorems due to Neshveyev, used to describe KMS states on some

groupoid C*-algebras. We state these theorems below and we refer to them as Neshveyev's

�rst theorem and Neshveyev's second theorem, respectively.

9

Page 12: Characterization of Extremal KMS States on Groupoid C

Theorem. [17, Theorem 1.1] Let G be a locally compact Hausdor� second countable étale

groupoid. There is a one-to-one correspondence between states on C*(G) with centralizer

containing C0(G(0)) and pairs (µ, {ϕx}x) consisting of a probability measure µ on G(0) and

a µ-measurable �eld of states ϕx on C*(Gxx). Namely, the state corresponding to (µ, {ϕx}x)

is given by

ϕ(f) =

∫G(0)

∑g∈Gxx

f(g)ϕx(ug)dµ(x) for f ∈ Cc(G).

Theorem. [17, Theorem 1.3] Let G be a locally compact second countable Hausdor� étale

groupoid. Let c be a continuous R-valued 1-cocycle on G and τ be the dynamics on C*(G)

de�ned by τt(f)(g) = eitc(g)f(g) for f ∈ Cc(G), g ∈ G. Fix β ∈ R. Then there exists a one-

to-one correspondence between KMSβ-states on C*(G) and pairs (µ, {ϕx}x∈G(0)) consisting

of a probability measure µ on G(0) and a µ-measurable �eld of states ϕx on C*(Gxx) such

that:

(i) µ is quasi-invariant with Radon-Nikodym derivative e−βc;

(ii) ϕx(ug) = ϕr(h)(uhgh−1) for every g ∈ Gxx and h ∈ Gx, for µ-a.e. x; in particular, ϕx is

tracial for µ-a.e. x;

(iii) ϕx(ug) = 0 for all g ∈ Gxx \ c−1(0), for µ-a.e. x.

Chapter 6: we de�ne the Renault-Deaconu groupoid, describe some of its properties, then

we characterize the extremal KMS-states proving the following theorem due to Thomsen:

Theorem. [26, Theorem 2.2] Let β ∈ R \ {0}. Assume that the periodic points of σ are

countable. The extremal KMSβ-states for τ are

1. States φm, where m is an extremal and continuous (non-atomic) eβF -conformal Borel

probability measure on X;

10

Page 13: Characterization of Extremal KMS States on Groupoid C

2. The states φλx, where λ ∈ C, |λ| = 1 and x is periodic with minimum period p, such that

p−1∑j=0

F (σj(x)) = 0 and∞∑n=1

∑y∈Yn

exp

(−β

n−1∑j=0

F (σj(y))

)<∞; (1.1)

3. The states φmz where z is aperiodic and β-summable.

11

Page 14: Characterization of Extremal KMS States on Groupoid C

Chapter 2

Measure Theory

The main theorems in this thesis, described in Chapters 5 and 6, shows that there is a relation

between a KMS state on a particular groupoid C*-algebra and a probability measure on a

subset of this groupoid. In order to understand these theorems, we should recall some results

from measure theory. We also generalize the notion of integral to functions from a measurable

space to a Banach space.

2.1 Radon-Nikodym Theorem

The Radon-Nikodym theorem proves that, under certain conditions, two measures ν, µ are

related by a non-negative measurable function f , denoted the Radon-Nikodym derivative.

In this case, ν can be interpreted as the integral of f with respect to µ. The results in this

section can be found in [14].

De�nition 2.1.1. Let X be a measurable space and let µ, ν be measures on X. We say ν

is absolutely continuous with respect to µ if

µ(A) = 0 implies ν(A) = 0, A measurable.

We denote ν � µ.

12

Page 15: Characterization of Extremal KMS States on Groupoid C

Note that � de�nes a partial order on the set of measures on X (assuming the σ-algebra

is �xed.)

Theorem 2.1.2. (Radon-Nikodym Theorem) Let X be a measurable space and ν, µ be

σ-�nite measures on X. If ν � µ then there exists a measurable nonnegative function f on

X such that f is �nite µ-a.e. and

ν(A) =

∫A

fdµ, A ⊂ X measurable.

Moreover, ν is �nite if and only if f is integrable.

The function f in Theorem 2.1.2 is called the Radon-Nikodym derivative of ν with respect

to µ and is denoted by

f =dν

dµ. (2.1)

Although we write 2.1 as an equality, the function f is not unique. If there exists a function

g satisfying 2.1, then f = g µ-a.e. We assume equality since we can neglect values of f on a

null set.

Remark 2.1.3. If the measure space X is locally compact Hausdor�, the Radon-Nikodym

derivative is a local property. That is, if we want to �nd the Radon-Nikodym derivative

dνdµ

(x) on a neighborhood of a point x, it is su�cient to study the relation between ν, µ on

this neighborhood.

In fact, let U be an open neighborhood of x and assume there exists a measurable function

∆ on U such that

∫U

f(y)dν(y) =

∫U

f(y)∆(y)dµ(y),

13

Page 16: Characterization of Extremal KMS States on Groupoid C

for every f ∈ Cc(U). Then using the de�nition of dν/dµ, we have

∫U

f(y)dν

dµ(y)dµ(y) =

∫U

f(y)∆(y)dµ(y).

Since f is arbitrary, we have

dµ(y) = ∆(y), for µ-a.e. y ∈ U .

Example 2.1.4. Let µ be the Lebesgue measure on R. De�ne the measure ν on R by

ν([a, b]) = a3 − b3, for every closed interval [a, b]. Then

ν([a, b]) =

∫ b

a

3x2µ(x).

Then we have, by Remark 2.1.3,

dµ(x) = 3x2.

Now we state some results on the Radon-Nikodym derivative which will be used through-

out the thesis.

Proposition 2.1.5. Let µ, ν be σ-�nite measures on X such that ν � µ. Then, for every

integrable function g with respect to ν we have

∫X

gdν =

∫X

gdν

dµdµ.

Proposition 2.1.6. If µ, ν are σ-�nite measures on X such that ν � µ and dν/dµ 6= 0

µ-a.e., then µ� ν and

dν=

(dν

)−1

.

14

Page 17: Characterization of Extremal KMS States on Groupoid C

Proposition 2.1.7. (Chain rule) If µ, ν, η are measures on X satisfying η � ν � µ, then

dµ=dη

dµ.

2.2 Pushforward Measure

Given a measurable function T : X → Y between two measurable spaces, assume X is en-

dowed with a measure µ. Then we can de�ne a measure on Y , referred to as the pushforward

measure. This notion is de�ned in [24].

This notion will be used to prove Theorem 6.3.21 on page 194:

Theorem. Let β ∈ R. A measure µ on G(0) is eβF -conformal if, and only if, µ is quasi-

invariant with Radon-Nikodym derivative e−βcF .

This theorem shows that one of the hypothesis of Neshveyev's second theorem holds for

every eβF -conformal measure on the unit space of the Renault-Deaconu groupoid. This will

be used to prove Thomsen's theorem.

De�nition 2.2.1. Let X, Y be measurable spaces. Let µ be a measure on X. Given a

measurable function σ : X → Y , we de�ne the pushforward measure σ∗µ on Y by

∫Y

fd(σ∗µ) =

∫X

f ◦ σdµ. (2.2)

Lemma 2.2.2. Equation (2.2) is equivalent to

σ∗µ(A) = µ(σ−1(A)), for every A ⊂ Y measurable. (2.3)

Proof. Assume (2.2) holds. Let A ⊂ Y be measurable. Then χA is a measurable function

15

Page 18: Characterization of Extremal KMS States on Groupoid C

on Y . σ is measurable, then χA ◦ σ is measurable on X. Note that, for x ∈ A,

χA ◦ σ(x) =

1 if σ(x) ∈ A

0 otherwise=

1 if x ∈ σ−1(A)

0 otherwise= χσ−1(A).

Hence,

σ∗µ(A) =

∫Y

χA(y)d(σ∗µ) =

∫X

χA ◦ σ(x)dµ(x) =

∫X

χσ−1(A)dµ(x) = µ(σ−1(A)).

Conversely, suppose (2.3) holds. Let ϕ be a simple nonnegative measurable function on

Y . There exist a1, . . . , an ≥ 0, A1, . . . , An measurable on Y such that ϕ =∑n

i=1 aiχAi . Then∫Y

ϕdσ∗µ =n∑i=1

aiσ∗µ(Ai) =n∑i=1

aiµ(σ−1(Ai))

=n∑i=1

ai

∫X

χσ−1(Ai)(x)dµ(x)

=n∑i=1

ai

∫X

χAi ◦ σ(x)dµ(x)

=

∫X

ϕ ◦ σ(x)dµ(x).

Let f be a measurable function on Y . Assume f is nonnegative. Then there exists a sequence

of simple nonnegative functions bounded by f and converging to f . Hence (2.2) holds for f .

Therefore (2.2) holds for every measurable function on Y .

Lemma 2.2.3. Let µ be a measure on X. Let σ2 : X → Y , σ1 : Y → Z be measurable.

Then σ1∗σ2∗µ = (σ1 ◦ σ2)∗µ.

Proof. Let A ⊂ X be measurable. Then,

σ1∗σ2∗µ(A) = σ1∗(σ2∗µ)(A) = σ2∗µ(σ−11 (A))

= µ(σ−12 (σ−1

1 (A))) = µ((σ1 ◦ σ2)−1(A))

16

Page 19: Characterization of Extremal KMS States on Groupoid C

= (σ1 ◦ σ2)∗(A).

Lemma 2.2.4. Let µ, ν be measures on X such that ν � µ. Let σ : X → Y be a measurable

bijection such that σ−1 is measurable. Then σ∗ν � σ∗µ and

dσ∗ν

dσ∗µ(y) =

dµ(σ−1(y)) y ∈ Y .

Proof. Let f be a σ∗ν-integrable function on Y . Then

∫Y

f(y)dσ∗ν(y) =

∫X

f ◦ σ(x)dν(x)

=

∫X

f ◦ σ(x)dν

dµ(x)dµ(x)

=

∫Y

f(y)dν

dµ(σ−1(y))dσ∗µ(y)

Then σ∗ν � σ∗µ and

dσ∗ν

dσ∗µ(y) =

dµ(σ−1(y)) y ∈ Y .

2.3 Purely Atomic and Non-Atomic Measures

In this section we recall that every �nite Borel measure can be decomposed uniquely as a

sum of two measures, one being purely atomic and the other one being continuous. As a

consequence, an extremal probability measure is either purely atomic or continuous.

In order to prove Thomsen's theorem, we will show in Chapter 6 that every extremal KMS

17

Page 20: Characterization of Extremal KMS States on Groupoid C

state corresponds to an extremal probability m, then m has one of the properties de�ned

below. The results in this section can be found in [12] and [25].

De�nition 2.3.1. A �nite Borel measure m on the topological space X is non-atomic or

continuous when m({x}) = 0 for every x ∈ X and purely atomic if there is a Borel set

A ⊂ X such that m(A) = m(X) and m({a}) > 0 for all a ∈ A.

Given a measure µ, we write µ(x) = µ({x}).

Proposition 2.3.2. If µ is a σ-�nite measure on a σ-algebra then there exist unique measures

µa and µc such that µ = µa + µc and such that µa is purely atomic and µc is non-atomic.

2.4 Measures on Locally Compact Spaces

The results in this section are presented in [6]. Here we introduce the notion of Radon

measures and we conclude that, if the topological space satis�es certain conditions, every

probability measure is Radon.

First we prove some properties of Hausdor� spaces.

Proposition 2.4.1. Let X be a Hausdor� space, and let K and L be disjoint compact

subsets of X. Then there are disjoint open subsets U and V of X such that K ⊂ U and

L ⊂ V .

Proof. We can assume that K and L are both non-empty (otherwise we could use ∅ as one

of our open sets and X as the other). Let us begin with the case where K contains exactly

one point, say x. We show that there are open disjoint sets Ux, Vx such that x ∈ Ux and

L ⊂ Vx.

Since X is Hausdor�, for each y ∈ L there is a pair Uy, Vy of disjoint open sets such that

x ∈ Uy and y ∈ Vy. Since L is compact, there is a �nite family y1, . . . , yn such that the sets

Vy1 , . . . , Vyn cover L. The sets Ux and Vx de�ned by Ux = ∩ni=1Uyi , Vx = ∪ni=1Vyi are then the

required sets.

18

Page 21: Characterization of Extremal KMS States on Groupoid C

Next consider the case where K has more than one element. We have just shown that for

each x ∈ K there are disjoint open sets Ux and Vx such that x ∈ Ux and L ⊂ Vx. Since K

is compact, there is a �nite family x1, . . . , xn such that Ux1 , . . . , Uxn cover K. The proof is

complete if we de�ne U = ∪ni=1Uxi , V = ∩ni=1Vxi .

Proposition 2.4.2. Let X be a locally compact Hausdor� space, x a point in X, and U an

open neighborhood of x. Then x has an open neighborhood whose closure is compact and

included in U .

Proof. Since X is locally compact, there is an open neighborhood of x, sayW , whose closure

is compact. By replacing W with W ∩U , we assume that W is included in U . The di�culty

is that W may extend outside U .

Use Proposition 2.4.1 to choose disjoint open sets V1 and V2 that separate the compact

sets {x} andW \W . Note that the closure of V1∩W is included inW . In fact, suppose there

exists y ∈ V1 ∩W such that y /∈ W . Then y ∈ W \W . By de�nition, V2 is a neighborhood

of y. Since y ∈ V1 ∩W , there exists x1 ∈ V1 ∩ W such that x1 ∈ V2. This leads to a

contradiction because V1 ∩ V2 = ∅.

Then V1 ∩W is compact and included in W , and hence in U ; thus V1 ∩W is the required

open neighborhood of x.

A subset of a topological space X is a Gδ if it is the intersection of a sequence of open

subsets of X, and Fσ if it is the union of a sequence of closed subsets of X.

Proposition 2.4.3. Let X be a locally compact Hausdor� space, let K be a compact subset

of X, and let U be an open subset of X that includes K. Then there is an open set V of X

that has a compact closure and satis�es K ⊂ V ⊂ V ⊂ U .

Proof. Proposition 2.4.2 implies that each point in K has an open neighborhood whose

closure is compact and included in U . Since K is compact, some �nite collection of these

neighborhoods covers K. Let V be the union of these sets in such a �nite collection; then V

is the required set.

19

Page 22: Characterization of Extremal KMS States on Groupoid C

Proposition 2.4.4. Let X be a locally compact Hausdor� second countable space. Then

each open subset of X is an Fσ, and is in fact the union of a sequence of compact sets.

Likewise, each closed subset is a Gδ.

Proof. Suppose that U is a countable basis for the topology of X. Let U be an open set in

X. Given x ∈ U , it follows from Proposition 2.4.2 that there exists an open neighborhood

Wx of x such that Wx is compact and Wx ⊂ U . Since U is the basis for the topology of X,

there exists Vx ∈ U such that x ∈ Vx ⊂ Wx. Then Vx is compact and Vx ⊂ U . Thus,

U =⋃x∈U

Vx.

Since each Vx ∈ U and U is countable, then U is a countable union of compact sets. Therefore

U is Fσ.

Let A ⊂ X be a closed set. Then Ac is open, and Ac is the union of a sequence {Fn}

consisting of closed sets. Hence,

A = (Ac)c =

(∞⋃n=1

Fn

)c

=∞⋂n=1

F cn.

Therefore A is Gδ.

Lemma 2.4.5. Let X be a locally compact Hausdor� second countable space. Given an

open subset U ⊂ X, there exists a sequence {Kn} of compact subsets such that Kn ⊂ Kn+1

for every n, and U = ∪∞n=1Kn.

Proof. Let U ⊂ X be an open set. It follows from Proposition 2.4.4 that there is a sequence

of compact sets {Fn} such that U = ∪nFn. De�ne, for each n ≥ 1, Kn = ∪ni=1Fi. Clearly

each Kn is compact and Kn ⊂ Kn+1. Moreover, U = ∪∞n=1Kn.

Let X be a Hausdor� topological space. Then B(X), the Borel σ-algebra on X, is the

σ-algebra generated by the open subsets of X; the Borel subsets of X are those that belong

to B(X).

20

Page 23: Characterization of Extremal KMS States on Groupoid C

We turn to terminology for measures. Let X be a Hausdor� topological space. A Borel

measure on X is a measure whose domain is B(X). Suppose that A is a σ-algebra on X

such that B(X) ⊂ A. A positive measure µ on A is Radon if

(a) each compact subset K of X satis�es µ(K) <∞,

(b) each set A in A satis�es

µ(A) = inf{µ(U) : A ⊂ U and U is open}, and

(c) each open set U of X satis�es

µ(U) = sup{µ(K) : K ⊂ U and K is compact}.

A Radon Borel measure on X is a Radon measure whose domain is B(X). A measure that

satis�es condition (b) is often called outer regular , and a measure that satis�es condition

(c), inner regular .

Now we de�ne the support of a Radon Borel measure. The following theorem is necessary

to show that the support is well-de�ned.

Proposition 2.4.6. Let X be a locally compact Hausdor� space, let µ be a Radon Borel

measure on X. Then the union of all open subsets of X that have measure zero under µ is

itself an open set that has measure zero under µ.

Proof. Let U be the collection of all open subsets of X that have measure zero under µ, and

let U be the union of the sets in U . Then U is open. If K is a compact subset of U , then K

can be covered by a �nite collection U1, . . . , Un of sets that belong to U , and so we have

µ(K) ≤n∑i=1

µ(Ui) = 0.

Since K is arbitrary, it follows from the de�nition of Radon measure that µ(U) = 0.

21

Page 24: Characterization of Extremal KMS States on Groupoid C

It follows from Proposition 2.4.6 that, for X, µ, there is the largest open subset U ⊂ X

with µ(U) = 0. Then we de�ne the support as follows.

De�nition 2.4.7. LetX be a locally compact Hausdor� space, and µ a Radon Borel measure

on X. We de�ne the support of µ as the complement of the largest open subset of X with

measure zero. We denote the support of X by supp(µ).

Note that supp(µ) is closed. Now we prove some properties of Radon measures.

Lemma 2.4.8. Let X be a Hausdor� space in which each open set is an Fσ, and let µ be a

�nite Borel measure on X. Then each Borel subset A of X satis�es

µ(A) = inf{µ(U) : A ⊂ U and U is open}, (2.4)

µ(A) = sup{µ(F ) : F ⊂ A and F is closed}. (2.5)

In particular, µ is Radon.

Proof. Let R denote the set of Borel sets A ⊂ X that satisfy conditions (2.4) and (2.5). We

prove that R contains all open subsets of X. Let U ⊂ X open. Clearly U satis�es (2.4). By

hypothesis there exists a sequence {Fn} of closed sets such that U = ∪nFn. We can assume

that Fn ⊂ Fn+1 for each n without loss of generality. Then µ(U) = limn µ(Fn). Therefore

(2.5) holds.

Now we show that conditions (2.4) and (2.5) hold for an arbitrary Borel set A if, and only

if, for every ε > 0 there are U open, F closed, such that

F ⊂ A ⊂ U and µ(U \ F ) < ε. (2.6)

In fact, assume (2.4) and (2.5) hold. Let A be measurable. Given ε > 0, by (2.4) there exists

U open such that A ⊂ U and

µ(U) < µ(A) + ε/2. (2.7)

22

Page 25: Characterization of Extremal KMS States on Groupoid C

Aplying (2.5), there exists F ⊂ A closed satisfying

µ(F ) > µ(A)− ε/2. (2.8)

Then, by (2.7) and (2.8), we have

µ(U \ F ) = µ(U)− µ(F ) <(µ(A) +

ε

2

)−(µ(A)− ε

2

)= ε.

Then (2.6) holds.

Conversely, assume (2.6) holds. Given ε > 0, there are U open, F closed, such that

F ⊂ A ⊂ U and µ(U \ F ) < ε. Hence,

µ(A) ≤ µ(U) = µ(A) + µ(U \ A)

≤ µ(A) + µ(U \ F ) , since F ⊂ A,

< µ(A) + ε,

and

µ(F ) ≤ µ(A) = µ(F ) + µ(A \ F )

≤ µ(F ) + µ(U \ F ) , since A ⊂ U ,

< µ(F ) + ε.

Since ε > 0 is arbitrary, it follows that conditions (2.4) and (2.5) are satis�ed.

We can now show that R is a σ-algebra. Clearly ∅ ∈ R, since ∅ is open. Given A ∈ R,

ε > 0, there are U open, F closed such that F ⊂ A ⊂ U and µ(U \ F ) < ε. Then

F c is open, U c is closed, and U c ⊂ Ac ⊂ F c. Since F c \ U c = U \ F , it follows that

µ(F c \ U c) = µ(U \ F ) < ε. Therefore Ac ∈ R.

Let {An} be a sequence of sets in R. Then, for every n ≥ 1, there exists Un open, Fn

23

Page 26: Characterization of Extremal KMS States on Groupoid C

closed such that

Fn ⊂ An ⊂ Un and µ(Un \ Fn) <ε

2n+1.

Let U = ∪nUn and F = ∪nFn. Then U , F satisfy the relations F ⊂ ∪nAn ⊂ U and

µ(U \ F ) ≤ µ

(∞⋃n=1

(Un \ Fn)

)≤

∞∑n=1

µ(Un \ Fn) <∞∑n=1

ε

2n+1=ε

2. (2.9)

The set U is open, but the set F can fail to be closed. However for each N the set ∪Nn=1Fn

is closed, and since

µ(U \ F ) = µ(U)− µ(F ) = µ(U)− limN→∞

µ

(N⋃n=1

Fn

),

we can choose N such that

µ

(U \

N⋃n=1

Fn

)< ε.

Thus U and ∪Nn=1Fk are the sets required in (2.6), and R is closed under countable unions.

We have now shown that R is a σ-algebra on X that contains the open sets. Since B(X) is

the smallest σ-algebra on X that contains the open sets, it follows that B(X) ⊂ R. Therefore

this lemma is proved.

Remark 2.4.9. Let X be a locally compact Hausdor� second countable space, then every

probability measure on X is Radon. In fact, it follows from Proposition 2.4.4 that every

open set is Gδ. Then, given a probability measure µ on X, it follows from Lemma 2.4.8 that

µ is Radon.

We assume in Lemma 2.4.8 that the measure µ is �nite, but this result can be generalized

to σ-�nite measures that are �nite on compact sets.

24

Page 27: Characterization of Extremal KMS States on Groupoid C

Proposition 2.4.10. Let X be a locally compact Hausdor� space that has a countable

basis, and let µ be a Borel measure on X that is �nite on compact sets. Then µ is Radon.

Proof. First consider the inner regularity of µ. Let U be an open subset of X. Lemma 2.4.5

implies that U is the union of a sequence {Kj} of compact subsets, then

µ(U) = limn→∞

µ

(n⋃j=1

Kj

).

The inner regularity follows.

Let {Un} be a sequence of open sets such that X = ∪nUn and such that µ(Un) <∞ holds

for each n (for instance, take a countable basis U for X, and arrange in a sequence those

sets U in U for which U is compact).

For each n de�ne a Borel measure µn on X by µn(A) = µ(A ∩ Un). The measures µn are

�nite, and so Lemma 2.4.8 implies that they are outer regular. Hence if A belongs to B(X)

and if ε is a positive number, then for each n there is an open set Vn that includes A and

satis�es µn(Vn) < µn(A) + ε/2n. Consequently,

µ((Un ∩ Vn) \ A) < ε/2n.

Then set V de�ned by V = ∪n(Un ∩ Vn) is open, includes A and satis�es

µ(V \ A) ≤∑n

µ((Un ∩ Vn) \ A) < ε.

Hence µ(V ) ≤ µ(A) + ε, and the outer regularity of µ follows.

Assume X is a locally compact second countable Hausdor� space. By de�nition, a Radon

measure µ on X is �nite on compact subsets of X. It follows from Proposition 2.4.10 that

a Borel measure on X is Radon if, and only if, it is �nite on the compact subsets of X.

Proposition 2.4.11. [6, Proposition 7.2.6] Let X be a Hausdor� space, let A be a σ-algebra

25

Page 28: Characterization of Extremal KMS States on Groupoid C

on X that includes B(X), and let µ be a Radon measure on A. If A belongs to A and is

σ-�nite under µ, then

µ(A) = sup{µ(K) : K ⊂ A and K is compact }. (2.10)

Remark 2.4.12. Let X be a locally compact second countable Hausdor� space, and µ a

Borel measure which is �nite on compact subsets of X. It follows from Propositions 2.4.10

and 2.4.11 that, for every A ⊂ X Borel, we have

µ(A) = inf{µ(U) : A ⊂ U and U is open},

µ(A) = sup{µ(K) : K ⊂ A and K is compact}.

Lemma 2.4.13. Let X be a locally compact Hausdor� second countable space, µ be a

Radon Borel measure on X, B ⊂ X a Borel set, U ⊂ X an open set satisfying B ⊂ U .

Given a continuous non-negative function f on X, we have

∫B

f(x)dµ(x) = infB⊂V⊂UV open

∫V

f(x)dµ(x).

Proof. Since f is continuous and non-negative, we can de�ne the Borel measure ν on X by

ν(A) =

∫A

f(x)dµ(x), A Borel set.

The function f is continuous and µ is �nite on compact subsets, then ν is �nite on compact

subsets of X. Hence, from Proposition 2.4.10, ν is Radon. Therefore, for every B ⊂ X

Borel,

∫B

f(x)dµ(x) = ν(B) = infB⊂VV open

ν(V ) = infB⊂VV open

∫V

f(x)dµ(x).

For every open set V such that B ⊂ V , it follows that µ(B) ≤ µ(V ∩ U) ≤ µ(U) and

26

Page 29: Characterization of Extremal KMS States on Groupoid C

V ∩U is open. Hence, we can take the in�mum over the open sets V such that B ⊂ V ⊂ U .

Therefore,

∫B

f(x)dµ(x) = ν(B) = infB⊂V⊂UV open

ν(V ) = infB⊂V⊂UV open

∫V

f(x)dµ(x).

Lemma 2.4.14. Let X be a locally compact Hausdor� second countable space and µ a

Radon measure on X. Given an open subset U ⊂ X, we have

µ(U) = supf∈Cc(X)0≤f≤χU

∫X

f(x)dµ(x).

Proof. • Assume µ(U) <∞.

Note that for every f ∈ Cc(X) such that 0 ≤ f ≤ χU , we have∫X

f(x)dµ(x) ≤ µ(U).

Given ε > 0, there is a compact set K ⊂ U satisfying µ(U \K) < ε by Remark 2.4.12.

Since X is locally compact Hausdor�, there is an open set V such that V is compact

and K ⊂ V ⊂ V ⊂ U by Proposition 2.4.3.

By Urysohn's lemma, there exists a continuous function f assuming values in the

interval [0, 1] such that f equals one on K and vanishes outside V ⊂ U . Then f ∈

Cc(X) and 0 ≤ f ≤ χU .

Using the fact that µ(U)− µ(K) ≤ ε, we have µ(K) ≥ µ(U)− ε. Hence,

∫X

f(x)dµ(x) = µ(K) +

∫U\K

f(x)dµ(x)

≥ µ(K)

≥ µ(U)− ε.

27

Page 30: Characterization of Extremal KMS States on Groupoid C

Since ε is arbitrary, we have

µ(U) = supf∈Cc(X)0≤f≤χU

∫X

f(x)dµ(x).

• Suppose µ(U) =∞.

Let n be a natural number. By Remark 2.4.12, there exists a compact set Kn ⊂ U

such that µ(Kn) ≥ n. From Urysohn's lemma, we can choose a continuous compactly

supported function fn assuming values in the interval [0, 1] such that fn(x) = 1 for

every x ∈ Kn and fn vanishes of U . Hence,

∫X

fn(x)dµ(x) ≥∫Kn

fn(x)dµ(x) = µ(Kn) ≥ n.

Therefore,

supf∈Cc(X)0≤f≤χU

∫X

f(x)dµ(x) ≥ supn∈N

∫X

fn(x)dµ(x) =∞.

Hence the result follows.

2.5 µ-Measurable Functions

In this section we de�ne the µ-completion of a σ-algebra. The de�nition here can be found

in [6]. This σ-algebra will be necessary to understand one of the conditions in Neshveyev's

�rst theorem.

De�nition 2.5.1. Let (X,A) be a measurable space and let µ be a measure on A. The

completion of A under µ is the collection Aµ of subsets A of X for which there are sets E

and F in A such that

28

Page 31: Characterization of Extremal KMS States on Groupoid C

E ⊂ A ⊂ F and µ(F \ E) = 0.

A set that belongs to Aµ is said to be µ-measurable.

In fact, Aµ is a σ-algebra on X. We say that a function f on X is µ-measurable if it is

measurable with respect to the σ-algebra Aµ. Note that if f is measurable with respect to

the σ-algebra A, then f is µ-measurable.

Lemma 2.5.2. LetX be a topological space and µ a Borel purely atomic probability measure

on X. Every complex-valued function on X is µ-measurable.

Proof. Let f be a complex valued function on X. Let I be the set of points x ∈ X such that

µ({x}) > 0, then I is countable µ(X \ I) = 0. Given V ⊂ C measurable, let A = f−1(V )

and J = I \ A. Then I, J are measurable since both are countable.

Note that I ∩ A ⊂ A ⊂ X \ J . Since (X \ J) \ I ⊂ X \ I, it follows that

µ((X \ J) \ I) ≤ µ(X \ I) = 0.

Then A is µ-measurable and, therefore, f is µ-measurable.

2.6 Vector-Valued Integration

Now we introduce the concept of vector-valued integral, that is, the integral of functions

f : R → B where B is a complex Banach space. This section is based on [21]. We will

need this notion to prove Proposition 5.1.19 on page 105, and then de�ne KMS states on a

arbitrary C*-algebra.

Recall that one of the main steps in the construction of the Lebesgue integral is the notion

of simple functions. A function ϕ : R→ R is simple if there are A1, · · · , An measurable sets

29

Page 32: Characterization of Extremal KMS States on Groupoid C

and a1, . . . , an real numbers such that

ϕ =n∑i=1

aiχAi .

Then we de�ne its integral by

∫Rϕ(t)dµ(t) =

n∑i=1

aiµ(Ai).

Then, under certain conditions, the integral of a measurable function can be approximated

by the integral of simple functions. We will try to de�ne the integral of vector-valued

functions similarly.

De�nition 2.6.1. Let µ be a Borel measure on R and B a Banach space. A function

ϕ : R→ B is simple if there are a1, . . . , an ∈ B, and A1, . . . , An Borel subsets of B such that

ϕ =n∑i=1

aiχAi , (2.11)

and each χAi : R→ R is de�ned by

χAi(x) =

1, if x ∈ Ai,

0, if x /∈ Ai,

for x ∈ R. We call (2.11) a representation of ϕ.

In the next example, we show that the representation (2.11) is not necessarily unique.

Example 2.6.2. Let ϕ : R→ R3 be de�ned by

ϕ(x) =

(1, 1, 1), if 0 < x ≤ 1,

(1, 2, 2), if 1 < x ≤ 2,

(1, 2, 3), if 2 < x ≤ 3.

30

Page 33: Characterization of Extremal KMS States on Groupoid C

Then,

ϕ = (1, 1, 1)χ(0,1] + (1, 2, 2)χ(1,2] + (1, 2, 3)χ(2,3]

= (1, 1, 1)χ(0,3] + (0, 1, 1)χ(1,3] + (0, 0, 1)χ(2,3].

Therefore ϕ is simple and can be written in at least two di�erent representations.

De�nition 2.6.3. Let µ be a Borel measure on R and B a Banach space. Given a simple

function ϕ with representation (2.11), we de�ne its integral by

∫Rϕ(x)dµ(x) =

n∑i=1

aiµ(Ai).

Lemma 2.6.4. The integral in De�nition (2.6.3) is well-de�ned, that is, it does not depend

on the representation.

Proof. Let B be a Banach space. Given a simple function ϕ : R → B, let a1, . . . , an ∈ B,

b1, . . . , bm ∈ B, and let A1, . . . , An, B1, . . . , Bm be Borel sets such that

ϕ =n∑i=1

aiχAi =m∑j=1

biχAi .

Let

x =n∑i=1

aiµ(Ai) and y =m∑j=1

bjµ(Bj).

Note that x, y ∈ B. Choose an arbitrary Λ ∈ B*. Then Λ ◦ ϕ : R→ C is a simple function

with

Λ ◦ ϕ =n∑i=1

Λ(ai)χAi =m∑j=1

Λ(bi)χAi .

Since the integral of complex-valued functions does not depend on the representation, we

31

Page 34: Characterization of Extremal KMS States on Groupoid C

have

∫R(Λ ◦ ϕ)(t)dµ(t) =

n∑i=1

Λ(ai)µAi =m∑j=1

Λ(bi)µBi

= Λ

(n∑i=1

aiµAi

)= Λ

(m∑j=1

bjµBj

)

= Λ(x) = Λ(y)

Since Λ is arbitrary and B* separates points in B, it follows that x = y.

Remark 2.6.5. Given a simple function ϕ : R → B, Λ ∈ B*, Λ ◦ ϕ : R → C is a simple

function. Note that we used the property

Λ

(∫Rϕ(t)dµ(t)

)=

∫R

Λ(ϕ(t))dµ(t)

in Lemma 2.6.4 to show that the integral is well-de�ned. Similarly, we will de�ne the integral

in such a way that this property holds when we replace ϕ by a Borel function f : R → B.

Given Λ ∈ B*, we denote Λf = Λ ◦ f . Note that both Λf and Λϕ are measurable functions.

De�nition 2.6.6. Given a Banach space B, a function f : R → B is weakly measurable if

Λf is measurable for every Λ ∈ X*.

Remark 2.6.7. Note that every Borel function f : R → B is weakly measurable. In

particular, every continuous function from R to B is weakly measurable.

De�nition 2.6.8. Let µ be a Borel measure on R. Given a Banach space B, let f : R→ B

be weakly measurable. If there exists y ∈ B such that for every Λ ∈ B*,

Λy =

∫R

Λf(t)dµ(t),

then we de�ne the integral of f by

∫Rf(t)dµ(t) = y. (2.12)

32

Page 35: Characterization of Extremal KMS States on Groupoid C

Remark 2.6.9. Note that there exists at most one y such that (2.12) holds. This follows

from the fact that B* separates points in B.

De�nition 2.6.10. Given a Banach space B, Cc(R, B) denotes the space of compactly

supported functions f : R → B which are continuous. Recall that the support of f is the

closure of the set {t ∈ R : f(t) 6= 0}.

Note that every function f ∈ Cc(R, B) is weakly measurable.

Given a Banach space B, we de�ne the norm on Cc(R, B) by

‖f‖∞ = supt∈R‖f(t)‖.

Lemma 2.6.11. Let µ be a Borel measure on R, B a Banach space and f ∈ Cc(R, B) such

that there exists y =∫R f(t)dµ(t). Then

‖y‖ ≤∫R‖f(t)‖dµ(t).

Proof. Since B is a Banach space, we have

‖y‖ = supΛ∈B*‖Λ‖≤1

|Λy|.

However, for every Λ ∈ B* such that ‖Λ‖ ≤ 1, we have

|Λy| =∣∣∣∣∫

RΛf(t)dµ(t)

∣∣∣∣ ≤ ∫R|Λf(t)|dµ(t) ≤

∫R‖Λ‖‖f(t)‖dµ(t) ≤

∫R‖f(t)‖dµ(t).

Lemma 2.6.12. Let µ be a Radon measure on R. Let f ∈ Cc(R, B). Then, for every ε > 0,

there are A1, . . . An disjoint Borel sets, t1, . . . , tn ∈ R, such that the function

ϕ =n∑i=1

f(ti)χAi

33

Page 36: Characterization of Extremal KMS States on Groupoid C

satis�es the following property: ‖ϕ− f‖∞ ≤ ε.

Proof. Let ε > 0. Since f is continuous, for every t ∈ R, there exists an open set Ut such

that for every s ∈ Ut, ‖f(s)− f(t)‖ < ε.

Let K be the support of f . Then there are t1, . . . tn ∈ K such that Ut1 , . . . , Utn is an open

cover for K. Let, for i = 1, . . . , n,

Ai =

Ut1 if i = 1

Uti \ Ai−1 if i = 2, . . . , n.

Then each Ai is Borel, and ∪ni=1Ai = ∪ni=1Uti . De�ne ϕ by

ϕ =n∑i=1

f(ti)χAi .

Now we prove that ‖ϕ − f‖∞ ≤ ε. Let t ∈ R. If t /∈ ∪ni=1Uti , then ϕ(t) = 0 by de�nition.

Moreover, t /∈ K, since Ut1 , . . . , Utn cover K. Then f(t) = 0 and ‖f(t)− ϕ(t)‖ = 0 ≤ ε.

Assume t ∈ ∪ni=1Uti . By de�nition of A1, . . . , An, there exists a unique i such that t ∈ Ai.

Hence ϕ(t) = f(ti). Since Ai ⊂ Uti , we have

‖ϕ(t)− f(t)‖ = ‖f(ti)− f(t)‖ < ε.

Therefore,

‖ϕ− f‖∞ = supt∈R‖ϕ(t)− f(t)‖ ≤ ε.

In order to prove the existence of the integral of functions in Cc(R, B), we will state

Theorem 2.6.14, which is an application of a theorem proved in [22].

De�nition 2.6.13. Let X be a normed vector space. Given a subset S of X, we de�ne

34

Page 37: Characterization of Extremal KMS States on Groupoid C

its convex hull as the smallest convex set containing S. We denote the convex hull of S by

co(S).

Theorem 2.6.14. [22, Theorem 3.25] Suppose H is the convex hull of a compact set K in

a Banach space B. Then H is compact.

Theorem 2.6.15. Let µ be a Radon measure on R. Let B be a Banach space. Given

f ∈ Cc(R, B), the integral of f :

y =

∫Rf(t)dµ(t)

exists.

Proof. Assume µ is a probability measure.

Let K = supp(f). Let L = f(K) ∪ {0}. This set is compact because f is continuous and

K is compact. De�ne H to be the closure of co(L). Then H is compact by Theorem 2.6.14.

Given k ∈ N with k ≥ 1, it follows from Lemma 2.6.12 that there is a simple function

ϕ(k) : R → B such that there are disjoint Borel sets A(k)1 , . . . , A

(k)nk , and t

(k)1 , . . . , t

(k)nk ∈ R

satisfying

ϕ(k) =

nk∑i=1

f(t(k)i )χ

A(k)i

and ‖ϕ(k) − f‖∞ <1

k.

Let

yk =

∫Rϕ(k)dµ(t) =

nk∑i=1

f(t(k)i )µ(A

(k)i ). (2.13)

Since the sets A(k)1 , . . . , A

(k)nk are disjoint and µ is a probability measure, we have

nk∑i=1

µ(A(k)i ) ≤ 1 and ‖yk‖ ≤ ‖f‖∞. (2.14)

35

Page 38: Characterization of Extremal KMS States on Groupoid C

Moreover, since each f(t(k)i ) ∈ H and 0 ∈ H, it follows from (2.13) and (2.14) that yk ∈ H.

H is compact, then {yk}k∈N has a subsequence {ykj}j∈N converging to some y ∈ H.

Let Λ ∈ B*. Assume Λ 6= 0 without loss of generality. Since Λ is continuous, we have

Λykj → Λy. However, by Remark 2.6.5,

Λykj = Λ

(∫Rϕ(kj)(t)dµ(t)

)=

∫R

Λϕ(kj)(t)dµ(t). (2.15)

By de�nition, each ϕ(kj) satis�es ‖ϕ(kj)(t)‖ ≤ ‖f‖∞. Hence, for t ∈ R,

|Λϕ(kj)(t)| ≤ ‖Λ‖‖ϕ(kj)(t)‖ ≤ ‖Λ‖‖f‖∞.

Moreover, Λϕ(kj) converges to Λf pointwise. Since µ is a probability measure, we can apply

the dominated convergence theorem, obtaining

∫R

Λf(t)dµ(t) = limj→∞

∫R

Λϕ(kj)(t)dµ(t) = limj→∞

Λykj = Λy.

Since Λ is arbitrary, we have

y =

∫Rf(t)dµ(t).

Now assume µ is an arbitrary Radon measure on R. Let K be the support of f .

Suppose µ(K) = 0, then for each Λ ∈ B*,

∫R

Λ(f(t))dµ(t) =

∫K

Λ(f(t))dµ(t) = 0.

Therefore∫R f(t)dµ(t) = 0.

Now suppose µ(K) > 0. De�ne the measure µ on R by

µ(I) =µ(I ∩K)

µ(K),

36

Page 39: Characterization of Extremal KMS States on Groupoid C

for every Borel set I ⊂ R. Then µ is a probability Borel measure on R.

Let y =∫R f(t)dµ(t). Then, for every Λ ∈ B*,

Λy =

∫R

Λ(f(t))dµ(t)

=

∫K

Λ(f(t))dµ(t)

=1

µ(K)

∫K

Λ(f(t))dµ(t)

=1

µ(K)

∫R

Λ(f(t))dµ(t).

Therefore,

µ(K)y =

∫Rf(t)dµ(t).

Proposition 2.6.16. Let µ be a Borel measure on R and let B be a Banach space. Let

f : R → B be a continuous function such that∫∞−∞ ‖f(t)‖dµ(t) < ∞. Then the integral∫

R f(t)dµ(t) exists.

Proof. Let, for each n, hn be a continuous function such that hn equals 1 in the closed

interval [−n, n] and vanishes outside ]− n− 1, n + 1[. For each n, fhn ∈ Cc(R, B). De�ne,

for every n,

yn =

∫Rhn(t)f(t)dµ(t).

In order to show that the sequence of yn converges, we only need to prove that {yn} is a

Cauchy sequence because B is complete. Given ε > 0, let n0 be such that

∫|t|≥n0

‖f(t)‖ < ε

2.

37

Page 40: Characterization of Extremal KMS States on Groupoid C

Given n,m ≥ n0, assume n ≥ m without loss of generality. By de�nition, we have

hn(t) = hm(t) = 1 for t satisfying |t| < n0. (2.16)

In this case, |hn(t)− hm(t)| = 0. Then,

‖yn − ym‖ =

∥∥∥∥∫Rhn(t)f(t)dµ(t)−

∫Rhm(t)f(t)dµ(t)

∥∥∥∥=

∥∥∥∥∫R(hn(t)− hm(t))f(t)dµ(t)

∥∥∥∥≤∫R|hn(t)− hm(t)|‖f(t)‖dµ(t)

=

∫|t|≥n0

|hn(t)− hm(t)|‖f(t)‖dµ(t), by (2.16),

≤∫|t|≥n0

(|hn(t)|+ |hm(t)|)‖f(t)‖dµ(t)

≤ 2

∫|t|≥n0

‖f(t)‖dµ(t), because hn, hm assume values in [0, 1],

≤ 2ε

2= ε.

Therefore yn → y for some y ∈ B.

Now we prove that∫R Λ(hn(t)f(t))dµ(t)→

∫R Λ(f(t))dµ(t). For every t, we have

|Λ(hn(t)f(t))| ≤ ‖λ‖‖f(t)‖.

By assumption, the function t 7→ ‖f(t)‖ is integrable. Then, the dominated convergence

theorem implies,

limn→∞

∫R

Λ(hn(t)f(t))dµ(t) =

∫R

(limn→∞

Λ(hn(t)f(t)))dµ(t) =

∫R

Λ(f(t))dµ(t).

