Abstract

For a locally compact groupoid 𝐺 with a fixed Haar system 𝜆 and quasi-invariant measure 𝜇, we introduce the notion of 𝜆-measurability and construct the space 𝐿1(𝐺, 𝜆, 𝜇) of absolutely integrable functions on 𝐺 and show that it is a Banach -algebra and a two-sided ideal in the algebra 𝑀(𝐺) of complex Radon measures on 𝐺. We find correspondences between representations of 𝐺 on Hilbert bundles and certain class of nondegenerate representations of 𝐿1(𝐺, 𝜆, 𝜇).

1. Introduction and Preliminaries

For a locally compact group 𝐺 with a Haar measure 𝜆, the Banach algebra 𝐿1(𝐺,𝜆) plays a central role in harmonic analysis on 𝐺 [1]. This motivated us to define a similar notion in the case where 𝐺 is a locally compact groupoid with a (fixed) Haar system 𝜆 and quasi-invariant measure 𝜇. This paper is devoted to the study of such a groupoid 𝐿1-algebra 𝐿1(𝐺,𝜆,𝜇). One may expect that as the group case, there is a full interaction between the properties of 𝐺 and that of 𝐿1(𝐺,𝜆,𝜇). This is not completely true. For instance, unlike the group case, not every nondegenerate representation of 𝐿1(𝐺,𝜆,𝜇) is integrated form a representation of 𝐺. In Section 2, we introduce the appropriate measurability notion used to define 𝐿1(𝐺,𝜆,𝜇). Sections 3 and 4 are devoted to the algebra structure of 𝐿1(𝐺,𝜆,𝜇) and its embedding into 𝑀(𝐺) as a closed ideal. In Section 5, we find the class of nondegenerate representations of 𝐿1(𝐺,𝜆,𝜇) which could be obtained by integrating a representation of 𝐺.

We start with some basic definitions. Our main reference for groupoids is the Renault's book [2]. In this paper, we frequently use the following version of Fubini's theorem for (not necessarily 𝜎-finite) Radon measures [1, Theorem  B.3.3].

Lemma 1.1. Let 𝜈 and 𝜌 be Radon measures on the Borel sets of the locally compact spaces 𝑋 and 𝑌, respectively. Then there exists a unique Radon measure 𝜈𝜌 on 𝑋×𝑌 such that (i)if 𝑓𝑋×𝑌𝐶 is 𝜈𝜌-integrable, then the partial integrals 𝑋𝑓(𝑥,𝑦)𝑑𝜈(𝑥),𝑌𝑓(𝑥,𝑦)𝑑𝜌(𝑦)(1.1) define integrable functions such that Fubini’s formula holds 𝑋×𝑌𝑓𝑑𝜈𝜌=𝑋𝑌𝑓𝑑𝜈𝑑𝜌=𝑌𝑋𝑓𝑑𝜌𝑑𝜈,(1.2)(ii)if 𝑓 is measurable such that 𝐴={(𝑥,𝑦)𝑋×𝑌𝑓(𝑥,𝑦)0} is 𝜎-finite and if one of the iterated integrals 𝑋𝑌|𝑓|𝑑𝜈𝑑𝜌 or 𝑌𝑋|𝑓|𝑑𝜌𝑑𝜈 is finite, then 𝑓 is integrable and the Fubini formula holds.

A groupoid is a set 𝐺 endowed with a subset 𝐺2 of 𝐺×𝐺, called the set of composable pairs, a product map: 𝐺2𝐺;(𝑥,𝑦)𝑥𝑦, and an inverse map: 𝐺𝐺;𝑥𝑥1, such that for each 𝑥,𝑦,𝑧𝐺,(i)(𝑥1)1=𝑥, (ii)if (𝑥,𝑦),(𝑦,𝑧)𝐺2 then (𝑥𝑦,𝑧),(𝑥,𝑦𝑧)𝐺2 and (𝑥𝑦)𝑧=𝑥(𝑦𝑧),(iii)(𝑥1,𝑥)𝐺2 and if (𝑥,𝑦)𝐺2 then 𝑥1(𝑥𝑦)=𝑦,(iv)(𝑥,𝑥1)𝐺2 and if (𝑧,𝑥)𝐺2 then (𝑧𝑥)𝑥1=𝑧.

If 𝑥𝐺, 𝑠(𝑥)=𝑥1𝑥 is called the source of 𝑥 and 𝑟(𝑥)=𝑥𝑥1 is called the range of 𝑥. The pair (𝑥,𝑦) is composable if and only if 𝑠(𝑥)=𝑟(𝑦). The set 𝐺0=𝑠(𝐺)=𝑟(𝐺) is the unit space of 𝐺, and its elements are called units in the sense that 𝑥𝑠(𝑥)=𝑥 and 𝑟(𝑥)𝑥=𝑥.

For 𝑢,𝑣𝐺0, 𝐺𝑢=𝑟1{𝑢}, 𝐺𝑣=𝑠1{𝑣}, 𝐺𝑢𝑣=𝐺𝑢𝐺𝑣, and 𝐺{𝑢}=𝐺𝑢𝑢. The latter is called the isotropy group at 𝑢. We define 𝑢𝑣 if 𝐺𝑢𝑣. It is checked that ~ is an equivalence relation on the unit space 𝐺0. The equivalence class of 𝑢 is denoted by [𝑢] and is called the orbit of 𝑢.

