• 検索結果がありません。

-algebras from self-similar actions

N/A
N/A
Protected

Academic year: 2022

シェア "-algebras from self-similar actions"

Copied!
20
0
0

読み込み中.... (全文を見る)

全文

(1)

New York Journal of Mathematics

New York J. Math.27(2021) 923–942.

On groupoids and 𝑪

-algebras from self-similar actions

Valentin Deaconu

Abstract. Given a self-similar groupoid action(𝐺, 𝐸)on the path space of a finite graph, we study the associated Exel-Pardo étale groupoid𝒢(𝐺, 𝐸)and its𝐶-algebra𝐶(𝐺, 𝐸). We review some facts about groupoid actions, skew products and semi-direct products and generalize a result of Renault about similarity of groupoids which resembles Takai duality. We also describe a general strategy to compute the𝐾-theory of𝐶(𝐺, 𝐸)and the homology of 𝒢(𝐺, 𝐸)in certain cases and illustrate with an example.

Contents

1. Introduction 923

2. Homology of étale groupoids 924

3. Groupoid actions and similarity 926

4. Self-similar groupoid actions and their𝐶-algebra 931 5. Exel-Pardo groupoids for self-similar actions 935

6. Example 937

References 941

1. Introduction

Informally, a self-similar action is given by isomorphisms between parts of an object (think fractals or Julia sets) at different scales. Self-similar actions were studied intensely after exotic examples of groups acting on rooted trees and generated by finite automata, like infinite residually finite torsion groups, and groups of intermediate growth were constructed by Grigorchuk in the 1980’s.

Using the Pimsner construction from a𝐶-correspondence, Nekrashevych in- troduced the𝐶-algebras associated with self-similar group actions in [13,14], where important results about their structure and their𝐾-theory were obtained.

Received December 29, 2020.

2020Mathematics Subject Classification. Primary 46L05.

Key words and phrases. Homology of groupoids; self-similar action; similarity of groupoids;

Cuntz-Pimsner algebra; K-theory.

ISSN 1076-9803/2021

923

(2)

Motivated by the construction of all Kirchberg algebras in the UCT class us- ing topological graphs given by Katsura, in [7] Exel and Pardo introduced self- similar group actions on graphs and realized their𝐶-algebras as groupoid𝐶- algebras.

In this paper, we are interested in self-similar actions of groupoids𝐺on the path space of finite directed graphs𝐸as introduced and studied in [10], where the main goal was to find KMS states on some resulting dynamical systems. We generalize certain results of Exel and Pardo, in particular we define a groupoid 𝒢(𝐺, 𝐸)and discuss the structure of the𝐶-algebra𝐶(𝐺, 𝐸), defined as a Cuntz- Pimsner algebra of a𝐶-correspondence over𝐶(𝐺). The𝐶-algebra𝐶(𝐺, 𝐸) has a natural gauge action and contains copies of𝐶(𝐸)and𝐶(𝐺). In general, its structure is rather intricate; in a particular case,𝐶(𝐺, 𝐸) ≅ 𝐶(𝐸) ⋊ 𝐺.

We begin with a review of étale groupoid homology, then we recall some facts about groupoid actions, skew products and semi-direct products and generalize a result of Renault about similarity of groupoids in the spirit of Takai duality.

We also describe a general strategy to compute the𝐾-theory of𝐶(𝐺, 𝐸)and the homology of𝒢(𝐺, 𝐸)in certain cases. We illustrate with an example.

We expect that many of our results will be true for self-similar actions of groupoids on the path space of infinite graphs. For the case when𝐺is a group, see [8,11].

2. Homology of étale groupoids

A groupoid𝐺is a small category with inverses. The set of objects is denoted by𝐺(0). We will use𝑑and𝑡for the domain and terminus maps𝑑, 𝑡 ∶ 𝐺 → 𝐺(0) to distinguish them from the range and source maps𝑟, 𝑠on directed graphs. For 𝑢, 𝑣 ∈ 𝐺(0), we write

𝐺𝑢= {𝑔 ∈ 𝐺 ∶ 𝑑(𝑔) = 𝑢}, 𝐺𝑣 = {𝑔 ∈ 𝐺 ∶ 𝑡(𝑔) = 𝑣}, 𝐺𝑢𝑣 = 𝐺𝑢∩ 𝐺𝑣. The set of composable pairs is denoted𝐺(2).

An étale groupoid is a topological groupoid where the terminus map𝑡(and necessarily the domain map𝑑) is a local homeomorphism (as a map from𝐺to 𝐺). The unit space𝐺(0)of an étale groupoid is always an open subset of𝐺. Definition 2.1. Let𝐺be an étale groupoid. A bisection is an open subset𝑈 ⊆ 𝐺such that𝑑and𝑡are both injective when restricted to𝑈.

Two units𝑥, 𝑦 ∈ 𝐺(0)belong to the same𝐺-orbit if there exists𝑔 ∈ 𝐺such that𝑑(𝑔) = 𝑥and𝑡(𝑔) = 𝑦. We denote by orb𝐺(𝑥)the𝐺-orbit of𝑥. When every 𝐺-orbit is dense in𝐺(0), the groupoid𝐺is called minimal. An open set𝑉 ⊆ 𝐺(0) is called𝐺-full if for every𝑥 ∈ 𝐺(0)we have orb𝐺(𝑥) ∩ 𝑉 ≠ ∅. We denote by𝐺𝑉 the subgroupoid{𝑔 ∈ 𝐺 | 𝑑(𝑔), 𝑡(𝑔) ∈ 𝑉}, called the restriction of𝐺to𝑉. When 𝐺is étale, the restriction𝐺𝑉 is an open étale subgroupoid with unit space𝑉.

The isotropy group of a unit𝑥 ∈ 𝐺(0)is the group 𝐺𝑥𝑥∶= {𝑔 ∈ 𝐺 | 𝑑(𝑔) = 𝑡(𝑔) = 𝑥},

(3)

and the isotropy bundle is

𝐺∶= {𝑔 ∈ 𝐺 | 𝑑(𝑔) = 𝑡(𝑔)} =

𝑥∈𝐺(0)

𝐺𝑥𝑥.

A groupoid𝐺is said to be principal if all isotropy groups are trivial, or equiva- lently,𝐺= 𝐺(0). We say that𝐺is effective if the interior of𝐺equals𝐺(0). Definition 2.2. A second countable groupoid𝐺is ample if it is étale and𝐺(0) is zero-dimensional; equivalently,𝐺 is ample if it has a basis of compact open bisections. An ample groupoid𝐺is elementary if it is compact and principal.

An ample groupoid is an𝐴𝐹groupoid if there exists an ascending chain of open elementary subgroupoids𝐺1⊆ 𝐺2⊆ ... ⊆ 𝐺such that𝐺 =⋃

𝑖=1𝐺𝑖.

We recall now the definion of homology of étale groupoids which was in- troduced by Crainic and Moerdijk in [3]. Let𝐴be an Abelian group and let 𝜋 ∶ 𝑋 → 𝑌be a local homeomorphism between two locally compact Haus- dorff spaces. Given any𝑓 ∈ 𝐶𝑐(𝑋, 𝐴), we define

𝜋(𝑓)(𝑦) ∶= ∑

𝜋(𝑥)=𝑦

𝑓(𝑥).

It follows that𝜋(𝑓) ∈ 𝐶𝑐(𝑌, 𝐴). Given an étale groupoid𝐺, let𝐺(1) = 𝐺and for𝑛 ≥ 2let𝐺(𝑛)be the space of composable strings of𝑛elements in𝐺with the product topology. For𝑛 ≥ 2and𝑖 = 0, ..., 𝑛, we let𝜕𝑖 ∶ 𝐺(𝑛) → 𝐺(𝑛−1)be the face maps defined by

𝜕𝑖(𝑔1, 𝑔2, ..., 𝑔𝑛) =

(𝑔2, 𝑔3, ..., 𝑔𝑛) if𝑖 = 0,

(𝑔1, ..., 𝑔𝑖𝑔𝑖+1, ..., 𝑔𝑛) if1 ≤ 𝑖 ≤ 𝑛 − 1, (𝑔1, 𝑔2, ..., 𝑔𝑛−1) if𝑖 = 𝑛.