38

Page 41: Characterization of Extremal KMS States on Groupoid C

Note that, for every n,

Λ

(∫Rhn(t)f(t)dµ(t)

)=

∫R

Λ(hn(t)f(t))dµ(t).

The left-hand side equals to Λyn and thus converges to Λy. As we already proved, the

right-hand side converges to∫R Λ(f(t))dµ(t). Therefore,

Λy =

∫Rλ(f(t))dµ(t).

Λ is arbitrary, then the integral

y =

∫Rf(t)dµ(t)

exists.

Corollary 2.6.17. Let B be a Banach space and µ a Borel measure on X. Let f : R→ B

be a continuous function such that t 7→ ‖f(t)‖ is integrable. Given a linear and bounded

operator L : B → B1 such that B1 is a Banach space, then

L

(∫Rf(t)dµ(t)

)=

∫RL(f(t))dµ(t).

Proof. Let y =∫R f(t)dµ(t). The function Lf : R → B1 is continous, moreover, t 7→

‖L(f(t))‖ is integrable, since

∫R‖L(f(t))‖dµ(t) ≤ ‖L‖

∫R‖f(t)‖dµ(t) <∞.

Let z =∫R L(f(t))dµ(t). Given Λ ∈ B1*, ΛL ∈ B*. Hence,

Λz = Λ

∫RL(f(t))dµ(t)

=

∫R

ΛL(f(t))dµ(t)

39

Page 42: Characterization of Extremal KMS States on Groupoid C

=

∫R(ΛL)(f(t))dµ(t)

= (ΛL)

∫R(f(t))dµ(t)

= (ΛL)y

= Λ(Ly).

Since Λ is arbitrary, it follows that z = Ly. Thus,

L

(∫Rf(t)dµ(t)

)=

∫RL(f(t))dµ(t).

40

Page 43: Characterization of Extremal KMS States on Groupoid C

Chapter 3

Groupoids

Groupoids can be understood as a generalization of groups where the unit is not unique

and not every pair of elements can be multiplied. Each groupoid G is endowed with two

functions r and s from G to the subset G(0) of units. We equip the groupoid with a topology

such that r, s are continuous.

If G has some nice topological properties, we can de�ne Cc(G), the space of continuous

and compactly supported complex functions on G. Moreover, we can endow this space with

an involution and a product which is not necessarily commutative. Then we de�ne a norm on

Cc(G) which depends on the ∗-representations of Cc(G). Then the full groupoid C*-algebra,

denoted C*(G) is the completion of Cc(G) with respect to this norm.

Most de�nitions and results in this chapter can be found in [9].

3.1 Introduction

In this section we de�ne groupoids and give some examples. The results in this section are

taken from [9] and [20].

De�nition 3.1.1. A groupoid is a set G together with a subset G(0) (called units , unit space

or objects), two surjective maps r, s : G → G(0) (called range and source, respectively) and

41

Page 44: Characterization of Extremal KMS States on Groupoid C

a law of composition

(g, h) ∈ G(2) 7→ gh = g · h ∈ G,

where G(2) = {(g, h) ∈ G × G : s(g) = r(h)} is called the set of composable elements or

composable pairs .

A groupoid satis�es the following properties for g, h, k ∈ G:

(i) s(gh) = s(h) and r(gh) = r(g) if (g, h) ∈ G(2);

(ii) r(x) = s(x) if x ∈ G(0);

(iii) gs(g) = g and r(g)g = g;

(iv) (gh)k = g(hk) if (g, h), (h, k) ∈ G(2);

(v) g has a two-sided inverse g−1 such that gg−1 = r(g) and g−1g = s(g).

The maps (g, h) ∈ G(2) 7→ gh and g 7→ g−1 are called product and inverse, respectively.

We can interpret groupoids as a collection of arrows attached to points on a plane. Two

arrows can be composed only if the end of the �rst arrow meets the start of the second.

Units are points with the null vector and the inverse of an element is obtained by reversing

the direction of the arrow. Figure 3.1 shows this idea.

s(g)

r(g)

s(h)

r(h)

gh

(a)

r(g)

s(g) = r(h)

s(h)g

h

gh

(b)

s(g)

r(g)

s(g)

r(g)

s(g) = g−1g

r(g) = gg−1

g g−1

(c)

Figure 3.1: Groupoids can be seen as arrows on a plane. s(g) and r(g) denote the sourceand range of g. (a) g and h are not composable, since s(g) 6= r(h); (b) The composition ofg and h is gh; (c) g−1 is the inverse of g. Note that g−1g = s(g) and gg−1 = r(g).

42

Page 45: Characterization of Extremal KMS States on Groupoid C

Example 3.1.2. Every group is a groupoid. Let G be a group with unit e. Let G(0) = {e},

G(2) = G×G and de�ne the range and source maps by r(g) = s(g) = e.

Since the range and source of each element is e and G is associative, one can easily show

that properties (i)-(v) are satis�ed.

Example 3.1.3. We show that a group action de�nes a groupoid. Let G be a group with

identity e and X a set. Recall that a group action [19] is a map G × X → X denoted by

(g, x) 7→ gx, satisfying the following properties:

(i) g(hx) = (gh)x for g, h ∈ G, x ∈ X,

(ii) ex = x for x ∈ X.

If G is an action, we say that G acts on X.

The cartesian productH = G×X has a groupoid structure with unit spaceH(0) = {e}×X.

The range and source maps are s(g, x) = (e, x) and r(g, x) = (e, gx) and the operations are

de�ned by

(g, hx)(h, x) = (gh, x) and (g, x)−1 = (g−1, gx).

H is a groupoid, called the action groupoid (or transformation groupoid [23]). In fact,

(i) s((g, x)(h, y)) = s(g, y) = (e, y) = s(h, y) for (g, x), (h, y) composable;

(ii) r(e, x) = (e, ex) = (e, x) = s(e, x) for x ∈ X;

(iii) Given (g, x) ∈ H,

(g, x)s(g, x) = (g, x)(e, x) = (ge, x) = (g, x)

r(g, x)(g, x) = (e, gx)(g, x) = (eg, x) = (g, x);

(iv) Let (g, x), (h, y), (k, z) ∈ H such that ((g, x), (h, y)), ((h, y), (k, z)) ∈ H(2). By hypoth-

43

Page 46: Characterization of Extremal KMS States on Groupoid C

esis, x = hy and y = kz. Hence, x = hkz. Then

[(g, x)(h, y)] (k, z) = (gh, y)(k, z)

= (ghk, z)

= (g, x)(hk, z)

= (g, x) [(h, y)(k, z)] .

(v) Given (g, x) ∈ H,

(g, x)(g, x)−1 = (g, x)(g−1, gx) = (gg−1, gx) = (e, gx) = r(g, x)

(g, x)−1(g, x) = (g−1, gx)(g, x) = (g−1g, x) = (e, x) = s(g, x).

Example 3.1.4. Let ∼ be an equivalence relation on a set X. Let

G = {(x, y) ∈ X ×X : x ∼ y},

G(0) = {(x, x) : x ∈ X}, and

G(2) = {((x, y), (y, z)) : x ∼ y, y ∼ z}.

De�ne the range and source maps by r(x, y) = (x, x) and s(x, y) = (y, y). Let (x, y)−1 =

(y, x) and (x, y)(y, z) = (x, z). The inverse and multiplication maps are well-de�ned by the

re�exivity and transitivity of ∼. Hence G is a groupoid.

Remark 3.1.5. Note that [9] and [20] de�ne groupoids di�erently. On the one hand, [9]

introduces a groupoid as in De�nition 3.1.1. On the other hand, Renuault [20] describes

groupoids as follows:

A groupoid is a set G endowed with a product map (g, h) 7→ gh : G(2) → G, where G(2) is

a subset of G×G called the set of composable pairs, and an inverse map g 7→ g−1 : G→ G

such that the following relations are satis�ed:

(i') (g−1)−1 = g;

44

Page 47: Characterization of Extremal KMS States on Groupoid C

(ii') If (g, h), (h, k) ∈ G(2), then (gh, k), (g, hk) ∈ G(2) and (gh)k = g(hk);

(iii') (g−1, g) ∈ G(2) and if (g, h) ∈ G(2), then g−1(gh) = h;

(iv') (g, g−1) ∈ G(2) and if (h, g) ∈ G(2), then (hg)g−1 = h.

Given g ∈ G, we de�ne r(g) = gg−1 and s(g) = g−1g. The unit space is de�ned by

G(0) = s(G) = r(G).

These de�nitions are equivalent.

First, suppose G is a groupoid as in De�nition 3.1.1. Note that r(x) = s(x) = x for each

x ∈ G(0). In fact, r(x) = s(x) by property (ii). Since s : G→ G(0) is surjective, there exists

g ∈ G such that x = s(g) = g−1g. Hence,

s(x) = s(s(g)) = s(g−1g) = s(g) = x.

Now we prove properties (i')�(iv').

(i') (g−1)−1 = g

Since s(g) = g−1g and s(s(g)) = r(s(g)), we have

s(s(g)) = s(g) = g−1g,

r(s(g)) = r(g−1g) = r(g−1) = g−1(g−1)−1.

Then g−1g = g−1(g−1)−1 and therefore g = (g−1)−1.

(ii') (gh)k = g(hk)

This holds by property (iv).

(iii') g−1(gh) = h

Note that s(g) = r(h). Then

g−1(gh) = (g−1g)h = s(g)h = r(h)h = h.

45

Page 48: Characterization of Extremal KMS States on Groupoid C

(iv') (hg)g−1 = h

Note that s(h) = r(g). Then

(hg)g−1 = h(gg−1) = hr(g) = hs(h) = h.

Then G is a groupoid as de�ned in [20].

Conversely, assume that G is a groupoid as in [20].

First we show that G(2) = {(g, h) : s(g) = r(h)}. Suppose (g, h) ∈ G(2). Then,

g−1(gh) = h by (iii'),

[g−1(gh)]h−1 = hh−1 by (iv'),

[(g−1g)h]h−1 = hh−1

g−1g = hh−1

s(g) = r(h).

Suppose g, h ∈ G are such that s(g) = r(g). Then

(h, h−1), (h−1, h) ∈ G(2) by (iii'), (iv'),

⇒ (hh−1, h) ∈ G(2)

⇒ (r(h), h) ∈ G(2)

⇒ (s(g), h) ∈ G(2)

⇒ (g−1g, h) ∈ G(2)

⇒ (g−1g, h) ∈ G(2).

46

Page 49: Characterization of Extremal KMS States on Groupoid C

On the other hand,

(g, g−1), (g−1, g) ∈ G(2) by (iii'), (iv'),

⇒ (g, g−1g) ∈ G(2).

Then (g, g−1g), (g−1g, h) ∈ G(2). Therefore (g, h) ∈ G(2).

Now we prove properties (i)�(v).

(i) s(gh) = s(h), r(gh) = r(g)

By assumption, (g, h) ∈ G(2). Also, (h, h−1) ∈ G(2) by (iii). Then s(h) = r(h−1) and

(gh, h−1) ∈ G(2). This implies s(gh) = r(h−1) = s(h).

The proof of r(gh) = r(g) is analogous.

(ii) r(x) = s(x) if x ∈ G(0)

Given x ∈ G(0), there exists g ∈ G such that x = r(g) = gg−1. Then

s(x) = s(gg−1) = s(g−1) = r(g) = x.

Note that s(gg−1) = s(g−1) by (i).

(iii) gs(g) = g and r(g)g = g

gs(g) = g(g−1g) = (g−1)−1(g−1g) = g,

r(g)g = (gg−1)g = (gg−1)(g−1)−1 = g.

(iv) (gh)k = g(hk)

This is equivalent to (ii')

(v) r(g) = gg−1, g−1g = s(g)

This follows from the de�nition of r, s.

47

Page 50: Characterization of Extremal KMS States on Groupoid C

Therefore the de�nitions are equivalent.

Remark 3.1.6. Given A,B subsets of a groupoid G, one may form the following subsets of

G:

A−1 = {g ∈ G : g−1 ∈ A}, AB = {gh ∈ G : g ∈ A, h ∈ B}.

Given x, y ∈ G(0):

Gx = r−1(x), Gy = s−1(y), and Gxy = Gx ∩Gy.

Gx (resp. Gy) is called the r-�ber of G over x (resp. s-�ber of G over y) as in [7].

Note that Gxx is a group. It is called the isotropy group at x. In fact,

(i) gh ∈ Gxx for g, h ∈ Gx

x;

(ii) gg−1 = g−1g = x for g ∈ Gxx. Hence x is the unity of Gx

x;

(iii) the product in Gxx associative.

Notation 3.1.7. Unless otherwise speci�ed, we will use the following notation in this thesis:

G denotes a groupoid; its units are denoted by the letters x, y, z; g, h are elements in G.

Subsets of G may be written as the uppercase letters U, V . The letters may be indexed or

marked with an accent or symbol.

3.2 Topological Groupoids

If G is a groupoid endowed with a topology, it is useful that its operations have interesting

topological properties. We de�ne the notion of topological groupoid, where its operations

are continuous. We also de�ne étale groupoids, where the range and source maps are local

homeomorphisms. This section is based on [9] and [20].

48

Page 51: Characterization of Extremal KMS States on Groupoid C

De�nition 3.2.1. A topological groupoid is a groupoid G with a topology such that G(2) has

the induced topology from G×G, and both the product and inverse maps are continuous.

Remark 3.2.2. Let G be a topological groupoid. Since r and s are de�ned by r(g) = gg−1

and s(g) = g−1g, it follows that these functions are continuous.

Now we de�ne the notion of étale groupoid. The main results in this thesis assume the

groupoid has this property.

De�nition 3.2.3. A topological groupoid is étale if the maps r and s are local homeomor-

phisms.

Example 3.2.4. Every discrete groupoid G is étale. In fact, for every g ∈ G, the subsets

{g}, {r(g)}, {s(g)} are open in G. Moreover, the maps r|{g} : {g} → {r(g)}, s|{g} : {g} →

{s(g)} are homeomorphimsms. In particular, every discrete group is étale.

Example 3.2.5. Let X be a topological space and let r, s : X → X be identity maps.

Moreover, de�ned for each x ∈ X, xx = x and x−1 = x. Then X is an étale groupoid

because r, s are homeomorphimsms.

Another example of étale groupoid is the transformation groupoid G×X when the group

G is discrete. We prove this in the following lemma:

Lemma 3.2.6. Let G be a group endowed with a topology. Let X be a topological space

and �x a continuous group action G×X → X. Suppose G×X is the action groupoid as in

Example 3.1.3 and equip this space with the product topology. Then G×X is étale if, and

only if, G is discrete.

Proof. • Suppose G is not discrete.

There exists g ∈ G such that for each neighborhood U of g, U \ {g} 6= ∅.

Fix x ∈ X and let V be an arbitrary open neighborhood of x. Let U be an open

neighborhood of g. Then there exists h 6= g such that h ∈ U . Then (g, x), (h, x) ∈ U×V

49

Page 52: Characterization of Extremal KMS States on Groupoid C

and s(g, x) = s(h, x) = (e, x). Since U, V are arbitrary, it follows that s is not a local

homeomorphimsm. Therefore, G is not étale.

• Now suppose that G is discrete

Then the product and inverse maps on the group G are continuous. Note that product

and inverse maps on the groupoid are continuous because they are compositions of

continuous functions. Then G×X is a topological groupoid.

For each g ∈ G the map X 7→ X de�ned by x 7→ gx is a homeomorphimsm with

inverse x 7→ g−1x. Then, for every open set U ⊂ X, the set gU = {gx : x ∈ U} is open

in X.

Now we show that G ×X is étale. Let (g, x) ∈ G ×X, U a neighborhood of x ∈ X.

Then {g} × U is an open neighborhood of (g, x). Then

r({g} × U) = {(e, gx) : x ∈ U} = {e} × gU,

s({g} × U) = {(e, x) : x ∈ U} = {e} × U.

Then r({g} × U), s({g} × U) are open sets in G×X.

The function s|{g}×U is injective. The function x ∈ U 7→ gx is injective, then r

is injective on {g} × U . Since g and U are arbitrary, it follows that r, s are open

bisections. Therefore, G×X is étale.

De�nition 3.2.7. An open subset U of an étale groupoid is an open bisection of G if

r(U), s(U) are open in G(0), and r|U : U → r(U) and s|U : U → s(U) are homeomorphisms.

Notation 3.2.8. We will usually denote an open bisection by the cursive letter U . This

letter may be indexed or marked with an accent or symbol.

Remark 3.2.9. Many times throughout the thesis, we will evaluate sums which take into

account values f(g) such that g ranges over Gx or Gx, assuming f ∈ Cc(G). However, if this

50

Page 53: Characterization of Extremal KMS States on Groupoid C

function is supported on an open bisection, we can consider only one term in the sum. This

element is usually denoted hx (resp. hx) and hx ∈ Gx ∩ U (resp. hx ∈ Gx ∩ U).

Later we prove that every f ∈ Cc(G) can be written as a �nite sum of continuous functions

supported on open bisections. Hence, in many cases, we can assume f is supported on an

open bisection without loss of generality.

Proposition 3.2.10. Let G be an étale groupoid. The set of open bisections of G forms an

open base for the topology of G.

Proof. Let U be an open set of G. We will show that for every g ∈ U there exists an open

bisection Ug such that g ∈ Ug ⊂ U . In fact, let g ∈ U . Since G is étale, r, s are local

homeomorphimsms. Then there exist Rg, Sg open neighborhoods of g such that r(Rg) and

s(Sg) are open in G(0), and r|Rg : Rg → r(Rg), s|Sg : Sg → s(Sg) are homeomorphisms.

Let Ug = Rg ∩ Sg ∩ U . r(Ug) is open in r(Rg), then r(Ug) is open in G(0). Hence,

r|Ug : Ug → r(Ug) is a homeomorphism. Analogously s|Ug : Ug → s(Ug) is a homeomorphism.

Therefore, for every open set U , we have U =⋃g∈U Ug.

Proposition 3.2.11. If G is an étale groupoid, then the subspace topology of Gx and Gx is

equivalent to the discrete topology for all x ∈ G(0). Furthermore, if G is second countable,

then Gx and Gx have a countable number of elements.

Proof. Let g ∈ Gx. There exists an open bisection Ug containing g. We show that Ug ∩Gx =

{g}. Suppose there exists h 6= g such that h ∈ Ug ∩Gx. Then r(h) = x. Contradiction, since

r is injective on Ug. Hence {g} is open in Gx. Therefore Gx is endowed with the discrete

topology.

Assume G is second countable. Then Gx is second countable. Since the sets {g}, g ∈ Gx,

form a family of disjoint open sets in Gx, it follows that Gx is countable. The proof for Gx

is analogous.

Proposition 3.2.12. If G is a locally compact Hausdor� étale groupoid, then G(0) is a

clopen subset of G. We assume G(0) is endowed with the subspace topology.

51

Page 54: Characterization of Extremal KMS States on Groupoid C

Proof. We divide the proof in two parts.

• G(0) is closed

Let xi be a net in G(0) converging to x ∈ G. The function r is continuous, then

r(xi)→ r(x). As xi ∈ G(0), we have r(xi) = xi. Hence x = r(x) ∈ G(0). Therefore G(0)

is closed.

• G(0) is open

Let x ∈ G(0). Let U ⊂ G be an open bisection containing x. Let V = G(0) ∩ U . Then

V is an open neighborhood in G(0) of x. Moreover, V ⊂ r(U), since r(y) = y for every

y ∈ V .

Since r|U : U → r(U) is a homeomorphimsm, r|−1U (V ) = V . Then V is open in G.

Therefore G(0) is open.

Let G′ = ∪x∈G(0)Gxx, called the isotropy bundle. The following lemma shows that G′ is

closed.

Lemma 3.2.13. Let G be a locally compact Hausdor� second countable étale groupoid.

Given g ∈ G such that r(g) 6= s(g), there exists an open bisection U including g such that

r(U) ∩ s(U) = ∅. Moreover, G′ ∩ U = ∅. In particular, G′ is closed.

Proof. Suppose this lemma is false. Then there exists g ∈ G \ G′ such that for every open

bisection U including g, we have

r(U) ∩ s(U) 6= ∅.

Since G is second countable and étale, we can choose a countable family {Un} of open

bisections containing g such that every neighborhood of g contains at least one Un. Hence,

52

Page 55: Characterization of Extremal KMS States on Groupoid C

for every n there are gn, hn ∈ Un satisfying

r(gn) = s(hn). (3.1)

By de�nition, both sequences {gn}n∈N, {hn}n∈N converge to g as n→∞. Then, by continuity

of r, s, we have r(gn)→ r(g) and s(hn)→ s(g). However, from (3.1), we have s(hn)→ r(g).

Hence r(g) = s(g). This leads to a contradiction because we assumed g /∈ G′.

Therefore we can choose U satisfying r(U) ∩ s(U) = ∅. Moreover, G′ ∩ U = ∅. Since

g ∈ G \G′ is arbitrary, it follows that G′ is closed.

Remark 3.2.14. Let G be a groupoid and V ⊂ G(0). We de�ne G|V = G∩r−1(V )∩s−1(V ).

Note that G|V is a groupoid. If G is a topological groupoid, then G|V is also a topological

groupoid. Analogously, if G is étale, so is G|V .

3.3 Groupoid C*-Algebras

Now we de�ne the full groupoid C*-algebra and prove some properties of this C*-algebra.

The results in this section can be found in [5], [9] and [23].

Let G be a locally compact second countable Hausdor� étale groupoid. Denote Cc(G) by

Cc(G) = {f : G→ C : f is continuous and supp(f) is compact}.

Recall that the support of f is de�ned by supp(f) = {g ∈ G : f(g) 6= 0}. We de�ne the

convolution and involution operations on Cc(G) by

(f1 · f2)(g) =∑g1g2=g

f1(g1)f2(g2) and f*(g) = f(g−1). (3.2)

Example 3.3.1. Let n be a positive integer and de�ne the groupoid G = {(i, j) : i, j =

53

Page 56: Characterization of Extremal KMS States on Groupoid C

1, . . . , n} such that

G(0) = {(i, i) : i = 1, . . . , n}

G(2) = {((i, k), (k, j)) : i, j, k = 1, . . . , n},

and de�ne the operations

(i, k)(k, j) = (i, j) and (i, j)−1 = (j, i).

Equip G with the discrete topology. Then G is locally compact Hausdor�. Moreover, G is

étale by Example 3.2.4.

Note that there is a bijection from Cc(G) toMn(C) given by f 7→ F such that Fi,j = f(i, j).

Moreover, we can identify Cc(G) with Mn(C). In fact, let f (1), f (2) ∈ Cc(G). Assume A is a

matrix corresponding to f (1) · f (2). Then, for every i, j = 1, . . . , n,

Ai,j = (f (1) · f (2))(i, j) =n∑k=1

f (1)(i, k)f (2)(k, j) =n∑k=1

F(1)i,k F

(2)k,j = (F (1)F (2))i,j,

then A = F (1)F (2).

Now assume F ∈Mn(C) corresponds to f ∈ Cc(G). Then F* corresponds to f*. In fact,

for i, j = 1, . . . , n,

F *i,j = Fj,i = f(j, i) = f((i, j)−1) = f*(i, j).

Therefore we can identify Mn(C) with Cc(G). Moreover, Cc(G) is not commutative.

Notation 3.3.2. We usually denote a function in Cc(G) by f . Note that, for an open

subset U ⊂ G, every function in Cc(U) can be extended uniquely to a function in Cc(G)

whose support lies in U . Thus, for every open set U ⊂ G, we will denote without loss of

54

Page 57: Characterization of Extremal KMS States on Groupoid C

generality,

Cc(U) = {f ∈ Cc(G) : supp(f) ⊂ U}.

Thus Cc(U) is a subspace of Cc(G).

The letter h sometimes denotes elements in Cc(G(0)). However, h is also used to indicate

elements in G.

Given f1, f2 ∈ Cc(G), f1f2 denotes the pointwise product of these functions, while f1 · f2

denotes the convolution product.

Lemma 3.3.3. Let G be a locally compact second countable Hausdor� étale groupoid.

Given f1, f2 ∈ Cc(G), g ∈ G,

(f1 · f2)(g) =∑

h∈Gs(g)

f1(gh−1)f2(h) (3.3)

=∑

h∈Gr(g)f1(h)f2(h−1g). (3.4)

Proof. Let g1g2 ∈ G such that g1g2 = g. This is equivalent to g1 = gg−12 . This equation

holds only if g2 ∈ Gs(g). Therefore, for every h ∈ Gs(g) we can choose g2 = h and g1 = gh−1.

Then,

(f1 · f2)(g) =∑

h∈Gr(g)f1(gh−1)f2(h).

This sum is �nite for every g, since Gr(g) is countable and f2 is compactly supported, then the

set of elements h ∈ Gr(g) such that f2(h) 6= 0 is �nite. The proof for (3.4) is analogous.

Lemma 3.3.4. Let f ∈ Cc(G), h ∈ Cc(G(0)), g ∈ G. Then

(h · f)(g) = h(r(g))f(g), and (f · h)(g) = f(g)h(s(g)).

55

Page 58: Characterization of Extremal KMS States on Groupoid C

Proof. It follows from Lemma 3.3.3 that

(h · f)(g) =∑

k∈Gr(g)h(k)f(k−1g).

Suppose h(k) 6= 0 for some k ∈ Gr(g). Then k ∈ G(0) and r(k) = r(g). Hence k = r(g) and

therefore,

(h · f)(g) = h(r(g))f(r(g)−1g) = h(r(g))f(r(g)g) = h(r(g))f(g).

The proof for f · h is analogous.

Now we show that every function in Cc(G) can be decomposed as a sum of continuous

compactly supported functions whose support are included in open bisections. This result

will be used many times in the thesis because many results are easier to prove when the

function is supported on an open bisection.

Lemma 3.3.5. Let G be a locally compact second countable étale Hausdor� groupoid.

Given f ∈ Cc(G), there are U1, . . . ,Un open bisections and f1, . . . , fn functions such that

f = f1 + . . . + fn and each fi ∈ Cc(Ui). Moreover, if f is non-negative, we can choose each

fi to be non-negative.

Proof. Let f ∈ Cc(G) with support K. From Proposition 3.2.10 the set of open bisections

forms an open base for G. Then there exists a �nite cover U1, . . .Un of K such that each Uiis an open bisection.

Let Un+1 = G \K. Then {Ui}n+1i=1 is an open cover of G. Let {αi}n+1

i=1 be the partition of

unit subordinate to the the open cover {Ui}n+1i=1 . Note that fαn+1 = 0 since αn+1 is supported

on Un+1 \K. Then,

f =n+1∑i=1

fαi =n∑i=1

fαi.

56

Page 59: Characterization of Extremal KMS States on Groupoid C

De�ne fi = fαi for i = 1, . . . , n. By de�nition of αi, each fi ∈ Cc(Ui). Moreover,

since αi assumes values in the interval [0, 1], if f is non-negative, it follows that each fi is

non-negative.

Lemma 3.3.6. Let G be a locally compact Hausdor� second countable étale groupoid. If

U ,V ⊂ G are open bisections, then

UV = {gh : g ∈ U , h ∈ V , (g, h) ∈ G(2)}

is an open bisection.

Proof. Before we prove UV is an open bisection, we will show that we can assume s(U) = r(V)

without loss of generality. Let W = s(U) ∩ r(V). Then W is an open set in G(0) because U ,

V are open bisections.

Let U0 = s|−1U (W ) and V0 = r|−1

V (W ). Both U0,V0 are open bisections, since they are

open subsets of open bisections. Moreover, we have

s(U0) = s ◦ s|−1U (W ) = W = r ◦ r|−1

V (W ) = r(V0).

Now we show that U×V∩G(2) = U0×V0∩G(2). In fact, given (g, h) ∈ U×V∩G(2), we have

g ∈ U , h ∈ V and s(g) = r(h). If we de�ne x = s(g), then x ∈ W . Moreover, g = s|−1U (x),

which implies g ∈ U0. Analogously, h ∈ V0. Then (g, h) ∈ U0 × V0 ∩ G(2). Therefore

U×V∩G(2) ⊂ U0×V0∩G(2). Since U0 ⊂ U and V0 ⊂ V , we have U×V∩G(2) = U0×V0∩G(2).

By de�nition UV , we have

UV = {gh : g ∈ U , h ∈ V , (g, h) ∈ G(2)}

= {gh : (g, h) ∈ U × V ∩G(2)}

= {gh : (g, h) ∈ U0 × V0 ∩G(2)}

= U0V0.

57

Page 60: Characterization of Extremal KMS States on Groupoid C

Therefore we can assume s(U) = r(V) without loss of generality.

Now we prove UV is an open bisection. Assume s(U) = r(V). Let φ : U → V be the

homeomorphism de�ned by φ = r|−1V ◦ s|U .

De�ne the map f from U to U × V by f(g) = (g, φ(g)). The image f(U) is included in

U × V ∩G(2), since

r(φ(g)) = r(r|−1V ◦ s(g)) = s(g).

We claim f(U) = U × V ∩G(2). Suppose (g, h) ∈ U × V ∩G(2). Then s(g) = r(g), g ∈ U ,

h ∈ V . Hence

h = r|−1V ◦ s|U(g) = φ(g).

Thus (g, h) = (g, φ(g)) = f(g). Therefore f(U) = U × V ∩G(2).

By de�nition of f , we have that f is injective, thus. We will show that f is a homeomor-

phism. Let π : U ×V ∩G(2) → U be the projection onto the �rst coordinate. π is continuous

by de�nition. So we will show that π is the inverse of f .

Given g ∈ U ,

π ◦ f(g) = π(g, φ(g)) = g.

Given (g, h) ∈ U × V ∩G(2), we have (g, h) = (g, φ(g)) = f(g) since f is a bijection. Then

(f ◦ π)(g, h) = f(g) = (g, φ(g)) = (g, h).

Therefore π is the inverse of f and f is a homeomorphism. Hence the set U × V ∩ G(2) is

open in U × V .

58

Page 61: Characterization of Extremal KMS States on Groupoid C

Now we can consider the product p : U × V ∩G(2) → UV and observe that

r|UV ◦ p = r|U ◦ π. (3.5)

In fact, given (g, h) ∈ U × V ∩G(2),

r|UV ◦ p(g, h) = r(gh) = r(g) = r|U ◦ π(g, h).

Equation (3.5) shows that r|UV ◦ p is a homeomorphism. Moreover, we conclude that p is

surjective and r|UV is injective.

In addition, p is injective because if p(g1, h1) = p(g2h2), we have by (3.5) the following

result,

r|U ◦ π(g1, h1) = r|U ◦ π(g2, h2)

r|U(g1) = r|U(g1)

g1 = g2, since U is an open bisection,

r(h1) = r(h2) because (g1, h1), (g2, h2) ∈ U × V ∩G(2),

h1 = h2 since V is an open bisection.

Therefore p is injective. Hence r|UV , p are continuous bijections such that their composi-

tion is a homeomorphism. Therefore p : U × V ∩G(2) → UV is a homeomorphism.

Since U ×V ∩G(2) is an open set, so is UV . We already proved that r|UV is injective. The

proof for s|UV is analogous. Therefore, UV is an open bisection.

Lemma 3.3.7. Let G be a locally compact Hausdor� second countable étale groupoid. If

U ⊂ G is an open bisection, then U−1 = {g−1 : g ∈ U} is an open bisection.

Proof. Let ι : G → G be the inverse map. ι is continuous and ι ◦ ι is the identity. Then

U−1 = ι(U) is open.

59

Page 62: Characterization of Extremal KMS States on Groupoid C

Let g1, g2 ∈ U−1 such that r(g1) = r(g2). There exist h1, h2 ∈ U such that gi = h−1i ,

i = 1, 2. Then

s(h1) = r(g1) = r(g2) = s(h2).

Since U is an open bisection, we have h1 = h2. Then g1 = g2. The proof for s is analogous.

Therefore U−1 is an open bisection.

Lemma 3.3.8. Let G be a locally compact Hausdor� second countable étale groupoid.

(i) Given U1,U2 open bisections, f1 ∈ Cc(U1), f2 ∈ Cc(U2), then f1 · f2 ∈ Cc(U1U2).

(ii) If U is an open bisection and f ∈ Cc(U), we have f* ∈ Cc(U−1).

Proof. (i) Note that U1U2 is an open bisection by Lemma 3.3.6.

Let g /∈ U1U2. Then f1 · f2(g) = 0 since f(g) 6= 0 implies that there are g1 ∈ U1, g2 ∈ U2

satisfying g1g2 = g. Therefore the support of f1 · f2 lies in U1U2.

Since U1U2 is an open bisection, the maps u1, u2 are homeomorphisms where u1 :

U1U2 → U1 is de�ned by u1 = r|−1U1 ◦ r and u2 : U1U2 → U2 is de�ned by u2 = s|−1

U2 ◦ s.

Given g ∈ U1U2, g1 = u1(g), g2 = u2(g) are the only elements satisfying g1 ∈ U1, g2 ∈

U2, g = g1g2. In fact, suppose there are (h1, h2) ∈ U1 × U2 ∩ G(2) such that g = h1h2.

Then r(h1) = r(g). Since U1 is an open bisection, we have h1 = r|−1U1 ◦ r(g) = g1.

Analogously h2 = g2.

Therefore, for every g ∈ U1U2,

(f1 · f2)(g) = f1(u1(g))f2(u2(g)).

For i = 1, 2, ui : U1U2 → Ui is continuous and fi : Ui → C is continuous. Hence f1 · f2

is continuous on U1U2. Since f1 · f2 vanishes outside U1U2, we have f1 · f2 ∈ Cc(U1U2).

60

Page 63: Characterization of Extremal KMS States on Groupoid C

(ii) Let ι : G → G be the inverse map. ι is continuous. Since f* is de�ned by f* = f ◦ ι,

then f* is continuous.

Let K = supp(f) and L = supp(f*). Then

L = {g ∈ G : f*(g) 6= 0}

= {g ∈ G : f(g−1) 6= 0}

= {g ∈ G : f(g−1) 6= 0}

= {g ∈ G : f(g) 6= 0}−1

= ({g ∈ G : f(g) 6= 0})−1

= ({g ∈ G : f(g) 6= 0})−1, since the inversion is continuous,

= K.

The inversion on G is continuous, thus L is compact. Moreover, L ⊂ U−1. Thus

f* ∈ Cc(U−1).

Theorem 3.3.9. Let G be a locally compact Hausdor� second countable étale groupoid.

Cc(G) with the operations (3.2) is a ∗-algebra.

Proof. Clearly Cc(G) is a vector space.

• The product is bilinear.

Let f1, f2, f ∈ Cc(G), λ ∈ C, g ∈ G. Then,

[(f1 + λf2) · f ](g) =∑g1g2=g

(f1 + λf2)(g1)f(g2)

=∑g1g2=g

f1(g1)f(g2) + λ∑g1g2=g

f2(g1)f(g2)

= [f1 · f ](g) + λ[f2 · f ](g).

61

Page 64: Characterization of Extremal KMS States on Groupoid C

The proof for f · (f1 + λf2) is analogous.

• The product is associative.

Let f1, f2, f3 ∈ Cc(G). Given g ∈ G,

[f1 · (f2 · f3)](g) =∑g1h=g

f1(g1)(f2 · f3)(h)

=∑g1h=g

∑g2g3=h

f1(g1)f2(g2)f3(g3)

=∑

g1g2g3=g

f1(g1)f2(g2)f3(g3)

=∑hg3=g

∑g1g2=h

f1(g1)f2(g2)f3(g3)

=∑hg3=g

( ∑g1g2=h

f1(g1)f2(g2)

)f3(g3)

=∑hg3=g

(f1 · f2)(h)f3(g3)

= [(f1 · f2) · f3](g).

• f1 · f2 ∈ Cc(G) if f1, f2 ∈ Cc(G).

Since the product is bilinear and, from Lemma 3.3.5, every function in Cc(G) can be

written as a �nite sum of continuous functions supported on open bisections, it su�ces

to show that f1 · f2 ∈ Cc(U1U2) for f1 ∈ Cc(U1), f2 ∈ Cc(U2) where U1,U2 are open

bisections. Note that U1U2 is an open bisection by Lemma 3.3.6. However, we already

proved f1 · f2 ∈ Cc(U1U2) in Lemma 3.3.8.

• For f ∈ Cc(G), f** = f .

Let g ∈ G, then

f**(g) = f*(g−1) = f((g−1)−1) = f(g).

• For f1, f2 ∈ Cc(G), (f1 · f2)* = f2* · f1*.

62

Page 65: Characterization of Extremal KMS States on Groupoid C

Let g ∈ G. Then,

(f1 · f2)*(g) = (f1 · f2)(g−1)

=∑

g1g2=g−1

f1(g1)f2(g2)

=∑

g1g2=g−1

f2*(g−12 )f1*(g−1

1 )

=∑

g−12 g−1

1 =g

f2*(g−12 )f1*(g−1

1 ),

making the change of variables h1 = g−12 , h2 = g−1

1 ,

=∑

h1h2=g

f2*(h1)f1*(h2)

= (f2* · f1*)(g).

• The involution is conjugate-linear.

Let f1, f2 ∈ Cc(G), λ ∈ C, g ∈ G. Then,

(f1 + λf2)*(g) = (f1 + λf2)(g−1) = f1(g−1) + λf2(g−1) = f1*(g) + λf2*(g).

• If f ∈ Cc(G), then f* ∈ Cc(G).

Since the involution is conjugate-linear, we can assume f ∈ Cc(U) for an open bisection.

It follows from Lemma 3.3.8 that f* ∈ Cc(U).

In Theorem 3.3.9 we proved Cc(G) is a ∗-algebra. Now we will equip this space with a

norm such that its completion is a C*-algebra.

Lemma 3.3.10. Let G be a locally compact Hausdor� second countable étale groupoid.

Cc(G(0)) is a sub-∗-algebra of Cc(G). Moreover, Cc(G(0)) is commutative with product given

by the pointwise multiplication and involution de�ned by f*(x) = f(x).

63

Page 66: Characterization of Extremal KMS States on Groupoid C

Proof. From Proposition 3.2.12 it follows that G(0) is open. Then Cc(G(0)) is a subspace of

Cc(G) as described in Notation 3.3.2.

• f* ∈ Cc(G(0)) if f ∈ Cc(G(0))

Let f ∈ Cc(G(0)). Then f* ∈ Cc(G) by Theorem 3.3.9. Let g ∈ G such that f(g) 6= 0.

Then f*(g) = f(g−1) 6= 0. Thus g−1 ∈ G(0) by assumption. Then g = g−1 ∈ G(0).

Therefore f* is also supported on G(0).

• f1 · f2 = f1f2 ∈ Cc(G(0))

Now let f1, f2 ∈ Cc(G(0)). Then f1 · f2 ∈ Cc(G) by by Theorem 3.3.9. Given g ∈ G, we

have

(f1 · f2)(g) =∑g1g2=g

f1(g1)f2(g2).

Suppose there are g1, g2 such that g = g1g2 and f1(g1)f2(g2) 6= 0. Then g1, g2 ∈ G(0).

Hence g ∈ G(0) and g = g1 = g2. Therefore, (f1 · f2)(g) = f1(g)f2(g).

Therefore Cc(G(0)) is commutative.

Proposition 3.3.11. Let G be a locally compact Hausdor� second countable étale groupoid.

For each f ∈ Cc(G), there is a constant Kf ≥ 0 such that ‖π(f)‖ ≤ Kf for every ∗-

representation π : Cc(G) → B(H) of Cc(G) on a Hilbert space H. If f is supported on an

open bisection, we can take Kf = ‖f‖∞.

Proof. Suppose π is a ∗-representation. Then π|Cc(G(0)) is a ∗-representation of the commu-

tative ∗-algebra Cc(G(0)), and so ‖π(h)‖ ≤ ‖h‖∞ for every h ∈ Cc(G(0)).

Let f ∈ Cc(G). There are f1, . . . , fn with f =∑n

i=1 fi such that each fi ∈ Cc(Ui) and Uiis an open bisection. Fix i, hence fi* ∈ Cc(U−1) and therefore fi* · fi ∈ Cc(U−1U). However,

64

Page 67: Characterization of Extremal KMS States on Groupoid C

UU−1 = s(U). In fact,

U−1U = {gh : g ∈ U−1, h ∈ U , s(g) = r(h)}

= {g−11 g2 : g1, g2 ∈ U , r(g1) = r(g2)}

= {g−1g : g ∈ U} since U is an open bisection

= s(U).

Thus fi*fi ∈ Cc(s(U)). So

‖π(fi)‖2 = ‖π(fi* · fi)‖ ≤ ‖fi* · fi‖∞ = ‖fi‖2.

Let Kf =∑n

i=1 ‖fi‖. Applying triangle inequality, we have ‖π(f)‖ ≤ Kf .

Proposition 3.3.12. Let f ∈ Cc(G) such that f 6= 0. There exists a ∗-representation of

Cc(G) such that π(f) 6= 0.

Proof. Let x ∈ G(0) such that f(hx) 6= 0 for some hx ∈ Gx. Since Gx is countable,

`2(Gx) =

{{ξg}g∈Gx : ξg ∈ C,

∑g∈Gx

|ξg|2 <∞

}

is a Hilbert space with inner product given by

〈ξ, ζ〉 =∑g∈Gx

ξgζg.

De�ne πx : Cc(G)→ B(`2(Gx)) by

(πx(f1)ξ)g =∑

h1h2=g

f1(h1)ξh2 .

65

Page 68: Characterization of Extremal KMS States on Groupoid C

Note that

(πx(f1)ξ)g =∑

h∈Gs(g)

f1(gh−1)ξh =∑

h∈Gr(g)f1(h)ξh−1g,

making the change of variables h2 = h, h1 = gh−1 in the �rst sum, and h1 = h, h2 = h−1g

in the second sum. Note that πx is linear.

• πx is well-de�ned.

We will prove that the image of πx is in B(`2(Gx)). Since πx is linear, it is su�cient

to show that πx(f1) ∈ B(`2(Gx)) for every f ∈ Cc(U) such that U ⊂ G is an open

bisection.

Let U ⊂ G be an open bisection, f1 ∈ Cc(U). Let L denote the set of g ∈ Gx such that

there exists h ∈ Gr(g) satisfying f(h) 6= 0. h is unique for every g ∈ L and it will be

denoted by hr(g). Then, for every ξ ∈ `2(Gx), g ∈ L,

(πx(f1)ξ)g = f1(hr(g))ξ(hr(g))−1g.

Note that (πx(f1)ξ)g = 0 if g ∈ Gx \ L.

Suppose there are g1, g2 ∈ L such that (hr(g1))−1g1 = (hr(g2))−1g2. Then s(hr(g1)) =

s(hr(g2)). Since hr(g1), hr(g2) ∈ U , we have hr(g1) = hr(g2). Then g1 = g2. Since hr(g)g ∈

Gx, the family {ξhr(g)g}g∈L has distinct elements. Hence,

‖πx(f1)ξ‖2 =∑g∈Gx

|(πx(f1)ξ)g|2

=∑g∈L

|(πx(f1)ξ)g|2

=∑g∈L

|f1(hr(g))ξ(hr(g))−1g|2

≤ ‖f1‖2∞

∑g∈L

|ξ(hr(g))−1g|2

66

Page 69: Characterization of Extremal KMS States on Groupoid C

≤ ‖f1‖2∞

∑h∈Gx

|ξh|2 , since {ξhr(g)g}g∈L has distinct elements,

≤ ‖f1‖2∞‖ξ‖2.