A topological groupoid consists of a groupoid 𝐺 and a topology compatible with the groupoid structure, such that the composition map is continuous on 𝐺2 in the induced product topology, and the inversion map is continuous on 𝐺. A locally compact groupoid is a topological groupoid 𝐺 which satisfies the following conditions:(i)𝐺0 is locally compact and Hausdorff in the relative topology.(ii)There is a countable family 𝐶 of compact Hausdorff subsets of 𝐺 whose interiors form a basis for the topology of 𝐺.(iii)Every 𝐺𝑢 is locally compact Hausdorff in the relative topology.

A locally compact groupoid is 𝑟-discrete if its unit space is an open subset. Let 𝐺 be a locally compact groupoid. The support of a function 𝑓𝐺 is the complement of the union of all open, Hausdorff subsets of 𝐺 on which 𝑓 vanishes. The space 𝐶𝑐(𝐺) consists of all continuous functions on 𝐺 with compact support. A left Haar system for 𝐺 consists of measures {𝜆𝑢,𝑢𝐺0} on 𝐺 such that(i)the support of 𝜆𝑢 is 𝐺𝑢,(ii)(continuity) for each 𝑓𝐶𝑐(𝐺),𝑢𝜆(𝑓)(𝑢)=𝑓𝑑𝜆𝑢 is continuous,(iii)(left invariance) for any 𝑥𝐺 and any 𝑓𝐶𝑐(𝐺), 𝑓(𝑥𝑦)𝑑𝜆𝑠(𝑥)(𝑦)=𝑓(𝑦)𝑑𝜆𝑟(𝑥)(𝑦).(1.3)

Let 𝜇 be a measure on 𝐺0. The measure on 𝐺 induced by 𝜇 is 𝜆𝜈=𝑢𝑑𝜇, defined by 𝐺𝑓𝑑𝜈=𝐺0𝑑𝜇(𝑢)𝐺𝑓𝑑𝜆𝑢, for 𝑓𝐶𝑐(𝐺). The measure on 𝐺2 induced by 𝜇 is 𝜈2=𝜆𝑢×𝜆𝑢𝑑𝜇(𝑢). The inversion of 𝜈 is 𝜈1=𝜆𝑢𝑑𝜇(𝑢). Note that the measures 𝜈,𝜈2,𝜈1 are Radon. A measure 𝜇 on 𝐺0 is said to be quasi invariant if its induced measure 𝜈 is equivalent to its inverse 𝜈1. Let 𝜇 be a quasi-invariant measure on 𝐺0. The Radon-Nikodym derivative 𝐷=𝑑𝜈/𝑑𝜈1 is called the modular function of 𝜇. An equivalent definition of modular function on 𝐺 is given in [3, Definition  2.3], defining the modular function Δ as a strictly positive continuous homomorphism on 𝐺 such that Δ|𝐺𝑢𝑢 is modular function for 𝐺𝑢𝑢.

A subset 𝐴 of 𝐺 is called a 𝐺-set if the restrictions of 𝑟 and 𝑠 to it are one to one. Equivalently, 𝐴 is a 𝐺-set if and only if 𝐴𝐴1 and 𝐴1𝐴 are contained in 𝐺0.

We introduce some notations from [4, 5] which is related to the representations of 𝐿1(𝐺,𝜆,𝜇). Let {𝜆𝑢}𝑢 be a fixed Haar system on 𝐺. Let 𝜇 be a quasi-invariant measure, Δ its modular function, 𝜈 be the measure induced by 𝜇 on 𝐺, and 𝜈0=Δ1/2𝜈. Let 𝐼𝐼𝜇(𝐺,𝜈,𝜇)=𝑓𝐿1𝐺,𝜈0𝑓𝐼𝐼,𝜇,<(1.4) where 𝑓𝐼𝐼,𝜇 is defined by 𝑓𝐼𝐼,𝜇||||=sup𝑓(𝑥)𝑗(𝑠(𝑥))𝑘(𝑟(𝑥))𝑑𝜈0||𝑗||(𝑥);2||𝑘||𝑑𝜇=2.𝑑𝜇=1(1.5) If 𝜇1 and 𝜇2 are two equivalent quasi-invariant measures, then 𝑓𝐼𝐼,𝜇1=𝑓𝐼𝐼,𝜇2,(1.6) because 𝑓𝐼𝐼,𝜇=𝐼𝐼𝜇(|𝑓|) for each quasi-invariant measure 𝜇, where 𝐼𝐼𝜇 is the one-dimensional trivial representation on 𝜇. Now define 𝑓𝐼𝐼=sup𝜇𝑓𝐼𝐼,𝜇.(1.7) The supremum is taken over the class of quasi-invariant measures.

Let 𝐼(𝐺,𝜈,𝜇)=𝑓𝐿1𝐺,𝜈0𝑓𝐼,𝜇,<(1.8) where 𝑓𝐼,𝜇||𝑓||=max𝑢𝑑𝜆𝑢,||𝑓𝑢||𝑑𝜆𝑢.(1.9) Under the canonical convolution and involution, 𝐼(𝐺,𝜈,𝜇) becomes a Banach -algebra [5, page  4]. Here 𝑓(𝑥)=𝑓(𝑥1). If we consider 𝑓𝐼,𝑟=sup𝑢||𝑓||𝑑𝜆𝑢,𝑓𝐼,𝑠=sup𝑢||𝑓||𝑑𝜆𝑢,(1.10) and put 𝑓𝐼=max(𝑓𝐼,𝑟,𝑓𝐼,𝑠), then 𝑓𝐼=𝑓𝐼,𝜇.

2. 𝜆-Measurability

For the rest of the paper, 𝐺 is a locally compact, Hausdorff, second countable groupoid which admits a left Haar system 𝜆={𝜆𝑢}.