We define the homomorphisms𝛿𝑛∶ 𝐶𝑐(𝐺(𝑛), 𝐴) → 𝐶𝑐(𝐺(𝑛−1), 𝐴)given by 𝛿1= 𝑑− 𝑡, 𝛿𝑛=

𝑛 𝑖=0

(−1)𝑖𝜕𝑖∗for𝑛 ≥ 2.

It can be verified that𝛿𝑛◦𝛿𝑛+1= 0for all𝑛 ≥ 1.

The homology groups𝐻𝑛(𝐺, 𝐴)are by definition the homology groups of the chain complex𝐶𝑐(𝐺(∗), 𝐴)given by

0⟵ 𝐶𝛿0 𝑐(𝐺(0), 𝐴)⟵ 𝐶𝛿1 𝑐(𝐺(1), 𝐴)⟵ 𝐶𝛿2 𝑐(𝐺(2), 𝐴) ⟵ ⋯ ,

i.e. 𝐻𝑛(𝐺, 𝐴) = ker 𝛿𝑛∕im𝛿𝑛+1, where𝛿0 = 0. If𝐴 = ℤ, we write𝐻𝑛(𝐺)for 𝐻𝑛(𝐺, ℤ).

The following HK-conjecture of Matui states that the homology of an étale groupoid refines the𝐾-theory of the reduced groupoid𝐶-algebra. Let𝐺 be a minimal effective ample Hausdorff groupoid with compact unit space. Then

(4)

𝐾𝑖(𝐶𝑟(𝐺)) ≅

𝑛=0

𝐻2𝑛+𝑖(𝐺), for𝑖 = 0, 1.

Recently, this conjecture was the source of intense research. It was con- firmed for several groupoids like𝐴𝐹-groupoids, transformation groupoids of Cantor minimal systems, groupoids of shifts of finite type and products of group- oids of shifts of finite type (see [12]). The homology of ample Hausdorff group- oids was investigated in [9], with emphasis on the Renault-Deaconu groupoids associated to𝑘pairwise-commuting local homeomorphisms of a zero dimen- sional space. It was shown that the homology of𝑘-graph groupoids can be com- puted in terms of the adjacency matrices, using spectral sequences and a chain complex developed by Evans in [6]. The HK-conjecture was also confirmed for groupoids on one-dimensional solenoids in [20]. Recently, counterexamples to the HK-conjecture of Matui were found by Scarparo in [18] and by Ortega and Sanchez in [16].

3. Groupoid actions and similarity

We recall the concept of a groupoid action on another groupoid from [1], page 122 and from [5].

Definition 3.1. A topological groupoid𝐺acts (on the right) on another topo- logical groupoid𝐻if there are a continuous open surjection𝑝 ∶ 𝐻 → 𝐺(0)and a continuous map𝐻 ∗ 𝐺 → 𝐻, write(ℎ, 𝑔) ↦ ℎ ⋅ 𝑔where

𝐻 ∗ 𝐺 = {(ℎ, 𝑔) ∈ 𝐻 × 𝐺 ∣ 𝑡(𝑔) = 𝑝(ℎ)}

such that

i)𝑝(ℎ ⋅ 𝑔) = 𝑑(𝑔)for all(ℎ, 𝑔) ∈ 𝐻 ∗ 𝐺,

ii)(ℎ, 𝑔1) ∈ 𝐻 ∗ 𝐺and(𝑔1, 𝑔2) ∈ 𝐺(2)implies that(ℎ, 𝑔1𝑔2) ∈ 𝐻 ∗ 𝐺and ℎ ⋅ (𝑔1𝑔2) = (ℎ ⋅ 𝑔1) ⋅ 𝑔2,

iii)(ℎ1, ℎ2) ∈ 𝐻(2) and(ℎ12, 𝑔) ∈ 𝐻 ∗ 𝐺implies(ℎ1, 𝑔), (ℎ2, 𝑔) ∈ 𝐻 ∗ 𝐺 and

(ℎ12) ⋅ 𝑔 = (ℎ1⋅ 𝑔)(ℎ2⋅ 𝑔), iv)ℎ ⋅ 𝑝(ℎ) = ℎfor allℎ ∈ 𝐻.

The action is called free ifℎ ⋅ 𝑔 = ℎimplies𝑔 = 𝑝(ℎ)and transitive if for all ℎ1, ℎ2∈ 𝐻there is𝑔 ∈ 𝐺withℎ2= ℎ1⋅ 𝑔.

Note that if𝐺acts on𝐻 on the right, we can define a left action of𝐺on𝐻 by taking𝑔 ⋅ ℎ ∶= ℎ ⋅ 𝑔−1and vice versa.

Example3.2. Given a topological groupoid𝐺, a𝐺-module in [19] is a topolog- ical groupoid𝐴with domain and terminus maps equal to𝑝 ∶ 𝐴 → 𝐺(0)such that𝐴𝑥𝑥is an abelian group for all𝑥 ∈ 𝐺(0),𝐺acts on𝐴as a space and such that for each𝑔 ∈ 𝐺 the action map𝛼𝑔 ∶ 𝐴𝑑(𝑔) → 𝐴𝑡(𝑔) is a group homomorphism.

(5)

In particular,𝐴can be a trivial group bundle𝐺(0)× 𝐷 for𝐷 an abelian group and𝛼𝑔= 𝑖𝑑𝐷for all𝑔 ∈ 𝐺.

Remark3.3. If𝐺 acts on the groupoid𝐻, then𝐺 acts on the unit space𝐻(0) using the restriction𝑝0∶= 𝑝|𝐻(0) ∶ 𝐻(0)→ 𝐺(0)and we have𝑝 = 𝑝0◦𝑡 = 𝑝0◦𝑑, where𝑑, 𝑡 ∶ 𝐻 → 𝐻(0). In particular,

𝑝(ℎ−1) = 𝑝0(𝑡(ℎ−1)) = 𝑝0(𝑑(ℎ)) = 𝑝(ℎ).

Using the fact thatℎ = ℎ𝑑(ℎ) = 𝑡(ℎ)ℎ, it follows that ℎ ⋅ 𝑔 = (ℎ ⋅ 𝑔)(𝑑(ℎ) ⋅ 𝑔) = (𝑡(ℎ) ⋅ 𝑔)(ℎ ⋅ 𝑔), so we deduce that𝑑(ℎ ⋅ 𝑔) = 𝑑(ℎ) ⋅ 𝑔and𝑡(ℎ ⋅ 𝑔) = 𝑡(ℎ) ⋅ 𝑔.

Definition 3.4. If𝐺acts on𝐻, then the semi-direct product groupoid𝐻 ⋊ 𝐺, also called the action groupoid, is defined as follows. As a set,

𝐻 ⋊ 𝐺 = 𝐻 ∗ 𝐺 = {(ℎ, 𝑔) ∈ 𝐻 × 𝐺 ∣ 𝑡(𝑔) = 𝑝(ℎ)}

and the multiplication is given by

(ℎ, 𝑔)(ℎ⋅ 𝑔, 𝑔) = (ℎℎ, 𝑔𝑔), when𝑡(𝑔) = 𝑑(𝑔)and𝑑(ℎ) = 𝑡(ℎ).

In a semi-direct product, the inverse is given by (ℎ, 𝑔)−1= (ℎ−1⋅ 𝑔, 𝑔−1) and we get

(ℎ, 𝑔)−1(ℎ, 𝑔) = (ℎ−1⋅ 𝑔, 𝑔−1)(ℎ, 𝑔) = ((ℎ−1⋅ 𝑔)(ℎ ⋅ 𝑔), 𝑑(𝑔)) = (𝑑(ℎ) ⋅ 𝑔, 𝑑(𝑔)), (ℎ, 𝑔)(ℎ, 𝑔)−1= (ℎ, 𝑔)(ℎ−1⋅ 𝑔, 𝑔−1) = (𝑡(ℎ), 𝑡(𝑔)).

Since𝑑(𝑔) = 𝑝(𝑑(ℎ) ⋅ 𝑔)and𝑡(𝑔) = 𝑝(𝑡(ℎ)), the unit space of𝐻 ⋊ 𝐺can be identified with𝐻(0)and then we make identifications

𝑑(ℎ, 𝑔) ≡ 𝑑(ℎ) ⋅ 𝑔, 𝑡(𝑔, ℎ) ≡ 𝑡(ℎ).