Therefore ‖πx(f1)‖ ≤ ‖f1‖∞ <∞.

• πx(f) 6= 0.

Let ξ ∈ `2(Gx) such that ξx = 1 and ξl = 0 if l 6= x. Let g ∈ Gx such that f(g) 6= 0.

Then, by de�nition of ξ,

(πx(f)ξ)g =∑

h∈Gr(g)f(h)ξh−1g = f(g)ξg−1g = f(g)ξx = f(g) 6= 0.

• πx(f1 · f2) = πx(f1)πx(f2)

Let f1, f2 ∈ Cc(G). Then,

[πx(f1)(πx(f2)ξ)]g =∑g1h=g

f1(g1)(πx(f2)ξ)h

=∑g1h=g

f1(g1)∑g2g3=h

f2(g2)ξg3

=∑

g1g2g3=g

f1(g1)f2(g2)ξg3

=∑hg3=g

( ∑g1g2=h

f1(g1)f2(g2)

)ξg3

=∑hg3=g

(f1 · f2)(h)ξg3

= [πx(f1 · f2)ξ]g .

• πx(f1*) = πx(f1)*

67

Page 70: Characterization of Extremal KMS States on Groupoid C

Let f1 ∈ Cc(G), ξ ∈ `2(Gx). Then

〈ξ, πx(f1*)ξ〉 =∑g∈Gx

ξg[πx(f1*)]g

=∑g∈Gx

ξg∑h∈Gx

f1*(gh−1)ξh

=∑g∈Gx

ξg∑h∈Gx

f1(hg−1)ξh

=∑h∈Gx

(∑g∈Gx

f1(hg−1)ξg

)ξh

=∑h∈Gx

[πx(f1)ξ]h ξh

= 〈πx(f1)ξ, ξ〉.

Therefore πx(f1*) = πx(f1)*.

It follows that πx is a ∗-representation of Cc(G) such that πx(f) 6= 0.

Theorem 3.3.13. AssumeG is a locally compact Hausdor� second countable étale groupoid.

There exists a C*-algebra C*(G) such that Cc(G) is dense in C*(G) and the norm on C*(G)

satis�es

‖f‖ = sup{‖π(f)‖: π : Cc(G)→ B(Hπ) is a ∗-representation of Cc(G)},

for every f ∈ Cc(G).

Proof. For every f ∈ Cc(G), Proposition 3.3.11 shows that the set

{‖π(f)‖ : π is a ∗-representation of Cc(G)}

is bounded above, and it is nonempty because of the zero representation. So we can de�ne

68

Page 71: Characterization of Extremal KMS States on Groupoid C

ρ : Cc(G)→ [0,∞) by

ρ(f) = sup{‖π(f)‖ : π : Cc(G)→ B(Hπ) is a ∗-representation}.

ρ is a norm on Cc(G). In fact, given λ ∈ C, f ∈ Cc(G),

ρ(λf) = supπ‖π(λf)‖ = |λ| sup

π‖π(f)‖ = |λ|ρ(f).

Given f1, f2 ∈ Cc(G),

ρ(f1 + f2) = supπ‖π(f1 + f2)‖ ≤ sup

π‖π(f1)‖+ sup

π‖π(f2)‖ = ρ(f1) + ρ(f2).

Given f ∈ Cc(G) such that f 6= 0, ρ(f) > 0 by Proposition 3.3.12.

The norm is submultiplicative. Given f1, f2 ∈ Cc(G),

ρ(f1 · f2) = supπ‖π(f1 · f2)‖ = sup

π‖π(f1)π(f2)‖ ≤ sup

π‖π(f1)‖ sup

π‖π(f2)‖ = ρ(f1)ρ(f2).

Given f ∈ Cc(G),

ρ(f*) = supπ‖π(f*)‖ = sup

π‖π(f)*‖ = sup

π‖π(f)‖ = ρ(f).

Moreover, ρ satis�es the C*-identity. Indeed, given f ∈ Cc(G),

ρ(f*f) = supπ‖π(f*f)‖ = sup

π‖π(f)*π(f)‖ = sup

π‖π(f)‖2 = sup

π‖π(f)‖ = ρ(f)2.

So we de�ne C*(G) to be the completion of Cc(G) with respect to the norm ρ. C*(G) is a

C*-algebra.

De�nition 3.3.14. Given a locally compact Hausdor� second countable étale groupoid G,

C*(G) is called the full C*-algebra of G.

Remark 3.3.15. In this thesis we also say C*(G) is the groupoid C*-algebra for G. However,

69

Page 72: Characterization of Extremal KMS States on Groupoid C

this is not the unique C*-algebra de�ned as the completion of Cc(G). For example, in

[9] the reduced C*-algebra is de�ned as the closure of Cc(G) with respect to the norm

‖f‖ = ‖πλ(f)‖, where πλ is a ∗-representation of Cc(G) called the regular representation of

Cc(G).

Lemma 3.3.16. Let G be a locally compact second countable Hausdor� étale groupoid.

Then C0(G(0)) is a sub-C*-algebra of C*(G) and the norm on C0(G(0)) is the uniform norm.

Moreover, Cc(G(0)) is dense in C0(G(0)).

Proof. From Proposition 3.2.12, we have that G(0) is clopen in G. Moreover, G(0) is an open

bisection because r and s are injective on G(0). From Lemma 3.3.10 we have that Cc(G(0))

is a sub-∗-algebra with product given by pointwise multiplication and involution de�ned by

f*(x) = f(x).

Let h ∈ Cc(G(0)). It follows from Proposition 3.3.11 that ‖π(h)‖ ≤ ‖h‖∞ for every

representation π of Cc(G). Then ‖h‖ ≤ ‖h‖∞.

We will show that ‖h‖ = ‖h‖∞ for every h ∈ Cc(G(0)). Given x ∈ G(0), let πx : Cc(G)→

B(`2(Gx)) be the ∗-representation as in the proof of Proposition 3.3.12. Then, for every

h ∈ Cc(G(0)), ξ ∈ `2(Gx), we have

[πx(h)ξ]g =∑g1g2=g

h(g1)ξg2 = h(r(g))ξg,

because h vanishes outside G(0). Let ζ ∈ `2(Gx) such that ζg = 0 if g 6= x and ζx = 1. Then

‖ζ‖ = 1 and

‖πx(h)‖2 ≥ ‖πx(h)ζ‖2 =∑g∈Gx

|πx(h)ζg|2 = |πx(h)|2 = |h(r(x))|2 = |h(x)|2.

Then,

‖h‖ = sup{‖π(h)‖ : π : Cc(G)→ B(Hπ) is a ∗-representation}

70

Page 73: Characterization of Extremal KMS States on Groupoid C

≥ supx∈G(0)

|πx(h)| = supx∈G(0)

|h(x)| = ‖h‖∞.

Then ‖h‖ = ‖h‖∞. Recall that C0(G(0)) is the closure of Cc(G(0)) with respect to the

norm ‖ · ‖∞. Therefore, C0(G(0)) is a sub-C*-algebra of C*(G).

Example 3.3.17. Let X be the groupoid of Example 3.2.5 and assume that X is locally

compact Hausdor� second countable. Then the operations on Cc(X) are the pointwise

multiplication and the complex conjugate by Lemma 3.3.10. It follows from Lemma 3.3.16

that C0(X) = C*(X) and the norm on this C*-algebra is the uniform norm.

Now we de�ne the inductive limit topology. Later we show that convergence with respect

to the inductive limit topology on Cc(G) implies convergence in the norm of C*(G). This

de�nition can be found in [10]. Then we will prove that C*(G) is separable.

De�nition 3.3.18. SupposeX is a locally compact Hausdor� second countable space. Given

a sequence {fn}n∈N on Cc(X) and f ∈ Cc(X), we say that fn → f with respect to the

inductive limit topology if, and only if, fn → f uniformly and there exists a compact set K

in X such that, eventually, all the fn and f vanish outside K. Given a topological space Y ,

we will say that a function F : Cc(X) → Y is continuous in the inductive limit topology if

F (fn)→ F (f) whenever fn → f with respect to the inductive limit topology.

Lemma 3.3.19. Let G be a locally compact Hausdor� second countable Hausdor� étale

groupoid. Let {fi}i∈N be a sequence in Cc(G) such that fi → f with respect to the inductive

limit topology. Then fi → f in C*(G).

Proof. Let K be a compact set such that fi eventually vanishes outside K. Let U1, . . . ,Unbe open bisections which cover K. Let p1, . . . , pn be a partition of unit subordinate to the

open cover.

Let π : Cc(G)→ B(Hπ) be a ∗-representation of Cc(G). Fix j = 1, . . . , n. Then pjf and

each pjfi are supported on the open bisection Uj. Hence, by Proposition 3.3.11, we have for

71

Page 74: Characterization of Extremal KMS States on Groupoid C

every i,

‖π(pjfi)− π(pjf)‖ ≤ ‖pjfi − pjf‖∞ ≤ ‖fi − f‖∞.

By taking the supremum on π, we have ‖pjfi − pjf‖ ≤ ‖fi − f‖∞. Since fi converges to f

uniformly, it follows that pjfi → pjf in C*(G). Since {pj} is a partition of unit subordinate

to the open cover of K, and the sequence fi is eventually supported on K, it follows that

fi → f in C*(G).

We state the Stone-Weierstrass theorem below, which can be found in [11] . This theorem

will be used to prove that C*(G) is separable.

Theorem 3.3.20. (Stone-Weierstrass theorem for complex-valued functions) Let K be a

compact space, A a subalgebra of C(X) which separates points in K, that is, for every

x1, x2 ∈ K, there is f ∈ A such that f(x1) 6= f(x2). Assume f ∈ A for every f ∈ A.

Moreover, suppose that for every x ∈ K there exists f ∈ A with f(x) 6= 0. Therefore A is

dense in C(K).

Lemma 3.3.21. Let X be a locally compact Hausdor� second countable space. Let U be

an open subset of X with compact closure. Then Cc(U) is separable with respect to the

supremum norm.

Proof. Let F = {Un} be a countable family of open sets such that Un is compact and

U = ∪nUn. Given n,m such that Un ∩ Um = ∅, let fn,m ∈ Cc(U) such that fn,m|Un = 1 and

fn,m|Um = 0.

Let A be the algebra generated by fn,m, and A0 be the set generated by sums and products

of fn,m, and also by multiplication of scalars in Q+ iQ. Note that A0 is countable and dense

in A. Moreover, if f ∈ A, then f ∈ A.

Note that A separates points in U . Let x1, x2 ∈ U . It follows from Propositions 2.4.1 and

2.4.2 that there are U1, U2 ∈ F such that U1 ∩ U2 = ∅, x1 ∈ U1, x2 ∈ U2. Then f1,2(x1) = 1

and f1,2(x2) = 0.

72

Page 75: Characterization of Extremal KMS States on Groupoid C

Therefore, by the Stone-Weierstrass theorem, A0 is dense in Cc(U).

Proposition 3.3.22. Let G be a locally compact Hausdor� second countable étale groupoid.

Then C*(G) is separable.

Proof. Let I be a countable family of open bisections with compact support that covers G.

From Lemma 3.3.21, there exists a countable subset AU of Cc(U) such that AU is dense in

Cc(U) with respect to the supremum norm. Let A0 be the set generated by �nite sums of

elements in ∪U∈IAU . Then A0 is countable.

Let f ∈ Cc(G). There is a �nite family U1, . . . ,Un ∈ I that covers the support of f . Let

p1, . . . , pn be a partition of unit subordinate to U1, . . . ,Un.

Let ε > 0. Given i = 1, . . . , n, pif ∈ Cc(Ui). Then there exists Fi ∈ AUi satisfying

‖Fi − pifi‖∞ < ε/n. It follows from Proposition 3.3.11 that ‖π(Fi) − π(pif)‖ < ε/n, for

every ∗-representation of Cc(G). Then ‖Fi − pif‖ < ε/n.

Let F =∑n

i=1 Fi. Then F ∈ A0. Moreover,

‖F − f‖ =

∥∥∥∥∥n∑i=1

Fi −n∑i=1

pifi

∥∥∥∥∥ ≤n∑i=1

‖Fi − pifi‖ < nε

n= ε.

Therefore A0 is dense in Cc(G). Since Cc(G) is dense in C*(G), it follows that C*(G) is

separable.

73

Page 76: Characterization of Extremal KMS States on Groupoid C

Chapter 4

Renault's Disintegration Theorem

Neshveyev's theorems describe KMS states ϕ on a groupoid C*-algebra. It is possible to

write ϕ(f) as an integral on G(0) with respect to a probability measure µ such that, for each

x, there is a state ϕx on C*(Gxx). Moreover, the family of states ϕx depends on µ.

Recall from the theory of C*-algebras [2], [16] that we can write ϕ(f) = 〈π(f)ξ, ξ〉 for a

representation π : C*(G)→ H, and ξ ∈ H. Analogously, we can write ϕx(f) = 〈πx(f)ξx, ξx〉

for a representation π : C*(Gxx)→ Hx, and ξx ∈ H

A fundamental step in the proof of Neshveyev's theorem is the Renault's disintegration

theorem which shows a relation between π and the family of πx. Moreover, it proves that

we can assume ξ = {ξx}x∈G(0) . In this case we say ξ ∈∫ ⊕G(0)Hxdµ(x).

The results in this chapter can be found in [4] and [10].

4.1 Haar Systems

Given a measure µ on G(0), it is possible to de�ne a measure ν on G if we have a family

of measures λx supported on Gx. The family {λx}x∈G(0) is called a Haar system. A Haar

system is a generalization of the notion of Haar measures on groups.

If we �x the Haar system, we can de�ne a family of measures λx supported on Gx and,

74

Page 77: Characterization of Extremal KMS States on Groupoid C

with this family, we can construct another measure ν−1 on G. When ν, ν−1 are equivalent,

we say µ is quasi-invariant.

Given a state ϕ on C*(G), it follows from Neshveyev's �rst theorem that there is corre-

sponding probability measure µ on G(0). In addition, if this state is KMS, then this measure

is necessarily quasi-invariant with respect to the Haar system given by counting measures

λx on Gx.

The results in this section are based on [4] and [10].

De�nition 4.1.1. Let G be a locally compact Hausdor� groupoid, a (left) Haar system

{λx}x∈G(0) for G is a family of Radon measures on G, such that the following conditions

hold:

(i) supp(λx) = Gx for every x ∈ G(0);

(ii) (continuously varying) for f ∈ Cc(G), the function

x 7→∫G

f(g)dλx(g)

is in Cc(G(0));

(iii) (left invariance) for f ∈ Cc(G), h ∈ G,

∫G

f(hg)dλs(h)(g) =

∫G

f(g)dλr(h)(g).

Now we prove that Haar systems are a generalization of Haar measures, as de�ned in

[6]. A topological group is a group G endowed with a topology such that its operations are

continuous functions. Given a locally compact Hausdor� group G, a (left) Haar measure is

a non-zero Radon Borel measure µ on G satisfying

µ(gA) = µ(A) for every g ∈ G and A ⊂ G measurable.

75

Page 78: Characterization of Extremal KMS States on Groupoid C

Lemma 4.1.2. Let G be a locally compact Hausdor� topological group. Let µ be a Radon

measure on G. Then µ is a Haar measure if, and only if, {λx}x∈G(0) is a Haar system assuming

λ1 = µ.

Proof. Note that a group G is a groupoid such that G(0) = {1}.

Assume µ is a Haar measure. First we show that supp(µ) = G1 = G. We prove that the

properties of Haar system hold for {λx}x∈G(0)

(i) supp(λ1) = G1 = G

Suppose there exists an open non-empty set U ⊂ G such that µ(U) = 0. Let h ∈ U .

Then h−1U is an open neighborhood of 1 and µ(h−1U) = µ(U) = 0. Let V = h−1U .

LetK be a compact set. For every g ∈ K, gV is an open neighborhood of g. Then there

are g1, . . . , gn ∈ K such that g1V, . . . , gnV is an open cover of K. However, µ(g1V ) = 0

for every i = 1, . . . , n. Hence µ(K) = 0. Then, by de�nition of Radon measure, we

have

µ(G) = sup{K ⊂ G : K is compact} = 0,

which leads to a contradiction. Therefore, µ is positive on all open subsets of G.

Now let F be the support of µ. Then, by de�nition, F is closed and µ(G \ F ) = 0.

Since G \ F is open, we have G \ F = ∅, that is, G = F .

(ii) Property (ii) in De�nition 4.1.1 since G(0) is singleton.

(iii) Left invariance in De�nition 4.1.1

Let f ∈ Cc(G). We can assume f ≥ 0 without loss of generality. Let A ⊂ G measurable

and a ≥ 0 such that ϕ = aχA ≤ f .

Given h ∈ G, the function g 7→ ϕ(hg) satis�es ϕ(hg) = χh−1A(g) ≤ f(hg) for every

76

Page 79: Characterization of Extremal KMS States on Groupoid C

g ∈ G. Hence,

∫G

ϕ(g)dµ(g) = aµ(A) = aµ(h−1A) =

∫G

ϕ(hg)dµ(g). (4.1)

Equation (4.1) holds for every simple function ϕ ≤ f by linearity of the integral. Then,

by taking the supremum over ϕ ≤ f ,

∫G

f(g)dµ(g) ≤∫G

f(hg)dµ(g).

The other inequality is proven analogously.

Conversely, assume {λx}x∈G(0) is a Haar system with λ1 = µ.

Let U ⊂ G be an open set and let h ∈ G. Note that there is a correspondence between

measurable functions 0 ≤ f ≤ χU and 0 ≤ f ≤ χh−1U given by the relation

f(g) = f(h−1g). (4.2)

In fact, given f satisfying f ≤ χU , g ∈ G such that f(g) 6= 0, then f(h−1g) 6= 0. Hence

h−1g ∈ U . Then g ∈ hU . Therefore 0 ≤ f ≤ χhU . Analogously, given 0 ≤ f ≤ χhU , there is

a unique 0 ≤ f ≤ χU such that (4.2) holds.

Hence, by Lemma 2.4.14,

µ(U) = supf∈Cc(G)0≤f≤χU

∫G

f(g)dµ(g) = supf∈Cc(G)0≤f≤χU

∫G

f(h−1g)dµ(g) = supf∈Cc(G)

0≤f≤χU

∫G

f(g)dµ(g) = µ(hU).

Let A be a Borel set. Since the map g 7→ hg is a homeomorphism, for every open set V ⊃ hA,

there exists a unique open set A ⊂ such that V = hU . Then,

µ(A) = infA⊂UU open

µ(U) = infA⊂UU open

µ(hU) = infhA⊂VV open

µ(V ) = µ(hA).

77

Page 80: Characterization of Extremal KMS States on Groupoid C

Therefore µ is a Haar measure.

Now we prove that the family λx of counting measures on Gx is a Haar system. In fact,

this will be the Haar system used in Chapter 5.

Proposition 4.1.3. Let G be a locally compact étale groupoid. If λx is the counting measure

on Gx, then {λx}x∈G(0) is a left Haar system.

Proof. Each Gx is countable, then the counting measure λx is well de�ned. We show the

properties of De�nition 4.1.1 hold.

(i) Given x ∈ G(0), Gx is closed in G. Moreover, by Proposition 3.2.11, Gx is countable.

Then Gx is the support of λx.

Since G is étale, for every g ∈ Gx there exists an open bisection Ug such that Ug∩Gx =

{g}. Let K ⊂ G be compact. Then K ′ = K ∩ Gx is compact and λx(K) = λx(K ′).

Moreover, there are g1, . . . gn ∈ Gx such that K ′ ⊂ Ug1 ∪ . . . ∪ Ugn . Hence,

λx(K) = λx(K ′) ≤ λx(Ug1 ∪ . . . ∪ Ugn) ≤n∑i=1

λx(Ugi) =n∑i=1

λx(gi) = n.

Then λx is �nite on compact subsets. Therefore, λx is Radon by Proposition 2.4.10.

(ii) Let f ∈ Cc(G). First assume f ∈ Cc(U), where U is an open bisection. Let K =

supp(f). De�ne the open set V = r(U) and the compact set L = r(K). Hence, we can

de�ne the function f ∈ Cc(r(V )) by

f(y) =

f(r|−1U (y)), if y ∈ L,

0, otherwise.

Note that, for every y ∈ G(0),

f(y) =

∫G

f(g)dλy(g).

78

Page 81: Characterization of Extremal KMS States on Groupoid C

Now let f ∈ Cc(G) arbitrary with support K = supp(f). Since K is compact and G

is étale, K has a �nite open cover U1, . . . ,Un of open bisections. By Proposition 3.3.5,

there are f1, . . . , fn such that f = f1 + . . .+fn and each fi ∈ Cc(Ui). As we have shown,

for every i = 1, . . . , n, the function

y 7→∫G

fi(g)dλy(g)

is continuous and compactly supported. Then the function

y 7→∫G

f(g)dλy(g)

is continuous and compactly supported.

(iii) Given f ∈ Cc(G), h ∈ G,

∫G

f(hg)dλs(h)(g) =∑

g∈Gs(h)f(hg).

We can use the change of variables g = hg because the function from Gs(h) to Gr(h)

de�ned by g 7→ hg is injective. Then we have g ∈ Gr(h) and

∑g∈Gs(h)

f(hg) =∑

g∈Gr(h)f(g) =

∫G

f(g)dλr(h)(g).

Given a Haar system {λx}, we de�ne for x ∈ G(0) the measure λx by λx(E) = λx(E−1),

for every E ⊂ G measureable.

Lemma 4.1.4. If G is a locally compact Hausdor� étale groupoid, and λx is the counting

measure on Gx, then λx is the counting measure on Gx.

Proof. Given a set A, |A| denotes the number of elements in this set. Let E ⊂ G measurable,

79

Page 82: Characterization of Extremal KMS States on Groupoid C

then

λx(E) = λx(E−1)

= |{g : g ∈ E−1 ∩Gx}|

= |{h−1 : h ∈ E ∩Gx}|, by the change of variables h−1 = g,

= |{h : h ∈ E ∩Gx}|,

since h 7→ h−1 is a bijection. Then λx is the counting measure on Gx.

De�nition 4.1.5. Let G be a locally compact Hausdor� groupoid with a Haar system

{λx}x∈G(0) . Given a Radon measure µ on G(0), we de�ne the induced measures on G by

ν(E) =

∫G(0)

λx(E)dµ(x), ν−1(E) =

∫G(0)

λx(E)dµ(x),

for every Borel set E. Or equivalently, for every f ∈ Cc(G),

∫f(g)dν(g) =

∫G(0)

∫Gxf(g)dλx(g)dµ(x),

∫f(g)dν−1(g) =

∫G(0)

∫Gx

f(g)dλx(g)dµ(x).

We denote ν =∫G(0) λ

xdµ(x), ν−1 =∫G(0) λxdµ(x).

If G a is locally compact Hausdor� étale groupoid and endowed with a Haar system {λx}

such that each λx is a counting measure on Gx, then we denote the induced measures ν and

ν−1 by µr and µs. In this case,

∫f(g)dµr(g) =

∫G(0)

∑g∈Gx

f(g)dµ(x),∫f(g)dµs(g) =

∫G(0)

∑g∈Gx

f(g)dµ(x).

De�nition 4.1.6. Suppose µ is a measure on G(0). We say µ is quasi-invariant if ν and ν−1

are equivalent measures. In this case we take ∆ : G → (0,∞) to be the Radon-Nikodym

derivative dν/dν−1.

Remark 4.1.7. It follows from [10, Theorem 3.72] that it is possible to choose ∆ to be a

80

Page 83: Characterization of Extremal KMS States on Groupoid C

homomorphism from G to (0,∞), where (0,∞) is a group with respect to the product.

Proposition 4.1.8. 1 Let G be a locally compact Hausdor� étale groupoid with Haar system

{λx} given by counting measures λx on Gx. If µ is a quasi-invariant measure on G(0), then

for µ-a.e. x and all g ∈ Gxx, we have ∆(g) = 1.

Proof. Let G′ be the isotropy bundle de�ned by G′ = ∪x∈G(0)Gxx. Note that G

′ is closed in G.

In fact, if {gi} is a sequence in G′ converging to some g in G, it follows from the continuity

of r and s that r(gi) → r(g) and s(gi) → s(g). Since r(gi) = s(gi) for every i, we have

r(g) = s(g), that is, g ∈ G′. Therefore G′ is measurable.

Let f be a positive measurable function whose support lies in G′. Then, by assumption,

∫G

f(g)dµr(g) =

∫G

f(g)∆(g)dµs(g). (4.3)

It follows from the de�nition of µr that∫G

f(g)dµr(g) =

∫G(0)

∑g∈Gx

f(g)dµ(x)

=

∫G(0)

∑g∈Gxx

f(g)dµ(x), (4.4)

since f is supported of G′. Note that f∆ is also supported on G′. Then, by de�nition of µs,

we have

∫G

f(g)∆(g)dµs(g) =

∫G(0)

∑g∈Gx

f(g)∆(g)dµ(x)

=

∫G(0)

∑g∈Gxx

f(g)∆(g)dµ(x). (4.5)

1Thanks Frausino!

81

Page 84: Characterization of Extremal KMS States on Groupoid C

Applying (4.4) and (4.5) in (4.3), we have

∫G(0)

∑g∈Gxx

(∆(g)− 1)f(g)dµ(x) = 0

De�ne B+ = {g ∈ G′ : ∆(g) ≥ 1} and B− = {g ∈ G′ : ∆(g) ≤ 1}. If we choose f = χB+ ,

we have

∫G(0)

∑g∈Gxx∩B+

(∆(g)− 1)dµ(x) = 0.

Since the function x 7→∑

g∈Gxx∩B+(∆(g)− 1) is non-negative, it follows that for µ-a.e. x,

∑g∈Gxx∩B+

(∆(g)− 1) = 0.

From the de�nition of B+, for µ-a.e. x and all g ∈ Gxx, if ∆(g) ≥ 1, it follows that ∆(g) = 1.

Analogously for B−, for µ-a.e. x and all g ∈ Gxx, if ∆(g) ≤ 1, it follows that ∆(g) = 1.

Therefore, for µ-a.e. x and all g ∈ Gxx, we have ∆(g) = 1.

Remark 4.1.9. When we study Neshveyev's second theorem, we will consider measures

µ such that ∆(g) = ec for a continuous R-valued 1-cocycle c : G → R. It follows from

Proposition 4.1.8 that for µ-a.e. x ∈ G(0), all g ∈ Gxx, we have c(g) = 0. In Remark 6.3.22

on page 200, we prove this result for extremal eβF -measures using the properties of the

Renault-Deaconu groupoid.

4.2 Borel Hilbert Bundles

Given a group G, [18] de�nes a unitary representation of G as a pair (L,H), where H is a

complex Hilbert space and L is homomorphism from G to the group of unitary operators on

H, usually written U(H), with product as group operation. We will denote Lg as the image

of g in G under the map L.

82

Page 85: Characterization of Extremal KMS States on Groupoid C

We want to extend this notion to groupoids. By de�nition of groupoids, two elements are

not necessarily composable. In order to de�ne a unitary representation of groupoids, we will

consider unitary operators Lg : Hs(g) → Hr(g), where {Hx}x∈G(0) is a family of Hilbert spaces

indexed by G(0) and satisfying certain conditions.

Given this family of Hilbert spaces, we de�ne a Hilbert space L2(X ∗ H, µ), also denoted

by∫ ⊕XHxdµ(x). By Renault's Disintegration Theorem, a representation π : Cc(G) → H

corresponds to a groupoid representation. Moreover, H can be identi�ed with∫ ⊕XHxdµ(x).

In this section we use results from [27], [10], [4], and [15].

Let H = {Hx}x∈X be a collection of separable (nonzero) complex Hilbert spaces indexed

by a locally compact Hausdor� second countable space X. Then the total space is de�ned

by

X ∗ H = {(x, h) : h ∈ Hx},

and we let p : X ∗ H→ X be de�ned by p(x, h) = x.

Remark 4.2.1. The total space X ∗ H is de�ned in [27] assuming X is analytic. However,

we will always assume that X is locally compact Hausdor� second countable without loss of

generality. Theorem 5.3 in [13] shows that every locally compact Hausdor� second countable

space is analytic. More details about analytic and Polish spaces can be found in [6].

A section of X ∗H is a function f : X → X ∗H such that (p ◦ f)(x) = x for each x ∈ X.

The total space can be seen as the union of Hilbert spaces such that each Hx is glued to a

point x ∈ X. A section maps x to some point in {x} ×Hx, as shown in Figure 4.1.

Example 4.2.2. Let Hx = C for every x ∈ X. Then X ∗ H = X × C. In this case, each

section f corresponds to a complex-valued function f such that f(x) = (x, f(x)). Then we

can identify f with f without loss of generality.

Given a Borel measure µ on X, we want to de�ne a space of square-integrable sections,

which we denote by L2(X ∗ H, µ), such that it is a Hilbert space and every section f ∈

83

Page 86: Characterization of Extremal KMS States on Groupoid C

x

f(x)

Hx

y

f(y)

Hy

z

f(z)

Hz

X

Figure 4.1: X ∗H can be seen as a union of Hilbert spaces Hx such that every Hx correspondsto a point x in X. A section is a function that maps x to an element of {x} ×Hx.

L2(X ∗ H, µ) satis�es

∫X

‖f(x)‖2dµ(x) <∞.

Here we identify functions that are equal µ-a.e.

If we endow X ∗ H with a Borel structure, we would naturally be interested in a subset

of measurable sections, denoted by Sec(X ∗ H) such that

x 7→ 〈f1(x), f2(x)〉 is measurable for every f1, f2 ∈ Sec(X ∗ H), (4.6)

since we want to de�ne the inner product on L2(X ∗ H) as

〈f1, f2〉 =

∫X

〈f1(x), f2(x)〉dµ(x).

We will replace condition (4.6) by simpler conditions which consider a countable family of

sections. Later we will see that for every pair of measurable functions, (4.6) holds.

De�nition 4.2.3. Let H = {Hx}x∈X be a family of separable Hilbert spaces indexed by a

locally compact Hausdor� second countable space X. Then (X ∗ H, p) is an analytic Borel

84

Page 87: Characterization of Extremal KMS States on Groupoid C

Hilbert bundle if X ∗ H has a Borel structure such that

(a) For every E ⊂ X, p−1(E) is a Borel in X ∗ H if, and only if, E is Borel in X;

(b) There is a sequence of sections {fn} such that

(i) the map fn : X ∗ H→ C de�ned by fn(x, h) = 〈fn(x), h〉Hx is Borel for each n,

(ii) for each n and m, x 7→ 〈fn(x), fm(x)〉Hx is Borel, and

(iii) the functions {fn} and p separate points of X ∗ H.

The sequence {fn} is called a fundamental sequence for (X ∗ H, p).

Remark 4.2.4. Given an analytic Borel Hilbert bundle (X, p), we let Sec(X ∗ H) denote

the set of sections f : X → X ∗ H such that

x 7→ 〈f(x), fn(x)〉

is Borel for every fn. It follows from [27, Remark F.3] that a section f of X ∗ H is in

Sec(X ∗ H) if, and only if, it is Borel.

Remark 4.2.5. Note that given a section f of X ∗ H, f(x) is an element of {x} × Hx.

Although this is an abuse of notation, we will identify {x} × Hx with Hx. Hence, the inner

product 〈f(x), h〉 is well-de�ned for h ∈ Hx. We can also write 〈f1(x), f2(x)〉 for sections

f1, f2 without further comments.

Example 4.2.6. Let X be a locally compact Hausdor� second countable space. Let H be

a separable Hilbert space. De�ne Hx = H for each x ∈ X. Then X ∗ H = X ×H. Endow

X ×H with the product topology. Given E ⊂ X, p−1(E) = E ×H. Thus p−1(E) is Borel

if, and only if, E is Borel. Therefore property (a) of De�nition 4.2.3 holds.

Assume {en} is an orthonormal basis for H. De�ne a sequence of sections of X ∗ H by

fn(x) = (x, en) ∈ {x} ×Hx. We will prove that {fn} is a fundamental sequence:

85

Page 88: Characterization of Extremal KMS States on Groupoid C

(i) fn : X ×H → C is de�ned by fn(x, h) = 〈fn(x), h〉 = 〈en, h〉. fn is continuous, then it

is Borel.

(ii) Given n,m, the function x 7→ 〈fn(x), fm(x)〉 = 〈en, em〉 is constant, therefore this

function is Borel.

(iii) Let (x, h), (y, k) ∈ X ×H. If x 6= y, then p(x, h) 6= p(y, k) by de�nition of p. Assume

x = y and h 6= k. There exists en such that 〈en, h〉 6= 〈en, k〉. Thus, fn(x, h) 6= fn(y, k).

Therefore (X ×H, p) is an analytic Borel Hilbert bundle.

Example 4.2.7. Given a locally compact Hausdor� second countable space X, let X =

X∞ ∪ X1 ∪ X2 ∪ . . . be a Borel partition of X, i.e., every Xd is Borel and the collection of

Xd is disjoint.

For every d = ∞, 1, 2, . . ., let H(d) be a Hilbert space of dimension d and basis {edn}n=dn=1.

Then

X ∗ H =d=∞⋃d=1

Xd ×H(d).

EndowX∗H with a Borel structure such that E ⊂ X∗H is Borel if, and only if, E∩(Xd×H(d))

is Borel for all d.

De�ne, for every n =∞, 1, 2, . . ., the section fn such that

fn(x) =

edn, if x ∈ Xd and 1 ≤ n ≤ d,

0, otherwise.

We prove that (X ∗ H, p) is an analytic Borel Hilbert bundle. Recall from Example 4.2.6

that (Xd×H(d), p|Xd) is an analytic Borel Hilbert bundle. Moreover, {fn|Xd} is a fundamental

sequence. Here we assume fn|Xd : Xd → Xd ×H(d) and p|Xd×H(d) : Xd ×H(d) → Xd without

loss of generality.

Now we show that all properties in De�nition 4.2.3 hold for (X ∗ H, p).

86

Page 89: Characterization of Extremal KMS States on Groupoid C

(a) Let E ⊂ X. Then

p−1(E) =d=∞⋃d=1

(E ∩Xd)×H(d). (4.7)

Assume E is Borel and �x d =∞, 1, 2, . . .. Then E ∩Xd is a Borel subset of Xd. Since

(Xd ×H(d), p|Xd×H(d)) is a Borel Hilbert bundle, it follows that

p|−1Xd×H(d)(E ∩Xd) = (E ∩Xd)×H(d) is Borel.

Note that d is arbitrary. From (4.7) we have that p−1(E) is Borel.

Conversely, assume p−1(E) is Borel. Fix d =∞, 1, 2, . . .. Then p−1(E) ∩ (Xd ×H(d)) is

Borel in Xd×H(d). However, from (4.7), p−1(E)∩ (Xd×H(d)) = (E∩Xd)×H(d). Hence

E ∩Xd is Borel in X. Since d is arbitrary, it follows that E is Borel.

(b) (i) Fix n. fn|Xd is Borel for every n, d from Example 4.2.6.

Since fn = fn|X∞ + fn|X1 + fn|X2 + . . ., it follows that fn is Borel.

(ii) Fix d and let n,m ∈ {∞, 1, 2, . . .}. {fn|Xd} is a fundamental sequence for the Borel

Hilbert bundle (Xd ×H(d), p|Xd×H(d)). Then the function from Xd to C de�ned by

x 7→ 〈fn(x), fm(x)〉 is Borel.

Since d is arbitrary, the function from X to C de�ned by x 7→ 〈fn(x), fm(x)〉 is

Borel.

(iii) Let (x, h), (y, k) ∈ X ∗ H. If x 6= y, p(x, h) = p(y, k).

Assume x = y. Let d be such that x ∈ Xd. Since (Xd × H(d), p|Xd×H(d)) is an

analytic Borel Hilbert bundle with fundamental sequence {fn|Xd×H(d)}n=∞n=1 , there

exists n such that fn|Xd×H(d)(x, h) 6= fn|Xd×H(d)(y, h). Thus fn(x, h) 6= fn(x, k).

Therefore (X ∗ H, p) is an analytic Borel Hilbert bundle.

Given f1, f2 ∈ Sec(X ∗H), [27, Proposition F.6] shows that the function x 7→ 〈f1(x), f2(x)〉

is Borel. If µ is a Borel measure on X, we can de�ne the normed vector space L2(X ∗ H, µ)

87

Page 90: Characterization of Extremal KMS States on Groupoid C

formed by the quotient of

L2(X ∗ H, µ) = {f ∈ Sec(X ∗ H) : x→ ‖f(x)‖2 is µ-integrable},

where functions agreeing µ-almost everywhere are identi�ed. L2(X ∗H, µ) is a Hilbert space

with inner product de�ned by

〈f1, f2〉 =

∫X

〈f1(x), f2(x)〉dµ(x).

Here we also denote L2(X ∗ H) by∫ ⊕XHxdµ(x).

Given a Borel Hilbert bundle X ∗ H, we can de�ne a groupoid, called the isomorphism

bundle, consisting of unitary operators between Hilbert spaces in the family {Hx}.

De�nition 4.2.8. GivenX a locally compact Hausdor� second countable space, let (X∗H, p)

be an analytic Borel Hilbert bundle. The isomorphism bundle of X ∗ H is the set

Iso(X ∗ H) = {(x, L, y) such that L : Hy → Hx is unitary}

endowed with the smallest Borel structure such that, for all f1, f2 ∈ Sec(X ∗ H),

φf1,f2(x, L, y) = 〈Lf1(y), f2(x)〉 (4.8)

de�ne Borel functions from Iso(X ∗ H) to C.

Lemma 4.2.9. Iso(X∗H) is a groupoid whose units are de�ned by (x, idHx , x). Two elements

(x, L, y) and (y′,M, z) are composable if y = y′. In this case, we de�ne their product by

(x, L, y)(y,M, z) = (x, LM, z). The inverse in Iso(X∗H) is de�ned by (x, L, y)−1 = (y, L*, x).

The range and source maps are de�ned by r(x, L, y) = (x, idHx , x) and s(x, L, y) = (y, idHy , y)

Proof. Given x ∈ X, idHx is unitary. Hence (x, idHx , x) ∈ Iso(X∗H). Then r, s are surjective.

Note that we can identify (x, idHx , x) with x without loss of generality. Thus we will assume

88

Page 91: Characterization of Extremal KMS States on Groupoid C

X = Iso(X ∗ H)(0).

Now we show that the product in Iso(X ∗ H) is well-de�ned. In fact, given (x, L, y),

(y,M, z) ∈ Iso(X ∗ H), then LM : Hz → Hx is a unitary operator. Thus (x, LM, z) ∈

Iso(X ∗H). The inverse is well-de�ned as well, since (x, L, y)−1 = (y, L*, x) by de�nition and

L* : Hx → Hy is unitary.

Now we prove the conditions in De�nition 3.1.1. Let g = (x, L, y), h = (y,M, z) and

k = (z,N,w) be elements of Iso(X ∗ H). Then

(i) s(gh) = s((x, L, y)(y,M, z)) = s(x, LM, z) = z = s(y,M, z) = s(h), and

r(gh) = r((x, L, y)(y,M, z)) = r(x, LM, z) = x = r(x, L, y) = r(g).

(ii) x = r(x, idHx , x) = r(x), and x = s(x, idHx , x) = s(x).

(iii) gs(g) = gy = (x, L, y)(y, idHy , y) = (x, LidHy , y) = (x, L, y) = g, and

r(g)g = xg = (x, idHx , x)(x, L, y) = (x, idHxL, y) = (x, L, y) = g.

(iv) (gh)k = [(x, L, y)(y,M, z)] (z,N,w) = (x, LM, z)(z,N,w) = (x, LMN,w), and

g(hk) = (x, L, y) [(y,M, z)(z,N,w)] = (x, L, y)(y,MN,w) = (x, LMN,w).

Then (gh)k = g(hk).

(v) gg−1 = (x, L, y)(y, L*, x) = (x, LL*, x) = (x, idHx , x) = x = r(g), and

g−1g = (y, L*, x)(x, L, y) = (y, L*L, y) = (y, idHy , y) = y = s(g).

Therefore Iso(X ∗ H) is a groupoid.

De�nition 4.2.10. Let G be a locally compact Hausdor� second countable étale groupoid

with Haar system {λx}x∈G(0) . A unitary representation of G is a triple (µ,G(0) ∗ H, L) such

that µ is a quasi-invariant measure on G(0), L = {Lg}g∈G is a family of unitary operators

Lg : Hs(g) → Hr(g), and L : G→ Iso(G(0) ∗H) is a Borel homomorphism such that g 7→ Lg =

(r(g), Lg, s(g)).

89

Page 92: Characterization of Extremal KMS States on Groupoid C

Now we de�ne a norm on Cc(G) which will be an upper bound for the norm of a rep-

resentation obtained with the Renault's disintegration theorem. This norm is de�ned in

[20].

De�nition 4.2.11. Let G be a locally compact Hausdor� second countable étale groupoid

with Haar system {λx}x∈G(0) . We de�ne the I-norm on Cc(G) by

‖f‖I = max

{supx∈G(0)

∫G

|f(g)|dλx(g), supx∈G(0)

∫G

|f(g)|dλx(g)

}.

If the Haar system is given by counting measures λx on Gx, then

‖f‖I = max

{supx∈G(0)

∑g∈Gx|f(g)|, sup

x∈G(0)

∑g∈Gx

|f(g)|

}.

Lemma 4.2.12. Let G be a locally compact Hausdor� second countable étale groupoid with

Haar system given by counting measures λx on Gx. Then ‖f‖I < ∞ for every f ∈ Cc(G).

Moreover, ‖ · ‖I de�nes a norm on Cc(G).

Proof. Let f1, f2, f ∈ Cc(G), λ ∈ C.

• ‖f‖I = 0 implies f = 0.

Assume ‖f‖I = 0. Let g ∈ G and x = r(g). Then,

|f(g)| ≤∑h∈Gx|f(h)| ≤ ‖f‖I = 0.

Therefore f = 0.

• ‖λf‖ = |λ|‖f‖.

Let x ∈ G(0). Then

∑g∈Gx|λf(g)| = |λ|

∑g∈Gx|f(g)|.

90

Page 93: Characterization of Extremal KMS States on Groupoid C

Thus,

supx∈G(0)

∑g∈Gx|λf(g)| = |λ| sup

x∈G(0)

∑g∈Gx|f(g)|.

Analogously,

supx∈G(0)

∑g∈Gx

|λf(g)| = |λ| supx∈G(0)

∑g∈Gx

|f(g)|.

Therefore ‖λf‖I = |λ|‖f‖I ..

• ‖f1 + f2‖I ≤ ‖f1‖I + ‖f2‖I .

‖f1 + f2‖I = max

{supx∈G(0)

∑g∈Gx|f1(g) + f2(g)|, sup

x∈G(0)

∑g∈Gx

|f1(g) + f2(g)|

}

≤ max

{supx∈G(0)

∑g∈Gx

(|f1(g)|+ |f2(g)|), supx∈G(0)

∑g∈Gx

(|f1(g)|+ |f2(g)|)

}

≤ max

{supx∈G(0)

∑g∈Gx|f1(g)|+ ‖f2‖I , sup

x∈G(0)

∑g∈Gx

|f1(g)|+ ‖f2‖I

}

= ‖f1‖I + ‖f2‖I .