Definition 2.1. A system of measures {𝜆𝑢}𝑢𝐺0 is said to be complete if for each 𝑢𝐺0, 𝜆𝑢 is complete on its 𝜎-algebra 𝔐𝜆𝑢. A Borel measurable set 𝐸𝐺 is called 𝜆-measurable if for each 𝑢𝐺0, 𝐸𝐺𝑢𝔐𝜆𝑢. A function 𝑓𝐺 is 𝜆-measurable if for every 𝑢𝐺0 and every open set 𝑂,𝑓1(𝑂)𝐺𝑢𝔐𝜆𝑢.

Proposition 2.2. If 𝜆={𝜆𝑢} is the completion of 𝜆={𝜆𝑢} and 𝑓𝐺[0,] is 𝜆-measureable, then there is a 𝜆-measurable function 𝑔 such that 𝑓=𝑔 on 𝐺𝑢(𝜆𝑢-a.e).

Proof. Since 𝑓 is 𝜆-measurable for each 𝑢𝐺0, 𝑓 is 𝜆𝑢-measurable on 𝐺𝑢, and since 𝜆𝑢 is the completion of 𝜆𝑢, there exists a 𝜆𝑢-measurable function 𝑔𝑢 such that 𝑓=𝑔𝑢(𝜆𝑢-a.e). Now define 𝑔=𝑔𝑢 on 𝐺𝑢 and zero, elsewhere. Since for every 𝑢𝐺0 and every open set 𝑂,𝑔1(𝑂)𝐺𝑢=𝑔𝑢1(𝑂)𝔐𝜆𝑢, 𝑔 is 𝜆-measurable, and since 𝜆𝑢(𝐺𝑢)𝑐=0, we have 𝑓=𝑔 on 𝐺𝑢 (𝜆𝑢-a.e).

From now on, we assume that the Haar system 𝜆 is complete.

Lemma 2.3. For each 𝑓𝐺, 𝜆-measurability of 𝑓 is equivalent to 𝜈-measurability of 𝑓.

Proof. We have supp𝜆𝑢=𝐺𝑢 and 𝑓1𝑓(𝑂)=1(𝑂)𝐺𝑢𝑓1(𝑂)(𝐺𝑢)𝑐.(2.1) Since 𝜆𝑢 is complete and (𝐺𝑢)𝑐 is 𝜆𝑢-null, (𝑓1(𝑂)(𝐺𝑢)𝑐)(𝐺𝑢)𝑐 is in 𝔐𝜆𝑢. Thus, for each 𝑢𝐺0 and open set 𝑂𝑓1(𝑂)𝔐𝜆𝑢𝑓1(𝑂)𝐺𝑢𝔐𝜆𝑢.(2.2) Now for 𝜆𝜈=𝑢𝑑𝜇(𝑢), we have 𝔐𝜈=𝑢𝐺0𝔐𝜆𝑢, hence 𝑓 is 𝜈-measurable if and only if 𝑓 is 𝜆-measurable.

If 𝑓𝐺 is 𝜆-measurable and 𝑔 is continuous, then 𝑔𝑓𝐺 is 𝜆-measurable. Also, if 𝑓,𝑔𝐺 is 𝜆-measurable, Φ𝑌 is continuous, and (𝑥)=Φ(𝑓(𝑥),𝑔(𝑥)), then is 𝜆-measurable. If 𝑓=𝑢+𝑖𝑣 then 𝑓 is 𝜆-measurable if and only if 𝑢,𝑣 are 𝜆-measurable. If 𝑓,𝑔𝐺 are 𝜆-measurable, so are 𝑓+𝑔 and 𝑓𝑔. Also, if {𝑓𝑗}1 is a sequence of -valued 𝜆-measurable functions, then the functions 𝑔1(𝑥)=sup𝑗𝑓𝑗(𝑥),𝑔2(𝑥)=inf𝑗𝑓𝑗(𝑥),𝑔3(𝑥)=limsup𝑗𝑓𝑗(𝑥), and 𝑔4(𝑥)=liminf𝑗𝑓𝑗(𝑥) are all 𝜆-measurable. If 𝑓(𝑥)=lim𝑗𝑓𝑗(𝑥) exists for every 𝑥𝐺, then 𝑓 is 𝜆-measurable. Thus if {𝑓𝑗}1 is a sequence of complex-valued 𝜆-measurable functions and 𝑓𝑗𝑓𝜆𝑢-a.e, then 𝑓 is 𝜆-measurable.

3. The Algebra 𝐿1(𝐺,𝜆,𝜇)

In this section, we define the space of integrable functions on 𝐺 with respect to a fixed Haar system 𝜆 and quasi-invariant measure 𝜇 and show that it is a Banach *-algebra under canonical convolution and involution.

Definition 3.1. Suppose 𝜇 is a quasi-invariant probability measure on 𝐺0, and 𝜈 is Radon measure induced by 𝜇, then we define 𝐿1(𝐺,𝜈)=𝐿1(𝐺,𝜆,𝜇)=𝑓𝐺𝑓is𝜆-measurable,𝑓1=𝐺||||.𝑓(𝑥)𝑑𝜈(𝑥)<(3.1)

It is clear that 𝐿1(𝐺,𝜆,𝜇) is a Banach space. We show that the Banach space 𝐿1(𝐺,𝜆,𝜇) is a Banach *-algebra under the following convolution product.(𝑓𝑔)(𝑥)=𝐺𝑟(𝑥)𝑦𝑓(𝑦)𝑔1𝑥𝑑𝜈(𝑦).(3.2)

Lemma 3.2. If 𝑓,𝑔𝐿1(𝐺,𝜆,𝜇), then 𝑓𝑔𝐿1(𝐺,𝜆,𝜇) and 𝑓𝑔1𝑓1𝑔1.