There is a groupoid homomorphism

𝜋 ∶ 𝐻 ⋊ 𝐺 → 𝐺, 𝜋(ℎ, 𝑔) = 𝑔

with kernel𝜋−1(𝐺(0)) = {(ℎ, 𝑝(ℎ)) ∣ ℎ ∈ 𝐻}isomorphic to𝐻.

Remark3.5. The notion of groupoid action on another groupoid includes the action of a groupoid on a space and the action of a group on another group by automorphisms. A particular situation is when𝐺1, 𝐺2are groupoids and𝑍 is a (𝐺1, 𝐺2)-space, i.e. 𝑍 is a left𝐺1-space, a right𝐺2-space and the actions commute. Then𝐺1⋉ 𝑍is a right𝐺2-groupoid and𝑍 ⋊ 𝐺2is a left𝐺1-groupoid.

(6)

Definition 3.6. Suppose now that𝐺, Γare étale groupoids and that𝜌 ∶ 𝐺 → Γ is a groupoid homomorphism, also called a cocycle. The skew product groupoid 𝐺 ×𝜌Γis defined as the set of pairs(𝑔, 𝛾) ∈ 𝐺 × Γsuch that(𝛾, 𝜌(𝑔)) ∈ Γ(2)with multiplication

(𝑔, 𝛾)(𝑔, 𝛾𝜌(𝑔)) = (𝑔𝑔, 𝛾)if(𝑔, 𝑔) ∈ 𝐺(2) and inverse

(𝑔, 𝛾)−1= (𝑔−1, 𝛾𝜌(𝑔)).

In a skew product, we have𝑑(𝑔, 𝛾) = (𝑑(𝑔), 𝛾𝜌(𝑔))and𝑡(𝑔, 𝛾) = (𝑡(𝑔), 𝛾). Its unit space is

𝐺(0) ∗ Γ = {(𝑢, 𝛾) ∈ 𝐺(0)× Γ ∶ 𝜌(𝑢) = 𝑑(𝛾)}.

In particular, if𝐺, 𝐻 are étale groupoids and 𝐺acts on𝐻 on the right, for the groupoid homomorphism𝜋 ∶ 𝐻 ⋊ 𝐺 → 𝐺, 𝜋(ℎ, 𝑔) = 𝑔we can form the skew product(𝐻 ⋊ 𝐺) ×𝜋𝐺 made of triples(ℎ, 𝑔, 𝑔) ∈ 𝐻 × 𝐺 × 𝐺such that 𝑝(ℎ) = 𝑡(𝑔)and(𝑔, 𝑔) ∈ 𝐺(2), with unit space𝐻(0)∗ 𝐺and operations

(ℎ, 𝑔, 𝑔)(ℎ, 𝑔′′, 𝑔𝑔) = (ℎ(ℎ⋅ 𝑔−1), 𝑔𝑔′′, 𝑔), (ℎ, 𝑔, 𝑔)−1= (ℎ−1⋅ 𝑔, 𝑔−1, 𝑔𝑔).

Remark3.7. Given a groupoid homomorphism𝜌 ∶ 𝐺 → Γ, there is a left action

̂

𝜌ofΓon the skew product𝐺 ×𝜌Γgiven by 𝛾⋅ (𝑔, 𝛾) = (𝑔, 𝛾𝛾).

Proof. We check all the properties defining a groupoid action. First, we define the continuous open map

𝑝 ∶ 𝐺 ×𝜌Γ → Γ(0), 𝑝(𝑔, 𝛾) = 𝑡(𝛾)

and note that𝑑(𝛾) = 𝑡(𝛾) = 𝑝(𝑔, 𝛾). Now,𝑝(𝑔, 𝛾𝛾) = 𝑡(𝛾)and if(𝛾1, 𝛾2) ∈ Γ(2), then

(𝛾1𝛾2) ⋅ (𝑔, 𝛾) = (𝑔, 𝛾1𝛾2𝛾) = 𝛾1⋅ (𝛾2⋅ (𝑔, 𝛾)).

Also, if(𝑔, 𝛾), (𝑔, 𝛾𝜌(𝑔)) ∈ 𝐺 ×𝜌Γare composable with product(𝑔𝑔, 𝛾), then 𝛾⋅ (𝑔𝑔, 𝛾) = (𝑔𝑔, 𝛾𝛾) = (𝑔, 𝛾𝛾)(𝑔, 𝛾𝛾𝜌(𝑔)) = (𝛾⋅ (𝑔, 𝛾))(𝛾⋅ (𝑔, 𝛾𝜌(𝑔))).

The last condition to check is𝑡(𝛾) ⋅ (𝑔, 𝛾) = (𝑔, 𝛾), which is obvious.

We define a right action ofΓon𝐺 ×𝜌Γby(𝑔, 𝛾) ⋅ 𝛾= (𝑔, 𝛾′−1𝛾), and we form the semi-direct product(𝐺 ×𝜌Γ) ⋊ Γmade of triples(𝑔, 𝛾, 𝛾) ∈ 𝐺 × Γ × Γsuch that(𝛾, 𝜌(𝑔)) ∈ Γ(2) and𝑡(𝛾) = 𝑝(𝑔, 𝛾) = 𝑡(𝛾), with unit space𝐺(0) ∗ Γand operations

(𝑔, 𝛾, 𝛾)(𝑔, 𝛾′−1𝛾𝜌(𝑔), 𝛾′′) = (𝑔𝑔, 𝛾, 𝛾𝛾′′),

(𝑔, 𝛾, 𝛾)−1= ((𝑔−1, 𝛾𝜌(𝑔)) ⋅ 𝛾, 𝛾′−1) = (𝑔−1, 𝛾′−1𝛾𝜌(𝑔), 𝛾′−1).

For the next result, which resembles Takai duality, see Definition 1.3 in [17], Proposition 3.7 in [12] and Definition 3.1 in [9].

(7)

Theorem 3.8. Let𝐺, 𝐻, Γ be étale groupoids such that𝐺 acts on𝐻 and such that𝜌 ∶ 𝐺 → Γis a groupoid homomorphism. Then, using the above notation, (𝐻 ⋊ 𝐺) ×𝜋𝐺is similar to𝐻and(𝐺 ×𝜌Γ) ⋊ Γis similar to𝐺.

Proof. Recall that two (continuous) groupoid homomorphisms𝜙1, 𝜙2∶ 𝐺1→ 𝐺2are similar if there is a continuous function𝜃 ∶ 𝐺1(0)→ 𝐺2such that

𝜃(𝑡(𝑔))𝜙1(𝑔) = 𝜙2(𝑔)𝜃(𝑑(𝑔))

for all𝑔 ∈ 𝐺1. Two topological groupoids𝐺1, 𝐺2are similar if there exist con- tinuous homomorphisms𝜙 ∶ 𝐺1 → 𝐺2 and𝜓 ∶ 𝐺2 → 𝐺1 such that 𝜓◦𝜙is similar to𝑖𝑑𝐺1 and𝜙◦𝜓is similar to𝑖𝑑𝐺2.

To show that(𝐻 ⋊ 𝐺) ×𝜋𝐺is similar to𝐻, we define 𝜙 ∶ (𝐻 ⋊ 𝐺) ×𝜋𝐺 → 𝐻, 𝜙(ℎ, 𝑔, 𝑔) = ℎ ⋅ 𝑔′−1, 𝜓 ∶ 𝐻 → (𝐻 ⋊ 𝐺) ×𝜋𝐺, 𝜓(ℎ) = (ℎ, 𝑝(ℎ), 𝑝(ℎ)) and

𝜃 ∶ 𝐻(0) ∗ 𝐺 → (𝐻 ⋊ 𝐺) ×𝜋𝐺, 𝜃(𝑢, 𝑔) = (𝑢 ⋅ 𝑔−1, 𝑔, 𝑝(𝑢)).

We check that

(𝜙◦𝜓)(ℎ) = 𝜙(ℎ, 𝑝(ℎ), 𝑝(ℎ)) = ℎ ⋅ 𝑝(ℎ)−1= ℎ and that

(∗) 𝜃[𝑡(ℎ, 𝑔, 𝑔)](ℎ, 𝑔, 𝑔) = (𝜓◦𝜙)(ℎ, 𝑔, 𝑔)𝜃[𝑑(ℎ, 𝑔, 𝑔)].