Now we show that ‖f‖I <∞ for every f ∈ Cc(G).

Let U be an open bisection of G and let f ∈ Cc(U). For every x ∈ r(U), there is a unique

hx ∈ U satisfying r(hx) = x. Note that x = r|−1U (x). Hence,

∑g∈Gx|f(g)| = |f(hx)|.

91

Page 94: Characterization of Extremal KMS States on Groupoid C

Since r|U is a homeomorphism, we have

supx∈r(U)

∑g∈Gx|f(g)| = sup

x∈r(U)

|f(hx)| = supx∈r(U)

|f ◦ r|−1U (x)| = sup

g∈U|f(g)| = ‖f‖∞.

If x /∈ r(U), then f(g) = 0 for every g ∈ Gx. Thus,

supx∈G(0)

∑g∈Gx|f(g)| = ‖f‖∞.

Analogously, we can prove that

supx∈G(0)

∑g∈Gx

|f(g)| = ‖f‖∞.

Therefore ‖f‖I = ‖f‖∞ for f ∈ Cc(U).

Now let f ∈ Cc(G). It follows from Proposition 3.3.5 that there exist f1, . . . fn continuous

functions supported on open bisections such that f = f1 + . . . + fn. Then ‖f‖I ≤ ‖f1‖ +

. . .+ ‖fn‖ <∞. Therefore ‖ · ‖I is a norm on Cc(G).

4.3 Renault's Disintegration Theorem

The results in this section can be found in [15]. The Renault's Disintegration Theorem can

also be found in [4] and [10]. However, both use the abstract notion of upper semi-continuous

C*-bundle and the theorem is stated in a more general case than the results presented here.

Proposition 4.3.1. Let G be a locally compact Hausdor� étale groupoid with Haar system

{λx}x∈G(0) . If (µ,G(0) ∗ H, L) is a unitary representation of G, then we obtain a ‖ · ‖I-

norm bounded representation π of Cc(G) on∫ ⊕G(0)Hxdµ(x), called the integrated form of

92

Page 95: Characterization of Extremal KMS States on Groupoid C

(µ,G(0) ∗ H, L), determined by

〈π(f)h, k〉 =

∫G

f(g)〈Lghs(g), kr(g)〉∆(g)−12dν(g), (4.9)

where ν =∫G(0) dλ

xdµ(x), ν−1 =∫G(0) dλxdµ(x) and ∆ = dν/dν−1 as described in De�nition

4.1.5.

Renault's disintegration theorem shows that a representation π : Cc(G) → H can be

identi�ed with a groupoid representation (µ,G(0) ∗ H, L) such that L(G) is measurable in

Iso(G(0) ∗ H). Condition (4.8) on page 88 will be necessary in order to write (4.9).

De�nition 4.3.2. Given a vector space H0, Lin(H0) denotes the set of linear operators

T : H0 → H0. Note that Lin(H0) is an algebra whose product is de�ned by composition of

operators.

Theorem 4.3.3. (Renault's disintegration theorem) Let G be a locally compact Hausdor�

étale groupoid. Suppose that H0 is a dense subspace of a complex Hilbert space H. Let

π : Cc(G)→ Lin(H0) be a homomorphism such that:

(a) {π(f)h : f ∈ Cc(G) and h ∈ H0} is dense in H;

(b) For each h, k ∈ H0,

f 7→ 〈π(f)h, k〉

is continuous in the inductive limit topology on Cc(G);

(c) For f ∈ Cc(G) and h, k ∈ H0, we have

〈π(f)h, k〉 = 〈h, π(f*)k〉.

Then each π(f) is bounded and extends to an operator π(f) on H of norm at most ‖f‖I .

Furthermore, π is a representation of Cc(G) on H and there is a unitary representation

93

Page 96: Characterization of Extremal KMS States on Groupoid C

(µ,G(0) ∗H, L) of G such that H ∼ L2(G(0) ∗H, µ) and π is equivalent to the integrated form

of (µ,G(0) ∗ H, L).

Remark 4.3.4. Recall that the inductive limit topology is introduced in De�nition 3.3.18

on page 71.

Lemma 4.3.5. Let G be a locally compact Hausdor� second countable étale groupoid. Let

ϕ be a state on C*(G) with GNS-triple (H, π, ξ). Then π satis�es the conditions of Renault's

disintegration theorem.

Proof. Assume ϕ is a state on C*(G). Let (H, π, ξ) be the corresponding GNS-triple.

(a) {π(f)h : f ∈ Cc(G) and h ∈ H} is dense in H

Since ξ is a cyclic vector, it follows that π(C*(G))ξ is dense in H. Recall that Cc(G) is

dense in C*(G), then π(Cc(G))ξ is dense in π(C*(G))ξ. Therefore, the property holds.

(b) For h, k ∈ H, f 7→ 〈π(f)h, k〉 is continuous in the inductive limit topology on Cc(G)

Let h, k ∈ H. Given f ∈ Cc(G), let {fi}i∈N be a sequence in Cc(G) such that fi → f

with respect to the inductive limit topology. It follows from Lemma 3.3.19 that fi → f

in C*(G). Then ‖π(fi)− π(f)‖ → 0 and therefore 〈(π(fi)− π(f))k, h〉 → 0.

(c) 〈π(f)h, k〉 = 〈h, π(f*)〉

Let h, k ∈ H, f ∈ Cc(G). Since π is an ∗-representation of C*(G), we have π(f*) =

π(f)*. Therefore the property holds.

94

Page 97: Characterization of Extremal KMS States on Groupoid C

Chapter 5

Neshveyev's Theorems

In this chapter we describe the KMS states of a groupoid C*-algebra. Neshveyev proves

this result in [17] in two theorems: the �rst theorem describes all states satisfying a certain

condition, and the second theorem describes all KMS states and it is a corollary of the �rst

result. These theorems show a correspondence between a KMS state ϕ on C*(G) and a pair

(µ, {ϕx}x∈G(0)) consisting of a probability measure µ on G(0) and a family of states ϕx on

C*(Gxx) satisfying a certain condition.

Before proving Neshveyev's theorems, we de�ne KMS states on a C*-algebra as described

in [2], [3] and [9].

5.1 KMS States

In this section we de�ne the notion of KMS states and prove some of their main properties.

The results in this section can be found in [2], [3] and [9]. KMS states characterizes the

equilibrium states in quantum statistical mechanics.

Before we de�ne KMS states, let us recall the de�nition of approximate unit and prove

some properties of states on a C*-algebra.

Let A be a C*-algebra. An approximate unit for A is an increasing net {uλ}λ∈Λ of positive

95

Page 98: Characterization of Extremal KMS States on Groupoid C

elements in the closed unit ball of A such that a = limλ auλ for all a ∈ A. Equivalently,

a = limλ uλa. From [16], it follows that every C*-algebra A contains an approximate unit.

Moreover, if A is separable, then it admits an approximate unit which is a sequence.

Recall the following theorems on positive linear functionals from [16]:

Theorem 5.1.1. If ϕ is a positive linear functional on a C*-algebra A, then it is bounded.

Theorem 5.1.2. Let ϕ be a bounded linear functional on a C*-algebra A. The following

conditions are equivalent:

(i) ϕ is positive.

(ii) For each approximate unit {uλ}λ∈Λ, ‖ϕ‖ = limλ ϕ(uλ).

(iii) For some approximate unit {uλ}λ∈Λ, ‖ϕ‖ = limλ ϕ(uλ).

Proof. Assume ϕ 6= 0. First we show the implication (i)⇒ (ii) holds.

Note that the map A2 → C de�ned by (a, b) 7→ ϕ(b*a) is a positive sesquilinear form on

A. Hence ϕ(b*a) = ϕ(a*b) and the Cauchy-Schwarz inequality |ϕ(b*a)| ≤ ϕ(b*b)1/2ϕ(a*a)1/2

holds.

Let (uλ)λ∈Λ be an approximate unit of A and let a ∈ A with a 6= 0. Then, for every λ ∈ Λ,

|ϕ(auλ)|2 ≤ ϕ(a*a)ϕ(uλ*uλ) = ϕ(a*a)ϕ(u2λ), (5.1)

by the Cauchy-Schwarz inequality. Since the net {uλ}λ∈Λ is increasing and its elements are in

the unit ball, then the net {u2λ}λ∈Λ is also increasing and included in the unit ball. Because

ϕ is positive and bounded, we have supλ∈Λ ϕ(u2λ) = limλ∈Λ ϕ(u2

λ). Then, for every i ∈ Λ, we

have by (5.1),

|ϕ(aui)|2 ≤ ϕ(a*a) limλ∈Λ

ϕ(u2λ) ≤ ‖ϕ‖‖a*a‖ lim

λ∈Λϕ(u2

λ) = ‖ϕ‖‖a‖2 limλ∈Λ

ϕ(u2λ).

96

Page 99: Characterization of Extremal KMS States on Groupoid C

Therefore, using the continuity of ϕ,

limλ∈Λ|ϕ(aui)|2 ≤ ‖ϕ‖‖a‖2 lim

λ∈Λϕ(u2

λ)

|ϕ(a)|2 ≤ ‖ϕ‖‖a‖2 limλ∈Λ

ϕ(u2λ)

|ϕ(a)|2

‖a‖2≤ ‖ϕ‖ lim

λ∈Λϕ(u2

λ).

Since a is arbitrary, ‖ϕ‖ ≤ limλ∈Λ ϕ(u2λ). On the other hand, for each λ ∈ Λ,

ϕ(u2λ) ≤ ‖ϕ‖‖u2

λ‖ ≤ ‖ϕ‖.

Then ‖ϕ‖ = limλ∈Λ ϕ(u2λ). Note that u

2λ−uλ ≤ 0 for every λ, thus ϕ(u2

λ) ≤ ϕ(uλ). Therefore,

‖ϕ‖ = limλ∈Λ

ϕ(u2λ) ≤ lim

λ∈Λϕ(uλ) ≤ ‖ϕ‖.

It is obvious that (ii)⇒ (iii).

Now we show that (iii) ⇒ (i). Suppose that {un}n∈N is an approximate unit such that

1 = limn→∞ τ(un). Let a be a selft-adjoint element of A such that ‖a‖ ≤ 1 and write

τ(a) = α + iβ where α, β are real numbers. To show that τ(a) ∈ R, we may suppose that

β ≤ 0. If k is a positive integer, then

‖a− iuλ‖2 = ‖(a+ ikun)(a− ikun)‖

= ‖a2 + k2u2n − ik(aun − una)‖

≤ 1 + k2 + k‖aun − una‖,

so |τ(a− ikun)|2 ≤ 1 + k2 + k‖aun − una‖ because ‖τ‖ = 1.

However, limk→∞ τ(a− ikun) = τ(a)− ik by hypothesis, and limk→∞(aun − una) = 0, so

97

Page 100: Characterization of Extremal KMS States on Groupoid C

in the limit as n→∞, we get

|α + iβ − ik|2 = |τ(a)− ik|2 = |α + iβ − ik|2 ≤ 1 + k2,

then

α2 + β2 − 2kβ + k2 ≤ 1 + n2

⇒− 2kβ ≤ 1− β2 − α2.

Since β is not positive and this inequality holds for all positive integers n, β must be zero.

Therefore, τ(a) is real if a is hermitian.

Now suppose that a is positive and ‖a‖ ≤ 1. Then un− a is hermitian and ‖un− a‖ ≤ 1,

so τ(un− a) ≤ 1. But then 1− τ(a) = limn→∞ τ(un− a) ≤ 1, and therefore τ(a) ≥ 0. Thus,

τ is positive and we have shown (iii)⇒ (i).

We want to show that, if the C*-algebra A is commutative and unital, then the extremal

states are precisely the characters on A. We need some results on convex spaces before

studying this.

De�nition 5.1.3. Given a commutative C*-algebra A, a character is a non-zero homomor-

phism ϕ : A→ C.

De�nition 5.1.4. Let X be a normed vector space. A functional η in a convex subset

C ⊂ X* is an extreme point in C (or extremal) if the condition η = tη1 + (1 − t)η2, where

η1, η2 ∈ C, 0 < t < 1, implies that η = η1 = η2.

De�nition 5.1.5. We say that a state ϕ on A is pure if it has the property that whenever

ρ is a positive linear functional on A such that ρ ≤ ϕ, necessarily there is a number t ∈ [0, 1]

such that ρ = tϕ.

Theorem 5.1.6. [3, Theorem 2.3.15] Let A be a unital C*-algebra. The set of states on A

is convex and its extremal points are the pure states.

98

Page 101: Characterization of Extremal KMS States on Groupoid C

Theorem 5.1.7. [16, Theorem 5.1.6] Let ϕ be a state on a commutative C*-algebra A.

Then ϕ is pure if, and only if, it is a character on A.

Lemma 5.1.8. Let X be a Borel space endowed with a probability measure µ. Let A be a

separable C*-algebra and A0 a dense subset of A. Let ψx be a family of states on A de�ned

for µ-a.e. x ∈ X such that for every a ∈ A0 the map x 7→ ψx(a) is µ-measurable. De�ne ϕ

by

ϕ(a) =

∫X

ψx(a)dµ(x).

Then ϕ is a state on A.

Proof. Let V ⊂ X be a conull set such that ψx is de�ned for every x ∈ V . De�ne for every

a ∈ A the function Fa : X → C by

Fa(x) =

ψx(a) if x ∈ V

0 otherwise.

By assumption, Fa is µ-measurable for every a ∈ A0.

Let a ∈ A. Since A0 is dense in A, there exists a sequence {an}n∈N in A0 converging to a.

Then, using the continuity of ψx, we have for µ-a.e. x,

Fa(x) = ψx(a) = limn→∞

ψx(an) = limn→∞

Fan(x).

Therefore Fa is measurable. In other words, the function x 7→ ψx(a) is µ-measurable for

every a ∈ A.

In order to show that ϕ is a state, we begin by proving that ϕ is a positive linear functional.

Since each ψx linear, it follows from the de�nition of ϕ that ϕ is also linear. Let a be a positive

99

Page 102: Characterization of Extremal KMS States on Groupoid C

element in A. Then ψx(a) ≥ 0 for µ-a.e. x. Thus,

ϕ(a) =

∫X

ψx(a)dµ(x) ≥ 0.

By Theorem 5.1.1, ϕ is bounded. It follows from Theorem 5.1.2 that, for some approximate

unit {un}n∈N of A,

‖ϕ‖ = limn→∞

ϕ(un) = limn→∞

∫X

ψx(un)dµ(x).

However,

|ψx(un)| ≤ ‖ψx‖‖un‖ = ‖un‖ ≤ 1,

for every n and for µ-a.e. x ∈ X because each ψx is a state and every un satis�es ‖un‖ ≤ 1.

Therefore we can apply the dominated convergence theorem, then

‖ϕ‖ = limn→∞

∫X

ψx(un)dµ(x)

=

∫X

(limn→∞

ψx(un))dµ(x)

=

∫X

1dµ(x), since each ψx is a state,

= 1.

Now we de�ne the notion of dynamical system on a C*-algebra and prove some of its

properties.

De�nition 5.1.9. We say that τ = {τt}t∈R is a one-parameter group of ∗-automorphisms of

a C*-algebra A if τt : A→ A is an ∗-automorphism and

(i) τt+s = τt ◦ τs for all t, s ∈ R;

100

Page 103: Characterization of Extremal KMS States on Groupoid C

(ii) τ0 = id.

Example 5.1.10. Consider the algebra of square matrices Mn(C) for some n. Let H ∈

Mn(C), and de�ne each τt : Mn(C) → Mn(C) by τt(A) = eitHAe−itH . Note that τt is linear

and τ0 is the identity. Each τt is a homomorphism because

τt(AB) = eitHABe−itH = eitHAe−itHeitHBe−itH = τt(A)τt(B).

Also, the equality τt+s = τt ◦ τs holds, since

τt+s(A) = ei(t+s)HAe−i(t+s)H = eitHeisHAe−isHe−itH = eitHτs(A)e−itH = τt(τs(A)).

De�nition 5.1.11. A C*-dynamical system is a pair (A, τ) where A is a C*-algebra and

τ = {τt}t∈R is a one-parameter group of ∗-automorphisms strongly continuous , i.e., t 7→ τt(a)

is continuous in the norm for all a ∈ A. If we �x τ , we say that τ is the dynamics on A.

Now we prove that, under certain conditions, if we de�ne a one-parameter group of ∗-

automorphisms τ on a dense ∗-algebra of a C*-algebra A, then we can extend the operators

uniquely on A. Moreover, τ de�nes a dynamics on A.

Lemma 5.1.12. Let A0 be a dense ∗-subalgebra of a C*-algebra A. Let τ = {τt}t∈R be a

family of ∗-automorphisms τt : A0 → A0 such that

(i) τt+s = τt ◦ τs for t, s ∈ R,

(ii) τ0 = id,

(iii) t 7→ τt(a) is continuous in the norm for each a ∈ A0.

Moreover, assume

‖τt(a)‖ ≤ ‖a‖, for every t ∈ R, a ∈ A0. (5.2)

101

Page 104: Characterization of Extremal KMS States on Groupoid C

Then τ can be extended uniquely to a dynamics on A, which we also denote by τ = {τt}t∈Rwithout loss of generality.

Proof. First we show that each τt can be extended uniquely to A. Let a ∈ A. Since A0 is

dense in A, there exists a sequence {an}n∈N converging to a.

Using the fact that A0 is a vector space and the equation (5.2), we have that {τt(an)}n∈Nis a Cauchy sequence, then we can de�ne τt(a) = limn→∞ τt(an).

Note that τt(a) is well-de�ned. In fact, let {bn}n∈N be an arbitrary sequence converging

to a, and let x = limn→∞ τt(bn). Then

‖τt(a)− x‖ = limn→∞

‖τt(an)− τt(bn)‖ ≤ limn→∞

‖an − bn‖ = 0.

It follows from the de�ntion of τt on A that ‖τt(a)‖ ≤ ‖a‖ for every a ∈ A.

Note that the extension τ also satis�es τ0 = id and τt+s = τt ◦ τs. In fact, given a ∈ A, let

{an}n∈N be a sequence in A0 converging to a. Then

τ0(a) = limn→∞

τ0(an) = limn→∞

an = a

τt+s(a) = limn→∞

τt+s(an) = limn→∞

τt ◦ τs(an) = τt ◦ τs(a).

Then τt is invertible with inverse τ−t.

Now we show that τt : A → A is an automorphism. Let a, b ∈ A and {an}n∈N, {bn}n∈Nsequences in A0 converging to a, b respectively. Then

τt(ab) = limn→∞

τt(anbn) = limn→∞

τt(an)τt(bn) = τt(a)τt(b),

τt(a*) = limn→∞

τt(an*) = limn→∞

τt(an)* = τt(a)*.

Finally, we show that τ is strongly continuous. Let a ∈ A and �x t0. Given ε, there exists

a0 ∈ A0 such that ‖a − a0‖ < ε/3. Let δ > 0 such that ‖τt(a0) − τs(a0)‖ < ε/3 for every

t ∈ R such that |t− t0| < δ.

102

Page 105: Characterization of Extremal KMS States on Groupoid C

Given t with |t− t0| < δ, we have

‖τt(a)− τt0(a)‖ ≤ ‖τt(a)− τt(a0)‖+ ‖τt(a0)− τt0(a0)‖+ ‖τt0(a0)− τt0(a)‖

≤ 2‖a− a0‖+ ‖τt(a0)− τt0(a0)‖

≤ 2ε

3+ε

3

= ε.

Let X be a complex Banach space and X* its dual. Let σ(X,X*) denote the topology

on X induced by the functionals on X. This topology is denoted the weak topology on X.

Remark 5.1.13. We say that a function f : R→ X is σ(X,X*)-continuous if η ◦f : R→ C

is continuous for every η ∈ X*.

De�nition 5.1.14. Given a Banach space X, let {τt}t∈R be a family of linear and bounded

operators τt : X → X. This family is called a one-parameter σ(X,X*)-continuous group of

isometries if

(i) τt+s = τt ◦ τs for all t, s ∈ R;

(ii) τ0 = id;

(iii) ‖τt‖ = 1 for all t ∈ R;

(iv) t 7→ τt(a) is σ(X,X*)-continuous for all a ∈ X.

Lemma 5.1.15. Let A be a C*-algebra and τ its dynamics. Then τ is a one-parameter

σ(A,A*)-continuous group of isometries.

Proof. Note that properties (i) and (ii) in De�nitions 5.1.9 and 5.1.14 are equal. In both,

every τ is a linear operator on A. So we need to show properties (iii) and (iv) of De�nition

5.1.14.

103

Page 106: Characterization of Extremal KMS States on Groupoid C

(iii) ‖τt‖ = 1

This follows from the fact that every ∗-automorphism on a C*-algebra is an isometry.

(iv) t 7→ τt(a) is σ(A,A*)-continuous for all a ∈ A

Let a ∈ A and t0 ∈ R. τ is strongly continuous by assumption. Then τt(a) → τt0(a)

in the norm as t → t0. Thus, for every η ∈ A*, η(τt(a)) → η(τt0(a)) as t → t0. Thus

property (iv) holds.

Therefore τ is a one-parameter σ(A,A*)-continuous group of isometries.

Note that, for every a ∈ X, the function t 7→ η ◦ τt(a) is continuous for all η ∈ X*. If we

can extend this function to an analytic function on a strip in C which we will de�ne later,

we say that a is analytic. We will prove that the set of analytic elements is dense in X.

De�nition 5.1.16. Let τ be a one-parameter σ(X,X*)-continuous group of isometries. An

element a ∈ X is analytic for τ if there exists λ > 0 and a function f : Iλ → X, where

Iλ = {z ∈ C : |Im(z)| < λ}, such that

(i) f(t) = τt(a) for all t ∈ R;

(ii) The function z 7→ η(f(z)) is analytic in Iλ for all η ∈ X*.

Under these conditions we write

τz(a) = f(z), for z ∈ Iλ.

If λ =∞, we say that a is entire analytic for τ .

Remark 5.1.17. Suppose a1, a2 ∈ X are entire analytic, and α ∈ C. Let f1, f2 be the

corresponding functions as in De�nition 5.1.16. Then, for every t ∈ R,

f1(t) + αf2(t) = τt(a1) + ατt(a2) = τt(a1 + αa2). (5.3)

104

Page 107: Characterization of Extremal KMS States on Groupoid C

Moreover, the function z 7→ η(f1(z) +αf2(z)) is analytic for every η ∈ X*. Thus a1 +αa2 is

entire analytic. Therefore the set of entire analytic elements in X, denoted Xτ , is a subspace

of X. For a1, a2 ∈ Xτ , α, z ∈ C, it follows from (5.3) that τz(a1 + αa2) = τz(a1) + τz(a2).

Later we will prove that Xτ is also dense in X, both in the σ(X,X*)-topology (see

Proposition 5.1.19) and with respect to the norm (see Corollary 5.1.24).

When τ de�nes a dynamics on a C*-algebra A, Aτ is not only a vector space, but a

∗-subalgebra. We will prove this result in Lemma 5.1.22 on page 110.

Lemma 5.1.18. Let n be a positive integer, and δ > 0. Then√n

π

∫ ∞−∞

e−nt2

dt = 1, (5.4)√n

π

∫|t|≥δ

e−nt2

dt =1√π

∫|t|≥√nδ

e−t2

dt. (5.5)

Moreover, the second integral converges to zero as n→∞.

Proof. Let δ ≥ 0. Since the function t 7→ e−nt2is even, we have√

n

π

∫|t|≥δ

e−nt2

dt = 2

√n

π

∫t≥δ

e−nt2

dt

=2√π

∫s≥√nδ

e−s2

ds, by the change of variables s =√nt,

=1√π

∫|s|≥√nδ

e−s2

ds.

Thus the equality (5.5) follows. If δ = 0, we have√n

π

∫ ∞−∞

e−nt2

dt =

√n

π

∫|t|≥0

e−nt2

dt =1√π

∫|s|≥0

e−s2

ds =1√π

∫ ∞−∞

e−s2

ds = 1.

Therefore the integral√

∫|t|≥δ e

−nt2dt converges to zero as n→∞.

Proposition 5.1.19. If τ is a one-parameter σ(X,X*)-continuous group of isometries, and

105

Page 108: Characterization of Extremal KMS States on Groupoid C

a ∈ X, de�ne for n = 1, 2, . . .,

an =

√n

π

∫ ∞−∞

τt(a)e−nt2

dt.

Then, for each n, an is an entire analytic element for τ and ‖an‖ ≤ ‖a‖. In addition,

τz(an) =

√n

π

∫ ∞−∞

τt(a)e−n(t−z)2dt,

for every z ∈ C. Moreover, an → a on the σ(X,X*) topology as n→∞. In particular, Xτ ,

is a σ(X,X*)-dense subspace of X.

Proof. Let a ∈ X. De�ne for each n = 1, 2, . . . the function fn : C→ X by

fn(z) =

√n

π

∫ ∞−∞

τt(a)e−n(t−z)2dt.

This function is well-de�ned since t 7→ e−n(t−z)2 is an integrable function, ‖τt(a)‖ ≤ ‖a‖ for

every t, and t 7→ e−n(t−z)2τt(a) is continuous, then we apply Proposition 2.6. Note that for

each s ∈ R, we have

fn(s) =

√n

π

∫ ∞−∞

τt(a)e−n(t−s)2dt

=

√n

π

∫ ∞−∞

τt+s(a)e−nt2

dt

=

√n

π

∫ ∞−∞

τs ◦ τt(a)e−nt2

dt

= τs

(√n

π

∫ ∞−∞

τt(a)e−nt2

dt

), by Corollary 2.6.17,

= τs(an).

Let η ∈ X*. Let z, z0 ∈ C such that z 6= z0. By de�nition of fn, it follows that∣∣∣∣η(fn(z))− η(fn(z0))

z − z0

−√n

π

∫ ∞−∞

2n(t− z)e−n(t−z)2η(τt(a))dt

∣∣∣∣106

Page 109: Characterization of Extremal KMS States on Groupoid C

equals to∣∣∣∣∣√n

π

∫ ∞−∞

η(τt(a))

(e−n(t−z)2 − e−n(t−z0)2

z − z0

)dt−

√n

π

∫ ∞−∞

2n(t− z)e−n(t−z)2η(τt(a))dt

∣∣∣∣∣=

√n

π

∣∣∣∣∣∫ ∞−∞

(e−n(t−z)2 − e−n(t−z0)2

z − z0

− 2n(t− z)e−n(t−z)2)η(τt(a))dt

∣∣∣∣∣≤√n

π

∫ ∞−∞

∣∣∣∣∣e−n(t−z)2 − e−n(t−z0)2

z − z0

− 2n(t− z)e−n(t−z)2∣∣∣∣∣ |η(τt(a))|dt

≤‖η‖‖a‖√n

π

∫ ∞−∞

∣∣∣∣∣e−n(t−z)2 − e−n(t−z0)2

z − z0

− 2n(t− z)e−n(t−z)2∣∣∣∣∣ dt,

since |η(τt(a))| ≤ ‖η‖‖τt(a)‖ ≤ ‖η‖‖a‖ for each t. This integral goes to zero when z → z0

and the entire analyticity follows. Also, fn(z) = τz(an) for every z ∈ C.

In addition, we have ‖an‖ ≤ ‖a‖. In fact,

‖an‖ =

∥∥∥∥√n

π

∫ ∞−∞

τt(a)e−nt2

dt

∥∥∥∥≤√n

π

∫ ∞−∞‖τt(a)‖e−nt2dt, from Lemma 2.6.11,

≤ ‖a‖√n

π

∫ ∞−∞

e−nt2

dt

= ‖a‖, from Lemma 5.1.18.

Now we prove that an converges to a on the σ(X,X*) topology. Let η ∈ X*. If η = 0

or a = 0, it follows that η(an) = η(a) = 0 for every n. Then we assume η 6= 0 and a 6= 0

without loss of generality.

Note that

η(an − a) = η(an)− η(a)

=

√n

π

∫ ∞−∞

η(τt(a))e−nt2

dt− η(a), by de�nition of an,

107

Page 110: Characterization of Extremal KMS States on Groupoid C

=

√n

π

∫ ∞−∞

(η(τt(a))− η(a))e−nt2

dt,

since√

∫∞−∞ e

−nt2dt = 1 by Lemma 5.1.18.

Let ε > 0. Since the map t 7→ τt(a) is continuous and η ∈ X*, there exists δ > 0 such

that

|t| < δ implies |η(τt(a))− η(a)| < ε/2. (5.6)

Moreover, from Lemma 5.1.18, there exists N > 0 such that, for every n ≥ N , we have√n

π

∫|t|≥δ

e−nt2

dt <ε

4‖η‖‖a‖. (5.7)

Then, for each n ≥ N ,

|η(an − a)| =∣∣∣∣√n

π

∫ ∞−∞

e−nt2

(η(τt(a))− η(a))dt

∣∣∣∣≤∣∣∣∣√n

π

∫|t|≥δ

e−nt2

(η(τt(a))− η(a))dt

∣∣∣∣+

∣∣∣∣√n

π

∫|t|<δ

e−nt2

(η(τt(a))− η(a))dt

∣∣∣∣≤√n

π

∫|t|≥δ

e−nt2|η(τt(a))− η(a))|dt+

√n

π

∫|t|<δ

e−nt2|η(τt(a))− η(a))|dt

≤√n

π

∫|t|≥δ

e−nt2|η(τt(a))− η(a))|dt+

ε

2

√n

π

∫|t|<δ

e−nt2

dt, from (5.6),

≤√n

π

∫|t|≥δ

e−nt2‖η‖(‖τt(a)‖+ ‖a‖)dt+

ε

2

√n

π

∫|t|<δ

e−nt2

dt

≤√n

π

∫|t|≥δ

e−nt2‖η‖(‖a‖+ ‖a‖)dt+

ε

2

√n

π

∫|t|<δ

e−nt2

dt, since ‖τt‖ = 1,

=2‖η‖‖a‖√n

π

∫|t|≥δ

e−nt2

dt+ε

2

√n

π

∫|t|<δ

e−nt2

dt

≤2‖η‖‖a‖ ε

4‖η‖‖a‖+ε

2

√n

π

∫|t|<δ

e−nt2

dt, from (5.7),

≤2‖η‖‖a‖ ε

4‖η‖‖a‖+ε

2

√n

π

∫ ∞−∞

e−nt2

dt,

108

Page 111: Characterization of Extremal KMS States on Groupoid C

2+ε

2, from Lemma 5.1.18,

=ε.

Hence an → a in the topology σ(X,X*).

Remark 5.1.20. Let τ be a one-parameter σ(X,X*)-continuous group of isometries. Given

λ, we say that a is analytic on the strip Iλ if the conditions in De�nition 5.1.16 hold for a

and λ.

We say that a ∈ X is strongly analytic on the strip Iλ if there exists f : Iλ → X such that

condition (i) in De�nition 5.1.16 is satis�ed and the limit

limh→0

f(z + h)− f(z)

h

exists for every z ∈ Iλ. We now show that these two notions are equivalent.

Lemma 5.1.21. Let τ be a one-parameter σ(X,X*)-continuous group of isometries, and

a ∈ Xτ . Given w ∈ C, τw(a) ∈ Xτ . Moreover, for every z ∈ C we have

τz+w(a) = τz ◦ τw(a).

Proof. Let a ∈ Xτ . Then there exists a function f : C→ X such that f(t) = τt(a) for every

t ∈ R, and η ◦ f is entire analytic for η ∈ X*.

Fix s ∈ R. Let fs : C → X be the function de�ned by fs(w) = f(s + w). Then η ◦ fs is

entire analytic for all η ∈ X*.

De�ne g : C→ X by g(w) = τs ◦τw(a). Note that η ◦g is entire analytic for every η ∈ X*.

In fact, given w ∈ C,

η ◦ g(w) = η ◦ τs ◦ τw(a) = (η ◦ τs) ◦ f(z).

Since η ◦ τs ∈ X* and a is entire analytic, it follows that (η ◦ τs) ◦ f is entire analytic. Thus

109

Page 112: Characterization of Extremal KMS States on Groupoid C

η ◦ g is entire analytic. Moreover, for t ∈ R, we have

η ◦ g(t) = η ◦ τs ◦ τt(a) = η ◦ τs+t(a) = η ◦ f(t+ s) = η ◦ fs(t).

Thus η ◦ g and η ◦ fs are entire analytic functions which agree on the real line. Thus, by

the uniqueness theorem for analytic functions [1, Theorem 6.9] we have η ◦ g = η ◦ fs. η is

arbitrary and X* separates poins in X, thus g = fs. Therefore, for every w ∈ C,

τs ◦ τw(a) = τs+w(a). (5.8)

Now we �x w ∈ C and assume s is a real variable. De�ne fw : C→ X by fw(z) = f(w+z).

Since a is entire analytic, the function η ◦fw is entire analytic for every η ∈ X*. By equation

(5.8), we have for all t ∈ R the equality

fw(t) = f(w + t) = τt+w(a) = τt ◦ τw(a).

Therefore τw(a) is entire analytic and

τz ◦ τw(a) = fw(z) = f(w + z) = τz+w(a).

Let τ de�ne a dynamics on a C*-algebra. We say that a subset A1 is τ -invariant if

τt(a) ∈ A1 for every t ∈ R, a ∈ A1.

Lemma 5.1.22. Let (A, τ) be a C*-dynamical system. Then Aτ is a ∗-subalgebra which is

τ -invariant.

Proof. Aτ is a vector space by Remark 5.1.17. Moreover, Aτ is τ -invariant by Lemma 5.1.21.

Given a, b ∈ Aτ , let f1, f2 : C → A be the functions such that η ◦ f1, η ◦ f2 are analytic

for all η ∈ A*, and such that f1(t) = τt(a), f2(t) = τt(b) for every t ∈ R.

110

Page 113: Characterization of Extremal KMS States on Groupoid C

Let f = f1f2. Then η ◦ f is analytic for all η ∈ A*. Moreover, given t ∈ R,

f(t) = f1(t)f2(t) = τt(a)τt(b) = τt(ab).

Therefore ab is entire analytic.

De�ne f1* by f1*(z) = f1(z)*. Then, for t ∈ R,

f1*(t) = f1(t)* = τt(a)* = τt(a*).

Given η ∈ A*, let η1 ∈ A* be de�ned by η1(b) = η(b*) for b ∈ A. Then,

η ◦ f1*(z) = η(f1(z)*) = η1 ◦ f(z).

Note that the function z 7→ η1 ◦ f1(z) is analytic. In fact,

limh→0

η1 ◦ f1(z + h)− η1 ◦ f1(z)

h=

(limh→0

η1 ◦ f1(z + h)− η1 ◦ f1(z)

h

).

This limit exists since η1 ◦ f1 is analytic. Then η ◦ f1* is analytic, and therefore a* is entire

analytic.

Proposition 5.1.23. Let τ be a one-parameter σ(X,X*)-continuous group of isometries.

Given λ > 0, a ∈ X is analytic on the strip Iλ if, and only if, a is strongly analytic on Iλ.

Proof. Let η ∈ X*. Let z0 ∈ Iλ. There exists r > 0 such that D(z0, r/2) ⊂ D(z0, r) ⊂ Iλ.

Let h, k ∈ D(0, r/2). Then z0 + h, z0 + k ∈ D(z0, r/2). Using Cauchy's integral formula

and assuming C = ∂D(z0, r), we have

η(f(z0)) =1

2πi

∮C

η(f(z))

z − z0

dz,

η(f(z0 + h)) =1

2πi

∮C

η(f(z))

z − z0 − hdz,

η(f(z0 + k)) =1

2πi

∮C

η(f(z))

z − z0 − kdz.

111

Page 114: Characterization of Extremal KMS States on Groupoid C

Then

η

(f(z0 + h)− f(z0)

h− f(z0 + k)− f(z0)

k

)=

1

2πi

∮C

η(f(z))

[1

h

(1

z − z0 − h− 1

z − z0

)− 1

k

(1

z − z0 − k− 1

z − z0

)]dz (5.9)

Note that 1a−b −

1a

= ba(a−b) for a, b complex numbers such that a 6= b and a 6= 0. Applying

this, we have that (5.9) equals to

1

2πi

∮C

η(f(z))

[1

h

h

(z − z0)(z − z0 − h)− 1

k

k

(z − z0)(z − z0 − k)

]dz

=1

2πi

∮C

η(f(z))

[1

(z − z0)(z − z0 − h)− 1

(z − z0)(z − z0 − k)

]dz

=1

2πi

∮C

η(f(z))

z − z0

[1

z − z0 − h− 1

z − z0 − k

]dz

=1

2πi

∮C

η(f(z))

z − z0

h− k(z − z0 − h)(z − z0 − k)

dz

Note that for every z such that |z − z0| = r, we have

|z − z0 − h| ≥ |z − z0| − |h| ≥ r − r

2=r

2

Thus

1

|z − z0 − h|≤ 2

h. (5.10)

Analogously, 1|z−z0−k| ≤

2k. Recall that C denotes the set of points z satisfying |z − z0| = r.

Therefore,

∣∣∣∣η(f(z0 + h)− f(z0)

h− f(z0 + k)− f(z0)

k

)∣∣∣∣ =

∣∣∣∣ 1

2πi

∮C

η(f(z))

z − z0

h− k(z − z0 − h)(z − z0 − k)

dz

∣∣∣∣≤ rmax

z∈C

|η(f(z))||z − z0|

|h− k||z − z0 − h||z − z0 − k|

112

Page 115: Characterization of Extremal KMS States on Groupoid C

≤ r4

r2maxz∈C

|η(f(z))||z − z0|

|h− k|, by (5.10),

=4

r2maxz∈C|η(f(z))||h− k|

≤ 4

r2‖η‖max

z∈C‖f(z)‖|h− k|.

Using the fact that ‖x‖ = sup‖η‖=1 ‖η(x)‖ in a Banach space, we obtain

∣∣∣∣f(z0 + h)− f(z0)

h− f(z0 + k)− f(z0)

k

∣∣∣∣ ≤ 4

r2maxz∈C|f(z)||h− k|.

Since X is complete, it follows that the limit

limh→0

f(z0 + h)− f(z0)

h

exists. In other words, a is strongly analytic on Iλ.

Conversely, suppose that a is strongly analytic on the strip Iλ. Let z ∈ Iλ. By hypothesis,

there exists x ∈ X such that

x = limh→0

f(z + h)− f(z)

h.

Let η ∈ X*. Then

η(x) = η

(limh→0

f(z + h)− f(z)

h

)= lim

h→0

η(f(z + h))− η(f(z))

h.

Hence η ◦ f is analytic at z. Since z and η are arbitrary, it follows that a is analytic on

Iλ.

Corollary 5.1.24. If τ = {τt}t∈R is a one-parameter σ(X,X*)-continuous group of isome-

tries, then τ is strongly continuous and Xτ is norm-dense in X.

Proof. Note that by strongly continuous, we mean that the map t 7→ τt(a) is continuous with

respect to the norm in X for every a ∈ X.

113

Page 116: Characterization of Extremal KMS States on Groupoid C

Suppose Xτ is not dense in X. Let H be the norm closure of Xτ . By hypothesis, H is a

proper subspace of X. Let y ∈ X \H. By the Hahn-Banach for the sets H and {y}, there

exists a ϕ ∈ X* such that

Re(ϕ(y)) < Re(ϕ(y)), h ∈ H. (5.11)

Since H is a proper subspace of X and ϕ is linear, Re(ϕ(H)) must be either {0} or R. By

equation (5.11), Re(ϕ(H)) = {0}.

Note that Im(ϕ(H)) = 0. In fact, let h ∈ H. Then there is λ ∈ R such that ϕ(h) = λi.

Since H is a complex vector space and ϕ is linear, ih ∈ H and ϕ(ih) = λ. It follows that

λ = 0. The choice of h is arbitrary, thus Im(ϕ(H)) = {0}. Hence ϕ is zero on H and

ϕ(y) 6= 0.

From Proposition 5.1.19, Xτ is dense in X with respect to the σ(X,X*)-topology. Since

H contains Xτ , H is also dense in X in this topology. Thus there exists a sequence {yn} in

H converging to y in the σ(X,X*)-topology. Since ϕ ∈ X*, we have ϕ(yn) → ϕ(y). This

leads to a contradiction because ϕ(y) 6= 0 and ϕ(yn) = 0 for every n. Therefore H = X,

that is, Xτ is norm dense in X.

Now we show that τ is strongly continuous. Given a ∈ Xτ , z ∈ C, it follows from

Proposition 5.1.23 that the limit

limh→0

∥∥∥∥τz+h(a)− τz(a)

h

∥∥∥∥exists. Thus ‖ τz+h(a)−τz(a)

h‖ converges to zero as h → 0, i.e., the function z 7→ τz(a) is

continuous with respect to the norm.

Now choose a ∈ X arbitrary, �x t ∈ R and let ε > 0. Since Xτ is norm dense in X, there

exists aε ∈ Xτ such that ‖a− aε‖ < ε/3.

Since aε ∈ Xτ , there exists δ > 0 such that for every h ∈ R with |h| < δ, we have

114

Page 117: Characterization of Extremal KMS States on Groupoid C

‖τt+h(aε)− τt(aε)‖ < ε/3. Then

‖τt+h(a)− τt(a)‖ ≤ ‖τt+h(a)− τt+h(aε)‖+ ‖τt+h(aε)− τt(aε)‖+ ‖τt(aε)− τt(a)‖

≤ ‖τt+h‖‖a− aε‖+ ‖τt+h(aε)− τt(aε)‖+ ‖τt‖‖aε − a‖

= ‖a− aε‖+ ‖τt+h(aε)− τt(aε)‖+ ‖aε − a‖

3+ ‖τt+h(aε)− τt(aε)‖+

ε

3

=2

3ε+ ‖τt+h(aε)− τt(aε)‖

≤ 2

3ε+

ε

3

= ε.

Then the map t 7→ τt(a) is continuous. Since a is arbitrary, τ is strongly continuous.

De�nition 5.1.25. Let (A, τ) be C*-dynamical system, ϕ a state on A and β ∈ R. We say

ϕ is a τ -KMSβ-state if

ϕ(aτiβ(b)) = ϕ(ba)

for all a, b in a ∗-subalgebra A0 composed of entire analytic elements such that A0 is norm-

dense and τ -invariant.

When τ is implicit, we just say that ϕ is a KMSβ-state in order to assert that ϕ is a

τ -KMSβ-state. Moreover, if β is �xed, we just say ϕ is a KMS state.

Remark 5.1.26. If ϕ is a KMS-state and β = 0, we have

ϕ(ab) = ϕ(ba),

for every a, b ∈ A0. From the continuity of ϕ, this equality holds for every a, b ∈ A. In this

case, we say the state ϕ is tracial .

De�nition 5.1.27. Let (A, τ) be a C*-dynamical system. We say that a ∈ A is τ -invariant

115

Page 118: Characterization of Extremal KMS States on Groupoid C

if τt(a) = a for every a ∈ A.

De�nition 5.1.28. Let A be a C*-algebra with dynamics τ . We say that a state ϕ on A is

τ -invariant if ϕ(τt(a)) = ϕ(a) for every a ∈ A and t ∈ R.

In this section we de�ned KMS-states as described in [3]. However, Neshveyev's theorem

assumes a diferent de�nition of KMS-states in [17]. The de�nition used is the item (ii) of

the next theorem. Now we show an equivalence in the de�nition of KMS states when β 6= 0.