Proof. For each 𝑓,𝑔, 𝑓𝑔1=𝐺||||𝑓𝑔(𝑥)𝑑𝜈(𝑥)𝐺𝐺𝑟(𝑥)||||||𝑔𝑦𝑓(𝑦)1𝑥||𝑑𝜈(𝑦)𝑑𝜈(𝑥)𝐺||||||𝑔𝑦𝑓(𝑦)1𝑥||=𝑑𝜈(𝑦)𝑑𝜈(𝑥)𝐺||𝑓||||𝑔||=(𝑦)(𝑥)𝑑𝜈(𝑦)𝑑𝜈(𝑦𝑥)𝐺0𝐺||||𝑔(𝑥)𝐺0𝐺||||𝑓(𝑦)𝑑𝜆𝑢(𝑦)𝑑𝜇(𝑢)𝑑𝜆𝑢=(𝑦𝑥)𝑑𝜇(𝑢)𝐺||||𝑔(𝑥)𝐺||||𝑓(𝑦)𝑑𝜈(𝑦)𝑑𝜈(𝑥)=𝑓1𝑔1.(3.3) The measurability of 𝑓𝑔 follows from 𝜆-measurability of 𝑓,𝑔.

Next, we define an involution on 𝐿1(𝐺,𝜆,𝜇). We say that the assertion 𝑃(𝑥) holds for 𝜆-a.e. 𝑥 if for 𝐸={𝑥¬𝑃(𝑥)}, 𝜇{𝑢𝜆𝑢(𝐸)>0}=0. Clearly an assertion holds 𝜆-almost everywhere if and only if it holds 𝜈-almost everywhere.

Lemma 3.3. Suppose 𝐷𝑢𝐺𝑢+ with 𝐷𝑢(𝑥)=𝑑𝜆𝑢(𝑥)/𝑑𝜆𝑢(𝑥)(𝑥𝐺). Then 𝐷=𝐷𝑢 on 𝐺𝑢(𝑎.𝑒.).

Proof. Suppose 𝐸𝐺𝑢. We have 𝜈(𝐸)=𝐺0𝜆𝑢(𝐸)𝑑𝜇(𝑢)=𝐺0𝐸𝑑𝜆𝑢(𝑥)𝑑𝜇(𝑢)=𝐺0𝐸𝐷𝑢(𝑥)𝑑𝜆𝑢(𝑥)𝑑𝜇(𝑢).(3.4) Also from 𝑑𝜈=𝐷𝑑𝜈1 we have 𝜈(𝐸)=𝐸𝐷(𝑥)𝑑𝜈1(𝑥)=𝐸𝐷(𝑥)𝐺0𝑑𝜆𝑢(𝑥)𝑑𝜇(𝑢)=𝐺0𝐸𝐷(𝑥)𝑑𝜆𝑢(𝑥)𝑑𝜇(𝑢).(3.5) Thus 𝐺0𝐸𝐷𝑢(𝑥)𝐷(𝑥)𝑑𝜆𝑢(𝑥)𝑑𝜇(𝑢)=0.(3.6) Now, let 𝐸𝑢={𝑥𝐺𝑢𝐷𝑢(𝑥)>𝐷(𝑥)}. If 𝜆𝑢(𝐸𝑢)>0 then 𝐸𝑢𝐷𝑢(𝑥)𝐷(𝑥)𝑑𝜆𝑢(𝑥)>0.(3.7) But 𝐸𝑢(𝐷𝑢(𝑥)𝐷(𝑥))𝑑𝜆𝑢(𝑥)=0(𝜇-a.e.), hence 𝜆𝑢(𝐸𝑢)=0(𝜇-a.e.). Thus 𝜇{𝑢𝜆𝑢(𝐸𝑢)>0}=0. Therefore, 𝐷𝑢(𝑥)𝐷(𝑥)(𝜆-a.e.). A similar argument leads to 𝐷𝑢(𝑥)𝐷(𝑥)(𝜆-a.e.).

Proposition 3.4. The map 𝐿1(𝐺,𝜆,𝜇)𝐿1(𝐺,𝜆,𝜇);𝑓𝑓, where 𝑓(𝑥)=𝑓(𝑥1)𝐷(𝑥1), is an isometric involution on 𝐿1(𝐺,𝜆,𝜇).

Note that from [5, page  9], we have 𝑓𝐿1(𝐺,𝜆,𝜇)=𝑓𝐿1(𝐺,𝜆,𝜇)=𝑓𝐿1(𝐺,𝜈0)=𝑓𝐿1(𝐺,𝜈0)𝑓𝐼𝐼,𝜇=𝑓𝐼𝐼,𝜇𝑓𝐼,𝜇=𝑓𝐼,𝜇.(3.8) Hence the space of 𝐿1(𝐺,𝜆,𝜇) is in general bigger than 𝐼(𝐺,𝜈,𝜇) and 𝐼𝐼𝜇(𝐺,𝜈,𝜇) with respect to 𝐼-norm and 𝐼𝐼-norm, indeed 𝐼(𝐺,𝜈,𝜇)𝐼𝐼𝜇(𝐺)𝐿1(𝐺,𝜆,𝜇).

According to [2, Lemma  1.4], 𝐼-norm topology is coarser than the inductive limit topology. Also [5, page  15] shows that 𝐶𝑐(𝐺) has a two-sided bounded approximate identity (𝑒𝑛)𝑛=1 with respect to the inductive limit topology with the following properties:(i)𝑒𝑛(𝑥)0 for all 𝑥𝐺,(ii)|𝑒𝑛(𝑥)𝑑𝜆𝑢(𝑥)1|<1/𝑛 for all 𝑢𝐾𝑛, where 𝑛𝐾𝑛=𝐺0 and 𝐾𝑛's are compact,(iii)𝑒𝑛(𝑥)=𝑒𝑛(𝑥1) for all 𝑥𝐺.