We have

𝑡(ℎ, 𝑔, 𝑔) = (ℎ, 𝑔, 𝑔)(ℎ, 𝑔, 𝑔)−1= (ℎ, 𝑔, 𝑔)(ℎ−1⋅ 𝑔, 𝑔−1, 𝑔𝑔) =

= (ℎ(ℎ−1⋅ 𝑔 ⋅ 𝑔−1), 𝑔𝑔−1, 𝑔) = (𝑡(ℎ), 𝑡(𝑔), 𝑔) ≡ (𝑡(ℎ), 𝑔), and

𝜃(𝑡(ℎ), 𝑔) = (𝑡(ℎ) ⋅ 𝑔′−1, 𝑔, 𝑝(𝑡(ℎ)).

The left-hand side of(∗)becomes

(𝑡(ℎ) ⋅ 𝑔′−1, 𝑔, 𝑝(𝑡(ℎ))(ℎ, 𝑔, 𝑔) = (ℎ ⋅ 𝑔′−1, 𝑔𝑔, 𝑝(𝑡(ℎ)) = (ℎ ⋅ 𝑔′−1, 𝑔𝑔, 𝑡(𝑔)).

Now,

(𝜓◦𝜙)(ℎ, 𝑔, 𝑔) = 𝜓(ℎ⋅𝑔′−1) = (ℎ⋅𝑔′−1, 𝑝(ℎ⋅𝑔′−1), 𝑝(ℎ⋅𝑔′−1) = (ℎ⋅𝑔′−1, 𝑡(𝑔), 𝑡(𝑔)), 𝑑(ℎ, 𝑔, 𝑔) = (ℎ, 𝑔, 𝑔)−1(ℎ, 𝑔, 𝑔) = (ℎ−1⋅ 𝑔, 𝑔−1, 𝑔𝑔)(ℎ, 𝑔, 𝑔) =

= ((ℎ−1⋅ 𝑔)(ℎ ⋅ 𝑔), 𝑔−1𝑔, 𝑔𝑔) = (𝑑(ℎ) ⋅ 𝑔, 𝑑(𝑔), 𝑔𝑔) ≡ (𝑑(ℎ) ⋅ 𝑔, 𝑔𝑔), and

𝜃(𝑑(ℎ) ⋅ 𝑔, 𝑔𝑔) = (𝑑(ℎ) ⋅ 𝑔 ⋅ (𝑔𝑔)−1, 𝑔𝑔), 𝑝(𝑑(ℎ) ⋅ 𝑔)) = (𝑑(ℎ) ⋅ 𝑔′−1, 𝑔𝑔), 𝑑(𝑔)).

The right-hand side becomes

(ℎ ⋅ 𝑔′−1, 𝑡(𝑔), 𝑡(𝑔))(𝑑(ℎ) ⋅ 𝑔′−1, 𝑔𝑔, 𝑑(𝑔)) = (ℎ ⋅ 𝑔′−1, 𝑔𝑔, 𝑡(𝑔)), so(∗)is verified.

(8)

To show that(𝐺 ×𝜌Γ) ⋊ Γis similar to𝐺, we define 𝜙 ∶ (𝐺 ×𝜌Γ) ⋊ Γ → 𝐺, 𝜙(𝑔, 𝛾, 𝛾) = 𝑔, 𝜓 ∶ 𝐺 → (𝐺 ×𝜌Γ) ⋊ Γ, 𝜓(𝑔) = (𝑔, 𝜌(𝑡(𝑔)), 𝜌(𝑔)) and

𝜃 ∶ 𝐺(0) ∗ Γ → (𝐺 ×𝜌Γ) ⋊ Γ, 𝜃(𝑢, 𝛾) = (𝑢, 𝜌(𝑢), 𝛾−1).

We have

(𝜙◦𝜓)(𝑔) = 𝜙(𝑔, 𝜌(𝑡(𝑔)), 𝜌(𝑔)) = 𝑔 and we need to verify that

(∗∗) 𝜃(𝑡(𝑔, 𝛾, 𝛾))(𝑔, 𝛾, 𝛾) = (𝜓◦𝜙)(𝑔, 𝛾, 𝛾)𝜃(𝑑(𝑔, 𝛾, 𝛾)).

We compute

𝑡(𝑔, 𝛾, 𝛾) = (𝑔, 𝛾, 𝛾)(𝑔, 𝛾, 𝛾)−1= (𝑔, 𝛾, 𝛾)(𝑔−1, 𝛾′−1𝛾𝜌(𝑔), 𝛾′−1) =

= (𝑡(𝑔), 𝛾, 𝑡(𝛾)) ≡ (𝑡(𝑔), 𝛾) and

𝜃(𝑡(𝑔), 𝛾) = (𝑡(𝑔), 𝜌(𝑡(𝑔), 𝛾−1), so the left-hand side of(∗∗)becomes

(𝑡(𝑔), 𝜌(𝑡(𝑔)), 𝛾−1)(𝑔, 𝛾, 𝛾) = (𝑔, 𝜌(𝑡(𝑔)), 𝛾−1𝛾).

Also

(𝜓◦𝜙)(𝑔, 𝛾, 𝛾) = (𝑔, 𝜌(𝑡(𝑔)), 𝜌(𝑔)),

𝑑(𝑔, 𝛾, 𝛾) = (𝑔, 𝛾, 𝛾)−1(𝑔, 𝛾, 𝛾) = (𝑔−1, 𝛾′−1𝛾𝜌(𝑔), 𝛾′−1)(𝑔, 𝛾, 𝛾) =

= (𝑑(𝑔), 𝛾′−1𝛾𝜌(𝑔), 𝑑(𝛾)) ≡ (𝑑(𝑔), 𝛾′−1𝛾𝜌(𝑔)), 𝜃(𝑑(𝑔), 𝛾′−1𝛾𝜌(𝑔)) = (𝑑(𝑔), 𝜌(𝑑(𝑔), 𝜌(𝑔)−1𝛾−1𝛾), and the right-hand side is

(𝑔, 𝜌(𝑡(𝑔)), 𝜌(𝑔))(𝑑(𝑔), 𝜌(𝑑(𝑔)), 𝜌(𝑔)−1𝛾−1𝛾) = (𝑔, 𝜌(𝑡(𝑔)), 𝛾−1𝛾),

so(∗∗)is verified.

For skew products and semi-direct products of groupoids, there are Lyndon–

Hochschild–Serre spectral sequences for the computation of their homology, see [3] and [12].

Theorem 3.9. Let𝐺, Γbe étale groupoids.

(1) Suppose that𝜌 ∶ 𝐺 → Γis a groupoid homomorphism. Then there exists a spectral sequence

𝐸2𝑝,𝑞= 𝐻𝑝(Γ, 𝐻𝑞(𝐺 ×𝜌Γ)) ⇒ 𝐻𝑝+𝑞(𝐺),

where𝐻𝑞(𝐺 ×𝜌Γ)is regarded as aΓ-module via the action𝜌 ∶ Γ ↷ 𝐺 ×̂ 𝜌Γ.

(2) Suppose that𝜑 ∶ Γ ↷ 𝐺is a groupoid action. Then there exists a spectral sequence

𝐸𝑝,𝑞2 = 𝐻𝑝(Γ, 𝐻𝑞(𝐺)) ⇒ 𝐻𝑝+𝑞(Γ ⋉ 𝐺), where𝐻𝑞(𝐺)is regarded as aΓ-module via the action𝜑.

(9)

Note that for𝐺an ample Hausdorff groupoid and𝜌 ∶ 𝐺 → ℤa cocycle, we have the following long exact sequence involving the homology of𝐺 and the homology of𝐺 ×𝜌ℤ,

0 ⟵ 𝐻0(𝐺) ⟵ 𝐻0(𝐺 ×𝜌ℤ)𝑖𝑑−𝜌⟵ 𝐻 0(𝐺 ×𝜌ℤ) ⟵ 𝐻1(𝐺) ⟵ ⋯

⋯ ⟵ 𝐻𝑛(𝐺) ⟵ 𝐻𝑛(𝐺 ×𝜌ℤ)𝑖𝑑−𝜌⟵ 𝐻 𝑛(𝐺 ×𝜌ℤ) ⟵ 𝐻𝑛+1(𝐺) ⟵ ⋯ Here𝜌is the map induced by the action𝜌 ∶ ℤ ↷ 𝐺 ×̂ 𝜌ℤ(see Lemma 1.3 in [15]).