Proposition 5.1.29. Let A be a C*-algebra with dynamics given by τ and let β ∈ R \ {0}.

Given a state ϕ on A, the following are equivalent:

(i) ϕ is a KMSβ-state;

(ii) ϕ is τ -invariant and

ϕ(aτiβ(b)) = ϕ(ba) (5.12)

for a dense set of analytic elements a, b ∈ A.

(iii) Equation (5.12) holds for every a, b ∈ Aτ .

Proof. Assume β > 0 without loss of generality.

(i)⇒ (ii) Let ϕ be a KMSβ-state. By de�nition, there exists a subalgebra A0 composed of

entire analytic elements such that A0 is norm-dense, τ -invariant and such that (5.12) holds

for every a, b ∈ A0. Thus, we only need to show that ϕ is τ -invariant.

Let a ∈ A0. De�ne f : C→ C by f(z) = ϕ(τz(a)). Since a is entire analytic and ϕ ∈ A*,

it follows that f is entire analytic.

Let {uλ}λ∈Λ be an approximate unit for A. Fix λ ∈ Λ. Since A0 is norm dense in A, for

every ε > 0 there exists u(ε)λ ∈ A0 satisfying ‖u(ε)

λ − uλ‖ < ε. Both u(ε)λ , a ∈ A0 and ϕ is a

KMSβ-state, then

ϕ(u(ε)λ τiβ(a)) = ϕ(au

(ε)λ ).

116

Page 119: Characterization of Extremal KMS States on Groupoid C

Using the continuity of ϕ, we have

ϕ(uλτiβ(a)) = ϕ(auλ),

for every λ ∈ Λ. Again, using the continuity of ϕ and the fact that {uλ}λ∈Λ is an approximate

unit, we have

ϕ(τiβ(a)) = limλϕ(uλτiβ(a)) = lim

λϕ(auλ) = ϕ(a).

De�ne g : C→ C by g(z) = ϕ(τz+iβ(a))−ϕ(τz(a)). Since a is entire analytic and ϕ ∈ A*,

it follows that g is entire analytic. Note that g is zero on the real line. In fact, given t ∈ R,

we have by Lemma 5.1.21

g(t) = ϕ(τt+iβ(a))− ϕ(τt(a)) = ϕ(τiβ(τt(a)))− ϕ(τt(a)) = 0.

Hence g(z) = 0 for every z ∈ C, that is,

ϕ(τz+iβ(a)) = ϕ(τz(a)). (5.13)

Let f : C → C be the analytic function de�ned by f(z) = ϕ(τz(a)). From (5.13) it follows

that f is periodic, that is, for every z ∈ C there exists s ∈ R and 0 < t ≤ β such that

f(z) = f(s+ it). Then, for every z ∈ C, we have

|f(z)| ≤ supw∈C|f(w)|

≤ sups∈R

0<t≤β

|f(s+ it)|

= sups∈R

0<t≤β

|ϕ(τs+it(a))|

≤ sups∈R

0<t≤β

‖τs+it(a)‖

117

Page 120: Characterization of Extremal KMS States on Groupoid C

= sups∈R

0<t≤β

‖τs(τit(a))‖, by Lemma 5.1.21,

≤ sup0<t≤β

‖τit(a)‖, by de�nition of τs for s ∈ R,

<∞,

since f is continuous. Then f is entire analytic and bounded. Therefore f is constant. Thus,

for every z ∈ C, we have

ϕ(a) = f(0) = f(z) = ϕ(τz(a)).

In particular, for every a ∈ A0 and t ∈ R, we have

ϕ(τt(a)) = ϕ(a). (5.14)

Since A0 is dense in A, it follows that (5.14) holds for every t ∈ R, a ∈ A. Therefore ϕ is

τ -invariant.

(ii) ⇒ (iii) Let ϕ be a τ -invariant state and suppose that there exists a dense set A1 of

analytic elements such that (5.12) holds for every a, b ∈ A1. Let

Dβ = {z ∈ C : 0 ≤ Im(z) ≤ β}.

Let a, b ∈ Aτ . De�ne η ∈ A* by η(x) = ϕ(ax). Then we can de�ne the entire analytic

function f : C→ C by f(z) = ϕ(aτz(b)) because f(z) = η(τz(a)).

Since A1 is dense in A, there are sequences {an}n∈N, {bn}n∈N in A1 such that an → a and

bn → b. Let a′ = supn ‖an‖, b′ = supn ‖bn‖. Both a′, b′ are �nite.

For every n, there exists λn > β such that bn is analytic on the strip Iλn . So we can de�ne

the analytic function fn by fn(z) = ϕ(anτz(bn)) for z ∈ Iλn . Let fn = fn|Dβ . We will show

that {fn}n∈N is a Cauchy sequence with respect to the supremum norm.

118

Page 121: Characterization of Extremal KMS States on Groupoid C

In fact, given n,m, de�ne λn,m = min{λn, λm}. Then fn|Iλn,m − fm|Iλn,m is analytic on

Iλn,m . Then the supremum of |fn − fm| is attained on the boundary of Dβ. Hence, for each

z ∈ Dβ, we have

|fn(z)− fm(z)| ≤ max

{supt∈R|fn(t)− fm(t)|, sup

t∈R|fn(t+ iβ)− fm(t+ iβ)|

}.

However, for every t ∈ R,

|fn(t)− fm(t)| = |ϕ(anτt(bn))− ϕ(amτt(bm))|

≤ |ϕ(anτt(bn))− ϕ(anτt(bm))|+ |ϕ(anτt(bm))− ϕ(amτt(bm))|

≤ ‖anτt(bn)− anτt(bm)‖+ ‖anτt(bm)− amτt(bm)‖

≤ ‖an‖‖τt(bn − bm)‖+ ‖an − am‖‖τt(bm)‖

≤ ‖an‖‖bn − bm‖+ ‖an − am‖‖bm‖

≤ a′‖bn − bm‖+ ‖an − am‖b′.

On the other hand,

|fn(t+ iβ)− fm(t+ iβ)| = |ϕ(anτt+iβ(bn))− ϕ(amτt+iβ(bm))|

= |ϕ(τt(bn)an)− ϕ(τt(bm)am)|

= |ϕ(τt(bn)an)− ϕ(τt(bn)am)|+ |ϕ(τt(bn)am)− ϕ(τt(bm)am)|

≤ ‖τt(bn)an − τt(bn)am‖+ ‖τt(bn)am − τt(bm)am‖

≤ ‖τt(bn)‖‖an − am‖+ ‖τt(bn)− τt(bm)‖‖am‖

≤ ‖bn‖‖an − am‖+ ‖bn − bm‖‖am‖

≤ b′‖an − am‖+ ‖bn − bm‖a′.

Therefore ‖fn− fm‖ ≤ b′‖an− am‖+ ‖bn− bm‖a′. Hence {fn}n∈N is a Cauchy sequence and

then this sequence converges to a continuous function f : Dβ → C which is analytic on the

119

Page 122: Characterization of Extremal KMS States on Groupoid C

interior of Dβ. Note that for every t ∈ R,

f(t) = ϕ(aτt(b)) = limn→∞

ϕ(anτt(bn)) = f(t),

by the continuity of ϕ and τt. Therefore f |Dβ = f by the uniqueness theorem for analytic

functions. Moreover,

ϕ(aτiβ(b)) = f(iβ)

= limn→∞

fn(iβ)

= limn→∞

ϕ(anτiβ(bn))

= limn→∞

ϕ(bnan)

= ϕ(ba),

for every a, b ∈ Aτ . Therefore ϕ is a KMSβ-state.

(iii) ⇒ (i) From Lemma 5.1.22, Aτ is a ∗-subalgebra and is τ -invariant. By Corollary

5.1.24, Aτ is dense in A. Hence ϕ is a KMSβ-state.

Remark 5.1.30. Let ϕ be a τ -invariant state on a C*-algebra A. In order to prove that ϕ is

a KMSβ-state, it su�ces to show that ϕ(aτiβ(b)) = ϕ(ba) in an arbitrary dense ∗-subalgebra

of A. In fact, if β 6= 0, this follows directly from Proposition 5.1.29. If β = 0, this follows

from the continuity of ϕ.

De�nition 5.1.31. Given a C*-algebra A and a state ϕ on A, we de�ne the centralizer of

ϕ as the set

{a ∈ A : ϕ(ab) = ϕ(ba) for every b ∈ A}.

Lemma 5.1.32. Let (A, τ) be a C*-dynamical system, β > 0, and ϕ a τ -KMSβ-state. Then,

for every a ∈ A such that a is τ -invariant, it follows that a is in the centralizer of ϕ.

120

Page 123: Characterization of Extremal KMS States on Groupoid C

Proof. Let a ∈ A be τ -invariant. Then τt(a) = a for every t. Then we can de�ne the function

f : C → A by f(z) = a. Then we have f(t) = τt(a) for every t ∈ R. The function f is

constant, then f is strongly analytic on C. Then, by Proposition 5.1.23, a is entire analytic.

Let b ∈ Aτ . Then,

ϕ(ba) = ϕ(bτiβ(a)) = ϕ(ab),

because ϕ is a τ -KMSβ-state.

Now we show that the set of KMS states, for a �xed β > 0 is convex. Moreover, in order

to describe all KMS states on a C*-algebra, it is su�cient to �nd only its extremal KMS

states.

Lemma 5.1.33. Let A be a C*-algebra with dynamics τ . Let β ∈ R. The set of KMSβ-states

is convex.

Proof. Let A be a C*-algebra with dynamics τ . Fix β ∈ R. Given ϕ1, ϕ2 KMS states,

0 < t < 1, de�ne ϕ = tϕ1 + (1− t)ϕ2. Now we show that ϕ is a KMS state.

(i) ϕ is a state.

Note that ϕ is positive. In fact, given a positive element a ∈ A,

ϕ(a) = tϕ1(a) + (1− t)ϕ2(a) ≥ 0,

since both ϕ1(a), ϕ2(a) ≥ 0. Moreover, we prove that ϕ has norm 1. Applying Lemma

5.1.2, we have for every approximate unit {uλ}λ∈Λ of A,

‖ϕ‖ = limλϕ(uλ)

= limλ

(tϕ1(uλ) + (1− t)ϕ2(uλ))

= t limλ

(ϕ1(uλ)) + (1− t) limλ

(ϕ2(uλ))

121

Page 124: Characterization of Extremal KMS States on Groupoid C

= t‖ϕ1‖+ (1− t)‖ϕ2‖

= t+ (1− t)

= 1.

(ii) ϕ is KMS

Given a, b ∈ Aτ ,

ϕ(aτiβ(b)) = tϕ1(aτiβ(b)) + (1− t)ϕ2(aτiβ(b)) = tϕ1(ba) + (1− t)ϕ2(ba) = ϕ(ba).

In order to prove that the extremal KMS-states are su�cient to describe all KMS-states,

we are going to use a the Krein-Milman theorem. But, before using this theorem, we show

that the set of KMSβ-states is compact with respect to a topology we de�ne below, the

weak* topology. Results used in this part can be found on [16] and [22].

Let X be a normed vector space. The weak*-topology on X* is generated by the family

of seminorms px : X*→ R such that η 7→ px(η) = |η(x)|.

Theorem 5.1.34. (The Banach-Alaoglu theorem) If V is a neighborhood of 0 in a normed

vector space X and if

K = {η ∈ X* : |η(x)| ≤ 1 for every x ∈ V },

then K is compact in the weak*-topology.

In this case, we can also say that K is weak*-compact .

Lemma 5.1.35. Let A be a C*-algebra with dynamics τ . Let β ∈ R. The set of KMSβ-states

is compact in the weak*-topology.

122

Page 125: Characterization of Extremal KMS States on Groupoid C

Proof. Let A be the unitization of A. Let

K = {η ∈ A* : |η(a)| ≤ 1 for a ∈ A with ‖a‖ ≤ 1}

= {η ∈ A* : ‖η‖ ≤ 1}.

It follows from the Banach-Alaoglu theorem that K is weak*-compact.

Let {ϕλ}λ∈Λ be a net of KMSβ-states on A. De�ne for each λ, ϕλ the extension of ϕλ on

A. Since each ϕλ ∈ K and K is compact, there exists a subnet {ϕλj}j∈J converging to some

ϕ ∈ K in the weak*-topology. Hence, for every a ∈ A,

ϕ(a) = limj∈J

ϕλj(a).

Also, since each ϕλj is a state

ϕ(1) = limj∈J

ϕλj(1) = 1.

It follows from Theorem 5.1.2 that ϕ is a state. Hence ϕ is a state. We will prove that ϕ is

a KMSβ-state.

Let a, b ∈ Aτ . Then

ϕ(aτiβ(b)) = limj∈J

ϕj(aτiβ(b)) = limj∈J

ϕj(ba) = ϕ(ba).

Therefore the set of KMSβ-states is weak*-compact.

We will use the fact that the set of KMS states is compact and convex to prove that the

extremal KMS states are su�cient to describe all KMS states. Before proving this, we de�ne

some concepts about convex sets and state then state the Krein-Milman theorem. Here all

results are de�ned for subsets of X*. For more general results, see [16].

De�nition 5.1.36. Let X be a normed vector space. Given a subset S of X*, co(S) denotes

the closure of co(S) with respect to the weak*-topology.

123

Page 126: Characterization of Extremal KMS States on Groupoid C

Now we state the Krein-Milman theorem for X*.

Theorem 5.1.37. Let X be a normed vector space. Let C ⊂ X* be a non-empty convex

weak*-compact subset. Then the set E of extreme points of C is non-empty and

C = co(E).

Moreover, if S is a closed subset of C with respect to the weak*-topology such that co(S) = C,

then S contains E.

Now we apply the Krein-Milman theorem to show that the extremal KMS states are

su�cient to describe all KMS states on A.

Corollary 5.1.38. Let A be a C*-algebra and τ its dynamics. Fix β > 0. Let C denote the

set of KMSβ-states on A and assume C 6= ∅. Let E be the set of extremal KMSβ-states.

Then C = co(E) and E 6= ∅.

Proof. It follows from Lemmas 5.1.33 and 5.1.35 that C is convex and compact. By hypoth-

esis, C 6= ∅. Then we can apply the Krein-Milman theorem and the result follows.

5.2 First Theorem

Let G be a locally compact Hausdor� second countable étale groupoid. Neshveyev's �rst

theorem shows that for every state ϕ on C*(G) with centralizer containing C0(G(0)) there

is a corresponding pair (µ, {ϕx}x), where µ is a probability measure on G(0), and {ϕx}x is a

µ-measurable �eld of states ϕx on C*(Gxx) for x ∈ G(0). The results in this section can be

found in [17].

In Section 5.3, we will de�ne a dynamics τ on C*(G) such that every function in C0(G(0))

is τ -invariant. Thus, by Lemma 5.1.32, the centralizer of every KMS state contains C0(G(0)).

Therefore, we can apply Neshveyev's �rst theorem to KMS states.

Now we describe µ-measurable �eld of states as described in [5].

124

Page 127: Characterization of Extremal KMS States on Groupoid C

De�nition 5.2.1. Let G be a locally compact Hausdor� second countable étale groupoid,

and let µ be a Radon Borel measure on G(0). For each x ∈ G(0), g ∈ Gxx, we let ug denote the

canonical unitary generators of C*(Gxx), i.e., ug ∈ Cc(Gx

x) is a function de�ned by ug(g) = 1

and ug(h) = 0 if h 6= g.

We call a collection {ϕx}x∈G(0) a µ-measurable �eld of states if each ϕx is a state on

C*(Gxx) and the function:

G(0) 3 x 7→∑g∈Gxx

f(g)ϕx(ug)

is µ-measurable for each f ∈ Cc(G).

Remark 5.2.2. Fix a probability measure µ on G(0). Given a conull set V ⊂ G(0), let

{ϕx}x∈V be a family of states ϕx on C*(Gxx) such that for every f ∈ Cc(G) the function

x 7→ χV (x)∑g∈Gxx

f(g)ϕx(ug) (5.15)

is µ-measurable. Then there exists a µ-measurable �eld of states {ϕx}x∈G(0) such that ϕx =

ϕx for every x ∈ V . In fact, any C*-algebra C*(Gxx) has at least on state, then you can just

choose any family of states ϕx on C*(Gxx) such that ϕx = ϕx for x ∈ V . Since the function

de�ned by (5.15) is µ-measurable, then {ϕx} is a µ-measurable �eld of states.

Later we will prove in Neshveyev's �rst theorem that if {ϕx}x∈G(0) , {ϕx}x∈G(0) are two

µ-measurable �eld of states whose states are equal on a conull subset of G(0), then (µ, {ϕx}x)

and (µ, {ϕx}x) de�ne the same state.

Remark 5.2.3. When there is no risk of confusion, we denote the µ-measurable �eld of

states {ϕx}x∈G(0) by {ϕx}x.

Lemma 5.2.4. Let G be a locally compact Hausdor� second countable étale groupoid and

let x ∈ G(0). Given g1, g2, g ∈ Gxx, ug1 · ug2 = ug1g2 and ug* = ug−1 .

125

Page 128: Characterization of Extremal KMS States on Groupoid C

Proof. Let h ∈ Gxx. Then

ug1 · ug2(h) =∑h1h2

ug1(h1)ug2(h2).

Note that for every h1, h2 such that h1 6= g1 or h2 6= g2, we have ug1 · ug2(h) = 0. Thus,

ug1 · ug2(h) =

1 if h = g1g2,

0 otherwise= ug1g2(h).

We also have

ug*(h) = ug(h−1)

= ug(h−1), since ug is real-valued,

=

1 if h−1 = g,

0 otherwise,

=

1 if h = g−1,

0 otherwise,

= ug−1(h).

Therefore the result follows.

Remark 5.2.5. If µ is purely atomic, it follows from Lemma 2.5.2 that that every family

{ϕx}x∈G(0) of states ϕx on C*(Gxx) is a µ-measurable �eld of states.

Given a state ϕ on C*(G), we obtain from Renault's disintegration theorem a unitary

representation of G. The following lemma uses this representation to �nd a representation

πx on C*(Gxx) for each x ∈ G(0). In the proof of Neshveyev's �rst theorem, we will use πx to

de�ne a state ϕx on C*(Gxx).

126

Page 129: Characterization of Extremal KMS States on Groupoid C

Lemma 5.2.6. Let G be a locally compact Hausdor� second countable étale groupoid. Let

(µ,G(0)∗H, L) be a unitary representation of G. Let x ∈ G(0). The linear map πx : C*(Gxx)→

B(Hx) de�ned by πx(ug) = Lg is a representation of C*(Gxx).

Proof. Let x ∈ G(0). Let g ∈ Gxx, then

πx(ug*) = πx(ug−1) = Lg−1 = Lg* = πx(ug)*.

Given g1, g2 ∈ Gxx,

πx(ug1ug2) = πx(ug1g2) = Lg1g2 = Lg1Lg2 = πx(ug1)πx(ug2).

Given a state ϕx on C*(Gxx), it corresponds to a representation πx : C*(Gx

x) → B(Hx).

We can use πx to de�ne a unitary map Lg : Hx → Hx by Lg = πx(ug), for every g ∈ Gxx.

Using this fact, we can apply Lemma 5.2.7 and obtain a Hilbert space Kx. We use this result

in the proof of Neshveyev's �rst theorem to show that there exists a state ψx corresponding

to Kx. Then we use Lemma 5.1.8 to �nd the state ϕ on C*(G).

Lemma 5.2.7. Let G be a locally compact Hausdor� second countable étale groupoid with

Haar system de�ned by the counting measures on Gx. Let {Lh}h∈G′ be a family of unitary

operators Lh : Hx → Hx on a Hilbert space Hx de�ned for all h ∈ Gxx and x ∈ G(0). Given

x ∈ G(0), let Kx be the family of functions v : Gx → Hx such that

v(gh) = Lh*v(g) for g ∈ Gx and h ∈ Gxx (5.16)

and

∑g∈Gx/Gxx

‖v(g)‖2 <∞. (5.17)

127

Page 130: Characterization of Extremal KMS States on Groupoid C

Then Kx is a Hilbert space.

Proof. Note that Gx/Gxx is de�ned by the following equivalence relation on Gx:

g1 ∼ g2 if g1 = g2h for some h ∈ Gxx.

In order to prove that (5.17) is well-de�ned, we will show that ‖v(g1)‖ = ‖v(g2)‖ in this case.

In fact, by de�nition of v,

‖v(g1)‖ = ‖v(g2h)‖ = ‖Lh*v(g2)‖ = ‖v(g2)‖,

since Lh is unitary.

Kx is a vector space. In fact, let u, v ∈ Kx, λ ∈ C. Then, for g ∈ Gx, h ∈ Gxx,

(u+ λv)(gh) = u(gh) + λv(gh)

= Lh*u(g) + Lh*λv(g)

= Lh* [(u+ λv)(g)] .

Now we can de�ne the inner product on Kx by

〈u, v〉 =∑

g∈Gx/Gxx

〈u(g), v(g)〉. (5.18)

Note that (5.18) de�nes an inner product on Kx. In fact, given g1, g2 ∈ Gx, h ∈ Gxx such

that g1 = g2h, we have for u, v ∈ Kx,

〈u(g1), v(g1)〉 = 〈u(g2h), v(g2h)〉 = 〈Lh*u(g2), Lh*v(g2)〉 = 〈u(g2), v(g2)〉.

It is easy to show this operation satis�es

〈u1 + λu2, v〉 = 〈u1, v〉+ λ〈u2, v〉 , 〈u, u〉 ≥ 0 and 〈u, v〉 = 〈v, u〉,

128

Page 131: Characterization of Extremal KMS States on Groupoid C

for u, u1, u2, v ∈ Kx, λ ∈ C.

Suppose 〈v, v〉 = 0. Let g ∈ Gx, then 〈v(g), v(g)〉 = 0 by de�nition. Therefore v = 0. We

denote the norm de�ned in (5.17) by ‖ · ‖.

Now we prove that Kx is Banach. Let {vn} be a Cauchy sequence on Kx. Given g ∈ Gx,

{vn(g)} is a Cauchy sequence on Hx, hence vn(g)→ v(g), where v : Gx → Hx is a function.

We show that v satis�es (5.16). Let g ∈ Gx, h ∈ Gxx. Then vn(gh)→ v(gh). But for each

n, vn(gh) = Lh*vn(g). Since Lh is unitary, it follows that

v(gh) = limn→∞

vn(gh) = limn→∞

Lh*vn(g) = Lh*v(g).

Now we prove that vn converges to v with respect to the norm. Since G is second countable

and étale, it follows from Proposition 3.2.11 that Gx is countable. Then Gx/Gxx is countable.

If Gx/Gxx is �nite, then pointwise convergence implies convergence in the norm. So we assume

Gx/Gxx is in�nite and denote its elements by a sequence {gk}. Given ε > 0, let n0 ≥ 0 such

that for every n,m ≥ n0, we have ‖vn − vm‖ < ε.

Fix k1. If n ≥ n0, we have

k1∑k=1

‖vn(gk)− v(gk)‖2 = limm→∞

k1∑k=1

‖vn(gk)− vm(gk)‖2 ≤ ε2.

Since k1 is arbitrary, we have

‖vn − v‖2 = limk1→∞

k1∑k=1

‖vn(gk)− v(gk)‖2 ≤ ε2.

Hence ‖vn − v‖ → 0. Therefore vn converges to v in the norm. Moreover, it is easy to see

that ‖v‖ < ∞. In fact, choose n such that ‖vn − v‖ < 1. Hence ‖v‖ ≤ ‖vn − v‖ + ‖vn‖ <

1 + ‖vn‖ <∞.

Proposition 5.2.8. Let G be a locally compact Hausdor� second countable étale groupoid.

Let f ∈ Cc(G) and assume there exists an open set U including the support of f such that

129

Page 132: Characterization of Extremal KMS States on Groupoid C

U ∩G′ = ∅. Then we can write f = f1 + . . . + fn where for every i = 1, . . . , n, fi ∈ Cc(Ui)

and Ui is an open bisection satisfying r(Ui) ∩ s(Ui) = ∅.

Proof. Let f ∈ Cc(G) and U an open set U containing its support such that U ∩ G′ = ∅.

Since K is compact, there exists an open set V whose closure is compact and K ⊂ V ⊂ U .

Then, by Lemma 3.2.13, for every g ∈ V there exists an open bisection Ug containing g

such that r(Ug) ∩ s(Ug) = ∅. However, V is compact, then there are g1, . . . , gn such that

Ug1 , . . . ,Ugn cover V . Denote Ui = Ugi for i = 1, . . . , n.

Let hii = 1n be the partition of unity subordinate to the open cover {Ui}ni=1. For each

i = 1, . . . , n, let fi = hif . Then fi ∈ Cc(Ui) and f = f1 + . . .+ fn.

Theorem 5.2.9. (Neshveyev) Let G be a locally compact Hausdor� second countable étale

groupoid. There is a one-to-one correspondence between states on C*(G) with centralizer

containing C0(G(0)) and pairs (µ, {ϕx}x) consisting of a probability measure µ on G(0) and

a µ-measurable �eld of states ϕx on C*(Gxx). Namely, the state corresponding to (µ, {ϕx}x)

is given by

ϕ(f) =

∫G(0)

∑g∈Gxx

f(g)ϕx(ug)dµ(x) for f ∈ Cc(G).

Proof. Endow G with the Haar system given by counting measures λx on Gx.

Assume ϕ is a state on C*(G) with centralizer containing C0(G(0)). Let (H, π, ξ) be the

corresponding GNS-triple. It follows from Lemma 4.3.5 that π satis�es the conditions of

Renault's disintegration theorem. Therefore there is a unitary representation (µ,G(0) ∗H, L)

of G such that H is isomorphic to L2(G(0) ∗H, µ) and π is equivalent to the integrated form

of (µ,G(0) ∗H, L). Here we identify H with L2(G(0) ∗H, µ) without loss of generality. Hence,

ϕ(f) = 〈π(f)ξ, ξ〉

=

∫G

f(g)〈Lgξs(g), ξr(g)〉Hr(g)∆(g)−12dµr(g)

130

Page 133: Characterization of Extremal KMS States on Groupoid C

=

∫G(0)

∑g∈Gx

f(g)〈Lgξs(g), ξx〉Hx∆(g)−12dµ(x). (5.19)

Now we prove that for every f ∈ Cc(G) and for µ-a.e. x ∈ G(0) we have

∑g∈Gx\Gxx

f(g)〈Lgξs(g), ξx〉Hx∆(g)−12 = 0. (5.20)

Using the lineariy of (5.20) with respect to f we can assume it is supported on an open

bisection U . Since the isotropy bundle G′ is closed from Lemma 3.2.13 and the sum in (5.20)

does not take into account elements in G′, we can assume U ∩G′ = ∅. Then, by Proposition

5.2.8, we can assume r(U) ∩ s(U) = ∅ without loss of generality.

Let h ∈ Cc(r(U)). Then f ·h = 0. In fact, given g ∈ G, (f ·h)(g) = f(g)h(s(g)). Suppose

f(g) 6= 0. Then g ∈ U . Thus s(g) ∈ s(U). Since s(U) ∩ r(U) = ∅ by hypothesis, it follows

that s(g) /∈ r(U). Thus h(s(g)) = 0. Therefore f · h = 0.

Since ϕ has centralizer containing C0(G(0)), we have ϕ(h · f) = ϕ(f · h) = 0. Applying

(5.19) for h · f , it follows that

0 = ϕ(h · f)

=

∫G(0)

∑g∈Gx

(h · f)(g)〈Lgξs(g), ξx〉∆(g)−12dµ(x)

=

∫G(0)

∑g∈Gx

h(r(g))f(g)〈Lgξs(g), ξx〉∆(g)−12dµ(x), by Lemma 3.3.4,

=

∫G(0)

h(x)∑g∈Gx

f(g)〈Lgξs(g), ξx〉∆(g)−12dµ(x)

=

∫r(U)

h(x)∑g∈Gx

f(g)〈Lgξs(g), ξx〉∆(g)−12dµ(x), since h ∈ Cc(r(U)),

=

∫r(U)

h(x)∑

g∈Gx\Gxx

f(g)〈Lgξs(g), ξx〉∆(g)−12dµ(x), since U ∩G′ = ∅.

Since h ∈ Cc(r(U)) is arbitrary, it follows that for µ-a.e x ∈ r(U), (5.20) holds. Since

131

Page 134: Characterization of Extremal KMS States on Groupoid C

f ∈ Cc(U), (5.20) holds for every x /∈ r(U). Indeed, given x /∈ r(U), g ∈ Gx, then g /∈ U .

Hence f(g) = 0. Therefore (5.20) holds for all f ∈ Cc(U), for µ-a.e. x. Therefore, (5.20) is

valid for every f ∈ Cc(G).

Let f ∈ Cc(G). Applying (5.20) on (5.19), we have

ϕ(f) =

∫G(0)

∑g∈Gx

f(g)〈Lgξs(g), ξx〉Hx∆(g)−12dµ(x)

=

∫G(0)

∑g∈Gxx

f(g)〈Lgξs(g), ξx〉Hx∆(g)−12dµ(x)

=

∫G(0)

∑g∈Gxx

f(g)〈Lgξx, ξx〉Hx∆(g)−12dµ(x).

From Lemma 5.2.6, we can de�ne for every x ∈ G(0) a representation πx : C*(Gxx) →

B(Hx) such that πx(ug) = Lg for all g ∈ Gxx. Then

ϕ(f) =

∫G(0)

∑g∈Gxx

f(g)〈πx(ug)ξx, ξx〉Hx∆(g)−12dµ(x).

By Proposition 4.1.8, ∆(g) = 1 for all g ∈ Gxx for µ-a.e. x ∈ G(0). Hence,

ϕ(f) =

∫G(0)

∑g∈Gxx

f(g)〈πx(ug)ξx, ξx〉Hxdµ(x).

Since ξ ∈ L2(X ∗ H, µ), the set V ⊂ G(0) of elements x such that ξx 6= 0, is measurable.

Let µ be a measure on G(0) such that dµ/dµ(x) = ‖ξx‖2. Note that µ is supported on V .

Moreover, µ is a probability measure. In fact,

∫G(0)

dµ(x) =

∫G(0)

‖ξx‖2dµ(x) = ‖ξ‖2 = 1.

Let ξ be de�ned by ξx = 0 if x /∈ V , and ξx = ξx/‖ξx‖ if x ∈ V . Then ξ ∈ L2(X ∗ H, µ) and

132

Page 135: Characterization of Extremal KMS States on Groupoid C

‖ξ‖ = 1. Moreover,

ϕ(f) =

∫G(0)

∑g∈Gxx

f(g)〈πx(ug)ξx, ξx〉Hxdµ(x)

=

∫V

∑g∈Gxx

f(g)〈πx(ug)ξx, ξx〉Hxdµ(x)

=

∫V

∑g∈Gxx

f(g)〈πx(ug)ξx, ξx〉Hx

‖ξx‖2‖ξx‖2dµ(x)

=

∫V

∑g∈Gxx

f(g)〈πx(ug)ξx, ξx〉Hx‖ξx‖2dµ(x)

=

∫V

∑g∈Gxx

f(g)〈πx(ug)ξx, ξx〉Hxdµ(x)

For every x ∈ V , there is a state ϕx on C*(Gxx) such that

ϕx(ug) = 〈πx(ug)ξx, ξx〉Hx , for every g ∈ Gxx. (5.21)

From Remark 5.2.2, we can choose a µ-measurable �eld of states {ϕx} such that ϕx is de�ned

by (5.21) for every x ∈ V . Therefore,

ϕ(f) =

∫V

∑g∈Gxx

f(g)ϕx(ug)dµ(x)

=

∫G(0)

∑g∈Gxx

f(g)ϕx(ug)dµ(x).

Conversely, let µ be a probability measure on G(0) and let a µ-measurable �eld of states

ϕx on C*(Gxx).

Given x, let (Hx, πx, ξx) be the GNS-triple for ϕx. De�ne for every h ∈ Gxx, Lh = πx(uh).

Each Lh is unitary, since

Lh* = πx(uh)* = πx(uh*) = πx(uh−1) = πx(u−1h ) = πx(uh)

−1 = L−1h .

133

Page 136: Characterization of Extremal KMS States on Groupoid C

Let Kx be the Hilbert space as in Lemma 5.2.7. Let ϑx : Cc(G)→ B(Kx) be de�ned by

(ϑx(f)v)(g) =∑

h∈Gr(g)f(h)v(h−1g) for f ∈ Cc(G), g ∈ Gx.

First we show that each ϑx(f) is in B(Kx). In fact, ϑx(f) is linear on Kx by de�nition.

Let v ∈ Kx, g ∈ Gx and k ∈ Gxx. Then

(ϑx(f)v)(gk) =∑

h∈Gr(g)f(h)v(h−1gk)

=∑

h∈Gr(g)f(h)Lk*v(h−1g)

= Lk*

∑h∈Gr(g)

f(h)v(h−1g)

= Lk* [(ϑx(f)v)(g)] .

Also, we need to show that ‖ϑx(f)‖2 < ∞ for every f ∈ Cc(G). Since ϑx is linear, we can

assume f ∈ Cc(U), where U ⊂ G is an open bisection.

Let g ∈ Gx. Assume r(g) /∈ r(U). Then, for every h ∈ Gr(g), it follows that h /∈ U , which

implies f(h) = 0. Hence (ϑx(f)v)(g) = 0. Now assume r(g) ∈ r(U). Since U is an open

bisection, there exists a unique hr(g) in U such that hr(g) ∈ Gr(g). Then

(ϑx(f)v)(g) =∑

h∈Gr(g)f(h)v(h−1g) = f(hr(g))v((hr(g))−1g).

Note that if g1, g2 ∈ Gx are equivalent, then r(g1) = r(g2). Indeed, there is k ∈ Gxx such

that g1 = g2k. Hence r(g1) = r(g2k) = r(g2). Moreover, if r(g1) ∈ r(U), it follows that

‖v((hr(g1))−1g1)‖ = ‖v((hr(g1))−1g2k)‖

= ‖v((hr(g2))−1g2k)‖, since r(g1) = r(g2),

= ‖Lk*v((hr(g2))−1g2)‖

134

Page 137: Characterization of Extremal KMS States on Groupoid C

= ‖v((hr(g2))−1g2)‖.

Therefore ‖v(hr(g)g)‖ is well-de�ned for g ∈ Gx/Gxx with r(g) ∈ r(U). Then,

‖ϑx(f)v‖2 =∑

g∈Gx/Gxxr(g)∈r(U)

‖(ϑx(f)v)(g)‖2

=∑

g∈Gx/Gxxr(g)∈r(U)

|f(hr(g))|2‖v((hr(g))−1g)‖2

≤ ‖f‖2∞

∑g∈Gx/Gxxr(g)∈r(U)

‖v((hr(g))−1g)‖2. (5.22)

Note that for g1, g2 with r(g1) ∈ U , g1, g2 are equivalent if, and only if (hr(g1))−1g1 and

(hr(g2))−1g2 are equivalent. In fact, suppose g1 and g2 are equivalent. Then g1 = g2k for

some k ∈ Gxx. Recall that r(g1) = r(g2). Then hr(g1) = hr(g2) and (hr(g1))−1g1 = (hr(g2))−1g2k.

Therefore (hr(g1))g1 and (hr(g2))g2 are equivalent.

Conversely, assume (hr(g1))−1g1 and (hr(g2))−1g2 are equivalent. Then there is k ∈ Gxx such

that (hr(g1))−1g1 = (hr(g2))−1g2k. Note that

s(hr(g1)) = r((hr(g1))−1) = r((hr(g1))−1g1) = r((hr(g2))−1g2) = r((hr(g2))−1) = s(hr(g2)).

Since hr(g1), hr(g2) ∈ U and U is an open bisection, it follows that hr(g1) = hr(g2). Then

g1 = g2k. Therefore g1 and g2 are equivalent.

Then the set of elements (hr(g))−1g for g ∈ Gx/Gxx with r(g) ∈ r(U) is a subset of the set

{g : g ∈ Gx/Gxx}. Hence, by (5.22),

‖ϑx(f)v‖2 ≤ ‖f‖2∞

∑g∈Gx/Gxxr(g)∈r(U)

‖v((hr(g))−1g)‖2

≤ ‖f‖2∞

∑g∈Gx/Gxx

‖v(g)‖2

135

Page 138: Characterization of Extremal KMS States on Groupoid C

= ‖f‖2∞‖v‖2

<∞.

Therefore, ‖ϑx(f)‖ <∞ for every f ∈ Cc(G).

Now we prove that ϑx is a representation of Cc(G) on Kx.

• ϑx(f1 · f2) = ϑx(f1)ϑx(f2).

Given f1, f2 ∈ Cc(G), v ∈ Kx, g ∈ Gx,

(ϑx(f1 · f2)v)(g) =∑

h∈Gr(g)(f1 · f2)(h)v(h−1g)

=∑

h∈Gr(g)

∑k∈Gr(h)

f1(k)f2(k−1h)

v(h−1g), by (3.4) on page 55,

=∑

h∈Gr(g)

∑k∈Gr(g)

f1(k)f2(k−1h)v(h−1g)

=∑

k∈Gr(g)f1(k)

∑h∈Gr(g)

f2(k−1h)v(h−1g)

=∑

k∈Gr(g)f1(k)

∑h∈Gr(g)

f2(k−1h)v(h−1kk−1g)

=∑

k∈Gr(g)f1(k)

∑h∈Gr(g)

f2(k−1h)v((k−1h)−1k−1g).

Making the change of variables h = k−1h, then h ∈ Gr(k−1) = Gr(k−1g). Then,

(ϑx(f1 · f2)v)(g) =∑

k∈Gr(g)f1(k)

∑h∈Gr(k−1g)

f2(h)v(h−1k−1g)

=∑

k∈Gr(g)f1(k)(ϑx(f2)v)(k−1g)

= ϑx(f1)(ϑx(f2)v)(g).

• ϑx(f*) = ϑx(f)*.

136

Page 139: Characterization of Extremal KMS States on Groupoid C

Given v, w ∈ Kx,

〈ϑx(f)v, w〉 =∑

g∈Gx/Gxx

〈(ϑx(f)v)(g), w(g)〉

=∑

g∈Gx/Gxx

∑h∈Gr(g)

f(h)〈v(h−1g), w(g)〉.

On the other hand,

〈v, ϑx(f*)w〉 =∑

g∈Gx/Gxx

〈v(g), (ϑx(f*)w)(g)〉

=∑

g∈Gx/Gxx

∑h∈Gr(g)

f*(h)〈v(g), w(h−1g)〉

=∑

g∈Gx/Gxx

∑h∈Gr(g)

f(h−1)〈v(g), w(h−1g)〉 (5.23)

Note that for every g ∈ Gx/Gxx, h ∈ Gr(g), there exist unique g ∈ Gx/G

xx, h ∈ Gr(g)

satisfying h = h−1

g = h−1g.

Then we can make the change of variables in (5.23), obtaining

〈v, ϑx(f*)〉 =∑

g∈Gx/Gxx

∑h∈Gr(g)

f(h)〈v(h−1g), w(g)〉

= 〈ϑx(f)v, w〉.

Thus ϑx(f*) = ϑ(f)*.

Let ζx be de�ned by ζx(g) = πx(ug*)ξx if g ∈ Gxx, and ζx(g) = 0 if g ∈ Gx \Gx

x. Note that

137

Page 140: Characterization of Extremal KMS States on Groupoid C

ζx is in Kx. Given h ∈ Gxx, g ∈ Gx, ζx(gh) = 0 if g /∈ Gx

x. If g ∈ Gxx, by de�nition of Kx,

ζx(gh) = πx(ugh*)ξx = πx(uh*ug*)ξx = πx(uh*)π(ug*)ξx = πx(uh*)ζx(g).

Moreover, ‖ζx‖ = ‖ζx(x)‖ = ‖ξx‖ = 1. Let ψx be the state on C*(G) de�ned by ψx(f) =

〈ϑx(f)ζx, ζx〉. Then, for f ∈ Cc(G),

ψx(f) =∑

g∈Gx/Gxx

〈ϑx(f)ζx(g), ζx(g)〉

= 〈(ϑx(f)ζx)(x), ζx(x)〉

=∑h∈Gx

f(h)〈ζx(h−1x), ζx(x)〉

=∑h∈Gx

f(h)〈ζx(h−1), ζx(x)〉

=∑g∈Gxx

f(g)〈πx(ug−1*)ξx, πx(ux*)ξx〉, by de�nition of ζx,

=∑g∈Gxx

f(g)〈πx(ug**)ξx, πx(ux*)ξx〉

=∑g∈Gxx

f(g)〈πx(ug)ξx, πx(ux)ξx〉

=∑g∈Gxx

f(g)ϕx(ug). (5.24)

Note that C0(G(0)) is in the centralizer of ψx. From Lemma 3.3.16, we have that Cc(G(0))

is dense in C0(G(0)). Using the continuity of ψx, it is su�cient to show that Cc(G(0)) is in

the centralizer of ϕ. Given f ∈ Cc(G) and h ∈ Cc(G(0)), we have

ψx(f · h) =∑g∈Gxx

(f · h)(g)ϕx(ug) =∑g∈Gxx

f(g)h(x)ϕx(ug) =∑g∈Gxx

(h · f)(g)ϕx(ug) = ψx(h · f).

By assumption, the map x 7→ ψx(f) is µ-measurable for every f ∈ Cc(G). By Lemma 5.1.8,

we can de�ne a state ϕ on C*(G) by ϕ(f) =∫G(0) ψx(f)dµ(x).

138

Page 141: Characterization of Extremal KMS States on Groupoid C

Finally we show that if (µ, {ϕx}x) and (µ, {ϕx}x) de�ne the same state ϕ, then µ = µ and

ϕx = ϕx for µ-a.e. x.

Recall from Proposition 3.2.12 that G(0) is clopen in G. Let f ∈ Cc(G(0)). Then,

∫G(0)

∑g∈Gxx

f(g)ϕx(ug)dµ(x) =

∫G(0)

∑g∈Gxx

f(g)ϕx(ug)dµ(x)∫G(0)

f(x)ϕx(ux)dµ(x) =

∫G(0)

f(x)ϕx(ux)dµ(x), since f ∈ Cc(G(0)),∫G(0)

f(x)dµ(x) =

∫G(0)

f(x)dµ(x), ux is the unity in C*(Gxx).

Since f is arbitrary, we have µ = µ.

We will prove that ϕx = ϕx for µ-a.e. x. Let W ⊂ G be the set of g ∈ G′ such that

x = r(g) and ϕx(ug) 6= ϕx(ug). Let V ⊂ G(0) be the set of x ∈ G(0) such that ϕx 6= ϕx. Note

that V = r(W ).

Given g ∈ W , let Ug be an open bisection containing g. Using the topological properties

of G, we can assume the family {Ug}g∈W is countable without loss of generality.

Given f ∈ Cc(r(Ug)), there exists F ∈ Cc(Ug) such that f = F ◦ r|−1Ug . We denote by

hx = r|−1Ug (x), hence F (hx) = f(x) for every x ∈ r(Ug). Hence,

∫r(Ug∩G′)

f(x) [ϕx(hx)− ϕx(hx)] dµ(x)

=

∫r(Ug∩G′)

F (hx) [ϕx(hx)− ϕx(hx)] dµ(x)

=

∫r(Ug∩G′)

∑g∈Gxx

F (g) [ϕx(hx)− ϕx(hx)] dµ(x), since hx ∈ G′,

=

∫r(Ug)

∑g∈Gxx

F (g) [ϕx(hx)− ϕx(hx)] dµ(x)

=ϕ(F )− ϕ(F )

=0.