An argument in [5, page  15] shows that there is 𝑀>0 such that 𝑒𝑛𝐼𝐼,𝜇𝑀 for all 𝑛. Since 𝐶𝑐(𝐺) is dense in 𝐿1(𝐺,𝜆,𝜇), thus 𝐿1(𝐺,𝜆,𝜇) has a two-sided bounded approximate identity.

For each 𝑓𝐿1(𝐺,𝜆,𝜇) define 𝐿𝑥𝑥𝑓(𝑦)=𝑓1𝑦,𝑅𝑥𝑓(𝑦)=𝑓(𝑦𝑥),(3.9) when the multiplications on the right hand sides are defined. It is easy to check that the maps 𝐿𝑥,𝑅𝑥 are homomorphisms.

Proposition 3.5. Let 𝐼 be a closed subspace of 𝐿1(𝐺,𝜆,𝜇). Then 𝐼 is a left ideal if and only if it is closed under left translation, and 𝐼 is a right ideal if and only if it is closed under right translation.

Proof. Note that since 𝑓𝑔=𝐺𝑟(𝑦)𝑓(𝑦)𝐿𝑦𝑔𝑑𝜈(𝑦), 𝐿𝑥(𝑓𝑔)=𝐺𝑟(𝑦)𝑓(𝑦)𝐿𝑥𝐿𝑦𝑔𝑑𝜈(𝑦)=𝐺𝑟(𝑦)𝑓(𝑦)𝐿𝑥𝑦=𝑔𝑑𝜈(𝑦)𝐺𝑟(𝑦)𝑓𝑥1𝑦𝐿𝑦𝑔𝑑𝜈(𝑦)=𝐺𝑟(𝑦)𝐿𝑥𝑓(𝑦)𝐿𝑦𝐿𝑔𝑑𝜈(𝑦)=𝑥𝑓𝑔.(3.10) Now suppose (𝑒𝑛)𝑛 is a bounded approximate identity for 𝐿1(𝐺,𝜆,𝜇). For the first assertion, if 𝑓𝐿1(𝐺,𝜆,𝜇) and 𝑔𝐼 and 𝐼 is a left ideal, then we have 𝐿𝑥𝑒𝑛𝑓=𝐿𝑥𝑒𝑛𝑓𝐿𝑥𝑓.(3.11) Conversely, if 𝐼 is closed under left translation and 𝑓𝐿1(𝐺,𝜆,𝜇) and 𝑔𝐼, 𝑓𝑔=𝐺𝑟(𝑦)𝑓(𝑦)𝐿𝑦𝑔𝑑𝜈(𝑦)(3.12) is in the closed linear span of the functions 𝐿𝑦𝑔 and hence in 𝐼. The other assertion is proved similarly.

4. 𝑀(𝐺) as Banach -Algebra

Let 𝑀(𝐺) be the space of complex Radon measures on 𝐺. If 𝜂,𝜃𝑀(𝐺), then the map 𝜓𝐼(𝜓) on 𝐶0(𝐺) defined by 𝐼(𝜓)=𝐺𝐺𝑟(𝑦)𝜓(𝑥𝑦)𝑑𝜂(𝑥)𝑑𝜃(𝑦) is a linear functional on 𝐶0(𝐺) satisfying |𝐼(𝜓)|𝜓sup𝜂𝜃, so by Riesz representation theorem, it is given by a measure shown as 𝜂𝜃 called the convolution of 𝜂,𝜃 with 𝜂𝜃𝜂𝜃. If we define 𝜂(𝐸)=𝜂(𝐸1), then 𝜂𝜂 is an involution on 𝑀(𝐺), and 𝑀(𝐺) is a Banach -algebra. In this section, we show that the space 𝐿1(𝐺,𝜆,𝜇) is a closed two-sided ideal of 𝑀(𝐺).

Proposition 4.1. The map 𝐿1(𝐺,𝜆,𝜇)𝑀(𝐺) with 𝜈𝑓(𝐸)=𝐸𝑓𝜒𝐸𝑑𝜈, for (𝐸𝐺); 𝑓𝜈𝑓 is an isometric embedding.

Proof. If 𝑓𝐿1(𝐺,𝜆,𝜇), then 𝑓 is 𝜆-measurable so the integral exists, and it is easy to check that 𝜈𝑓 is a measure on 𝐺. We show that 𝜈𝑓 is Radon. If 𝑓=𝑢+𝑖𝑣, then 𝜈𝑓=𝜈𝑢+𝑖𝜈𝑣, so 𝜈𝑓 is Radon if and only if 𝜈𝑢 and 𝜈𝑣 are Radon. Since 𝐺 is LCH and second countable, we have 𝜈𝑢(𝐾)=𝐾𝑢𝑑𝜈𝐺|𝑢|𝑑𝜈=𝑢1<, for each compact set 𝐾, thus 𝜈𝑢 is Radon. Similarly 𝜈𝑣 is Radon, and so is 𝜈𝑓.
By definition, 𝜈𝑓=sup{𝑛1|𝜈𝑓(𝐸𝑖)|𝑛𝑁,𝐺=𝑛1𝐸𝑖}, so for each 𝜖>0, there exists a partition {𝐸𝑖}𝑛1 of 𝐺 such that 𝜈𝑓𝜖<𝑛1||𝜈𝑓𝐸𝑖||=𝑛1||||𝐺𝑓𝜒𝐸𝑖||||𝑑𝜈𝐺||𝑓||𝑛1𝜒𝐸𝑖𝑑𝜈=𝑓1.(4.1) Thus, 𝜈𝑓𝑓1. Conversely, suppose 𝑓0, then 𝜈𝑓0 and for every partition {𝐸𝑖}𝑛1 of 𝐺 we have, 𝜈𝑓𝑛1𝜈𝑓𝐸𝑖=𝑛1𝐺𝑓𝜒𝐸𝑖𝑑𝜈=𝐺𝑓𝑑𝜈=𝑓1.(4.2) If 𝑓=𝑢+𝑖𝑣=(𝑓1𝑓2)+𝑖(𝑓3𝑓4), where 𝑓𝑖0 then 𝜈𝑓=𝜈𝑓1+𝜈𝑓2+𝜈𝑓3+𝜈𝑓4𝑓11+𝑓21+𝑓31+𝑓41𝑓1. Hence 𝜈𝑓𝑓1 and equality holds.