4. Self-similar groupoid actions and their𝑪-algebra

We recall some facts about self-similar groupoid actions and their Cuntz- Pimsner algebras from [10]. Let 𝐸 = (𝐸0, 𝐸1, 𝑟, 𝑠) be a finite directed graph with no sources. For𝑘 ≥ 2, define the set of paths of length𝑘in𝐸as

𝐸𝑘 = {𝑒1𝑒2⋯ 𝑒𝑘 ∶ 𝑒𝑖 ∈ 𝐸1, 𝑟(𝑒𝑖+1) = 𝑠(𝑒𝑖)}.

The maps𝑟, 𝑠are naturally extended to𝐸𝑘by taking

𝑟(𝑒1𝑒2⋯ 𝑒𝑘) = 𝑟(𝑒1), 𝑠(𝑒1𝑒2⋯ 𝑒𝑘) = 𝑠(𝑒𝑘).

We denote by𝐸 ∶=⋃

𝑘≥0𝐸𝑘 the space of finite paths (including vertices) and by𝐸the infinite path space of𝐸with the usual topology given by the cylinder sets𝑍(𝛼) = {𝛼𝜉 ∶ 𝜉 ∈ 𝐸}for𝛼 ∈ 𝐸.

We can visualize the set𝐸as indexing the vertices of a union of rooted trees or forest𝑇𝐸 given by𝑇𝐸0 = 𝐸and with edges

𝑇1𝐸 = {(𝜇, 𝜇𝑒) ∶ 𝜇 ∈ 𝐸, 𝑒 ∈ 𝐸1and𝑠(𝜇) = 𝑟(𝑒)}.

Example4.1. For the graph

𝑢 𝑣 𝑤

𝑒2

𝑒3

𝑒4

𝑒5 𝑒6 𝑒1

the forest𝑇𝐸 looks like

𝑢 𝑣 𝑤

𝑒1 𝑒3

𝑒1𝑒1

𝑒1𝑒3

𝑒3𝑒2

𝑒3𝑒6

𝑒2 𝑒6

𝑒2𝑒1

𝑒2𝑒3

𝑒6𝑒4

𝑒6𝑒5

𝑒4 𝑒5

𝑒4𝑒2

𝑒4𝑒6

𝑒5𝑒2

𝑒5𝑒6

(10)

Recall that a partial isomorphism of the forest𝑇𝐸 corresponding to a given directed graph𝐸consists of a pair(𝑣, 𝑤) ∈ 𝐸0× 𝐸0and a bijection𝑔 ∶ 𝑣𝐸 → 𝑤𝐸such that

∙ 𝑔|𝑣𝐸𝑘 ∶ 𝑣𝐸𝑘 → 𝑤𝐸𝑘 is bijective for all𝑘 ≥ 1.

∙ 𝑔(𝜇𝑒) ∈ 𝑔(𝜇)𝐸1for𝜇 ∈ 𝑣𝐸and𝑒 ∈ 𝐸1with𝑟(𝑒) = 𝑠(𝜇).

The set of partial isomorphisms of𝑇𝐸 forms a groupoid PIso(𝑇𝐸)with unit space𝐸0. The identity morphisms are𝑖𝑑𝑣 ∶ 𝑣𝐸 → 𝑣𝐸, the inverse of𝑔 ∶ 𝑣𝐸 → 𝑤𝐸is𝑔−1 ∶ 𝑤𝐸 → 𝑣𝐸, and the multiplication is composition. We often identify𝑣 ∈ 𝐸0with𝑖𝑑𝑣∈PIso(𝑇𝐸).

Definition 4.2. Let𝐸be a finite directed graph with no sources, and let𝐺be a groupoid with unit space𝐸0. Aself-similar action(𝐺, 𝐸)on the path space of 𝐸is given by a faithful groupoid homomorphism𝐺 →PIso(𝑇𝐸)such that for every𝑔 ∈ 𝐺and every𝑒 ∈ 𝑑(𝑔)𝐸1there exists a uniqueℎ ∈ 𝐺denoted by𝑔|𝑒 and called the restriction of𝑔to𝑒such that

𝑔 ⋅ (𝑒𝜇) = (𝑔 ⋅ 𝑒)(ℎ ⋅ 𝜇) for all 𝜇 ∈ 𝑠(𝑒)𝐸.

Remark4.3. It is possible that𝑔|𝑒 = 𝑔for all𝑒 ∈ 𝑑(𝑔)𝐸1, in which case 𝑔 ⋅ (𝑒1𝑒2⋯ 𝑒𝑛) = (𝑔 ⋅ 𝑒1) ⋯ (𝑔 ⋅ 𝑒𝑛).

We have

𝑑(𝑔|𝑒) = 𝑠(𝑒), 𝑡(𝑔|𝑒) = 𝑠(𝑔 ⋅ 𝑒) = 𝑔|𝑒⋅ 𝑠(𝑒), 𝑟(𝑔 ⋅ 𝑒) = 𝑔 ⋅ 𝑟(𝑒).

In particular, the source map may not be equivariant as in [7]. It is shown in Appendix A of [10] that a self-similar group action(𝐺, 𝐸)as in [7] determines a self-similar groupoid action(𝐸0⋊ 𝐺, 𝐸) as in Definition4.2, where𝐸0⋊ 𝐺 is the semi-direct product or the action groupoid of the group𝐺acting on𝐸0. Note that not any self-similar groupoid action comes from a self-similar group action, as seen in our example below.

Proposition 4.4. A self-similar groupoid action (𝐺, 𝐸)as above extends to an action of𝐺on the path space𝐸and determines an action of𝐺on the graph𝑇𝐸, in the sense that𝐺 acts on both the vertex space𝑇𝐸0 and the edge space𝑇1𝐸 and intertwines the range and the source maps of𝑇𝐸, see Definition 4.1 in[5].

Proof. Indeed, the vertex space𝑇0𝐸 = 𝐸is fibered over𝐺(0) = 𝐸0via the map 𝜇 ↦ 𝑟(𝜇). For (𝜇, 𝜇𝑒) ∈ 𝑇𝐸1 we set 𝑠(𝜇, 𝜇𝑒) = 𝜇𝑒 and 𝑟(𝜇, 𝜇𝑒) = 𝜇. Since 𝑟(𝜇𝑒) = 𝑟(𝜇), the edge space𝑇𝐸1 is also fibered over𝐺(0). The action of𝐺on𝑇𝐸1 is given by

𝑔 ⋅ (𝜇, 𝜇𝑒) = (𝑔 ⋅ 𝜇, 𝑔 ⋅ (𝜇𝑒)) when 𝑑(𝑔) = 𝑟(𝜇).

Since

𝑠(𝑔 ⋅ (𝜇, 𝜇𝑒)) = 𝑠(𝑔 ⋅ 𝜇, 𝑔 ⋅ (𝜇𝑒)) = 𝑔 ⋅ (𝜇𝑒) = 𝑔 ⋅ 𝑠(𝜇, 𝜇𝑒) and

𝑟(𝑔 ⋅ (𝜇, 𝜇𝑒)) = 𝑟(𝑔 ⋅ 𝜇, 𝑔 ⋅ (𝜇𝑒)) = 𝑔 ⋅ 𝜇 = 𝑔 ⋅ 𝑟(𝜇, 𝜇𝑒),

the actions on𝑇0𝐸and𝑇𝐸1 are compatible.

(11)

The faithfulness condition ensures that for each𝑔 ∈ 𝐺and each𝜇 ∈ 𝐸with 𝑑(𝑔) = 𝑟(𝜇), there is a unique element𝑔|𝜇 ∈ 𝐺satisfying

𝑔 ⋅ (𝜇𝜈) = (𝑔 ⋅ 𝜇)(𝑔|𝜇⋅ 𝜈) for all𝜈 ∈ 𝑠(𝜇)𝐸.