139

Page 142: Characterization of Extremal KMS States on Groupoid C

Since f is arbitrary, it follows that ϕx(uhx) = ϕx(uhx) for µ-a.e. x ∈ r(Ug ∩ G′). Then

µ(r(W∩Ug)) = 0. The family of Ug indexed by g ∈ W is countable, and V = ∪g∈W r(W∩Ug).

Then µ(V ) = 0.

Remark 5.2.10. If we de�ne a dynamics τ on C*(G) such that every function in C0(G(0)) is

τ -invariant, it follows from Lemma 5.1.32, that every KMS state ϕ on C*(G) has centralizer

containing C0(G(0)). In this case, we can apply the �rst Neshveyev's theorem.

5.3 Second Theorem

The second theorem shows the conditions that the pair (µ, {ϕx}x) satis�es if its corresponding

state is KMS. We begin by de�ning a dynamics τ on C*(G).

De�nition 5.3.1. Let G be a topological groupoid. A continuous R-valued 1-cocyle on G

is a continuous function c : G→ R such that c(gh) = c(g) + c(h) for every (g, h) ∈ G(2).

Now we will prove a lemma which will help us to show that the dynamics de�ned by the

cocycle is well-de�ned.

Lemma 5.3.2. Let G be a locally compact Hausdor� second countable étale groupoid. Let

K be R or C. Let F : K×G→ C be a continuous function. Let U ⊂ G be an open bisection

and suppose there exists a compact set K ⊂ U such that F (z, g) = 0 for every z ∈ K, g /∈ K.

De�ne for every z ∈ K the function Fz : G→ C by Fz(g) = F (z, g). Then the map from

K to Cc(G) de�ned by z 7→ Fz is continuous with respect to the norm of C*(G).

Proof. We can assume K = C without loss of generality. Note that each Fz is continuous

and its support is included in the compact set K.

Fix z0 ∈ C. Note that F is continuous at (z0, g) for every g ∈ K. Thus, for each g ∈ K

there exists δg > 0 and an open neighborhood Ug of g such that

|F (z, h)− F (z0, g)| ≤ ε

2,

140

Page 143: Characterization of Extremal KMS States on Groupoid C

for every (z, h) satisfying z ∈ Bδg(z0) = {w ∈ C : |w − z0| < δg}, h ∈ Ug.

The family {Bδg(z0)× Ug}g∈K forms an open cover for the compact set {z0} ×K. Then

there are g1, . . . , gn ∈ K such that {Bδgi(z0)× Ugi}ni=1 covers {z0} ×K. Let

δ =1

2min{δg1 , . . . , δgn}.

Now let z ∈ C such that |z − z0| < δ. Let g ∈ K. There exists i ∈ {1, . . . , n} such that

(z0, g) ∈ Bδgi(z0)× Ugi .

Since δ < δgi , we have (z, g) ∈ Bδgi(z0)× Ugi . Thus,

|F (z, g)− F (z0, g)| ≤ |F (z, g)− F (z0, gi)|+ |F (z0, gi)− F (z0, g)|

≤ ε

2+ε

2

≤ ε,

because (z, g), (z0, g) ∈ Bδgi× Ugi . Since g ∈ K is arbitrary, we have

‖Fz − Fz0‖ ≤ ‖Fz − Fz0‖∞, from Proposition 3.3.11,

= supg∈K|Fz(g)− Fz0(g)|

= supg∈K|F (z, g)− F (z0, g)|

≤ ε.

Therefore the map z 7→ Fz is continuous.

The next lemma is useful to prove that the elements of Cc(G) are entire analytic for our

dynamics.

Corollary 5.3.3. Using the same conditions of Lemma 5.3.2 for K = C, suppose that for

141

Page 144: Characterization of Extremal KMS States on Groupoid C

every g ∈ C the function z 7→ F (z, g) is di�erentiable. De�ne F ′ : C×G→ C by

F ′(z, g) =∂F

∂z(z, g) = lim

h→0

F (z + h, g)− F (z, g)

h.

Suppose F ′ is continuous. For every z ∈ C, de�ne F ′z : G → C by Fz(g) = F ′(z, g). Then

we have the limit

limh→0

∥∥∥∥Fz+h − Fzh− F ′z

∥∥∥∥ = 0,

for every z ∈ C.

Proof. Fix z ∈ C. By de�nition, F ′z is supported on K. De�ne the function H : C×G→ C

by

H(h, g) =

Fz+h(g)− Fz(g)

h− F ′z(g) if h 6= 0,

0 if h = 0.

This function continuous. Moreover, using the notation from of 5.3.2, we have that Hh is

supported on K for every h. Then by Lemma 5.3.2,

0 = limh→0‖Hh −H0‖

= limh→0‖Hh‖

= limh→0

∥∥∥∥Fz+h − Fzh− F ′z

∥∥∥∥ .

Now we de�ne a dynamics on C*(G) for a continuous R-valued cocycle. Throughout this

section, the dynamics is �xed.

Lemma 5.3.4. Let G be a locally compact Hausdor� second countable étale groupoid and

c : G → R a continuous R-valued cocycle. De�ne τ = {τt}t∈R by τt(f)(g) = eitc(g)f(g) for

142

Page 145: Characterization of Extremal KMS States on Groupoid C

every g ∈ G, f ∈ Cc(G). Then (C*(G), τ) is a C*-dynamical system.

Proof. Fix t ∈ R. Note that τt(f) ∈ Cc(G) for every f ∈ Cc(G). By de�nition, τt is a linear

map on Cc(G).

• τt is a ∗-homomorphism.

Given f1, f2 ∈ Cc(G), g ∈ G,

τt(f1 · f2)(g) = eitc(g)(f1 · f2)(g)

= eitc(g)∑g1g2=g

f1(g1)f2(g2)

=∑g1g2=g

eitc(g1g2)f1(g1)f2(g2)

=∑g1g2=g

(eitc(g1)f1(g1))(eitc(g2)f2(g2))

=∑g1g2=g

τt(f1)(g1)τt(f2)(g2)

= (τt(f1) · τt(f2))(g).

Given f ∈ Cc(G), g ∈ G, we have

τt(f*)(g) = eitc(g)f*(g) = e−itc(g)f(g−1) = eitc(g−1)f(g−1) = τt(f)(g−1) = τt(f)*(g).

Thus τt is a ∗-homomorphism.

• τt ◦ τs = τt+s and τ0 is the identity

Given f ∈ Cc(G), g ∈ G, t, s ∈ R,

τt ◦ τs(f)(g) = τt(τs(f))(g)

= eitc(g)τs(f)(g)

= eitc(g)eisc(g)f(g)

143

Page 146: Characterization of Extremal KMS States on Groupoid C

= ei(t+s)c(g)f(g)

= τt+s(f)(g).

By de�nition, τ0 is the identity.

Each τt is invertible. Therefore every τt is a ∗-automorphism.

• τ is strongly continuous

Let f ∈ Cc(G). Let K be the support of f . Assume there exists and open bisection U

such that K ⊂ U .

De�ne F : R×G→ C by F (t, g) = τt(f)(g) = eitc(g)f(g). Both c and f are continuous

functions, then F is continuous. Let K be the compact support of f . Then F (t, g) = 0

for every t ∈ R, g /∈ K. Using the notation of Lemma 5.3.2, we have τt(f) = Ft.

Therefore, by Lemma 5.3.2, the function t 7→ τt(f) is continuous.

Now let f ∈ Cc(G) be arbitrary. There are open bisections U1, . . . ,Un and functions

f1, . . . , fn such that each fk ∈ Cc(Uk) and f = f1 + . . . + fn. Since the function

t 7→ τt(fk) is continuous for every k and each τt is linear, it follows that the function

t 7→ τt(f) is continuous.

Note that, for every f ∈ Cc(G), ‖τ(f)‖ ≤ ‖f‖. In fact, let π be a ∗-representation of

Cc(G). Then π ◦ τ is an ∗-representation of Cc(G). Then. by de�ntion of the norm on

Cc(G), we have

‖π(τt(f))‖ = ‖π ◦ τt(f)‖f‖.

Since π is arbitrary, using the de�ntion of ‖π(τt(f))‖, we have

‖τt(f)‖ ≤ ‖f‖.

Therefore, from Lemma 5.1.12, τ de�nes a dynamics on C*(G).

144

Page 147: Characterization of Extremal KMS States on Groupoid C

Lemma 5.3.5. Let G be a locally compact second countable étale groupoid. Fix β > 0.

Assume ϕ is a KMSβ-state. Then the centralizer of ϕ contains C0(G(0)).

Proof. Let h ∈ C0(G(0)). Note that for every x ∈ G(0), c(x) = c(xx) = 2c(x) = 0. Then

c(x) = 0. Since h is supported on G(0), we have for every t ∈ R,

τt(h)(g) = eitc(g)h(g) = h(g) for g ∈ G.

Then τt(h) = h. Therefore, by Lemma 5.1.32, h is in the centralizer of ϕ.

Lemma 5.3.5 shows that we can apply Theorem 5.2.9 for every KMS state ϕ on C*(G).

Then ϕ corresponds to a pair (µ, {ϕy}y∈G(0)) as shown in that theorem.

Lemma 5.3.6. Let G be a locally compact Hausdor� second countable étale groupoid. Then

every f ∈ Cc(G) is entire analytic for τ .

Proof. Let f ∈ Cc(G). Since the set of entire analytic elements in C*(G) forms a vector space

and each function in Cc(G) can be decomposed as a sum of functions in Cc(G) supported on

open bisections, we can assume f is supported on an open bisection without loss of generality.

Let K = supp(f).

De�ne F : C × G → C by F (z, g) = eizc(g)f(g). Then F is continuous. De�ne for every

z ∈ C, Fz : G→ C by

Fz(g) = F (z, g) = eizc(g)f(g).

Note that Ft(f) = τt(f) for every t ∈ R. For every z ∈ C, Fz is continuous and supported

on K. Analogously, de�ne F ′ : C×G→ by

F ′(z, g) =∂F

∂z(z, g) = ic(g)eizc(g)f(g).

145

Page 148: Characterization of Extremal KMS States on Groupoid C

Then F ′ is continuous. From Corollary 5.3.3 we have the limit

limh→0

∥∥∥∥Fz+h − Fzh− F ′z

∥∥∥∥ = 0.

Therefore f is entire analytic. In this case, we can write τz(f) = Fz for z ∈ C.

The following lemma proves some properties of compactly supported functions. These

properties will be used in the proof of Neshneyev's theorem.

Lemma 5.3.7. Let G be a locally compact Hausdor� second countable étale groupoid. Let

U be a an open bisection and let f1 ∈ Cc(U), f2 ∈ Cc(G). Then, given g ∈ G,

(f1 · f2)(g) =

f1(hx)f2((hx)−1g), for x ∈ r(U), g ∈ Gx,

0, if r(g) /∈ r(U).(5.25)

(f2 · f1)(g) =

f1(hx)f2(g(hx)−1), for x ∈ s(U), g ∈ Gx,

0, if s(g) /∈ s(U).(5.26)

For x ∈ r(U), hx denotes the unique element in U ∩Gx. Analogously, for x ∈ s(U), hx is the

unique element in U ∩Gx.

Proof. Let f1 ∈ Cc(U), f2 ∈ Cc(G).

Equation (5.25): Let x ∈ G(0), g ∈ Gx. Note that Gx ∩ U = ∅ if x /∈ r(U). From Lemma

3.3.3, we have

(f1 · f2)(g) =∑h∈Gx

f1(h)f2(h−1g)

=∑

h∈Gx∩U

f1(h)f2(h−1g), since f1 ∈ Cc(U),

=

f1(hx)f2((hx)−1g), for x ∈ r(U), g ∈ Gx,

0, if r(g) /∈ r(U).

146

Page 149: Characterization of Extremal KMS States on Groupoid C

Equation (5.26): Let x ∈ G(0), g ∈ Gx. Note that Gx ∩ U = ∅ if x /∈ s(U). From Lemma

3.3.3, we have

(f2 · f1)(g) =∑h∈Gx

f2(gh−1)f1(h)

=∑

h∈Gx∩U

f2(gh−1)f1(h), since f1 ∈ Cc(U),

=

f2(g(hx)−1)f1(hx), if x ∈ s(U),

0, otherwise.

Lemma 5.3.8. Let G be a locally compact Hausdor� second countable étale groupoid. Let

c be an R-valued 1-cocyle. A measure µ on G(0) is quasi-invariant with Radon-Nikodym

derivative ec if, and only if, for every open bisection U ⊂ G, we have

dT∗µ

dµ(x) = ec(hx), (5.27)

for x ∈ s(U), where hx ∈ U is the unique element such that s(hx) = x and T : r(U)→ s(U)

is the homeomorphism de�ned by T (r(hx)) = x. In particular, T = s|U ◦ r|−1U .

Proof. Let U be an open bisection in G. Then hx = s|−1U (x) for every x ∈ s(U). Let

T : r(U)→ s(U) such that T (r(hx)) = x for every x ∈ s(U). Then, for every x,

x = T (r(hx)) = T (r|U(hx)) = T (r|U(s|−1U (x))).

Therefore T = (r|U ◦ s|−1U )−1 = s|U ◦ r|−1

U .

First we show a formula which holds if, and only if, condition (5.27) is satis�ed. Note

that equation (5.27) holds if, and only if, for every f ∈ Cc(s(U)),

∫s(U)

f(x)d(T∗µ)(x) =

∫s(U)

ec(hx)f(x)dµ(x). (5.28)

147

Page 150: Characterization of Extremal KMS States on Groupoid C

Recall that a measure µ on G(0) is quasi-invariant with Radon-Nikodym derivative ec if,

and only if, for every f ∈ Cc(G),

∫G

f(g)dµr(g) =

∫G

ec(g)f(g)dµs(g).

From the de�nition of µr, µs, this is equivalent to∫r(U)

∑g∈Gx

f(g)dµ(x) =

∫s(U)

∑g∈Gx

f(g)dµ(x). (5.29)

Since f ∈ Cc(U), we can consider only g ∈ U in the integrals. Recall that U ∩ Gx = {hx},

if x ∈ r(U), where hx = r|−1U (x). Analogously, U ∩ Gx = {hx} if x ∈ s(U). Then we can

rewrite (5.29) as

∫r(U)

f(hx)dµ(x) =

∫s(U)

ec(hx)f(hx)dµ(x). (5.30)

Therefore µ is quasi-invariant with Radon-Nikodym derivative ec if, and only if, (5.30) holds

for every open bisection U , f ∈ Cc(U).

Note that for every x ∈ r(U), hx = r|−1U (x) = s|−1

U ◦ s|U ◦ r|−1U (x) = s|−1

U (Tx) = hTx. Then

(5.30) is equivalent to

∫r(U)

f(hTx)dµ(x) =

∫s(U)

ec(hx)f(hx)dµ(x). (5.31)

There is a bijection from Cc(s(U)) to Cc(U) given by f 7→ f = f ◦s|U . This holds because

s|U : U → s(U) is a homeomorphism. Therefore, for every x ∈ s(U), f(x) = f(s(hx)) =

f(hx). In the rest of this proof, given f ∈ Cc(s(U)), we denote by f its corresponding function

in Cc(U). Analogously, given f ∈ Cc(U), f is the corresponding function in Cc(s(U)).

Assume µ is quasi-invariant with Radon-Nikodym derivative ec. Let U be an open bisec-

148

Page 151: Characterization of Extremal KMS States on Groupoid C

tion of G. Let f ∈ Cc(s(U)). Then

∫s(U)

f(x)d(T∗µ)(x) =

∫s(U)

f(hx)d(T∗µ)(x)

=

∫r(U)

f(hTx)dµ(x), from the de�nition of T∗µ,

=

∫s(U)

ec(hx)f(hx)dµ(x), by (5.31)

=

∫s(U)

ec(hx)f(x)dµ(x).

Hence, (5.28) holds, then (5.27) holds.

Conversely, suppose (5.28) holds. Given an open bisection U , f ∈ Cc(U), we have

∫r(U)

f(hTx)dµ(x) =

∫r(U)

f(Tx)dµ(x)

=

∫s(U)

f(x)d(T∗µ)(x)

=

∫s(U)

ec(hx)f(x)dµ(x)

=

∫s(U)

ec(hx)f(hx)dµ(x).

Then (5.31) holds for f . Therefore, µ is quasi-invariant with Radon-Nikodym derivative

ec.

Lemma 5.3.9. Let ϕ be a state on C*(G) with centralizer containing C0(G(0)). Assume ϕ

corresponds to the pair (µ, {ϕx}x). Then ϕ is τ -invariant if, and only if,

ϕx(ug) = 0 for every g ∈ Gxx \ c−1(0), µ-a.e. x. (5.32)

Proof. Assume ϕ is τ -invariant. It follows from the continuity of c that c−1(0) is closed. Let

g ∈ G such that c(g) 6= 0. Let t ∈ R such that tc(g) ∈ (0, 2π). Then 1 − eitc(g) 6= 0. There

exists an open bisection U containing g such that 1− eitc(g) 6= 0 for every g ∈ U .

149

Page 152: Characterization of Extremal KMS States on Groupoid C

Let f ∈ Cc(s(U)) and de�ne f = f ◦ s|U ∈ Cc(U). Let x ∈ G(0). Suppose g ∈ Gxx ∩ U .

Then x ∈ s(U) and g = hx, where hx = s|−1U (x) denotes the unique element in U ∩ Gx.

Hence, we can write

ϕ(f) =

∫G(0)

∑g∈Gxx

f(g)ϕx(ug)dµ(x)

=

∫G(0)

∑g∈Gxx∩U

f(g)ϕx(ug)dµ(x)

=

∫s(U)

∑g∈Gxx∩U

f(g)ϕx(ug)dµ(x)

=

∫s(U)

∑g∈Gx∩U

χG′(g)f(g)ϕx(ug)dµ(x)

=

∫s(U)

χG′(hx)f(hx)ϕx(uhx)dµ(x)

=

∫s(U)

χG′(hx)ϕx(uhx)f(x)dµ(x), by de�nition of f . (5.33)

Analogously,

ϕ(τt(f)) =

∫s(U)

eitc(hx)χG′(hx)ϕx(uhx)f(x)dµ(x). (5.34)

Since ϕ(τt(f)) = ϕ(f) by hypothesis, then from equations (5.33) and (5.34) we have

∫s(U)

[1− eitc(hx)]χG′(hx)ϕx(uhx)f(x)dµ(x) = 0.

The function f ∈ Cc(s(U)) is arbitrary and each 1− eitc(hx) 6= 0. Then, for µ-a.e. x ∈ s(U),

χG′(hx)ϕx(uhx) = 0,

or equivalently, for µ-a.e. x ∈ s(U), g ∈ U ∩Gx

χG′(g)ϕx(ug) = 0. (5.35)

150

Page 153: Characterization of Extremal KMS States on Groupoid C

Note that G \ c−1(0) can be covered by a countable family of open bisections {Un}n∈Nsuch that (5.35) holds for each Un. Therefore, for µ-a.e.x ∈ G(0), for all g ∈ Gx \ c−1(0),

χG′(g)ϕx(ug) = 0.

Thus, for µ-a.e. x ∈ G(0), for all g ∈ Gxx, ϕx(ug) = 0.

Conversely, assume (5.32) holds. Let f ∈ Cc(G). Then, for µ-a.e. x ∈ G(0) and all g ∈ Gxx,

ϕ(ug) 6= 0 implies c(g) = 0. Then, for µ-a.e. x ∈ G(0),

∑g∈Gxx

f(g)ϕx(ug) =∑g∈Gxx

eitc(g)f(g)ϕx(ug). (5.36)

Therefore,

ϕ(τt(f)) =

∫G(0)

∑g∈Gxx

eitc(g)f(g)ϕx(ug)dµ(x)

=

∫G(0)

∑g∈Gxx

f(g)ϕx(ug)dµ(x)

= ϕ(f).

Since ϕ and ϕ ◦ τt are continuous functions and Cc(G) is dense in C*(G), it follows that

ϕ(τt(a)) = ϕ(a) for every a ∈ C*(G). In other words, ϕ is τ -invariant.

Now we prove Neshveyev's second theorem. Note that in this theorem we assume a

di�erent de�nition for KMS-states. Given β ∈ R, a state ϕ on a C*-algebra A is a KMSβ-

state if ϕ is τ -invariant and ϕ(aτiβ(b)) = ϕ(ba) for a dense subset of analytic elements

a, b ∈ A. This de�nition corresponds to item (ii) in Proposition 5.1.29, so it is equivalent to

the de�nition introduced in Section 5.1 when β 6= 0.

Theorem 5.3.10. (Neshveyev) Let G be a locally compact second countable Hausdor� étale

groupoid. Let c be a continuous R-valued 1-cocycle on G and τ be the dynamics on C*(G)

de�ned by τt(f)(g) = eitc(g)f(g) for f ∈ Cc(G), g ∈ G. Fix β ∈ R. Then there exists a one-

151

Page 154: Characterization of Extremal KMS States on Groupoid C

to-one correspondence between KMSβ-states on C*(G) and pairs (µ, {ϕx}x∈G(0)) consisting

of a probability measure µ on G(0) and a µ-measurable �eld of states ϕx on C*(Gxx) such

that:

(i) µ is quasi-invariant with Radon-Nikodym derivative e−βc;

(ii) ϕx(ug) = ϕr(h)(uhgh−1) for every g ∈ Gxx and h ∈ Gx, for µ-a.e. x; in particular, ϕx is

tracial for µ-a.e. x;

(iii) ϕx(ug) = 0 for all g ∈ Gxx \ c−1(0), for µ-a.e. x.

Proof. From Lemma 5.3.5, the centralizer of any τ -KMS-state ϕ contains C0(G(0)). By

Theorem 5.2.9, ϕ is de�ned by a pair (µ, {ϕx}x) consisting of a probability measure µ on

G(0) and a µ-measurable �eld of states ϕx on C*(Gxx). It follows from Lemma 5.3.9 that

property (iii) is satis�ed if and only if ϕ is τ -invariant.

Therefore we have to prove that properties (i) and (ii) are satis�ed if, and only if,

ϕ(f1 · f2) = ϕ(f2 · τiβ(f1)), for every f1, f2 ∈ Cc(G). (5.37)

Each function in Cc(G) can be decomposed as a �nite sum of continuous functions sup-

ported on open bisections from Lemma 3.3.5. Every function in Cc(G) is entire analytic by

by Lemma 5.3.6. Note that τiβ is linear on Cc(G) by de�nition. Therefore (5.37) holds if,

and only if, for every open bisection U we have

ϕ(f1 · f2) = ϕ(f2 · τiβ(f1)), for every f1 ∈ Cc(U), f2 ∈ Cc(G). (5.38)

We will show (5.38) is equivalent to another equation and we will use this to prove the

equivalence between (5.38) and conditions (i) and (ii). Given an open bisection U , de�ne

hx = r|−1U (x) for every x ∈ r(U). Analogously, de�ne hx = s|−1

U (x) for x ∈ s(U). Then, for

152

Page 155: Characterization of Extremal KMS States on Groupoid C

every x ∈ r(U),

Tx = s|U ◦ r|−1U (x) = s(hx), and

hTx = s|−1U (Tx) = s|−1

U (s|U ◦ r|−1U (x)) = r|−1

U (x) = hx.

Suppose (5.37) holds. Let f1 ∈ Cc(U), f2 ∈ Cc(G). It follows that

ϕ(f1 · f2) =

∫G(0)

∑g∈Gxx

(f1 · f2)(g)ϕx(ug)dµ(x)

=

∫r(U)

∑g∈Gxx

f1(hx)f2((hx)−1g)ϕx(ug)dµ(x), from Lemma 5.3.7,

=

∫r(U)

f1(hx)∑g∈Gxx

f2((hx)−1g)ϕx(ug)dµ(x).

Since τiβ(f1) ∈ Cc(U), we can apply Lemma 5.3.7 and obtain

ϕ(f2 · τiβ(f1)) =

∫G(0)

∑g∈Gxx

(f2 · τiβ(f1))(g)ϕx(ug)dµ(x)

=

∫s(U)

∑g∈Gxx

τiβ(f1)(hx)f2(g(hx)−1)ϕx(ug)dµ(x), from Lemma 5.3.7,

=

∫s(U)

∑g∈Gxx

e−βc(hx)f1(hx)f2(g(hx)−1)ϕx(ug)dµ(x)

=

∫s(U)

e−βc(hx)f1(hx)∑g∈Gxx

f2(g(hx)−1)ϕx(ug)dµ(x).

Therefore (5.38) is equivalent to the following equation for f1 ∈ Cc(U), f2 ∈ Cc(G), U open

bisection.

∫r(U)

f1(hx)∑g∈Gxx

f2((hx)−1g)ϕx(ug)dµ(x) =

∫s(U)

e−βc(hx)f1(hx)∑g∈Gxx

f2(g(hx)−1)ϕx(ug)dµ(x).

(5.39)

153

Page 156: Characterization of Extremal KMS States on Groupoid C

Suppose (5.39) holds. Let f ∈ Cc(s(U)), then we can de�ne f1 ∈ Cc(U) such that

f |U = f ◦ s|U . Let f2 = f1*. Then f2 ∈ Cc(U−1) by Lemma 3.3.8.

Let x ∈ r(U), g ∈ Gxx such that f2((hx)−1g) 6= 0. Then (hx)−1g ∈ U−1 which implies

g−1hx ∈ U . Recall that U is an open bisection and hx ∈ U . Moreover, s(g−1hx) = s(hx),

then g−1hx = hx, hence g = x. Therefore, for all x ∈ r(U), we have

∑g∈Gxx

f2((hx)−1g)ϕx(ug) = f2((hx)−1x)ϕx(ux) = f2((hx)−1).

Therefore, for f1, f2, we can rewrite (5.39) as

∫s(U)

e−βc(hx)f1(hx)f2((hx)−1)dµ(x) =

∫r(U)

f1(hx)f2((hx)−1)dµ(x)

=

∫r(U)

f1(hTx)f2((hTx)−1)dµ(x), (5.40)

because hTx = hx for x ∈ r(U). Using the de�nition of f1 and f2, we have for every x ∈ s(U),

f1(hx) = f ◦ s|U(hx) = f ◦ s|U(s|−1U (x)) = f(x),

f2((hx)−1) = f1*((hx)

−1) = f1(hx) = f(x).

Then we can replace the values in the integrals in (5.40) and obtain

∫s(U)

e−βc(hx)|f(x)|2dµ(x) =

∫r(U)

|f(Tx)|2dµ(x).

Since f is arbitrary, it follows that for every f ∈ Cc(s(U)),

∫s(U)

e−βc(hx)f(x)dµ(x) =

∫r(U)

f(Tx)dµ(x).

HencedT∗µ

dµ(x) = e−βc(hx). It follows from Lemma 5.3.8 that property (i) holds.

Now we show that property (ii) is satis�ed. Let U be an open bisection and f ∈ Cc(r(U)),

154

Page 157: Characterization of Extremal KMS States on Groupoid C

de�ne f1 ∈ Cc(U) such that f1|U = f ◦r|U . Then f1(hx) = f(x). Given f2 ∈ Cc(G) arbitrary,

de�ne the function F : G(0) → C by

F (x) =∑g∈Gxx

(f2 · f1)(g)ϕx(ug).

This function is µ-measurable because f2 · f1 ∈ Cc(G) and {ϕx}x is a µ-measurable �eld of

states. Moreover, by Lemma 5.3.7,

F (x) =

f1(hx)∑

g∈Gxxf2(g(hx)

−1)ϕx(ug), if x ∈ s(U),

0, otherwise.

Then

∫r(U)

f(x)∑g∈GTxTx

f2(g(hx)−1)ϕTx(ug)dµ(x) =

∫r(U)

f1(hx)∑g∈GTxTx

f2(g(hx)−1)ϕTx(ug)dµ(x)

=

∫r(U)

f1(hTx)∑g∈GTxTx

f2(g(hTx)−1)ϕTx(ug)dµ(x)

=

∫r(U)

F (Tx)dµ(x)

=

∫s(U)

F (x)dT∗µ(x)

=

∫s(U)

e−βc(hx)F (x)dµ(x),

by property (i) that we already proved. Then,

=

∫s(U)

e−βc(hx)f1(hx)∑g∈Gxx

f2(g(hx)−1)ϕx(ug)dµ(x),

=

∫r(U)

f1(hx)∑g∈Gxx

f2((hx)−1g)ϕx(ug)dµ(x),

155

Page 158: Characterization of Extremal KMS States on Groupoid C

by (5.39). Then,

=

∫r(U)

f(x)∑g∈Gxx

f2((hx)−1g)ϕx(ug)dµ(x),

if we make the change of variables g = hxg(hx)−1, we get

=

∫r(U)

f(x)∑g∈GTxTx

f2(g(hx)−1)ϕx(u(hx)−1ghx)dµ(x).

Assume U has compact closure. Let V be an open bisection with compact closure such

that s(V) ⊂ s(U). We can assume s(V) = s(U) without loss of generality. In fact, if this is

not the case, just replace U by s|−1U ◦ s(V). VU−1 is an open bisection from Lemmas 3.3.6

and 3.3.7.

Assume f2 ∈ Cc(G) is such that f2 is positive on VU−1 and vanishes outside this set.

Since f ∈ Cc(r(U)) is arbitrary, we have for µ-a.e. x,

∑g∈GTxTx

f2(g(hx)−1)[ϕx(u(hx)−1ghx)− ϕTx(ug)] = 0. (5.41)

Let x ∈ r(U) be such that (5.41) holds. Let g ∈ V ∩ GTx. Then g(hx)−1 is the unique

element in VU−1. Thus we can write (5.41) as

ϕx(u(hx)−1ghx) = ϕTx(ug) for g ∈ GTxTx ∩ V , µ-a.e. x ∈ r(U). (5.42)

Since the preimage of s(U) under s can be covered by a countable family of open bisections

V with compact closure such that s(V) ⊂ s(U), it follows that for µ-a.e. x ∈ r(U), g ∈ GTxTx,

we have

ϕx(u(hx)−1ghx) = ϕTx(ug) = ϕs(hx)(ug).

156

Page 159: Characterization of Extremal KMS States on Groupoid C

Note that the set

{x ∈ G(0) : ϕx(ug) 6= ϕr(h)(uhgh−1) for some g ∈ Gxx, h ∈ Gx}

={x ∈ G(0) : ϕx(ug) 6= ϕr(h−1)(uh−1gh) for some g ∈ Gxx, h ∈ Gx}

={x ∈ G(0) : ϕx(ug) 6= ϕs(h)(uh−1gh) for some g ∈ Gxx, h ∈ Gx}

=⋃U

{x ∈ G(0) : ϕx(ug) 6= ϕs(h)(uh−1gh) for some g ∈ Gxx, h ∈ Gx ∩ U}

=⋃U

{x ∈ r(U) : ϕx(ug) 6= ϕs(h)(uh−1gh) for some g ∈ Gxx, h ∈ Gx ∩ U}

=⋃U

{x ∈ r(U) : ϕx(ug) 6= ϕs(hx)(u(hx)−1ghx) for some g ∈ Gxx}

=⋃U

{x ∈ r(U) : ϕx(ug) 6= ϕTx(u(hx)−1ghx) for some g ∈ Gxx}

has measure zero. Here U ranges over a countable open cover of G such that r, s are injective

on U . Therefore property (ii) holds.

Conversely, assume properties (i), and (ii) are satis�ed. Given an open bisection U , let

f1 ∈ Cc(U), f2 ∈ Cc(G). Then

∫r(U)

f1(hx)∑g∈Gxx

f2((hx)−1g)ϕx(ug)dµ(x)

=

∫r(U)

f1(hx)∑g∈GTxTx

f2(g(hx)−1)ϕx(uhxg(hx)−1)dµ(x) , making g = (hx)−1ghx,

=

∫r(U)

f1(hx)∑g∈GTxTx

f2(g(hx)−1)ϕr(hx)(uhxg(hx)−1)dµ(x)

=

∫r(U)

f1(hTx)∑g∈GTxTx

f2(g(hTx)−1)ϕr(hTx)(uhTxg(hTx)−1)dµ(x)

=

∫r(U)

f1(hTx)∑g∈GTxTx

f2(g(hTx)−1)ϕTx(ug)dµ(x), from property (ii),

=

∫s(U)

e−βc(hx)f1(hx)∑g∈Gxx

f2(g(hx)−1)ϕx(ug)dµ(x), from (i) and Lemma 5.3.8.

157

Page 160: Characterization of Extremal KMS States on Groupoid C

Then (5.39) holds. However, we already proved this is equivalent to equation (5.37).

158

Page 161: Characterization of Extremal KMS States on Groupoid C

Chapter 6

Renault-Deaconu Groupoid

In this chapter we prove a theorem due to Thomsen [26] which characterizes the extremal

KMS states on the full C*-algebra of the Renault-Deaconu groupoid. The de�nition of this

groupoid depends on a local homeomorphism σ : X → X such that X has some topological

properties. We can identify the subset of units with X.

We can apply Neshveyev's theorems to this groupoid C*-algebra in order to describe its

KMS states. In this chapter we show that, on this groupoid, quasi-invariant measures are

the same as conformal measures. Moreover, the corresponding measure of an extremal KMS

state is either continuous or supported on an orbit.

Under certain conditions, Thomsen's theorem gives an explicit formula for the extremal

KMS states. The results in this chapter are based on [9] and [26].

6.1 Introduction

Now we de�ne the Renault-Deaconu groupoid and prove some of its topological properties.

De�nition 6.1.1. Let X be a locally compact second countable Hausdor� space. Let σ :

159

Page 162: Characterization of Extremal KMS States on Groupoid C

X → X be a local homeomorphism. The Renault-Deaconu groupoid is the groupoid

G = {(x, k, y) ∈ X × Z×X : ∃n,m ∈ N, k = n−m,σn(x) = σm(y)},

such that

G(2) = {((x1, k1, y1), (x2, k2, y2)) ∈ G × G : y1 = x2},

with the following multiplication and inversion laws

(x, k1, y)(y, k2, z) = (x, k1 + k2, z) and (x, k, y)−1 = (y,−k, x),

and unit space de�ned by G(0) = {(x, 0, x) : x ∈ X}.

The range and source maps are de�ned by r(x, k, y) = (x, 0, x) and s(x, k, y) = (y, 0, y).

Since the map x 7→ (x, 0, x) is a bijection from X to G(0), we identify X with G(0).

Remark 6.1.2. We assume, by convention, that σ0 is the identity.

The groupoid G can be understood intuitively as follows: given x ∈ X, we can interpret

the sequence {σn(x)}n∈N as a trajectory starting at x, as shown in Figure 6.1.

x σ(x) σ1(x) σ2(x) σ3(x) . . .

Figure 6.1: The sequence {σn(x)}n∈N, x ∈ X, can be interpreted as the trajectory of x.

If, for some y ∈ X, there is some n such that σn(y) is an element of the trajectory of

x, we can say the trajectories eventually meet. In other words, there exists m such that

σm(x) = σn(y). k = m − n is the delay of one trajectory with respect to the other. Hence

160

Page 163: Characterization of Extremal KMS States on Groupoid C

(x, k, y) ∈ G if the trajectories of x and y eventually meet. This idea is shown in Figure 6.2.

y

x · σ(y) ·

σ(x) σ2(y)

σ2(x) = σ3(y)

Figure 6.2: If (x, k, y) ∈ G then the trajectories {σl(x)}l∈N and {σl(y)}l∈N eventually meet.k can be seen as the delay of one trajectory with respect to the other. In this �gure, k = −1,since σ2(x) = σ3(y).

Proposition 6.1.3. The Renault-Deaconu groupoid is a groupoid.

Proof. Let G be the Renault-Deaconu groupoid. Clearly the maps r and s are surjective.

The product is well-de�ned: In fact, let g, h ∈ G be composable. Then g = (x, k, y),

h = (y, l, z) for some x, y, z ∈ X, k, l ∈ Z.

By de�nition of G, there exist m,n ∈ N such that k = m− n and σm(x) = σn(y). There

are p, q ∈ N satisfying l = p− q and σp(y) = σq(z). Hence,

σm+p(x) = σp(σm(x)) = σp(σn(y)) = σp+n(y) = σn(σp(y)) = σn(σq(z)) = σn+q(z).

The inverse is well-de�ned: Given (x, k, y) ∈ G, there exist m,n ∈ N such that

k = m−n and σm(x) = σn(y). Then −k = n−m and σn(y) = σm(x). Hence (y,−k, x) ∈ G.

Now we show G satis�es properties (i)-(v) of De�nition 3.1.1:

(i) Let g = (x, k, y) ∈ G, h = (y, l, z) ∈ G, then

s(gh) = s(x, k + l, z)= z = s(h)

r(gh) = r(x, k + l, z)= x = r(g).

(ii) Given x ∈ X, r(x, 0, x) = (x, 0, x) = s(x, 0, x).

161

Page 164: Characterization of Extremal KMS States on Groupoid C

(iii) Given g = (x, k, y) ∈ G,

gs(g) = (x, k, y)(y, 0, y)= (x, k, y) = g

r(g)g = (x, 0, x)(x, k, y)= (x, k, y) = g.

(iv) Let (g1, g2), (g2, g3) ∈ G(2). Then g1 = (x, k1, y), g2 = (y, k2, z), g3 = (z, k3, w). Hence,

(g1g2)g3 = (x, k1 + k2, z)(z, k3, w) = (x, k1 + k2 + k3, w)

g1(g2g3) = (x, k1, y)(y, k2 + k3, w)= (x, k1 + k2 + k3, w).

(v) Given g = (x, k, y) ∈ G,

gg−1 = (x, k, y)(y,−k, x)= (x, 0, x) = r(g)

g−1g = (y,−k, x)(x, k, y)= (y, 0, y) = s(g).

Example 6.1.4. Let X = {x = {xn}n∈N : xn ∈ {0, 1}}. Endow X with the metric

d(x, y) = 2−min {n∈N : xn 6=yn}.

Moreover, σ is a local homeomorphism. Indeed, let x ∈ X. De�ne the map ρ : X → X by

ρ(y)n =

x0, if n = 0

yn−1 if n ≥ 1.

Then ρ is continuous, since for y, z ∈ X,

d(σ(y), σ(z)) = 2−min{n≥1:yn−1 6=zn−1} = 2−1d(y, z)

162

Page 165: Characterization of Extremal KMS States on Groupoid C

Moreover, for every y ∈ B(x, 1), y0 = x0. Hence y = ρ ◦ σ(y) = σ ◦ ρ(y). Therefore σ is

invertible on B(x, 1) with inverse ρ. That is, ρ is a local homeomorphism.

Then (x, k, y) ∈ G if there are n,m ∈ N such that k = n −m and σn(x) = σm(y). For

instance, if x = (0, 1, 0, 0, . . .) and y = (1, 1, 1, 1, 0, 0, . . .) then (x,−2, y) ∈ G, as shown in

Figure 6.3.

x = (0, 1, 0, 0, . . .)

y = (1, 1, 1, 1, 0, 0, . . .)

σ(x)

σ3(y)

Figure 6.3: It follows from the equality σ(x) = σ3(y) = (1, 0, 0, . . .) that (x,−2, y) ∈ G.

Now we de�ne a topology on G which makes the Renault-Deaconu groupoid an étale

groupoid.

Given A,B open subsets of X, m,n ∈ N, let

Un,mA,B = {(x, n−m, y) ∈ G : σn(x) = σm(y), x ∈ A, y ∈ B}.

These sets form a basis of the topology on G.

Proposition 6.1.5. The family of sets Un,mA,B , for A,B open subsets of X and n,m ∈ N,

generates a topology on G. Moreover, G is second countable.

Proof. First we show that G is the union of these sets. Let (x, k, y) ∈ G. Then there exist

n,m such that σn(x) = σm(y). Hence (x, n−m, y) ∈ Un,mX,X .

Now we prove that for every (x, k, y) ∈ Un1,m1

A1,B1∩ Un2,m2

A2,B2there exists Un,mA,B such that

(x, k, y) ∈ Un,mA,B ⊂ Un1,m1

A1,B1∩ Un2,m2

A2,B2.

Let (x, k, y) ∈ Un1,m1

A1,B1∩ Un2,m2

A2,B2. Then,

x ∈ A1 ∩ A2 , y ∈ B1 ∩B2 , σn1(x) = σm1(y) , σn2(x) = σm2(y).

163

Page 166: Characterization of Extremal KMS States on Groupoid C

Let p1 = n2, p2 = n1. Note that m1 − n1 = m2 − n2 implies m1 + p1 = m2 + p2. From the

de�nition of p1, p2, we have n1 + p1 = n2 + p2.

For i = 1, 2, let Ui be an open neighborhood of σni(x) such that σpi is injective on Ui. Let

n = n1 + p1 = n2 + p2

m = m1 + p1 = m2 + p2

A = A1 ∩ A2 ∩ σ−n1(U1) ∩ σ−n2(U2)

B = B1 ∩B2 ∩ σ−m1(U1) ∩ σ−m2(U2).

From continuity of σni , σmi , it follows that A and B are open sets.

We show that (x, k, y) ∈ Un,mA,B . Clearly n−m = n1 +p1− (m1 +p1) = n1−m1 = k. Then,

σn(x) = σp1+n1(x) = σp1(σn1(x)) = σp1(σm1(y)) = σp1+m1(y) = σm(y).

For i = 1, 2, σni(x) ∈ Ui, σmi(y) = σni(x) ∈ Ui. Then x ∈ A, y ∈ B. Therefore (x, k, y) ∈

Un,mA,B .

Now we show that Un,mA,B ⊂ Un1,m1

A1,B1∩ Un2,m2

A2,B2. Let (u, k, v) ∈ Un,mA,B .

For i = 1, 2, we have

σn(u) = σm(v)

σni+pi(u) = σmi+pi(v)

σpi(σni(u)) = σpi(σmi(u)).

Since σni(u), σmi(v) ∈ U and σpi is injective on this set, it follows that σni(u) = σmi(v).

Hence (u, k, v) ∈ Uni,miAi,Bi. Therefore (u, k, v) ∈ Un1,m1

A1,B1∩ Un2,m2

A2,B2.

Now we show that G is second countable.

164

Page 167: Characterization of Extremal KMS States on Groupoid C

Let B be a countable base of X. Then the family of sets of the form

Un,mA,B such that n,m ∈ N and A,B ∈ B

is also countable. We show this family form a base for G. Let Un,mA,B with A,B arbitrary open

sets in X and n,m ∈ N. Let (x, n−m, y) ∈ Un,mA,B . There exists A, B such that x ∈ A ⊂ A,

y ∈ B ⊂ B. Then (x, n−m, y) ∈ Un,mA,B⊂ Un,mA,B .

Lemma 6.1.6. Let {(xi, ki, yi)}i∈N be a sequence in G converging to (x, k, y). Then xi → x,

yi → y and there exists i0 such that ki = k for every i ≥ i0. Hence we can assume, without

loss of generality, that ki is constant.

Proof. Let n,m ∈ N such that σn(x) = σm(y) and k = n −m. Let A,B be neighborhoods

of x, y, respectively. Then there exists i0 such that for every i ≥ i0, (xi, ki, yi) ∈ Un,mA,B , then,

x ∈ A, y ∈ B, ki = n−m = k.