Corollary 4.2. 𝐿1(𝐺,𝜆,𝜇) is a closed subspace of 𝑀(𝐺).

Lemma 4.3. If 𝑓,𝑔𝐿1(𝐺,𝜆,𝜇), then 𝜈(𝑓𝑔)=𝜈𝑓𝜈𝑔.

Proof. For each compact set 𝐾, we have 𝜈𝑓𝜈𝑔(𝐾)=𝐺𝜒𝐾(𝜈𝑥)𝑑𝑓𝜈𝑔(𝑥)=𝐺𝐺𝑟(𝑥)𝜒𝐾(𝑦𝑥)𝑑𝜈𝑓(𝑦)𝑑𝜈𝑔(=𝑥)𝐺𝑓(𝑦)𝐺𝑠(𝑦)𝜒𝐾(𝑦𝑥)𝑑𝜈𝑔(𝑥)𝑑𝜈(𝑦)=𝐺𝐺𝑠(𝑦)𝑦𝑓(𝑦)𝑔1𝑥𝜒𝐾=(𝑥)𝑑𝜈(𝑦)𝑑𝜈(𝑥)𝐺𝜒𝐾(𝑥)𝐺𝑟(𝑥)𝑦𝑓(𝑦)𝑔1𝑥𝑑𝜈(𝑦)𝑑𝜈(𝑥)=𝐺(𝑓𝑔)(𝑥)𝜒𝐾(𝑥)𝑑𝜈(𝑥)=𝜈𝑓𝑔(𝐾).(4.3) Since 𝜈𝑓𝑔 and 𝜈𝑓𝜈𝑔 are regular measures, the equality holds for each open set and then for each measurable set.

If 𝑓𝐿1(𝐺,𝜆,𝜇) and 𝜂𝑀(𝐺), we will define 𝜂𝑓 such that 𝜈𝜂𝑓=𝜂𝜈𝑓. Suppose 𝜑𝐶0(𝐺), we put 𝜂𝑓(𝜑)=𝐺𝜑(𝑥)𝑑𝜈𝜂𝑓(𝑥)=𝐺𝜑(𝑥)(𝜂𝑓)(𝑥)𝑑𝜈(𝑥).(4.4) On the other hand, 𝜂𝜈𝑓(𝜑)=𝐺𝐺𝑟(𝑥)𝜑(𝑦𝑥)𝑑𝜂(𝑦)𝑑𝜈𝑓(𝑥)=𝐺𝐺𝑟(𝑥)=𝜑(𝑦𝑥)𝑑𝜂(𝑦)𝑓(𝑥)𝑑𝜈(𝑥)𝐺𝐺𝑠(𝑦)𝑓𝑦1𝑥𝑦𝜑(𝑥)𝑑𝜂(𝑦)𝑑𝜈1𝑥=𝐺𝜑(𝑥)𝐺𝑠(𝑦)𝑓𝑦1𝑥𝑑𝜂(𝑦)𝑑𝜈(𝑥).(4.5)

Comparing these equalities, (𝜂𝑓)(𝑥)=𝐺𝑟(𝑥)𝑓𝑦1𝑥𝑑𝜂(𝑦).(4.6) If 𝑓𝐿1(𝐺,𝜆,𝜇), then it is easy to check that 𝐺𝑢𝑅𝑦𝑓(𝑥)𝑑𝜆𝑢(𝑥)=𝐺𝑢𝑓(𝑥)𝑑𝜆𝑢𝑥𝑦1𝑦=𝐷1𝐺𝑢𝑓(𝑥)𝑑𝜆𝑢(𝑥).(4.7) Thus 𝐺𝑅𝑦𝑦𝑓(𝑥)𝑑𝜈(𝑥)=𝐷1𝐺𝑓(𝑥)𝑑𝜈(𝑥).(4.8) Similarly, we want to define 𝑓𝜂 in such a way that the equality 𝜈(𝑓𝜂)=𝜈𝑓𝜂 holds. Again suppose 𝜑𝐶0(𝐺). We have 𝜈(𝑓𝜂)(𝜑)=𝐺𝜑(𝑥)𝑑𝜈𝑓𝜂(𝑥)=𝐺𝜑(𝑥)(𝑓𝜂)(𝑥)𝑑𝜈(𝑥)=𝐺0𝐺𝜑(𝑥)(𝑓𝜂)(𝑥)𝑑𝜆𝑢(=𝑥)𝑑𝜇(𝑢)𝐺0𝐺𝑢𝜑(𝑥)(𝑓𝜂)(𝑥)𝑑𝜆𝑢(𝑥)𝑑𝜇(𝑢)=𝐺𝑢𝜑(𝑥)(𝑓𝜂)(𝑥)𝑑𝜈(𝑥).(4.9) On the other hand, 𝜈𝑓(𝜂𝜑)=𝐺𝐺𝑟(𝑦)𝜑(𝑥𝑦)𝑑𝜈𝑓(=𝑥)𝑑𝜂(𝑦)𝐺𝐺𝑟(𝑦)=𝜑(𝑥𝑦)𝑓(𝑥)𝑑𝜈(𝑥)𝑑𝜂(𝑦)𝐺𝐺𝑟(𝑦)𝜑(𝑥)𝑓𝑥𝑦1𝑑𝜈𝑥𝑦1=𝑑𝜂(𝑦)𝐺𝐺𝑟(𝑦)𝜑(𝑥)𝑓𝑥𝑦1𝐷𝑦1=𝑑𝜈(𝑥)𝑑𝜂(𝑦)𝐺𝑠(𝑥)𝜑(𝑥)𝐺𝑓𝑥𝑦1𝐷𝑦1𝑑𝜂(𝑦)𝑑𝜈(𝑥).(4.10) Comparing the above equalities, we have (𝑓𝜂)(𝑥)=𝐺𝑓𝑥𝑦1𝐷𝑦1𝑑𝜂(𝑦).(4.11)