By Proposition 3.6 of [10], self-similar groupoid actions have the following prop- erties: for𝑔, ℎ ∈ 𝐺, 𝜇 ∈ 𝑑(𝑔)𝐸, and𝜈 ∈ 𝑠(𝜇)𝐸,

(1)𝑔|𝜇𝜈 = (𝑔|𝜇)|𝜈; (2)𝑖𝑑𝑟(𝜇)|𝜇 = 𝑖𝑑𝑠(𝜇);

(3) if(ℎ, 𝑔) ∈ 𝐺(2), then(ℎ|𝑔⋅𝜇, 𝑔|𝜇) ∈ 𝐺(2)and(ℎ𝑔)|𝜇 = (ℎ|𝑔⋅𝜇)(𝑔|𝜇); (4)𝑔−1|𝜇 = (𝑔|𝑔−1⋅𝜇)−1.

Definition 4.5. The𝐶-algebra𝐶(𝐺, 𝐸)of a self-similar action(𝐺, 𝐸)is de- fined as the Cuntz-Pimsner algebra of the𝐶-correspondence

ℳ = ℳ(𝐺, 𝐸) = 𝒳(𝐸) ⊗𝐶(𝐸0)𝐶(𝐺)

over𝐶(𝐺). Here𝒳(𝐸) = 𝐶(𝐸1)is the𝐶-correspondence over𝐶(𝐸0)associ- ated to the graph𝐸and𝐶(𝐸0) = 𝐶(𝐺(0)) ⊆ 𝐶(𝐺). The right action of𝐶(𝐺)on ℳis the usual one and the left action is determined by the representation

𝑊 ∶ 𝐺 → ℒ(ℳ), 𝑊𝑔(𝑖𝑒⊗ 𝑎) = {𝑖𝑔⋅𝑒⊗ 𝑖𝑔|

𝑒𝑎 if𝑑(𝑔) = 𝑟(𝑒) 0 otherwise,

where𝑔 ∈ 𝐺, 𝑖𝑒 ∈ 𝐶(𝐸1)and𝑖𝑔 ∈ 𝐶𝑐(𝐺)are point masses and𝑎 ∈ 𝐶(𝐺). The inner product ofℳis given by

⟨𝜉 ⊗ 𝑎, 𝜂 ⊗ 𝑏⟩ = ⟨⟨𝜂, 𝜉⟩𝑎, 𝑏⟩ = 𝑎⟨𝜉, 𝜂⟩𝑏 for𝜉, 𝜂 ∈ 𝐶(𝐸1)and𝑎, 𝑏 ∈ 𝐶(𝐺).

Remark4.6. Recall that the operations on𝒳(𝐸)are given by (𝜉 ⋅ 𝑎)(𝑒) = 𝜉(𝑒)𝑎(𝑠(𝑒)), ⟨𝜉, 𝜂⟩(𝑣) = ∑

𝑠(𝑒)=𝑣

𝜉(𝑒)𝜂(𝑒), (𝑎 ⋅ 𝜉)(𝑒) = 𝑎(𝑟(𝑒))𝜉(𝑒)

for𝑎 ∈ 𝐶(𝐸0)and𝜉, 𝜂 ∈ 𝐶(𝐸1). The elements𝑖𝑒⊗ 1for𝑒 ∈ 𝐸1form a Parseval frame forℳand every𝜁 ∈ ℳis a finite sum

𝜁 = ∑

𝑒∈𝐸1

𝑖𝑒⊗ ⟨𝑖𝑒⊗ 1, 𝜁⟩.

In particular, if𝒳(𝐸)denotes the dual𝐶-correspondence, then

ℒ(ℳ) = 𝒦(ℳ) ≅ 𝒳(𝐸) ⊗𝐶(𝐸0)𝐶(𝐺) ⊗𝐶(𝐸0)𝒳(𝐸) ≅ 𝑀𝑛⊗ 𝐶(𝐺), where𝑛 = |𝐸1|. The isomorphism is given by

𝑖𝑒𝑗⊗ 𝑖𝑔⊗ 𝑖𝑒𝑘 ↦ 𝑒𝑗𝑘⊗ 𝑖𝑔

for𝐸1= {𝑒1, ..., 𝑒𝑛}and for matrix units𝑒𝑗𝑘 ∈ 𝑀𝑛. There is a unital homomor- phism𝒦(𝒳(𝐸)) → 𝒦(ℳ)given by

𝑖𝑒⊗ 𝑖𝑓→ 𝑖𝑒⊗ 1 ⊗ 𝑖𝑓.

(12)

Since our groupoids have finite unit space𝐸0, the orbit space for the canon- ical action of 𝐺 on 𝐸0 is finite, and 𝐶(𝐺) is the direct sum of 𝐶-algebras of transitive groupoids. Each such transitive groupoid will be isomorphic to a groupoid of the form𝑉 × 𝐻 × 𝑉 with the usual operations, for some sub- set𝑉 ⊆ 𝐸0 and isotropy group𝐻, hence its𝐶-algebra will be isomorphic to 𝐶(𝐻) ⊗ 𝑀|𝑉|.

We recall the following result, see Propositions 4.4 and 4.7 in [10].

Theorem 4.7. If𝑈𝑔, 𝑃𝑣and𝑇𝑒 are the images of𝑔 ∈ 𝐺, 𝑣 ∈ 𝐸0 = 𝐺(0)and of 𝑒 ∈ 𝐸1in the Cuntz-Pimsner algebra𝐶(𝐺, 𝐸), then

∙ 𝑔 ↦ 𝑈𝑔is a representation by partial isometries of𝐺with𝑈𝑣 = 𝑃𝑣 for 𝑣 ∈ 𝐸0;

∙ 𝑇𝑒are partial isometries with𝑇𝑒𝑇𝑒= 𝑃𝑠(𝑒)and

𝑟(𝑒)=𝑣

𝑇𝑒𝑇𝑒 = 𝑃𝑣;

∙ 𝑈𝑔𝑇𝑒 = {𝑇𝑔⋅𝑒𝑈𝑔|𝑒 if𝑑(𝑔) = 𝑟(𝑒)

0, otherwise and 𝑈𝑔𝑃𝑣 = {𝑃𝑔⋅𝑣𝑈𝑔 if𝑑(𝑔) = 𝑣 0, otherwise.

There is a gauge action𝛾of𝕋on𝐶(𝐺, 𝐸)such that𝛾𝑧(𝑈𝑔) = 𝑈𝑔,and𝛾𝑧(𝑇𝑒) = 𝑧𝑇𝑒for𝑧 ∈ 𝕋.

Given𝜇 = 𝑒1⋯ 𝑒𝑛 ∈ 𝐸with𝑒𝑖 ∈ 𝐸1, we let𝑇𝜇 ∶= 𝑇𝑒1⋯ 𝑇𝑒𝑛. Then𝐶(𝐺, 𝐸) is the closed linear span of elements𝑇𝜇𝑈𝑔𝑇𝜈, where𝜇, 𝜈 ∈ 𝐸and𝑔 ∈ 𝐺𝑠(𝜈)𝑠(𝜇).

For each𝑘 ≥ 1, considerℱ𝑘the closed linear span of elements𝑇𝜇𝑈𝑔𝑇𝜈with 𝜇, 𝜈 ∈ 𝐸𝑘 and𝑔 ∈ 𝐺𝑠(𝜈)𝑠(𝜇). Then the fixed point algebraℱ(𝐺, 𝐸) ∶= 𝐶(𝐺, 𝐸)𝕋 under the gauge action is isomorphic tolim

,,→ℱ𝑘. We have

𝑘 ≅ ℒ(ℳ⊗𝑘) ≅ 𝒳(𝐸)⊗𝑘𝐶(𝐸0)𝐶(𝐺) ⊗𝐶(𝐸0)𝒳(𝐸)∗⊗𝑘

using the map𝑇𝜇𝑈𝑔𝑇𝜈 ↦ 𝑖𝜇⊗ 𝑖𝑔⊗ 𝑖𝜈, where𝑖𝜇 ∈ 𝒳(𝐸)⊗𝑘 = 𝐶(𝐸𝑘)are point masses. The embeddingsℱ𝑘 ↪ ℱ𝑘+1are determined by the map

𝜙 = 𝜙𝑊 ∶ 𝐶(𝐺) → ℒ(ℳ), 𝜙𝑊(𝑖𝑔) = 𝑊𝑔. In particular, for𝑎 ∈ 𝐶(𝐺)we get

𝜙(𝑎) = ∑

𝑒∈𝐸1

𝜃𝑖𝑒⊗1,𝑎(𝑖𝑒⊗1),

where𝜃𝜉,𝜂(𝜁) = 𝜉⟨𝜂, 𝜁⟩. The embeddingsℱ𝑘 ↪ ℱ𝑘+1are then

𝜙𝑘(𝑖𝜇⊗ 𝑖𝑔⊗ 𝑖𝜈) =

𝑥∈𝑑(𝑔)𝐸1

𝑖𝜇𝑦⊗ 𝑖𝑔|

𝑥 ⊗ 𝑖𝜈𝑥, if𝑔 ∈ 𝐺𝑠(𝜇)

𝑠(𝜈) and𝑔 ⋅ 𝑥 = 𝑦 0, otherwise.