Therefore ki is eventually constant, xi → x and yi → y.

Lemma 6.1.7. Fix n0,m0 ∈ N. Given a sequence {(xi, k, yi)}i∈N a net assume that for all

neighborhoods A of x and B of y there exists i0 such that (xi, k, yi) ∈ Un0,m0

A,B for i ≥ i0.

Then for every A,B open neighborhoods of x, y, respectively, n,m such that (x, k, y) ∈

Un,mA,B . There exists i0 such that

(xi, k, yi) ∈ Un,mA,B for i ≥ i0.

Then (xi, k, yi)→ (x, k, y).

Proof. Note that k = n0−m0. Let n,m ∈ N such that σn(x) = σm(y) and k = n−m. Then

n0 +m = n+m0.

Let V be an open neighborhood of σn(x) = σm(y) where σn0 and is injective. Since xi → x

and yi → y and σ is continuous, there exists i0 such that for every i ≥ i0, σn(xi), σ

m(yi) ∈ V .

165

Page 168: Characterization of Extremal KMS States on Groupoid C

Then, given i ≥ i0,

σn+n0(xi) = σn(σn0(xi)) = σn(σm0(yi)) = σn+m0(yi) = σn0+m(yi).

Then σn0(σn(xi)) = σn0(σm(yi)). Since σn(xi), σm(yi) ∈ V and σn0 is injective on V , it

follows that σn(xi) = σm(yi) for every i ≥ i0.

Note that (x, k, y) /∈ Un,mA,B if n−m 6= k.

Let A,B be open neighborhoods of x, y respectively. Let m,n such that m−n = k. Since

xi → x and yi → y, there exists i0 such that if i ≥ i0,

xi ∈ A, yi ∈ B, σn(xi) = σm(yi),

or equivalently, (xi, k, yi) ∈ Un,mA,B for every i ≥ i0.

Corollary 6.1.8. Let (x, n −m, y) ∈ G. Let {xi}i∈N, {yi}i∈N be sequences in X such that

xi → x and yi → y. If σn(xi) = σm(yi) for each i, then (xi, n−m, yi)→ (x, n−m, y) in G.

Proof. Let A be an open neighborhood of x, B an open neighborhood of y. There exists

i0 such that, for every i ≥ i0, xi ∈ A, yi ∈ B. By hypothesis, σn(xi) = σm(yi). Then

(xi, n−m, yi) ∈ Un,mA,B . From Lemma 6.1.7, it follows that (xi, n−m, yi) → (x, n−m, y) in

G.

Theorem 6.1.9. The Renault-Deaconu groupoid G, with topology generated by Un,mA,B is a

topological groupoid, locally compact Hausdor�, second countable and étale.

Proof. • G is a topological groupoid

(i) G(2) is closed in G × G.

Assume {(gi, hi)}i∈N is a sequence in G(2) converging to (g, h) ∈ G × G. Then

gi = (xi, ki, yi), hi = (yi, li, zi) for each i. Assume g = (x, k, y) and h = (y, l, z).

It follows from Lemma 6.1.6 that yi → y and yi → y in X. Since X is Hausdor�,

we have y = y. Therefore (g, h) ∈ G(2).

166

Page 169: Characterization of Extremal KMS States on Groupoid C

(ii) The inverse is continuous

Denote the inverse map by ι. Let Um,nB,A be an open set in the topological base of

G. Then,

ι−1(Um,nB,A) = {(x, k, y) : (y,−k, x) ∈ Um,nB,A}

= {(x, k, y) : y ∈ B, x ∈ A, σm(y) = σn(x), k = n−m}

= Un,mA,B .

Therefore the inverse map is continuous.

(iii) The product is continuous

Let (gi, hi)→ (g, h) in G(2). We can assume g = (x, k, y), h = (y, l, z) and for each

i, gi = (xi, k, yi) and hi = (yi, l, zi).

Let n1,m1, n2,m2 ∈ N such that

k = n1 −m1 , σn1(x) = σm1(y)

l = n2 −m2 , σn2(y) = σm2(z).

Since (xi, k, yi) → (x, k, y) and (yi, l, zi) → (y, l, z), it follows that for A, B

neighborhoods of x and z, there exists i0 such that

(xi, k, yi) ∈ Un1,m1

A,X and (yi, k, zi) ∈ Un1,m1

X,B for i ≥ i0.

Hence, if i ≥ i0,

σn1+n2(xi) = σn2(σn1(xi)) = σn2(σm1(yi)) = σn2+m1(yi)

= σm1(σn2(yi))= σm1(σm2(zi)) = σm1+m2(zi).

167

Page 170: Characterization of Extremal KMS States on Groupoid C

Then (xi, k + l, zi) ∈ Un1+n2,m1+m2

A,B . It follows from Lemma 6.1.7 that

(xi, k + l, zi)→ (x, k + l, z).

• G is Hausdor�

Let gi = (xi, ni −mi, yi) ∈ G such that σni(xi) = σmi(yi), i = 1, 2. Assume g1 6= g2.

(i) If n1 −m1 6= n2 −m2,

then g1 ∈ Un1,m1

X,X , g2 ∈ Un2,m2

X,X and Un1,m1

X,X ∩ Un2,m2

X,X = ∅.

(ii) If n1 −m1 = n2 −m2,

Then x1 6= x2 or y1 6= y2. Assume x1 6= x2. Since X is Hausdor�, we can choose

A1, A2 open neighborhoods of x1, x2 respectively, such that A1 ∩ A2 = ∅. Then

gi ∈ Uni,miAi,X, i = 1, 2, and Un1,m1

A1,X∩ Un2,m2

A2,X= ∅.

The proof for y1 6= y2 is analogous.

• G is locally compact.

Let A,B be open sets of X such that A,B are compact, and let n,m ∈ N. Then

Un,mA,B ⊂ Un,m

A,B, where

UA,B = {(x, n−m, y) ∈ G : σn(x) = σm(y), x ∈ A, y ∈ B}.

Let {(xi, n−m, yi)}i∈N be a sequence in UA,B. Then {(xi, yi)}i∈N is a sequence in the

compact set A×B. Then there exists a subsequence {(xij , yij)}j∈N such that xij → x

for some x ∈ A and yij → y for some y ∈ B. By continuity of σ, σn(x) = σm(y).

Therefore UA,B is compact.

• G is étale.

Let (x, n − m, y) ∈ G such that σn(x) = σm(y). Since σ is a local homeomorphism,

168

Page 171: Characterization of Extremal KMS States on Groupoid C

there are A,B open neighborhoods of x, y, respectively, satisfying

σn(A) is open and σn|A : A→ σn(A) is a homeomorphism

σm(B) is open and σm|B : B → σm(B) is a homeomorphism.

Then (x, n−m, y) ∈ Un,mA,B . Since G is a topological groupoid, r is continuous. In order

to prove r is a local homeomorphism, we will show r is injective on Un,mA,B , r(Un,mA,B ) is

open and r|−1Un,mA,B

is continuous.

(i) r is injective.

Suppose there exist x1, x2 ∈ A, y1, y2 ∈ B such that r(x1, n −m, y1) = r(x2, n −

m, y2). Then y1 = y2. Moreover,

x1 = σ−n|A(σm(y1)) = σ−n|A(σm(y2)) = x2.

Therefore (x1, n−m, y1) = (x2, n−m, y2).

(ii) r(Un,mA,B ) is open.

r(Un,mA,B ) = {(y, 0, y) ∈ G : (x, n−m, y) ∈ Un,mA,B}

= {(y, 0, y) ∈ G : x ∈ A, y ∈ B, σn(x) = σm(y)}

= {(y, 0, y) ∈ G : y ∈ B, x = σ|−nA (σm(y)) ∈ A}

= {(y, 0, y) ∈ G : y ∈ B, σm(y) ∈ σn(A)} note that σn(A) is open

= {(y, 0, y) ∈ G : y ∈ B, y ∈ σ|−mB (σn(A))}

= U0,0C,C ,

where C = B ∩ σ|−mB (σn(A)).

(iii) r|−1Un,mA,B

is continuous.

Let {(yi, 0, yi)}i∈N be a sequence in r(Un,mA,B ) converging to some (y, 0, y) in r(Un,mA,B ).

Then yi → y ∈ B. De�ne the sequence xi = σ|−nA (σm(yi)). Then xi → x =

169

Page 172: Characterization of Extremal KMS States on Groupoid C

σ|−nA (σm(y)).

Note that xi is the only element in A satisfying σn(xi) = σm(yi), hence,

(xi, n−m, yi) = r|−1Un,mA,B

(yi, 0, yi).

Analogously, (x, n−m, y) = r|−1Un,mA,B

(y, 0, y). Then, it follows from Corollary 6.1.8

that (xi, n−m, yi)→ (x, n−m, y).

Analogously, we can show s is a local homeomorphism.

The next lemma shows that we can identify X wth the unit space of G. In this chapter,

we �x X and σ, and we assume G is the Renault-Deaconu groupoid.

Lemma 6.1.10. We can identify the unit space G(0) with the set X. In fact, both have the

same topology.

Proof. Clearly the map ι : G(0) → X de�ned by (x, 0, x)→ x is a bijection. Denote this map

by ι. We will show ι is a homeomorphism.

Let n,m ∈ N, A,B ⊂ X open sets. If Un,mA,B ∩ G(0) 6= ∅, then n = m. In this case,

Un,mA,B ⊂ G(0). Hence,

ι(Un,mA,B ) = ι({(x, 0, x) : x ∈ A ∩B}) = A ∩B

is open in X. On the other hand, for any open set A ⊂ X,

ι−1(A) = {(x, 0, x) : x ∈ A} = U0,0A,A,

Therefore ι is a homeomorphism.

170

Page 173: Characterization of Extremal KMS States on Groupoid C

6.2 Full orbits

Given x ∈ X, the full orbit of x denotes the set of elements inX whose trajectories eventually

meet the trajectory of x. There are two types of orbits, periodic and aperiodic. In this section

we will study their properties.

De�nition 6.2.1. Let x ∈ X, the full orbit of x is the set

O(x) = {y ∈ X : there exists k ∈ Z such that (x, k, y) ∈ G}.

Lemma 6.2.2. Given x, y ∈ X, O(x) = O(y) if, and only if, y ∈ O(x).

Proof. Let y ∈ O(x). There exists (x, k, y) ∈ G.

• O(y) ⊂ O(y) Let z ∈ O(y). There exists (y, l, z) ∈ G. Hence (x, k + l, z) ∈ G. Then

z ∈ O(x). Therefore O(y) ⊂ O(x).

• O(x) ⊂ O(y) Let z ∈ O(x). There exists (x, l, z) ∈ G. Then (z, k − l, y) ∈ G. Thus

z ∈ O(y). Then O(x) ⊂ O(y).

Therefore O(x) = O(y).

Conversely, assume O(x) = O(y). By de�nition, (y, 0, y) ∈ G. Then y ∈ O(y) = O(x).

Remark 6.2.3. Lemma 6.2.2 is a fact of general groupoids. Given a groupoid G, if we de�ne

the set O(x) = r(Gx) for every x ∈ G(0), then the Lemma 6.2.2 holds.

Lemma 6.2.4. Given x ∈ X, O(x) is countable.

Proof.

O(x) = {y ∈ X : there exists k ∈ Z such that (x, k, y) ∈ G}

= {y ∈ X : Gyx 6= ∅}

171

Page 174: Characterization of Extremal KMS States on Groupoid C

= {r(g) : g ∈ Gx 6= ∅} identifying X with G(0)

= r(Gx).

G is second countable étale from Theorem 6.1.9. It follows from Proposition 3.2.11 that Gxis countable. Therefore O(x) is countable.

There are two types of full orbits: periodic and aperiodic orbits. This di�erence will be

fundamental when we de�ne the extremal conformal measures later.

De�nition 6.2.5. Given x ∈ X, we say it is periodic or σ-periodic if there is a positive

integer p such that

σp(x) = x. (6.1)

The minimum positive natural number such that (6.1) holds is called the minimal period of

x.

De�nition 6.2.6. A point z ∈ X is called aperiodic if O(z) does not contain periodic points.

De�nition 6.2.7. Let y ∈ X. IfO(y) has periodic points, O(y) is called periodic. Otherwise,

O(y) is aperiodic.

When represented graphically, periodic and aperiodic orbits look di�erent. An aperiodic

orbits look like a tree while a periodic orbits has a single cycle, the trajectory of a periodic

172

Page 175: Characterization of Extremal KMS States on Groupoid C

element. Figure 6.4 shows these two types of orbits.

x

(a)

z

. . .

(b)

Figure 6.4: Periodic and aperiodic orbits look di�erent. The periodic orbit (a) has a cycle,while the aperiodic orbit looks like a tree. Circles in green represent the trajectories of xand z.

Given y in the orbit of x, the trajectory of y eventually meets the trajectory of x. If x is

periodic, we can see in Figure 6.4 that y eventually meets the point x, that is, there exists

an n such that σn(y) = x. Now we will prove this result.

Lemma 6.2.8. Let x ∈ X be a periodic point with minimum period p. Given y ∈ O(x),

there exists n ∈ N such that σn(y) = x.

Proof. Let y ∈ O(x). Let n1,m1 ∈ N such that σn1(x) = σm1(x). Let N be a natural number

such that N +m1 ∈ pN. De�ne n = N + n1. Then

σn(y) = σN+n1(y) = σN+m1(x) = x.

Lemma 6.2.9. Assume the set of periodic points in X is countable. Let I denote the set of

aperiodic points. Then I is countable

Proof. Let N be the set of points with periodic orbits. Then I = X \N and we can write

N =⋃x∈X

x periodic

O(x).

173

Page 176: Characterization of Extremal KMS States on Groupoid C

O(x) is countable for every x ∈ X by Lemma 6.2.4. Moreover, the set of periodic points is

countable by hypothesis. Then N is countable and µ(N) = 0.

Lemma 6.2.10. Let z ∈ X be aperiodic. Then Gzz = {(z, 0, z)}. If y ∈ X is such that O(y)

has a periodic point with minimum period p, then Gyy = {(y, kp, y) : k ∈ Z}.

Proof. Let z ∈ X be aperiodic. Let k ∈ Z such that (z, k, z) ∈ G. Then there exists n,m ∈ N

such that σn(z) = σm(z), k = n−m.

Assume without loss of generality that n > m. Denote x = σm(z) ∈ O(z). Then x is

periodic, since

σn−m(x) = σn−m(σm(z)) = σn(z) = σm(z) = x.

Contradiction. Then k = 0 and therefore Gzz = {(z, 0, z)}.

Let y ∈ X such that O(y) is periodic. Then there exist x ∈ X periodic with minimum

period p, and l ∈ Z such that (x, l, y) ∈ G. Then (y,−l, x)(x, p, x)(x, l, p) = (y, p, y) ∈ G.

Therefore, by induction, (y, kp, y) ∈ G for every k ∈ Z.

Suppose there exists k /∈ pZ such that (y, k, y) ∈ Gyy . Let k1, k2 ∈ N such that k = k1+k2p,

0 ≤ k1 < p. Note that k1 6= 0 by hypothesis.

Then (x, k, x) ∈ G since (x, k, x) = (y,−l, x)(y, k, y)(y, l, x). Hence there exist n,m ∈ N

such that n−m = k and σn(x) = σm(x).

Assume n > m without loss of generality. Let N be an integer such that N + m ∈ pN.

Then,

x = σm+N(x) = σn+N(x) = σn−m(σm+N(x)) = σn−m(x)

= σk(x) = σk1+k2p(x) = σk1(σk2p(x)) = σk1(x).

Contradiction, since x has minimum period p. Therefore Gyy = {(y, kp, y) : k ∈ Z}.

174

Page 177: Characterization of Extremal KMS States on Groupoid C

6.3 Conformal Measures

We will show explicitly all the extremal atomic eβF -conformal probability measures on G.

for β 6= 0.

Now we de�ne conformal measures as described in [8].

De�nition 6.3.1. Consider a measurable function T : X → X on a measurable space

(X,F) and a measurable nonnegative function f on X. A measure µ on (X,F) is called

f -conformal if

µ(T (A)) =

∫A

f(x)dµ(x),

whenever A ⊂ X is a measurable set, for which T (A) is measurable and T : A → T (A) is

invertible.

A set A as in De�nition 6.3.1 is called special .

The set of f -conformal probability measures µ forms a convex set. We say µ is extremal

if µ is an extremal point in this set.

De�nition 6.3.2. Let µ be a f -conformal probability measure. We say that µ is extremal

if for all µ1, µ2, t ∈ (0, 1) such that µ = tµ1 + (1− t)µ2, it follows that µ1 = µ2 = µ.

Lemma 6.3.3. Let µ be a �nite measure on the topological space X with µ = µa + µc, µa

purely atomic and µc non-atomic. Then µ is f -conformal if and only if µa, µc are f -conformal.

Proof. Since µ = µa + µc, both µa, µc are �nite measures.

Assume µ is f -conformal. LetXa be the Borel set such that µa(Xa) = µa(X) and µ(x) > 0

for every x ∈ Xa. Since µa is �nite, then Xa is countable and µc(Xa) = 0.

Let A be a special set. Since Xa is countable, T (A∩Xa) is also countable, then T (A∩Xa)

is measurable. Note that T (A) = T (A∩Xa)∪ T (A \Xa) and T (A∩Xa)∩ T (A \Xa) = ∅.

Hence T (A \Xa) is measurable.

µc(T (A ∩Xa)) = 0 since T (A ∩Xa) is countable. Now we show µa(T (A \Xa)) = 0. Let

y ∈ T (A \ Xa). There exists a ∈ A \ Xa such that y = T (a). Using that µ is f -conformal

175

Page 178: Characterization of Extremal KMS States on Groupoid C

and a /∈ Xa, we have

µa(T (a)) = µ(T (a)) = f(a)µ(a) = f(a)µa(a) = 0.

Since y is arbitrary and µa is atomic, we have µa(T (A \Xa)) = 0. Then

µa(T (A)) = µa(T (A ∩Xa)) + µa(T (A \Xa))

= µa(T (A ∩Xa))

= µa(T (A ∩Xa)) + µc(T (A ∩Xa))

= µ(T (A ∩Xa))

=

∫A∩Xa

f(x)dµ(x)

=

∫A

f(x)dµa(x).

Since µ = µa on Xa and µa = 0 outside Xa, then µa is f -conformal. The proof for µc is

analogous.

Conversely, assume µa, µc are f -conformal. Let A be a Borel set such that T : A→ T (A)

is invertible. Then

µ(T (A)) = µa(T (A)) + µc(T (A))

=

∫A

f(x)dµa(x) +

∫A

f(x)dµc(x)

=

∫A

f(x)dµ(x).

Then µ is f -conformal.

Remark 6.3.4. Fix a continuous function F : X → R and assume β 6= 0. It follows

from Lemma 6.3.3 that every extremal eβF -conformal measure is either purely atomic or

non-atomic. In fact, let µ be an extremal eβF -conformal probability measure and assume is

neither purely atomic nor non-atomic. By Lemma 6.3.3, µ = µa + µc, µa is purely atomic

176

Page 179: Characterization of Extremal KMS States on Groupoid C

and µc is non-atomic.

Let t = µa(X). Then t > 0 (otherwise µ = µc) and t < 1 (otherwise µ = µa). De�ne

µ1 = t−1µa, µ2 = (1− t)−1µc. Note that µ1, µ2 are probability measures. In fact,

µ1(X) = t−1µa(X) = t−1t = 1

and

µ2(X) = (1− t)−1µc(X)

= (1− t)−1(µ(X)− µa(X)) , since µ = µa + µc,

= (1− t)−1(1− t) = 1.

Then µ1, µ2 are probability measures. Now let A be a special set. Then

µ1(T (A)) = t−1µa(T (A)) = t−1

∫A

eβF (x)dµa(x) =

∫A

eβF (x)dµ1(x)

and

µ2(T (A)) = (1− t)−1µc(T (A)) = (1− t)−1

∫A

eβF (x)dµc(x) =

∫A

eβF (x)dµ2(x).

Therefore µ1, µ2 are eβF -conformal.

In this section we want to �nd all extremal eβF -conformal purely atomic probability

measures on X. Given µ eβF -conformal, then µ(σ(x)) = eβF (x)µ(x).

Lemma 6.3.5. Let β ∈ R and µ an eβF -conformal measure on X, n a positive natural

number. Given y ∈ X, we have

µ(σn(y)) = exp

n−1∑k=0

F (σk(y))

)µ(y). (6.2)

177

Page 180: Characterization of Extremal KMS States on Groupoid C

Proof. We prove this by induction. Assume 6.2 holds for n. Then

µ(σn+1(y)) = µ(σ(σn(y))) = eβF (σn(y))µ(σn(y)) , since µ is eβF -conformal,

= eβF (σn(y)) exp

n−1∑k=0

F (σk(y))

)µ(y) , by hypothesis,

= exp

n∑k=0

F (σk(y))

)µ(y)

= exp

β (n+1)−1∑k=0

F (σk(y))

µ(y).

Then (6.2) holds for n+ 1. Let n = 1. Using the fact that µ is eβF -conformal, we have

µ(σn(y)) = µ(σ(y)) = eβF (y)µ(y) = exp

(βn−1∑k=0

F (σk(y))

)µ(y).

Proposition 6.3.6. Given β ∈ R, let µ be an eβF -conformal measure on X, x ∈ X. Then,

for every y ∈ O(x),

µ(y) = exp

(−β

(m−1∑j=0

F (σj(y))−n−1∑j=0

F (σj(x))

))µ(x),

where σn(x) = σm(y).

Proof. Let y ∈ O(x). There exist m,n > 0 such that σn(x) = σm(y). Then

µ(σn(x)) = exp

n−1∑j=0

F (σj(x))

)µ(x) and µ(σm(y)) = exp

m−1∑j=0

F (σj(y))

)µ(y).

Since σn(x) = σm(y), it follows that

exp

m−1∑j=0

F (σj(y))

)µ(y) = exp

n−1∑j=0

F (σj(x))

)µ(x)

178

Page 181: Characterization of Extremal KMS States on Groupoid C

µ(y) = exp

(βn−1∑j=0

F (σj(x))

)µ(x) exp

(−β

m−1∑j=0

F (σj(y))

)

µ(y) = exp

(−β

(m−1∑j=0

F (σj(y))−n−1∑j=0

F (σj(x))

))µ(x).

It follows from Proposition 6.3.6 that if two eβF -conformal measures µ1, µ2 are equal on

a point x ∈ X, then µ1, µ2 are equal on O(x). Moreover, if µ is an eβF -conformal measure

such that µ(x) > 0, then µ(y) > 0 for every y ∈ O(x).

Corollary 6.3.7. Let x ∈ X and β ∈ R. There exists at most one eβF -conformal probability

measure that vanishes outside O(x). In particular, if µ is an eβF -conformal probability

measure that vanishes outside O(x), then µ is extremal.

Proof. Let µ1, µ2 be a eβF -conformal probability measures vanishing outside O(x). It follows

from Proposition 6.3.6 that µ1(x), µ2(x) > 0. In fact, let i = 1, 2. Since O(x) is countable

and µi is a probability measure whose support lies in O(x), there exists y ∈ O(x) such that

µi(y) > 0. Then µi(x) > 0 by Proposition 6.3.6.

Suppose that µ1(x) < µ2(x). Then µ1(y) < µ2(y) for every y ∈ O(x) by Proposition 6.3.6.

Therefore

1 = µ1(X) = µ1(O(x)) = µ2(O(x)) = µ2(X) = 1,

which is a contradiction. is not a probability measure. The proof is analogous for µ1(x) >

µ2(x). Then µ1(x) = µ2(x). Then µ1 and µ2 are equal on O(x) by Proposition 6.3.6.

Therefore µ1 = µ2.

Now we show µ is extremal. Let µ1, µ2 be two eβF -conformal probability measures such

that µ = tµ1 + (1 − t)µ2, 0 < t < 1. Let A = X \ O(x). A is measurable. Since µ(A) = 0

and µ1(A), µ2(A) ≥ 0, we have µ1(A) = µ2(A) = 0. By previous arguments, µ = µ1 = µ2.

Therefore µ is extremal.

179

Page 182: Characterization of Extremal KMS States on Groupoid C

Corollary 6.3.8. Assume β ∈ R \ {0}. Let µ be an eβF -conformal measure. Let x ∈ X

periodic with minimum period p such that µ(x) > 0. It follows that

p−1∑j=0

F (σj(x)) = 0.

Proof. σp(x) = x. Then,

µ(x) = µ(σp(x)) = exp

p−1∑k=0

F (σk(x))

)µ(x).

µ(x) > 0, hence

1 = exp

p−1∑k=0

F (σk(x))

).

Note that β 6= 0. Thereforep−1∑k=0

F (σk(x)) = 0.

Lemma 6.3.9. Let µ be a purely atomic extremal eβF -conformal probability measure, β ∈ R.

Then there exists x ∈ X such that for y ∈ X, µ(y) > 0 if, and only if, y ∈ O(x).

Proof. Let x ∈ X such that µ(x) > 0.

Suppose there exists y ∈ X \ O(x) such that µ(y) > 0. Since µ is a probability measure,

µ(x) > 0 and x /∈ O(y), we have 0 < µ(O(y)) < 1. Let t = µ(O(y)). De�ne µ1 by

µ1(z) =

t−1µ(z) if z ∈ O(y)

0 otherwise.

Then µ1 is a probability measure. Note that σ(z) ∈ O(y) if, and only if, z ∈ O(y). In fact,

O(σ(z)) = O(z) from Lemma 6.2.2. Then

µ1(σ(z)) = t−1µ(σ(z)) = eβF (z)t−1µ(z) = eβF (z)µ1(z).

180

Page 183: Characterization of Extremal KMS States on Groupoid C

Therefore µ1 is eβF -conformal. De�ne µ2 by

µ2(z) =

(1− t)−1µ(z) if z /∈ O(y)

0 if z ∈ O(y)

.

µ2 is an eβF -conformal probability measure as well. In fact,

µ2(X) = µ2(X \ O(y))

= (1− t)−1µ(X \ O(y))

= (1− t)−1 [µ(X)− µ(O(y))]

= (1− t)−1 [1− t] , since t = µ(O(x)),

= 1.

Then µ2 is a probability measure. Given z /∈ O(y), σ(z) /∈ O(y). In fact, suppose σ(z) ∈

O(y). Then O(σ(z)) = O(y) by Lemma 6.2.2, but O(z) = O(σ(z)). Then O(σ(z)) = O(y).

Therefore z ∈ O(y). Contradiction. Then,

µ2(σ(z)) = (1− t)−1µ(σ(z))

= eβF (z)(1− t)−1µ(z) , since µ is eβF -conformal

= eβF (z)µ2(z).

Therefore µ2 is eβF -conformal.

Moreover, µ = tµ1 +(1−t)µ2. Therefore µ is not extremal. Contradiction. Then µ(y) = 0

if y /∈ O(x).

Let y ∈ O(x). It follows from Proposition 6.3.6 that µ(y) > 0.

For β 6= 0, each extremal atomic eβF -conformal probability measure corresponds to an

orbit O(x). However, an orbit O(x) does not necessarily have a correspondent extremal

atomic eβF -conformal probability measure.

181

Page 184: Characterization of Extremal KMS States on Groupoid C

Proposition 6.3.10. Let β ∈ R \ {0}. Let x ∈ X be a periodic point with minimum period

p. There exists an extremal atomic eβF -conformal probability measure with support O(x)

if, and only if,

p−1∑j=0

F (σj(x)) = 0, (6.3)

M =∞∑n=1

∑y∈Yn

exp

(−β

n−1∑j=0

F (σj(y))

)<∞. (6.4)

where Yn = σ−n(x) \⋃n−1j=0 σ

−j(x) for n ≥ 1, and Y0 = {x}.

In this case, the measure is denoted by

mx = (1 +M)−1

[δx +

∞∑n=1

∑y∈Yn

exp

(−β

n−1∑j=0

F (σj(y))

)δy

]. (6.5)

Proof. Let µ be an extremal atomic eβF -conformal probability measure with support O(x).

It follows from Corollary 6.3.8 that (6.3) holds.

Let y ∈ O(x), by Lemma 6.2.8 there exists a minimum natural number n such that

σn(y) = x. Hence y ∈ Yn. Therefore

O(x) =∞⋃n=0

Yn and µ(O(x)) =∞∑n=0

µ(Yn).

Let y ∈ Yn. From Proposition 6.3.6,

µ(y) = exp

(−β

n−1∑j=0

F (σj(y))

)µ(x).

182

Page 185: Characterization of Extremal KMS States on Groupoid C

O(x) is countable by Lemma 6.2.4. Thus,

µ(Yn) =∑y∈Yn

µ(y) =∑y∈Yn

exp

(−β

n−1∑j=0

F (σj(y))

)µ(x).

Hence,

1 = µ(O(x)) =∞∑n=0

µ(Yn)

= µ(x) +∞∑n=1

∑y∈Yn

exp

(−β

n−1∑j=0

F (σj(y))

)µ(x)

= µ(x)

[1 +

∞∑n=1

∑y∈Yn

exp

(−β

n−1∑j=0

F (σj(y))

)]

= µ(x)(1 +M).

Then M <∞ and µ(x) = (1 +M)−1. Given y ∈ Yn, n ≥ 1,

µ(y) = (1 +M)−1 exp

(−β

n−1∑j=0

F (σj(y))

).

Therefore (6.5) holds.

Conversely, assume (6.3), (6.4) hold. Given n > 1,

mx(Yn) = (1 +M)−1∑y∈Yn

exp

(−β

n−1∑j=0

F (σj(y))

).

Hence,

mx(O(x)) = mx(x) +∞∑n=1

mx(Yn)

= (1 +M)−1

[1 +

∞∑n=1

∑y∈Yn

exp

(−β

n−1∑j=0

F (σj(y))

)]

183

Page 186: Characterization of Extremal KMS States on Groupoid C

= (1 +M)−1(1 +M) = 1.

Then mx is a probability measure. Now we prove mx is eβF -conformal.

Let y ∈ Yn, n ≥ 2. Then σ(y) ∈ Yn−1. In fact, σn−1(σ(y)) = σn(y) = x. Let 0 ≤ l < n−1,

then σl(σ(y)) = σl+1(y) 6= x, since 1 ≤ l < n. Let y′ = σ(y), then

mx(σ(y)) = (1 +M)−1 exp

(−β

n−2∑j=0

F (σ(y′))

)

= (1 +M)−1 exp

(−β

n−1∑j=1

F (σ(y))

)

= eβF (y)(1 +M)−1 exp

(−β

n−1∑j=0

F (σ(y))

)

= eβF (y)mx(y).

Let y ∈ Yn, n = 1. Then σ(y) = x. Hence,

mx(σ(y)) = mx(x) = (1 +M)−1

= eβF (y)(1 +M)−1 exp (−βF (y))

= eβF (y)(1 +M)−1 exp

(−β

n−1∑j=0

F (σj(y))

)

= eβF (y)mx(y).

Note that σ(x) ∈ Yp−1. In fact, σp−1(σ(x)) = σp(x) = x. Given 0 ≤ l < p − 1, σl(σ(x)) =

σl+1(x) 6= x since 1 ≤ l + 1 < p. Then, if x′ = σ(x), we have

mx(σ(x)) = (1 +M)−1 exp

(−β

p−2∑j=0

F (σj(x′))

)

= (1 +M)−1 exp

(−β

p−1∑j=1

F (σj(x))

)

184

Page 187: Characterization of Extremal KMS States on Groupoid C

= eβF (x)(1 +M)−1 exp

(−β

p−1∑j=0

F (σj(x))

)

= eβF (x)(1 +M)−1 from (6.3)

= eβF (x)mx(x).

Therefore mx is eβF -conformal. mx has support O(x) by de�nition. From Corollary 6.3.7,

mx is extremal.

De�nition 6.3.11. Given a continuous function F : X → R we de�ne cF : G → R by

cF (x, k, y) =n−1∑j=0

F (σj(x))−m−1∑j=0

F (σj(y)),

for n,m ∈ N such that k = n−m and σn(x) = σm(y).

Lemma 6.3.12. cF is well-de�ned.

Proof. Let (x, k, y) ∈ G. For i = 1, 2, let ni,mi ∈ N such that σni(x) = σmi(y), k = ni −mi.

Then n2 − n1 = m2 −m1. Assume n2 > n1 without loss of generality. Then m2 > m1 and

n2−1∑j=0

F (σj(x))−m2−1∑j=0

F (σj(y))

=

(n1−1∑j=0

F (σj(x)) +

n2−1∑j=n1

F (σj(x))

)−

(m1−1∑j=0

F (σj(y)) +

m2−1∑j=m1

F (σj(y))

)

=

(n1−1∑j=0

F (σj(x)) +

n2−n1−1∑j=0

F (σj+n1(x))

)−

(m1−1∑j=0

F (σj(y)) +

m2−m1−1∑j=0

F (σj+m1(y))

)

=

(n1−1∑j=0

F (σj(x))−m1−1∑j=0

F (σj(y))

)+

(n2−n1−1∑j=0

F (σj+n1(x))−m2−m1−1∑

j=0

F (σj+m1(y))

)

=

(n1−1∑j=0

F (σj(x))−m1−1∑j=0

F (σj(y))

)+

(n2−n1−1∑j=0

F (σj(σn1(x)))−n2−n1−1∑j=0

F (σj(σm1(y)))

)

=

(n1−1∑j=0

F (σj(x))−m1−1∑j=0

F (σj(y))

)+

(n2−n1−1∑j=0

F (σj(σm1(y)))−n2−n1−1∑j=0

F (σj(σm1(y)))

)

185

Page 188: Characterization of Extremal KMS States on Groupoid C

=

(n1−1∑j=0

F (σj(x))−m1−1∑j=0

F (σj(y))

).

Therefore cF (x, k, y) does not depend on the choice of n,m satisfying k = n−m.

Proposition 6.3.13. cF is a continuous R-valued 1-cocycle on G.

Proof. Let (x, k, y), (y, l, z) ∈ G. There exist m,n, p, q ∈ N such that k = m− n, l = p− q,

σm(x) = σn(y) and σp(y) = σq(z).

Then m+ p− n− q = k + l and σm+p(x) = σn+p(y) = σn+q(z). Hence,

cF (x, k + l, z) =

m+p−1∑j=0

F (σj(x))−n+q−1∑j=0

F (σj(z))

=m−1∑j=0

F (σj(x)) +

m+p−1∑j=m

F (σj(x))−q−1∑j=0

F (σj(z))−n+q−1∑j=q

F (σj(z))

=m−1∑j=0

F (σj(x)) +

p−1∑j=0

F (σj+m(x))−q−1∑j=0

F (σj(z))−n−1∑j=0

F (σj+q(z))

=m−1∑j=0

F (σj(x)) +

p−1∑j=0

F (σj+n(y))−q−1∑j=0

F (σj(z))−n−1∑j=0

F (σj+p(y))

=m−1∑j=0

F (σj(x)) +

p+n−1∑j=n

F (σj(y))−q−1∑j=0

F (σj(z))−p+n−1∑j=p

F (σj(y))

=m−1∑j=0

F (σj(x))−n−1∑j=0

F (σj(y)) +

p+n−1∑j=0

F (σj(y))

−q−1∑j=0

F (σj(z)) +

p−1∑j=0

F (σj(y))−p+n−1∑j=0

F (σj(y))

= cF (x, k, y) + cF (y, l, z).

Now we prove the continuity of cF . Let (xi, ki, yi) → (x, k, y). Then xi → x and yi → y.

There exists m,n ∈ N with m−n = k and i0 such that σm(xi) = σn(yi) for every i ≥ i0. By

186

Page 189: Characterization of Extremal KMS States on Groupoid C

continuity of σ, σm(x) = σn(y).

Hence, for every i ≥ i0,

cF (xi, ki, yi) =m−1∑j=0

F (σj(xi))−n−1∑j=0

F (σj(yi)).

Since F and σ are continuous on X, we have cF (xi, ki, yi)→ cF (x, k, y).

The following lemma will be used to prove that every continuous R-valued 1-cocycle

corresponds to a unique cF .

Lemma 6.3.14. Let (x, n −m, y) ∈ G such that σn(x) = σm(y). Suppose that n,m ≥ 1.

Then

(x, n−m, y) =(x, 1, σ(x))(σ(x), 1, σ2(x)) · · · (σn−1(x), 1, σn(x))

(σm(y),−1, σm−1(y))(σm−1(y),−1, σm−2(y)) · · · (σ(y),−1, y). (6.6)

Proof. First we show that for every natural number N ≥ 1,

(x,N, σN(x)) = (x, 1, σ(x)) · · · (σN−1(x), 1, σN(x)). (6.7)

Clearly (6.7) holds for N = 1. Suppose that (6.7) is satis�ed for an arbitrary N . Then

(x,N + 1, σN+1(x)) = (x,N, σN(x))(σN(x), 1, σN+1(x))

= (x, 1, σ(x)) · · · (σN(x), 1, σN+1(x)).

Hence, (6.7) is satis�ed for every N . In particular, this equality holds for N = n. By the

same argument,

(y,m, σm(y)) = (y, 1, σ(y)) · · · (σm−1(y), 1, σm(y)).

187

Page 190: Characterization of Extremal KMS States on Groupoid C

Then, applying the inverse on both sides, we have

(σm(y),−m, y) = (σm(y),−1, σm−1(y)) · · · (σ(y),−1, y).

Since σn(x) = σm(y), we have

(x, n−m, y) =(x, n, σn(x))(σm(y),−m, y)

=(x, 1, σ(x))(σ(x), 1, σ2(x)) · · · (σn−1(x), 1, σn(x))

(σm(y),−1, σm−1(y))(σm−1(y),−1, σm−2(y)) · · · (σ(y),−1, y).

Now we will prove that there exists a bijection between continuous R-valued 1-cocycles

on G and the continuous functions from X to R. Recall from Proposition 6.3.13 that every

cF is a continuous R-valued 1-cocycle.

Proposition 6.3.15. For every continuous R-valued 1-cocycle c on G, there exists a unique

continuous function F : X → R such that c = cF .

Proof. Let c : G → R be a continuous R-valued 1-cocycle. De�ne the function F : X → G

by F (x) = c(x, 1, σ(x)).

The function F is continuous. In fact, let {xn}n∈N be a sequence inX converging to x ∈ X.

By de�nition of G, (x, 1, σ(x)) ∈ G and each (xn, 1, σ(xn)) ∈ G. Since σ is continuous, we

have σ(xn) → σ(x). It follows from Corollary 6.1.8 that (xn, 1, σ(xn)) → (x, 1, σ(x)). Since

c is continuous, we have F (xn)→ F (x). Therefore F is continuous.

Let (x, n − m, y) ∈ G such that σn(x) = σm(y). Assume that n,m ≥ 1 without loss of

generality. Then, by Lemma 6.3.14,

(x, n−m, y) =(x, 1, σ(x))(σ(x), 1, σ2(x)) · · · (σn−1(x), 1, σn(x))

=(σm(y),−1, σm−1(y))(σm−1(y),−1, σm−2(y)) · · · (σ(y),−1, y).

188

Page 191: Characterization of Extremal KMS States on Groupoid C

Then

c(x, n−m, y) =c(x, 1, σ(x)) + · · ·+ c(σn−1(x), 1, σn(x))

+ c(σm(y),−1, σm−1(y)) + · · ·+ c(σ(y),−1, y).

=n−1∑i=0

c(σi(x), 1, σi+1(x)) +m−1∑i=0

c(σi+1(y),−1, σi(y))

=n−1∑i=0

c(σi(x), 1, σi+1(x))−m−1∑i=0

c(σi(y), 1, σi+1(y))

=n−1∑i=0

F (σi(x))−m−1∑i=0

F (σi(y))

=cF (x, n−m, y).

Finally we prove that cF is unique. Suppose that there exists a continuous function

H : X → R such that cH = cF . Then, for every x ∈ X,

H(x) = cH(x, 1, σ(x)) = cF (x, 1, σ(x)) = F (x).

Therefore H = F .

The following lemma shows a equality for cF on Gyy where O(y) is periodic.

Lemma 6.3.16. Let x ∈ X be a periodic point with minimum period p. Then

cF (y, kp, y) = k

p−1∑j=0

F (σj(x)),

for every y ∈ O(x), k ∈ Z.

Proof. Since σp(x) = x, we have

cF (x, p, x) =

p−1∑j=0

F (σj(x)).

189

Page 192: Characterization of Extremal KMS States on Groupoid C

Let y ∈ O(x). Then, from Lemma 6.2.8, there exists n such that σn(y) = x. Thus

(y, n, x) ∈ G. Then

cF (y, kp, y) = kcF (y, p, y)

= kcF ((y, n, x)(x, p, x)(x,−n, y))

= k[cF (y, n, x) + cF (x, p, x) + cF (x,−n, y)]

= k[cF (y, n, x) + cF (x, p, x)− cF (x,−n, y)]

= kcF (x, p, x)

= k

p−1∑j=0

F (σj(x)).

Proposition 6.3.17. Let β 6= 0. Let z ∈ X be aperiodic. There exists an extremal atomic

eβF -conformal probability measure with support O(z) if, and only if,

M =∑y∈O(z)

e−βF(y) <∞, (6.8)

where

F(y) =m−1∑j=0

F (σj(y))−n−1∑j=0

F (σj(z)), (6.9)

with σm(y) = σn(z). In this case the measure is de�ned by

mz = M−1∑y∈O(z)

e−βF(y)δy. (6.10)

In particular, if z satis�es condition (6.8), we say z is β-summable.

Proof. First we prove F is well-de�ned. Let y ∈ O(z). There exists a unique k ∈ Z such

that (z, k, y) ∈ G. In fact, suppose there are k1, k2 ∈ Z satisfying (z, k1, y), (z, k2, y) ∈ G.

190

Page 193: Characterization of Extremal KMS States on Groupoid C

Then (z, k2 − k1, z) ∈ G. k2 − k1 = 0 from Lemma 6.2.10.

Let m,n ∈ N such that σm(y) = σn(z). Then (z, n−m, y) ∈ G. Hence k = n−m and

m−1∑j=0

F (σj(y))−n−1∑j=0

F (σj(z)) = −cF (z, k, y).

Therefore F(y) does not depend on the choice of m,n.

Let µ be an extremal atomic eβF -conformal probability measure with support O(z). Given

y ∈ O(z) there exist m,n ∈ N such that σm(y) = σn(z). Then by Proposition 6.3.6,

µ(y) = exp

(−β

(m−1∑j=0

F (σj(y))−n−1∑j=0

F (σj(z))

))µ(z) = e−βF(y)µ(z).

Since µ is a probability measure, it follows that

1 = µ(O(z)) =∑y∈O(z)

µ(y) =∑y∈O(z)

e−βF(y)µ(z).

Hence,

µ(z) =

∑y∈O(z)

e−βF(y)

−1

= M−1.

Therefore M <∞ and µ(y) = M−1e−βF(y) for y ∈ O(z).