Lemma 4.4. 𝐿1(𝐺,𝜆,𝜇) is a two-sided ideal of 𝑀(𝐺).

Proof. Suppose 𝑓𝐿1(𝐺,𝜆,𝜇) and 𝜂𝑀(𝐺). Then we have 𝜂𝑓1=𝐺||||𝜂𝑓(𝑥)𝑑𝜈(𝑥)𝐺𝐺𝑟(𝑥)||𝑓𝑦1𝑥||𝑑||𝜂||(=𝑦)𝑑𝜈(𝑥)𝐺𝐺𝑟(𝑥)||||𝑑||𝜂||𝑓(𝑥)(𝑦)𝑑𝜈(𝑦𝑥)=𝐺||||𝑓(𝑥)𝐺𝑟(𝑥)𝑑||𝜂||(𝑦)𝑑𝜈(𝑥)𝜂𝑓1<.(4.12) Thus 𝜂𝑓𝐿1(𝐺,𝜆,𝜇). Also 𝑓𝜂1=𝐺||(||𝑓𝜂)(𝑥)𝑑𝜈(𝑥)𝐺||𝑓𝑥𝑦1𝐷𝑦1||𝑑||𝜂||(=𝑦)𝑑𝜈(𝑥)𝐺||||𝐷𝑦𝑓(𝑥)1||𝜂||=𝐷(𝑦)𝑑𝜈(𝑥)𝑑(𝑦)𝐺||𝑓||||𝜂||𝑑𝜈𝑑=𝑓1𝜂<.(4.13) Hence, 𝑓𝜂𝐿1(𝐺,𝜆,𝜇).

5. Representation Theory of Locally Compact Groupoids

Let us briefly review some basic facts of representation theory on locally compact and Hausdorff groupoids. Recall that the measure on 𝐺 defined by 𝜆𝜈=𝑢𝑑𝜇(𝑢) is a Radon measure. Let 𝜈0 be the measure on 𝐺 given by 𝑑𝜈0=𝐷1/2𝑑𝜈,(5.1) where 𝐷 is modular function of 𝜇. Then for each Borel subset 𝐸 of 𝐺, we have 𝜈0(𝜒𝐸)=𝐸(𝑥)𝐷1/2(𝜒𝑥)𝑑𝜈(𝑥)=𝐸𝑥1𝐷1/2𝑥1𝑑𝜈1(=𝜒𝑥)𝐸1(𝑥)𝐷1/2(𝑥)𝐷1(𝑥)𝑑𝜈(𝑥)=𝜈0𝐸1.(5.2) Hence, 𝜈0 is symmetric under inversion.

Definition 5.1. A representation of the locally compact groupoid 𝐺 is a triple (𝜇,{𝐻𝑢}𝑢,𝜋) consisting of a Hilbert bundle (𝐺0,{𝐻𝑢}𝑢𝐺0,𝜇), where 𝜇 is a quasi-invariant measure on 𝐺0 (with associated Radon measures 𝜈,𝜈1,𝜈2,𝜈0) and for each 𝑥𝐺, a unitary element 𝜋(𝑥)(𝐻𝑠(𝑥),𝐻𝑟(𝑥)) such that(i)𝜋(𝑢) is the identity map on 𝐻𝑢 for all 𝑢,(ii)𝜋(𝑥𝑦)=𝜋(𝑥)𝜋(𝑦) for 𝜈2-a.e. (𝑥,𝑦)𝐺2,(iii)𝜋(𝑥1)=𝜋1(𝑥) for 𝜈-a.e. 𝑥𝐺,(iv)for any 𝜉,𝜂𝐿2(𝐺0,{𝐻𝑢}𝑢,𝜇), the map𝑥𝜋(𝑥)𝜉(𝑠(𝑥)),𝜂(𝑟(𝑥))(5.3) is 𝜈-measurable on 𝐺.

Definition 5.2. A representation Π of 𝐿1(𝐺,𝜆,𝜇) on a Hilbert space, 𝐻 is a -homomorphism Π𝐿1(𝐺,𝜆,𝜇)(𝐻).(5.4) It is called nondegenerate if Π(𝑓)𝜉𝑓𝐿1,𝜉𝐻 is dense in 𝐻.

Continuity of Π automatically holds, because each -homomorphism from a Banach -algebra to a 𝐶-algebra is norm decreasing, namely, Π(𝑓)𝑓1, for each 𝑓𝐿1.