Remark4.8. The𝐶-algebra𝐶(𝐺, 𝐸)can be described as the crossed product ofℱ(𝐺, 𝐸)by an endomorphism and in many cases, knowledge about𝐾(ℱ𝑘) is sufficient to determine𝐾(ℱ(𝐺, 𝐸))and𝐾(𝐶(𝐺, 𝐸)). For the case when𝐺 is a group, see section 3 in [14]. In the particular case when𝑔|𝑒= 𝑔for all𝑔 ∈ 𝐺 and𝑒 ∈ 𝑑(𝑔)𝐸1we have𝐶(𝐺, 𝐸) ≅ 𝐶(𝐸) ⋊ 𝐺.

(13)

5. Exel-Pardo groupoids for self-similar actions

In this section, we generalize results from [7] and we define the groupoid associated to a self-similar action of a groupoid𝐺on the path space of a finite directed graph𝐸with no sources.

As in [7], we first define the inverse semigroup

𝒮(𝐺, 𝐸) = {(𝛼, 𝑔, 𝛽) ∶ 𝛼, 𝛽 ∈ 𝐸, 𝑔 ∈ 𝐺𝑠(𝛽)𝑠(𝛼)} ∪ {0}

associated to the self-similar action(𝐺, 𝐸), with operations

(𝛼, 𝑔, 𝛽)(𝜆, ℎ, 𝜔) =

(𝛼, 𝑔(ℎ|−1⋅𝜇), 𝜔(ℎ−1⋅ 𝜇)) if𝛽 = 𝜆𝜇 (𝛼(𝑔 ⋅ 𝜇), 𝑔|𝜇ℎ, 𝜔) if𝜆 = 𝛽𝜇

0 otherwise

and(𝛼, 𝑔, 𝛽) = (𝛽, 𝑔−1, 𝛼)for𝛼, 𝛽, 𝜆, 𝜔 ∈ 𝐸. These operations make sense since

𝑑(𝑔) = 𝑠(𝛽) = 𝑡(ℎ|−1⋅𝜇)and𝑑(𝑔(ℎ|−1⋅𝜇)) = 𝑠(𝜔(ℎ−1⋅ 𝜇))when𝛽 = 𝜆𝜇, 𝑑(𝑔|𝜇) = 𝑠(𝜇) = 𝑡(ℎ)and𝑑(𝑔|𝜇ℎ) = 𝑠(𝜔) when𝜆 = 𝛽𝜇.

Note that(𝛼, 𝑔, 𝛽)(𝛽, ℎ, 𝜔) = (𝛼, 𝑔ℎ, 𝜔)and the nonzero idempotents are of the form𝑧𝛼= (𝛼, 𝑠(𝛼), 𝛼).

The inverse semigroup𝒮(𝐺, 𝐸)acts on the infinite path space𝐸by partial homeomorphisms. The action of(𝛼, 𝑔, 𝛽) ∈ 𝒮(𝐺, 𝐸)on𝜉 = 𝛽𝜇 ∈ 𝛽𝐸is given by

(𝛼, 𝑔, 𝛽) ⋅ 𝛽𝜇 = 𝛼(𝑔 ⋅ 𝜇) ∈ 𝛼𝐸.

The action of𝐺on𝐸is defined by𝑔 ⋅ 𝜇 = 𝜂, where for all𝑛we have𝜂1⋯ 𝜂𝑛= 𝑔 ⋅ (𝜇1⋯ 𝜇𝑛). Note that𝑟(𝑔 ⋅ 𝜇) = 𝑔 ⋅ 𝑟(𝜇) = 𝑔 ⋅ 𝑠(𝛽) = 𝑠(𝛼), so the action is well defined.

The groupoid of germs associated with(𝒮(𝐺, 𝐸), 𝐸)is

𝒢(𝐺, 𝐸) = {[𝛼, 𝑔, 𝛽; 𝜉] ∶ 𝛼, 𝛽 ∈ 𝐸, 𝑔 ∈ 𝐺𝑠(𝛽)𝑠(𝛼), 𝜉 ∈ 𝛽𝐸}.

Two germs[𝛼, 𝑔, 𝛽; 𝜉], [𝛼, 𝑔, 𝛽; 𝜉]in𝒢(𝐺, 𝐸)are equal if and only if𝜉 = 𝜉 and there exists a neighborhood𝑉of𝜉such that(𝛼, 𝑔, 𝛽) ⋅ 𝜂 = (𝛼, 𝑔, 𝛽) ⋅ 𝜂for all𝜂 ∈ 𝑉. We obtain that𝜉 = 𝛽𝜆𝜁 for𝜆 ∈ 𝐸 and𝜁 ∈ 𝐸, with𝑟(𝜆) = 𝑠(𝛽) and𝑟(𝜁) = 𝑠(𝜆). Moreover,

𝛼 = 𝛼(𝑔 ⋅ 𝜆), 𝛽= 𝛽𝜆, and𝑔= 𝑔|𝜆. The unit space of𝒢(𝐺, 𝐸)is

𝒢(𝐺, 𝐸)(0) = {[𝛼, 𝑠(𝛼), 𝛼; 𝜉] ∶ 𝜉 ∈ 𝛼𝐸}, identified with𝐸by the map[𝛼, 𝑠(𝛼), 𝛼; 𝜉] ↦ 𝜉.

The terminus and domain maps of the groupoid𝒢(𝐺, 𝐸)are given by 𝑡([𝛼, 𝑔, 𝛽; 𝛽𝜇]) = 𝛼(𝑔 ⋅ 𝜇), 𝑑([𝛼, 𝑔, 𝛽; 𝛽𝜇]) = 𝛽𝜇.

(14)

If two elements𝛾1, 𝛾2 ∈ 𝒢(𝐺, 𝐸)are composable, then

𝛾1 = [𝛼1, 𝑔1, 𝛼2; 𝛼2(𝑔2⋅ 𝜉)], 𝛾2= [𝛼2, 𝑔2, 𝛽; 𝛽𝜉]

for some𝛼1, 𝛼2, 𝛽 ∈ 𝐸, 𝜉 ∈ 𝐸, (𝑔1, 𝑔2) ∈ 𝐺(2)and in this case 𝛾1𝛾2= [𝛼1, 𝑔1𝑔2, 𝛽; 𝛽𝜉].

In particular,

[𝛼, 𝑔, 𝛽; 𝛽𝜇]−1= [𝛽, 𝑔−1, 𝛼; 𝛼(𝑔 ⋅ 𝜇)].

The topology on𝒢(𝐺, 𝐸)is generated by the compact open bisections of the form

𝐵(𝛼, 𝑔, 𝛽; 𝑉) = {[𝛼, 𝑔, 𝛽; 𝜉] ∈ 𝒢(𝐺, 𝐸) ∶ 𝜉 = 𝛽𝜁 ∈ 𝑉},

where𝛼, 𝛽 ∈ 𝐸, 𝑔 ∈ 𝐺𝑠(𝛽)𝑠(𝛼)are fixed, and𝑉 ⊆ 𝑍(𝛽) = 𝛽𝐸is an open subset.

Definition 5.1. A self-similar groupoid action(𝐺, 𝐸)is called pseudo free if for every𝑔 ∈ 𝐺and every𝑒 ∈ 𝑑(𝑔)𝐸1, the condition𝑔 ⋅ 𝑒 = 𝑒and𝑔|𝑒 = 𝑠(𝑒)implies that𝑔 = 𝑟(𝑒).