Conversely, assume (6.8) holds. We prove that mz is an extremal eβF -conformal proba-

bility measure.

mz(O(z)) = M−1∑y∈O(z)

e−βF(y) = M−1M = 1,

thenmz is a probability measure. Now we showmz is eβF -conformal. Let y ∈ O(z), n,m ∈ N

191

Page 194: Characterization of Extremal KMS States on Groupoid C

such that σm(y) = σn(z), m ≥ 2. Then σm−1(σ(y)) = σn(z) and

F(σ(y)) =m−2∑j=0

F (σj(σ(y)))−n−1∑j=0

F (σj(z)) =m−1∑j=1

F (σj(y))−n−1∑j=0

F (σj(z))

= −F (y) +m−1∑j=0

F (σj(y))−n−1∑j=0

F (σj(z))

= −F (y) + F(y).

Then,

mz(σ(y)) = M−1e−βF(σ(y)) = M−1e−β[−F (y)+F(y)] = M−1eβF (y)e−βF(y) = eβF (y)mz(y).

Therefore mz is eβF -conformal. By de�nition of mz, its support is O(z). From Corollary

6.3.7, mz is extremal.

Lemma 6.3.18. Let X be a locally compact second countable Hausdor� topological space,

µ a Borel measure which is �nite on compact subsets of X. Given a local homeomorphism

σ : X → X, and a non-negative function f on X, µ is f -conformal if, and only if,

µ(σ(A)) =

∫A

f(x)dµ(x), (6.11)

for every open set A such that σ|A is injective.

Proof. The measure µ is Radon by Proposition 2.4.10. Assume µ is f -conformal, then (6.11)

holds by de�nition.

Conversely, suppose (6.11) holds. Let A be a measurable subset such that σ is injective

on A. First we assume there exists an open set U such that A ⊂ U and σ is injective on U .

Given an open set W including σ(A), there exists an open set W ′ = W ∩ σ(U) such that

σ(A) ⊂ W ′ ⊂ σ(U) and µ(W ′) ≤ µ(W ). Since µ is a Radon measure and W is arbitrary, we

192

Page 195: Characterization of Extremal KMS States on Groupoid C

have

µ(σ(A)) = infσ(A)⊂WW open

µ(W ) = infσ(A)⊂W ′⊂σ(U)

W ′ open

µ(W ′).

The function σ is injective on U , then for every open set W satisfying σ(A) ⊂ W ⊂ σ(U),

there is a unique open set V such that A ⊂ V ⊂ U and W = σ(V ). Clearly σ is injective on

each V . Then,

µ(σ(A)) = infA⊂V⊂UV open

µ(σ(V ))

= infA⊂V⊂UV open

∫V

f(x)dµ(x), by hypothesis,

=

∫A

f(x)dµ(x), by Lemma 2.4.13.

Now let A be an arbitrary measurable set A such that σ|A is injective. Since σ is a local

homeomorphism and X is second countable, there exists a countable open cover {Un}n∈N of

A such that σ is injective on each Un. De�ne A1 = A ∩ U1 and, for every n,

An+1 = A ∩ Un+1 \n⋃j=1

Aj.

Then A = ∪∞n=1An and the family {An}n∈N is disjoint. Moreover, An ⊂ Un for every n.

Then,

µ(σ(An)) =

∫An

f(x)dµ(x).

Since σ is injective on A, we have,

µ(σ(A)) =∞∑n=1

µ(σ(An)) =∞∑n=1

∫An

f(x)dµ(x) =

∫A

f(x)dµ(x).

193

Page 196: Characterization of Extremal KMS States on Groupoid C

Therefore µ is f -conformal.

Lemma 6.3.19. Let σ : X → X be a local homeomorphism. Let n ∈ N*, x ∈ X. Then

there exists an open neighborhood U of x such that for j = 1, . . . , n,

σ|σj−1(U) : σj−1(U)→ σj(U) is a homeomorphism and σj(U) is open. (6.12)

Proof. We prove this by induction. Let n = 1. There exists an open neighborhood U of x

such that σ(U) is open and σ|U : U → σ(U) is a homeomorphism. Then the result holds for

n = 1.

Now assume the result holds for n ≥ 1. Since σ is a local homeomorphism, there exists

an open neighborhood W of σn(x) in σn(U) such that σ(W ) is open and σ|W : W → σ(W )

is a homeomorphism.

Let V = σ|−nU (W ). V is an open neigborhood of x in U . Let j = 1, . . . , n. Note that

σ|jU : U → σj(U) is a homeomorphism. Since V ⊂ U , then σj(V ) is open and σj(V ) ⊂ σj(U).

Hence σ|σj−1(V ) : σj−1(V )→ σj(V ) is a homeomorphism for i = 1, . . . , n.

Note that σn+1(V ) = σ(σ|nU ◦ σ|−nU (W )) = σ(W ). Then σn+1(V ) is open. It follows that

σ|σn(V ) : σn(V )→ σn+1(V ) is a homeomorphism, therefore the result holds for n+ 1.

Remark 6.3.20. Let x, y ∈ X such that σn(x) = σm(y) for n,m ∈ N. Assume A is an

open neigborhood of x such that (6.12) holds for j = 1, . . . , n, replacing U by A. By the

same argument, suppose there exists B, an open neigborhood of y such that (6.12) holds for

j = 1, . . . ,m. Then we can assume without loss of generality that σn(A) = σm(B).

In fact, let V = σn(A) ∩ σm(B). V is an open neigborhood of σn(x). Let A0 = σ|−nA (V ),

B0 = σ|−mB (V ). Clearly A0 and B0 are open neigborhoods of x, y, respectively. Then, for

j = 1, . . . , n, (6.12) holds for A0, since σj−1(A0) ⊂ σj−1(A) and σj(A0) ⊂ σj(A) are open

sets. Analogously (6.12) holds for B0.

Moreover,

σn(A0) = σn(σ|−nA (V )) = V = σm(σ|−mB (V )) = σm(B0).

194

Page 197: Characterization of Extremal KMS States on Groupoid C

Theorem 6.3.21. Let β ∈ R. A measure µ on G(0) is eβF -conformal if, and only if, µ is

quasi-invariant with Radon-Nikodym derivative e−βcF .

Proof. Assume µ is eβF -conformal.

Let U ⊂ G be an open bisection. Let T : r(U)→ s(U) be the homeomorphism de�ned by

T = s|U ◦ r|−1U . Given y ∈ s(U), let hy = s|−1

U (y) = (x, k, y). There are A,B ⊂ X open sets,

n,m ∈ N such that hy ∈ Un,mA,B ⊂ U .

By Lemma 6.3.19 there exists an open neigborhood of x, A0 ⊂ A, such that for j =

1, . . . , n,

σ|σj−1(A0) : σj−1(A0)→ σj(A0) is a homeomorphism and σj(A0) is open.

By the same argument, there exists an open neigborhood B0 of y, with B0 ⊂ B, such that

for j = 1, . . . ,m,

σ|σj−1(B0) : σj−1(B0)→ σj(B0) is a homeomorphism and σj(B0) is open.

We can assume σn(A0) = σm(B0) without loss of generality. Then hy ∈ Un,mA0,B0⊂ Un,mA,B .

Note that s(Un,mA0,B0) = B0 and r(Un,mA0,B0

) = A0. In fact, since σ|nA0, σ|mB0

are homeomor-

phisms, it follows that

Un,mA0,B0= {(x′, n−m, y′) : x′ ∈ A0, y

′ ∈ B0, σn(x′) = σm(y′)}

= {(σ|−nA0(σ|mB0

(y′)), n−m, y′) : y′ ∈ B0}

= {(x′, n−m,σ|−mB0(σ|nA0

(x′))) : x′ ∈ A0}.

Moreover, T |A0 = σ|−mB0◦ σ|nA0

. Note that

σ|−mB0◦ σ|nA0

= σ|−1B0◦ σ|−1

σ(B0) ◦ . . . ◦ σ|−1σm−1(B0) ◦ σ|

nA0

= σ|−mB0◦ σ|σn−1(A0) ◦ σ|σn−2(A0) ◦ . . . σ|σ1(A0) ◦ σ|A0 .

195

Page 198: Characterization of Extremal KMS States on Groupoid C

Given j = 0, . . . , n+m, let

Tj =

σ|−mB0◦ σ|n−j

σj(A0), if 0 ≤ j ≤ n,

σ|−m−n+jB0

, if n ≤ j ≤ n+m.

Then T0 = T |A0 and Tn+m = id|B0 . Note that Tj : σj(A0) → B0 if j ≤ n and Tj :

σn+m−j(B0)→ B0 if j ≥ n. Moreover,

• if 0 ≤ j ≤ n− 1,

Tj = σ|−mB0◦ σ|n−j

σj(A0)= σ|−mB0

◦ σ|n−j−1σj+1(A0)

◦ σ|σj(A0) = Tj+1 ◦ σ|σj(A0).

Then Tj+1 = Tj ◦ σ|−1σj(A0)

. Also,

T−1j y = σ|j−n

σj(A0)◦ σ|mB0

(y)

= (σ|n−jσj(A0)

)−1 ◦ σ|nA0(x)

= (σ|nA0◦ σ|−jA0

)−1 ◦ σ|nA0(x)

= σ|jA0◦ σ|−nA0

◦ σ|nA0(x)

= σ|jA0(x).

Let B1 ⊂ B0 be measurable. Then,

Tj+1∗µ(B1) = Tj∗σ|−1σj(A0)∗

µ(B1)

= σ|−1σj(A0)∗

µ(Tj−1(B1))

= µ(σ(Tj−1(B1))).

Tj−1(B1) is measurable, since Tj is continuous. Since Tj

−1(B1) ⊂ σj(A0), σ is injective

196

Page 199: Characterization of Extremal KMS States on Groupoid C

on σj(A0) and µ is eβF -conformal, we have

µ(σ(Tj−1(B1))) =

∫T−1j (B0)

eβF (u)dµ(u)

=

∫T−1j (B0)

eβF (T−1j Tju)dµ(u)

=

∫B0

eβF (T−1j u)dTj∗µ(u) by (2.2) on page 15

ThenTj+1∗µ

Tj∗µ(y) = eβF (T−1

j y) = eβF (σj(x)).

Therefore,

Tj∗µ

Tj+1∗µ(y) = e−βF (σj(x)). (6.13)

• if n ≤ j ≤ n+m− 1,

Tj = σ|−m−n+jB0

= (σ|m+n−jB0

)−1

= (σ|σm+n−j−1(B0) ◦ σ|m+n−j−1B0

)−1

= (σ|m+n−j−1B0

)−1 ◦ σ|−1σm+n−j−1(B0)

= (σ|−m−n+j+1B0

)−1 ◦ σ|−1σm+n−j−1(B0)

= Tj+1 ◦ σ|−1σm+n−j−1(B0)

.

Moreover,

T−1j y = σ|m+n−j

B0(y) = σm+n−j(y).

197

Page 200: Characterization of Extremal KMS States on Groupoid C

Let B1 ⊂ B0 measurable. Then,

Tj∗µ(B1) = Tj+1∗σ|−1σm+n−j−1(B0)∗

µ(B1)

= σ|−1σm+n−j−1(B0)∗

µ(T−1j+1(B1))

= µ(σ(T−1j+1(B1))).

Note that T−1j+1(B1) is measurable by continuity of Tj+1. Since

Tj+1−1(B1) ⊂ σm+n−j−1(B0),

σ is injective on σm+n−j−1(B0) and µ is eβF -conformal, we have

µ(σ(T−1j+1(B1))) =

∫T−1j+1(B1)

eβF (u)dµ(u)

=

∫T−1j+1(B1)

eβF (T−1j+1Tj+1u)dµ(u)

=

∫B1

eβF (T−1j+1u)dTj+1∗µ(u) by (2.2) on page 15.

Then,

dTj∗µ

dTj+1∗µ(y) = eβF (T−1

j+1(y)) = eσm+n−j−1(y). (6.14)

Therefore,

dT∗µ

dµ(y) =

dT0∗µ

dTm+n∗µ(y)

=dT0∗µ

dT1∗µ(y)

dT1∗µ

dT2∗µ(y) . . .

dTn+m−1∗µ

dTn+m∗µ(y)

=

(n−1∏j=0

dTj∗µ

dTj+1∗µ(y)

)(n+m−1∏j=n

dTj∗µ

dTj+1∗µ(y)

)

198

Page 201: Characterization of Extremal KMS States on Groupoid C

=

(n−1∏j=0

e−βF (σj(x))

)(n+m−1∏j=n

eβF (σm+n−j−1(y))

)from (6.13) and (6.14),

=

(n−1∏j=0

e−βF (σj(x))

)(m−1∏j=0

eβF (σm−j−1(y))

).

Making the change of variables j 7→ m− j − 1 in the second product, we have

=

(n−1∏j=0

e−βF (σj(x))

)(m−1∏j=0

eβF (σj(y))

)

= exp

(−β

n−1∑j=0

F (σj(x)) + β

m−1∑j=0

F (σj(y))

)

= exp

(−β

(n−1∑j=0

F (σj(x))−m−1∑j=0

F (σj(y))

))

= e−βcF (x,k,y)

= e−βcF (hy).

Since y ∈ s(U) is arbitrary, the equality holds for every y ∈ s(U). U is any open bisection.

Therefore, by Lemma 5.3.8, µ is quasi-invariant with Radon-Nikodym derivative e−βcF .

Conversely, assume µ is quasi-invariant with Radon-Nikodym derivative e−βcF .

Let A ⊂ X be an open set such that σ|A is invertible. U0,1σ(A),A is an open bisection. In

fact,

U0,1σ(A),A = {(x,−1, y) : x ∈ σ(A), y ∈ A, x = σ(y)}

= {(σ(y),−1, y) : y ∈ A}.

Since σ is injective on A, it follows that r, s are injective on U0,1σ(A),A. Moreover, A = s(U0,1

σ(A),A),

σ(A) = r(U0,1σ(A),A), and T : σ(A)→ A is given by T = σ|−1

A . From Lemma 5.3.8,

dT∗µ

dµ(y) = e−βcF (hy)

= exp (−βcF (σ(y),−1, y))

199

Page 202: Characterization of Extremal KMS States on Groupoid C

= exp

(−β

(0∑j=0

F (σj(σ(y))−1∑j=0

F (σj(y))

))

= exp (−β [F (σ(y))− F (y)− F (σ(y))])

= eβF (y).

Hence,

µ(σ(A)) = µ(σ|A(A)) = µ(T−1(A)) = T∗µ(A) =

∫A

eβF (y)dµ(y).

It follows from Lemma 6.3.18 that µ is eβF -conformal.

Remark 6.3.22. Let β 6= 0. Given a eβF -conformal measure, it follows from Theorem 6.3.21

that µ is quasi-invariant with Radon-Nikodym derivative e−βcF . From Remark 4.1.9 we have

−βcF (g) = 0 for µ-a.e. x ∈ G(0) and all g ∈ Gxx, (6.15)

then

cF (g) = 0 for µ-a.e. x ∈ G(0) and all g ∈ Gxx, (6.16)

We can show this fact for extremal eβF -conformal probability measures using the properties

of G when the set of periodic points in X is countable.

Let µ be an extremal eβF -conformal probability measure. From Lemma 6.3.3 and Lemma

6.3.9, the measure falls in one of these cases: µ is continuous; µ is purely atomic and

supported on a periodic orbit; or µ is purely atomic and supported on an aperiodic orbit.

Suppose µ is continuous. Since the set of periodic points is countable, it follows that for

µ-a.e. y, Gyy = {(y, 0, y)}. However, cF (y, 0, y) = 0. Then the result holds for µ continuous.

If µ is purely atomic and supported on an aperiodic orbit O(z), then Gyy = {(y, 0, y)} for

200

Page 203: Characterization of Extremal KMS States on Groupoid C

every y ∈ O(z). Then (6.16) holds.

If µ is purely atomic and supported on a periodic orbit O(x) for a periodic point x with

minimum period p, it follows from 6.3.8 that

p−1∑i=0

F (σi(x)) = 0.

Given y ∈ O(x), Gyy = {(y, kp, y) : k ∈ Z} from Lemma 6.2.10. From Lemma 6.3.16, we have

cF (y, kp, y) = 0 for every k ∈ Z. Then (6.16) holds.

Hence, the result follows for every extremal eβF -conformal probability measure.

6.4 KMS States on the Renault-Deaconu Groupoid

In this section we �nd all extremal KMS states on C*(G) using Neshveyev's Theorems.

Since the continuous R-valued 1-cocycle on G is given by cF as in De�nition 6.3.11, we �x

the dynamics τ on C*(G), given by τt(f)(g) = eitcF (g)f(g).

We will show that all extremal KMS states are φm, φmz , φλx, de�ned below.

Let m be an extremal eβF -conformal non-atomic probability measure on X. We de�ne

φm by

φm(f) =

∫X

f(y, 0, y)dm(y),

where f ∈ Cc(G).

Given an aperiodic β-summable point z, by Proposition 6.3.17 mz is an extremal eβF -

conformal probability measure. We de�ne φmz by

φmz(f) =

∫X

f(y, 0, y)dmz(y),

with f ∈ Cc(G).

201

Page 204: Characterization of Extremal KMS States on Groupoid C

Given a periodic point x with period p with x satisfying the conditions of Proposition

6.3.10, λ ∈ C with |λ| = 1, we de�ne φλx by

φλx(f) =

∫X

∑k∈Z

λkf(y, kp, y)dmx(y),

for f ∈ Cc(G).

Lemma 6.4.1. Let x ∈ G(0). Given g1, g2 ∈ Gxx , ug1 · ug2 = ug1g2 . As a consequence, C*(Gxx)

is a commutative C*-algebra with identity ux.

Proof. Let g ∈ G, then

(ug1 · ug2)(g) =∑ab=g

ug1(a)ug2(b)

=

1 if g1g2 = g

0 otherwise

= ug1g2(g).

Lemma 6.4.2. Let x ∈ X, y ∈ O(x), h ∈ Gxy . There exists an ∗-isomorphism P : C*(Gyy )→

C*(Gxx) given by P (ug) = uhgh−1 . Moreover, P is an isometry.

Proof. Let P : C*(Gyy ) → C*(Gxx) be the linear map de�ned by P (ug) = uhgh−1 . Note that

P is invertible with inverse given by P−1 : C*(Gxx)→ C*(Gyy ), de�ned by P−1(ug) = uh−1gh.

First we show that P is a homomorphism. Given g1, g2, g ∈ Gyy ,

P (ug1 · ug2) = P (ug1g2) = uhg1g2h−1 = uhg1h−1hg2h−1 = uhg1h−1 · uhg2h−1 = P (ug1) · P (ug2),

P (ug*) = P (ug−1) = uhg−1h−1 = u(hgh−1)−1 = u(hgh−1)* = P (ug)*.

202

Page 205: Characterization of Extremal KMS States on Groupoid C

Since P is linear, we have that P is a ∗-homomorphism. So is P−1 by the same arguments.

Then P is an ∗-isomorphism between C*-algebras and therefore P is an isometry.

Remark 6.4.3. Lemma 6.4.2 can be generalized for a locally compact Hausdor� second

countable étale groupoid G. Given x, y ∈ G(0) such that Gxy 6= ∅, there exists an ∗-

isomorphism P :: C*(Gyy)→ C(Gx

x) given by P (ug) = uhgh−1 .

Proposition 6.4.4. Suppose the set of periodic points in X is countable. Let β ∈ R \ {0}.

Let φ be the KMSβ-state corresponding to the pair (µ, {ϕy}y). Then φ is extremal if and

only if µ is an extremal eβF -conformal measure and ϕy is a character for µ-a.e. y.

Proof. It follows from Theorem 5.3.10 that µ is quasi-invariant with Radon-Nikodym deriva-

tive e−βcF . It follows from Theorem 6.3.21 that µ is eβF -conformal.

Assume φ is extremal.

• Suppose µ is not extremal, then µ = tµ1 + (1 − t)µ2 where µ1, µ2 are eβF -conformal

probability measures, 0 < t < 1 and µ1 6= µ2.

Then µ1 and µ2 are quasi-invariant with Radon-Nikodym derivative e−βcF . Since

µ1, µ2 � µ, conditions (ii) and (iii) of Theorem 5.3.10 are satis�ed. Then (µ1, {ϕy}y)),

(µ2, {ϕy}y)) correspond to the KMS states φ1, φ2, respectively. Note that φ1 6= φ2.

Then

tφ1(f) + (1− t)φ2(f) = t

∫X

∑g∈Gxx

f(g)ϕx(ug)dµ1(x) + (1− t)∫X

∑g∈Gxx

f(g)ϕx(ug)dµ2(x)

=

∫X

∑g∈Gxx

f(g)ϕx(ug)d(tµ1 + (1− t)µ2)(x)

=

∫X

∑g∈Gxx

f(g)ϕx(ug)dµ(x)

= φ(f).

Then φ is not extremal. Contradiction.

203

Page 206: Characterization of Extremal KMS States on Groupoid C

• Suppose µ is extremal and non-atomic. Let I be the set aperiodic points in X. Then

µ(I) = 1 by Lemma 6.2.9. Given y ∈ I, Gyy = {y} by Lemma 6.2.10. Then every

element in C*(Gyy ) in the form auy, where a ∈ C. In addition ϕy(uy) = 1 because ϕ is

a state and uy is the unit of C*(Gyy ). Therefore ϕy is a character.

• Assume µ is extremal and atomic. Then there exists an orbit I with µ(I) = 1. Suppose

there is x ∈ I such that ϕx is not a character. Then, by Lemma 6.4.1, there are states

ϕ(1)x , ϕ

(2)x , t ∈ (0, 1) such that ϕ(1)

x 6= ϕ(2)x and ϕx = tϕ

(1)x + (1− t)ϕ(2)

x .

De�ne the �eld of states {ϕ(1)y }y, {ϕ(2)

y }y by

ϕ(i)y (ug) =

ϕy(ug) if y /∈ I

ϕ(i)x (uhgh−1) if y ∈ I, and h ∈ Gx

y arbitrary,

for i = 1, 2. Note that hgh−1 does not depend on the choice of g. In fact, given

y ∈ O(x), g ∈ Gyy , h ∈ Gxy , there exists k, l ∈ N such that g = (y, k, y) and h = (x, l, y).

Then hgh−1 = (x, k, x).

It is clear that ϕ(i)y is a state for y /∈ I. We will show that ϕ(i)

y is also a state when

y ∈ I. Given h ∈ Gxy , let P : C*(Gyy )→ C*(Gxx) be de�ned by P (ug) = uhgh−1 for every

g ∈ Gyy . It follows from Lemma 6.4.2 that P is an isometry. Moreover, by de�nition of

ϕ(i)y , we have ϕ(i)

y = ϕ(i)x ◦ P . ϕ(i)

y is linear and bounded because

‖ϕ(i)y ‖ ≤ ‖ϕ(i)

x ‖‖P‖ = ‖ϕ(i)x ‖.

Since ϕ(i)x = ϕ

(i)y ◦ P−1, we can show analogously that ‖ϕ(i)

x ‖ ≤ ‖ϕ(i)y ‖. Then ‖ϕ(i)

y ‖ =

‖ϕ(i)x ‖ = 1. Moreover,

ϕ(i)y (uy) = ϕ(i)

x (uhyh−1) = ϕ(i)x (uhh−1) = ϕ(i)

x (ux) = 1.

Note that uy is the unit of C*(Gyy ). It follows from Theorem 5.1.2 that ϕ(i)y is a state.

204

Page 207: Characterization of Extremal KMS States on Groupoid C

Since µ is atomic, {ϕ(i)y }y de�nes a µ-measurable �eld of states by Remark 5.2.5. For

i = 1, 2, let φ(i) be the state de�ned by (µ, {ϕ(i)y }y). We will show that φ(i) is a KMS

state. In order to prove this, we have show that properties (i)-(iii) in Theorem 5.3.10

hold.

We already know that µ is quasi-invariant with Radon-Nikodym derivative e−βcF , so

property (i) holds.

We will prove property (ii). Let y ∈ I, then there exists k ∈ Gxy . Let h ∈ Gy. Then

r(k) ∈ I and kh−1 ∈ Gxr(h). For every g ∈ Gyy , we have

ϕ(i)r(h)(uhgh−1) = ϕ(i)

x (ukh−1(hgh−1)hk−1), by de�nition of ϕ(i)r(h),

= ϕ(i)x (ukgk−1)

= ϕ(i)y (uy), by de�nition of ϕ(i)

y .

Then property (ii) holds.

For µ-a.e. y ∈ X, all g ∈ Gyy , we have cF (g) = 0 by Remark 6.3.22. Then property (iii)

holds.

By de�nition of ϕ(i)y , ϕy(ug) = tϕ

(1)y (ug) + (1− t)ϕ(2)

y (ug) if y /∈ I. If y ∈ I, there exists

h ∈ Gxy . Then,

ϕy(ug) = ϕr(h)(uhgh−1)

= ϕx(uhgh−1)

= tϕ(1)x (uhgh−1) + (1− t)ϕ(2)

x (uhgh−1)

= tϕ(1)y (ug) + (1− t)ϕ(2)

y (ug).

Hence ϕy = tϕ(1)y + (1− t)ϕ(2)

y and ϕ(1)y 6= ϕ

(2)y for every y ∈ I. By de�nition of {ϕ(i)

y }y,

(µ, {ϕ(i)y }y) de�nes a state φ(i), i = 1, 2. Note that φ(1) 6= φ(2).

205

Page 208: Characterization of Extremal KMS States on Groupoid C

Then, for f ∈ Cc(G),

φ(f) =

∫X

∑g∈Gyy

f(g)ϕy(ug)dµ(y)

=

∫X

∑g∈Gyy

f(g)[tϕ(1)

y (ug) + (1− t)ϕ(2)y (ug)

]dµ(y)

= t

∫X

∑g∈Gyy

f(g)ϕ(1)y dµ(y) + (1− t)

∫X

∑g∈Gyy

f(g)ϕ(2)y dµ(y)

= tφ(1)(f) + (1− t)φ(2)(f).

Then φ is not extremal. Contradiction. Therefore µ is extremal and ϕy is a character

for µ-a.e. y.

Conversely, suppose µ is extremal and ϕy is a character for µ-a.e. y. Suppose there exist

KMS states φ(1), φ(2), t ∈ (0, 1) such that φ = tφ(1) + (1− t)φ(2) and each φ(i) corresponds to

the pair (µi, {ϕ(i)y }y). Since X is clopen G, we have for every f ∈ Cc(X),

φ(f) = tφ(1)(f) + (1− t)φ(2)(f)∫X

f(x)dµ(x) =

∫X

f(x)dµ1(x) + (1− t)∫X

f(x)dµ2(x).

Then µ = tµ1 + (1 − t)µ2. Since µ is extremal, we have µ1 = µ2 = µ. Then, for every

f ∈ Cc(G),

φ(f) = tφ(1)(f) + (1− t)φ(2)(f)

= t

∫X

∑g∈Gxx

f(g)ϕ(1)x (ug)dµ(x) + (1− t)

∫X

∑g∈Gxx

f(g)ϕ(2)x (ug)dµ(x)

=

∫X

∑g∈Gxx

f(g)[tϕ(1)

x (ug) + (1− t)ϕ(2)x (ug)

]dµ(x).

Each tϕ(1)x +(1−t)ϕ(2)

x is a state on C*(Gxx). Moreover, {tϕ(1)x +(1−t)ϕ(2)

x }x is a µ-measurable

�eld of states. Then the pair (µ, {tϕ(1)x +(1−t)ϕ(2)

x }x) also de�nes φ. It follows from Theorem

206

Page 209: Characterization of Extremal KMS States on Groupoid C

5.2.9 that

ϕx = tϕ(1)x + (1− t)ϕ(2)

x , for µ-a.e. x

Since ϕx is a character µ-a.e., it follows that ϕx = ϕ(1)x = ϕ

(2)x for µ-a.e. x. Then, by Theorem

5.2.9, φ = φ(1) = φ(2). Therefore φ is extremal.

Lemma 6.4.5. Let y ∈ X be aperiodic. There is a unique state ϕy on C*(Gyy ). In particular,

ϕy is a character.

Proof. Since Gyy = {y}, it follows that C*(Gyy ) is isomorphic to C. Hence there is a unique

state on C*(Gyy ), which is a character.

Lemma 6.4.6. Let y ∈ X such that O(y) is periodic with period p. Let λ ∈ C such that

|λ| = 1. De�ne the linear functional ϕλy on C*(Gyy ) by ϕλy(u(y,kp,y)) = λk. Then ϕλy is a

character on C*(Gyy ). In fact, ϕλy are the only characters de�ned on C*(Gyy ).

Proof. Note that we can identify Gyy with Z by the isomorphism (y, kp, y) 7→ k. Moreover,

C*(Z) is isomorphic to C(S1), the set of continuous functions on the complex unit circle. In

fact, for k ∈ Z, let uk : Z→ C be de�ned by

uk(l) =

1 if l = k

0 otherwise,

and let pk : S1 → C be de�ned by pk(z) = zk.

There exists an isomorphism from C*(Z) to C(S1) given by uk 7→ pk. Since {pk}k∈Zgenerates the commutative C*-algebra C(S1), it follows that all characters of C(S1) corre-

spond to elements on the unit circle S1. Hence, each character ϕ on C*(Gyy ) corresponds to

a character λ on C(S1) such that

ϕ(u(y,kp,y)) = pk(λ) = λk.

207

Page 210: Characterization of Extremal KMS States on Groupoid C

Now we describe all extremal KMSβ-states on C*(G). Recall that the dynamics on C*(G)

is given by τt(f)(g) = eitcF (g)f(g).

Theorem 6.4.7. [26, Theorem 2.2] Let β ∈ R \ {0}. Assume that the periodic points of σ

are countable. The extremal KMSβ-states for τ are

1. States φm, where m is an extremal and continuous (non-atomic) eβF -conformal Borel

probability measure on X;

2. The states φλx, where λ ∈ C, |λ| = 1 and x is periodic with minimum period p, such that

p−1∑j=0

F (σj(x)) = 0 and∞∑n=1

∑y∈Yn

exp

(−β

n−1∑j=0

F (σj(y))

)<∞; (6.17)

3. The states φmz where z is aperiodic and β-summable.

Proof. Let φ be an extremal KMSβ-state corresponding to the pair (µ, {ϕy}y). From Propo-

sition 6.4.4, µ is an extremal eβF -conformal probability measure, and ϕy is a character for

µ-a.e. y.

It follows from Theorem 6.3.21 that µ is quasi-invariant with Radon-Nikodym derivative

e−βcF . Since µ is extremal, µ is either atomic or non-atomic.

(i) Suppose µ = m is non-atomic or µ = mz for z aperiodic.

If µ = mz, then z is β-summable by Proposition 6.3.17.

Let φ be an extremal KMSβ-state corresponding to the pair (m, {ϕy}y). De�ne I by

• the set of aperiodic points, i.e. points X whose orbits are aperiodic, if µ is non-

atomic. From Lemma 6.2.9 we have m(I) = 1.

• I = O(z) if µ = mz.

208

Page 211: Characterization of Extremal KMS States on Groupoid C

Then µ(I) = 1 and y ∈ I is aperiodic for every y ∈ I. Then by Lemma 6.4.5, ϕy

de�ned by ϕy(uy) = 1 is the unique character de�ned on C*(Gyy ).

Hence φ is de�ned by

φ(f) =

∫X

∑g∈Gyy

f(g)dµ(y) =

∫I

f(y, 0, y)dµ(y) =

∫X

f(y, 0, y)dµ(y).

Then φ = φm if µ = m and φ = φmz if µ = mz.

(ii) Suppose µ = mx with x periodic with minimum period p.

From Proposition 6.3.10 conditions (6.17) must hold.

Let I = O(x). Let φ be an extremal KMSβ-state with corresponding pair (mx, {ϕy}y).

It follows from Proposition 6.4.4 and Lemma 6.4.6 that ϕx = ϕλx for some λ ∈ S1.

Let y ∈ O(x), then there exists h = (x, l, y) ∈ G for some l ∈ Z. Let g ∈ Gyy ,

then g = (y, kp, y) for some k ∈ Z by Lemma 6.2.10. Then applying property (iii) of

Theorem 5.3.10,

ϕy(ug) = ϕr(h)(uhgh−1)

= ϕx(u(x,l,y)(y,kp,y)(y,−l,x))

= ϕx(u(x,kp,x))

= ϕλx(u(x,kp,x))

= λk

= ϕλy(u(y,kp,y))

= ϕλy(ug).

Then ϕy = ϕλy for every y ∈ I. Then for f ∈ Cc(G),

φ(f) =

∫X

∑g∈Gyy

f(g)ϕy(ug)dmx(y)

209

Page 212: Characterization of Extremal KMS States on Groupoid C

=

∫I

∑g∈Gyy

f(g)ϕy(ug)dmx(y)

=

∫I

∑g∈Gyy

f(g)ϕλy(ug)dmx(y)

=

∫I

∑k∈Z

f(y, kp, y)ϕλy(uy,kp,y)dmx(y)

=

∫I

∑k∈Z

f(y, kp, y)λkdmx(y)

=

∫X

∑k∈Z

f(y, kp, y)λkdmx(y)

= φλx(f).

Hence every extremal KMS state has the form φm, φmz or φλx.

Now we prove that φm, φmz , φλx satisfying the conditions of the theorem always de�ne

extremal KMS states.

Note that, for an extremal eβF -conformal probability measure µ on X, it follows from

Remark 6.3.22 that for µ-a.e. x ∈ X, all g ∈ Gxx , we have cF (g) = 0. Then in order to

prove that a state de�ned by (µ, {ϕy}y) is KMSβ, we only need to show that property

(ii) of Theorem 5.3.10 holds.

• φm

Let m be an extremal and continuous eβF -conformal Borel probability measure

on X. Let {ϕy}y∈X be a family of states ϕy on C*(Gyy ).

Let I be the set of aperiodic points. m(X \I) = 0 by Lemma 6.2.9. Let f ∈ Cc(G).

Given y ∈ I, Gyy = {y}. Then

∑g∈Gyy

f(g)ϕ(ug) = f(y, 0, y).

Thus {ϕy}y is a m-measurable �eld of states. Moreover, it follows from Lemma

6.4.5 that ϕy is a character for every y ∈ I.

210

Page 213: Characterization of Extremal KMS States on Groupoid C

Now we show that (m, {ϕy}y) de�nes a KMS state. Let y ∈ I, then its orbit is

aperiodic. Let h ∈ Gy. g = y is the unique element in Gyy . Then

ϕy(ug) = ϕy(uy) = 1, and

ϕr(h)(uhgh−1) = ϕr(h)(uhyh−1) = ϕr(h)(uhh−1) = ϕr(h)(ur(h)) = 1.

Then property (ii) of Theorem 5.3.10 and, therefore, φm is a KMS state. Since m

is extremal and ϕy is extremal for m-a.e. y, it follows from Proposition 6.4.4 that

φm is extremal.

• φmz

This case is analogous to the proof for m continuous if we de�ne I = O(z). Note

that mz is de�ned only if z is β-summable by Proposition 6.3.17.

• φλx

Let x ∈ X be periodic with minimum period p such that (6.17) holds. Then

the extremal eβF -conformal probability measure mx supported on O(x) exists by

Proposition 6.3.10. Let {ϕy}y∈X be a family os states such that ϕy = ϕλy for every

y ∈ O(x). Then {ϕy}y∈X de�nes a mx-measurable �eld of states. Moreover, ϕy is

a character for µ-a.e. y by Lemma 6.4.6.

Now we show that (mx, {ϕy}y) de�nes a KMS state, so we will show that property

(ii) of Theorem 5.3.10 holds. Let h ∈ Gx. There is l ∈ Z, such that h = (r(h), l, x).

Let g ∈ Gxx . Then there exists k ∈ Z such that g = (x, kp, x). Hence,

ϕx(ug) = ϕx(u(x,kp,x)) = λk, and

ϕr(h)(uhgh−1) = ϕr(h)(u(r(h),l,x)(x,kp,x)(x,−l,r(h))) = ϕr(h)(u(r(h),kp,r(h))) = λk.

Then (mx, {ϕy}y) de�nes a KMS state. Moreover, this state is extremal by Propo-

211

Page 214: Characterization of Extremal KMS States on Groupoid C

sition 6.4.4. Given f ∈ Cc(G), this state is de�ned by

∫X

∑g∈Gyy

f(g)ϕx(ug)dmx =

∫I

∑g∈Gyy

f(g)ϕλx(ug)dmx

=

∫I

∑k∈Z

f(y, kp, y)ϕλx(u(y,kp,y))dmx

=

∫I

∑k∈Z

f(y, kp, y)λkdmx

=

∫X

∑k∈Z

f(y, kp, y)λkdmx

= φλx(f).

Therefore the extremal KMS states on C*(G) are precisely φm, φmz and φλx.

Corollary 6.4.8. Let β ∈ R \ {0}. Assume F is positive on X. Then there exists a

correspondence between the extremal KMSβ-states and the extremal and continuous eβF -

conformal probability measures on X.

Proof. Let m be an extremal and continuous eβF -conformal probability measure on X. Then

φm is a KMSβ-state on C*(G) by Theorem 6.4.7.

Conversely, suppose φ is an extremal KMSβ-state. We will see that φ is not in cases 2

and 3 of theorem Theorem 6.4.7:

2. Suppose φ = φλx for some λ ∈ C with |λ| = 1 and x ∈ X periodic with minimum period

p > 0. Then

p−1∑j=0

F (σj(x)) = 0

Contradiction, since F assumes positive values.

3. Suppose φ = φmz for some z ∈ X aperiodic.

212

Page 215: Characterization of Extremal KMS States on Groupoid C

De�ne the sequence {yn}n∈N by yn = σn(z). Then

F(yn) = −n−1∑j=0

F (σj(z)) = −n−1∑j=0

F (yj).

Then, for each n, F(yn+1) < F(yn). Hence, e−βF(yn+1) > e−βF(yn). So,

∞∑n=0

e−βF(yn) =∞.

Then

∑y∈O(z)

e−βF(y) =∞.

Therefore z is not β-summable. Contradiction, since we assumed φ = φmz .

Therefore φ = φm for an extremal and continuous eβF -conformal Borel probability measure

on X.

213

Page 216: Characterization of Extremal KMS States on Groupoid C

Chapter 7

Concluding Remarks

In this thesis we described KMS states on groupoid C*-algebras for locally compact Hausdor�

second countable étale groupoids using Neshveyev's theorems. Then we studied a theorem

due to Thomsen which characterizes the extremal KMS states for the Renault-Deaconu

groupoid.

Neshveyev's theorem proved to be a useful tool to �nd an explicit formula for all KMS

states. However, the proof of this theorem depends on the fact that the groupoid is étale.

When the groupoid is not étale, it is possible to de�ne groupoid C*-algebras which are similar

to the C*-algebras studied in this thesis. For instance, [10] de�nes the crossed product of

a C*-algebra by a groupoid G where this groupoid is locally compact Hausdor� and is

endowed with a Haar system. This space is a closure of a space of continuous and compactly

supported functions f on X such that, for every x ∈ X, f(x) is an element of a C*-algebra.

The operations in this space are analogous to the operations in a full C*-algebra. So, one

challenge is to extend Neshveyev's theorem to non-étale groupoids.

Christensen [5] generalized the theorem of Neshveyev to describe KMS weights. This

result can be applied to describe KMS weights for di�erent groupoids.

214

Page 217: Characterization of Extremal KMS States on Groupoid C

Bibliography

[1] Joseph Bak and Donald J. Newman. Complex Analysis. Springer, 2010.

[2] Ola Bratteli and Derek W. Robinson. Operator Algebras and Quantum Statistical Me-

chanics: Volume 1: C*-and W*-Algebras. Symmetry Groups. Decomposition of States.

Springer-Verlag, 1979.

[3] Ola Bratteli and Derek W. Robinson. Operator Algebras and Quantum Statistical Me-

chanics. Vol. 2: Equilibrium states. Models in quantum statistical mechanics. Springer-

Verlag, 1997.

[4] Jonathan H. Brown. Proper Actions of Groupoids on C*-algebras. PhD thesis, Dart-

mouth College, Hanover, New Hampshire, May 2009.

[5] Johannes Christensen. KMS Weights on Groupoid C*-algebras: With an Emphasis on

Graph C*-algebras. PhD thesis, Aarhus University, Aarhus, Denmark, July 2018.

[6] Donald L. Cohn. Measure Theory. Birkhäuser, 1980.

[7] Claire Debord and Jean-Marie Lescure. Index theory and Groupoids. In Hernan

Ocampo, Eddy Pariguan, and Sylvie Paycha, editors, Geometric and Topological Meth-

ods for Quantum Field Theory, pages 86�158. Cambridge University Press, 2010.

[8] Manfred Denker and Mariusz Urba«ski. On the Existence of Conformal Measures.

Transactions of the American Mathematical Society, 328(2):563�587, 1991.

215

Page 218: Characterization of Extremal KMS States on Groupoid C

[9] Rodrigo S. Frausino. Groupoid C*-algebras, Conformal Measures and Phase Transi-

tions. Master's thesis, University of São Paulo, São Paulo, May 2018.

[10] Geo� Goehle. Groupoid Crossed Products. PhD thesis, Dartmouth College, Hanover,

New Hampshire, May 2009.

[11] Chaim S. Hönig. Aplicações da Topologia à Análise. Instituto de Matemática Pura e

Aplicada, CNPq, 1976.

[12] Roy A. Johnson. Atomic and Nonatomic Measures. Proceedings of the American Math-

ematical Society, 25(3):650�655, 1970.

[13] Alexander Kechris. Classical descriptive set theory. Springer-Verlag, 1995.

[14] Mauro S. de F. Marques. Teoria da medida. UNICAMP, 2009.

[15] Paul S. Muhly and Dana P. Williams. Renault's equivalence theorem for groupoid

crossed products. New York Journal of Mathematics, 3:1�87, 2008.

[16] Gerard J. Murphy. C*-Algebras and Operator Theory. Academic press, 1990.

[17] Sergey Neshveyev. KMS States on the C*-Algebras of Non-Principal Groupoids. Journal

of Operator Theory, 70(2):513�530, 2013.

[18] Ian F. Putnam. Lecture Notes on C*-Algebras, January 2016.

[19] Clive Reis. Abstract Algebra: An Introduction to Groups, Rings and Fields. World

Scienti�c Publishing Company, 2011.

[20] Jean Renault. A Groupoid Approach to C*-Algebras, volume 793. Springer-Verlag, 1980.

[21] Jadevilson C. Ribeiro and Rafael P. Lima. Funções holomorfas a valores em um C-

espaço vetorial topológico. https : / / www . ime . usp . br / ~cordaro / wp -

content/uploads/2017/09/Jadevilson-Rafael.pdf, 2017. Final project for a course

on Locally Convex Spaces.

216

Page 219: Characterization of Extremal KMS States on Groupoid C

[22] Walter Rudin. Functional Analysis. McGraw-Hill, 1973.

[23] Aidan Sims. Hausdor� Étale Groupoids and Their C*-Algebras. arXiv preprint

arXiv:1710.10897, 2017.

[24] Gerald Teschl. Topics in Real and Functional Analysis. unpublished, available online at

http: // www. mat. univie. ac. at/ ~gerald/ ftp/ book-fa/ fa. pdf , 1998.

[25] Klaus Thomsen. KMS States and Conformal Measures. Communications in Mathemat-

ical Physics, 316(3):615�640, 2012.

[26] Klaus Thomsen. Phase transition in O2. Communications in Mathematical Physics,

349(2):481�492, 2017.

[27] Dana P. Williams. Crossed Products of C*-Algebras. In Mathematical Surveys and

Monographs, volume 134. American Mathematical Society, 2007.

217