Remark5.2. If(𝐺, 𝐸)is pseudo free, then𝑔1⋅ 𝛼 = 𝑔2⋅ 𝛼and𝑔1|𝛼 = 𝑔2|𝛼 for some𝛼 ∈ 𝐸implies𝑔1= 𝑔2.

Proof. Indeed, since𝑔−12 𝑔1 ⋅ 𝛼 = 𝛼and𝑔−12 𝑔1|𝛼 = 𝑔−12 |𝑔1⋅𝛼𝑔1|𝛼 = 𝑑(𝑔1|𝛼) = 𝑠(𝛼), it follows that𝑔2−1𝑔1= 𝑟(𝛼), so𝑔1 = 𝑔2. Theorem 5.3. If the action of𝐺on𝐸 is pseudo free, then the groupoid𝒢(𝐺, 𝐸) is Hausdorff and its𝐶-algebra𝐶(𝒢(𝐺, 𝐸))is isomorphic to the Cuntz-Pimsner algebra𝐶(𝐺, 𝐸).

Proof. Since(𝐺, 𝐸) is pseudo free, it follows that[𝛼, 𝑔, 𝛽; 𝜉] = [𝛼, 𝑔, 𝛽; 𝜉]if and only if𝑔 = 𝑔. Moreover, the groupoid𝒢(𝐺, 𝐸)is Hausdorff, see Proposi- tion 12.1 in [7]. Using the properties given in Theorem 4.7and the groupoid multiplication, the isomorphism𝜙 ∶ 𝐶(𝐺, 𝐸) → 𝐶(𝒢(𝐺, 𝐸))is given by

𝜙(𝑃𝑣) = 𝜒𝐵(𝑣,𝑣,𝑣;𝑍(𝑣)), 𝜙(𝑇𝑒) = 𝜒𝐵(𝑒,𝑠(𝑒),𝑠(𝑒);𝑍(𝑠(𝑒))), 𝜙(𝑈𝑔) = 𝜒𝐵(𝑡(𝑔),𝑔,𝑑(𝑔);𝑍(𝑑(𝑔)))

for𝑣 ∈ 𝐸0, 𝑒 ∈ 𝐸1and𝑔 ∈ 𝐺. Here𝜒𝐴is the indicator function of𝐴. Recall that ample Hausdorff groupoids which are similar or Morita equiva- lent have isomorphic homology, see Lemma 4.3 and Theorem 3.12 in [9]. The general strategy of computing the homology of the ample groupoid𝒢(𝐺, 𝐸)is the following.

There is a cocycle𝜌 ∶ 𝒢(𝐺, 𝐸) → ℤgiven by[𝛼, 𝑔, 𝛽; 𝜉] ↦ |𝛼| − |𝛽|with kernel

ℋ(𝐺, 𝐸) = {[𝛼, 𝑔, 𝛽; 𝜉] ∈ 𝒢(𝐺, 𝐸) ∶ |𝛼| = |𝛽|}.

It follows from Theorem3.9that we have a spectral sequence 𝐸𝑝,𝑞2 = 𝐻𝑝(ℤ, 𝐻𝑞(ℋ(𝐺, 𝐸))) ⇒ 𝐻𝑝+𝑞(𝒢(𝐺, 𝐸)).

(15)

Nowℋ(𝐺, 𝐸) =⋃

𝑘≥1𝑘(𝐺, 𝐸)where

𝑘(𝐺, 𝐸) = {[𝛼, 𝑔, 𝛽; 𝜉] ∈ 𝒢(𝐺, 𝐸) ∶ |𝛼| = |𝛽| = 𝑘}.

There are groupoid homomorphisms

𝜏𝑘 ∶ ℋ𝑘(𝐺, 𝐸) → 𝐺, 𝜏𝑘([𝛼, 𝑔, 𝛽; 𝜉]) = 𝑔

andker 𝜏𝑘 is AF for all𝑘 ≥ 1. Indeed, consider𝑅𝑘 the equivalence relation on𝐸𝑘 such that(𝛼, 𝛽) ∈ 𝑅𝑘 if there is𝑔 ∈ 𝐺with𝑔 ⋅ 𝑠(𝛽) = 𝑠(𝛼). Then the map[𝛼, 𝑔, 𝛽; 𝜉] ↦ ((𝜉, 𝑔), (𝛼, 𝛽))gives an isomorphism betweenℋ𝑘(𝐺, 𝐸)and (𝐸 ⋊ 𝐺) × 𝑅𝑘, soker 𝜏𝑘 is isomorphic to𝐸 × 𝑅𝑘. It follows that we have another spectral sequence

𝐸𝑝,𝑞2 = 𝐻𝑝(𝐺, 𝐻𝑞(ker 𝜏𝑘)) ⇒ 𝐻𝑝+𝑞(ℋ𝑘(𝐺, 𝐸)).

It is known that 𝐻0(ker 𝜏𝑘) ≅ 𝐾0(𝐶(ker 𝜏𝑘)) and 𝐻𝑞(ker 𝜏𝑘) = 0 for𝑘 ≥ 1. Also, ker 𝜏𝑘 is similar with ℋ𝑘(𝐺, 𝐸) ×𝜏𝑘 𝐺. Assuming that we computed 𝐻𝑞(ℋ𝑘(𝐺, 𝐸))for all𝑘, then

𝐻𝑞(ℋ(𝐺, 𝐸)) = lim

𝑘→∞,,→

𝐻𝑞(ℋ𝑘(𝐺, 𝐸)) can be computed using the inclusion maps

𝑗𝑘 ∶ ℋ𝑘(𝐺, 𝐸) ↪ ℋ𝑘+1(𝐺, 𝐸), 𝑗𝑘([𝛼, 𝑔, 𝛽; 𝛽𝑥𝜇]) = [𝛼𝑦, 𝑔|𝑥, 𝛽𝑥; 𝛽𝑥𝜇], where𝑥 ∈ 𝐸1and𝑔 ⋅ 𝑥 = 𝑦.

6. Example

Consider again the graph from Example4.1

𝑢 𝑣 𝑤

𝑒2

𝑒3

𝑒4 𝑒5

𝑒6

𝑒1

with𝐸0= {𝑢, 𝑣, 𝑤}and𝐸1 = {𝑒1, 𝑒2, 𝑒3, 𝑒4, 𝑒5, 𝑒6}.

Consider the groupoid𝐺 with unit space 𝐺(0) = {𝑢, 𝑣, 𝑤}and generators 𝑎, 𝑏, 𝑐where𝑑(𝑎) = 𝑢, 𝑡(𝑎) = 𝑑(𝑏) = 𝑣, 𝑑(𝑐) = 𝑡(𝑏) = 𝑤.

𝑢 𝑣 𝑤

𝑎

𝑏

𝑐

We define the action of𝐺by

𝑎 ⋅ 𝑒1= 𝑒2, 𝑎|𝑒1 = 𝑢, 𝑎 ⋅ 𝑒3= 𝑒6, 𝑎|𝑒3 = 𝑏, 𝑏 ⋅ 𝑒2 = 𝑒5, 𝑏|𝑒2 = 𝑎, 𝑏 ⋅ 𝑒6= 𝑒4, 𝑏|𝑒6 = 𝑐,

参照

関連したドキュメント

Furthermore, we will investigate unbounded conditional-expectations in case that ᏹ and ᏺ are generalized von Neumann algebras which are unbounded generalization of von Neumann

In [RS1] the authors study crossed product C ∗ –algebras arising from certain group actions on ˜ A 2 -buildings and show that they are generated by two families of partial

This paper investigates the problem of existence and uniqueness of positive solutions under the general self-similar form of the degenerate parabolic partial di¤erential equation

First, this property appears in our study of dynamical systems and group actions, where it was shown that some information about orbits can be detected from C ∗ -reflexivity of

based on variational methods established the existence of an unbounded sequence of weak solutions for a class of differential equations with p(x)-Laplacian and subject to

BoL Soc. From the observation that self-similar solutions of conservation laws in two space dimensions change type, it follows that for systems of more than two equations,

This result is applicable to most typical nested fractals (but not to the d-dimensional standard Sierpi´ nski gasket with d ≥ 2 at this moment) and all generalized Sierpi´ nski

In [18] we introduced the concept of hypo-nilpotent ideals of n-Lie algebras, and proved that an m-dimensional simplest filiform 3-Lie algebra N 0 can’t be a nilradical of