Our main aim here is to find a correspondence between unitary representations of 𝐺 and nondegenerate representations of 𝐿1(𝐺,𝜆,𝜇). Unfortunately, this is impossible in general. Such a correspondence exists between representations of 𝐺 and those of 𝐶𝑐(𝐺), when 𝐻 is separable, 𝐺 is second countable and admits sufficiently many nonsingular 𝐺-sets [2, Theorem  1.21].

Proposition 5.3. Let 𝐺 be a second countable locally compact groupoid with Haar system and with sufficiently many nonsingular Borel 𝐺-sets. Then, every representation of 𝐶𝑐(𝐺) on a separable Hilbert space is the integrated form of a representation of 𝐺.

These assumptions satisfied in the case of 𝑟-discrete groupoids and transformation groups. The main problem is that a continuous representation of (𝐿1(𝐺,𝜆,𝜇),𝐼) is not necessarily continuous in the 𝐿1-norm. To get a partial result, we use the following result [2, Proposition  1.7].

Proposition 5.4. Suppose (𝜋,{𝐻𝑢}𝑢,𝜇) is a representation of 𝐺. For 𝜉,𝜂𝐿2(𝐺0,{𝐻𝑢}𝑢,𝜇) and 𝑓𝐶𝑐(𝐺), Π(𝑓)𝜉,𝜂=𝐺𝑓(𝑥)𝜋(𝑥)𝜉(𝑠(𝑥)),𝜂(𝑟(𝑥))𝑑𝜈0(𝑥)(5.5) defines a bounded nondegenerate -representation of 𝐶𝑐(𝐺) on 𝐿2(𝐺0,{𝐻𝑢}𝑢,𝜇) such that two equivalent representations of 𝐺 give two equivalent representations of 𝐶𝑐(𝐺).

The equation above is called the integrated form of a unitary representation. If Π is a representation of 𝐶𝑐(𝐺), the above proposition says that Π should be of the form Π(𝑓)𝜉,𝜂=𝐺𝑓(𝑥)𝜋(𝑥)𝜉(𝑠(𝑥)),𝜂(𝑟(𝑥))𝑑𝜈0(𝑥).(5.6) For each 𝜉,𝜂𝐿2(𝐺0,{𝐻𝑢}𝑢,𝜇) and 𝑓𝐶𝑐(𝐺).

Next, we define a representation of 𝐼(𝐺,𝜆,𝜇), denoted by Π𝐼 as Π𝐼(𝑓)𝜉,𝜂=𝐺𝑓(𝑥)𝜋(𝑥)𝜉(𝑠(𝑥)),𝜂(𝑟(𝑥))𝑑𝜈0(𝑥)(5.7) for each 𝜉,𝜂𝐿2(𝐺0,{𝐻𝑢}𝑢,𝜇) and 𝑓𝐼(𝐺,𝜆,𝜇)

Lemma 5.5. Π𝐼 is a bounded representation of 𝐼(𝐺,𝜆,𝜇) on (𝐿2(𝐺0,{𝐻𝑢}𝑢,𝜇)).

Proof. We have Π𝐼(𝑓)𝜉,𝜂=𝐺𝑓(𝑥)𝜋(𝑥)𝜉(𝑠(𝑥)),𝜂(𝑟(𝑥))𝑑𝜈0(=𝑥)𝐺0𝐺𝑢𝑓(𝑥)𝜋(𝑥)𝜉(𝑠(𝑥)),𝜂(𝑟(𝑥))𝑑1/2(𝑥)𝑑𝜆𝑢=(𝑥)𝑑𝜇(𝑢)𝐺0𝐺𝑢𝑓(𝑥)𝜋(𝑥)𝜉(𝑠(𝑥))𝐷1/2(𝑥)𝑑𝜆𝑢(𝑥),𝜂(𝑢)𝑑𝜇(𝑢).(5.8) Thus, we may define for 𝜇-a.e. 𝑢𝐺0, Π𝐼(𝑓)𝜉(𝑢)=𝐺𝑢𝑓(𝑥)𝜋(𝑥)𝜉(𝑠(𝑥))𝐷1/2(𝑥)𝑑𝜆𝑢(𝑥).(5.9) Now, we have ||Π𝐼||=||||(𝑓)𝜉,𝜂𝐺𝑓(𝑥)𝜋(𝑥)𝜉(𝑠(𝑥)),𝜂(𝑟(𝑥))𝑑𝜈0||||(𝑥)𝐺||𝑓||(𝑥)𝜉𝑠(𝑥)𝜂𝑟(𝑥)𝑑𝜈0(𝑥)𝐺||||𝑓(𝑥)𝜉𝑠(𝑥)2𝑑𝜈1(𝑥)1/2𝐺||||𝑓(𝑥)𝜂𝑟(𝑥)2𝑑𝜈(𝑥)1/2𝐺0𝐺𝑢||||𝑓(𝑥)𝑑𝜆𝑢(𝑥)𝜉(𝑢)2𝑑𝜇(𝑢)1/2×𝐺0𝐺𝑢||||𝑓(𝑥)𝑑𝜆𝑢(𝑥)𝜂(𝑢)2𝑑𝜇(𝑢)1/2𝑓1/2𝐼,𝑠𝜉𝑓1/2𝐼,𝑟𝜂𝑓𝐼𝜉𝜂.(5.10) Therefore, the map 𝑥𝑓(𝑥)𝜋(𝑥)𝜉(𝑠(𝑥)),𝜂(𝑟(𝑥)) is 𝜈0-integrable, and Π𝐼(𝑓) is a bounded operator of norm Π𝐼(𝑓)𝑓𝐼. We have to check that Π𝐼 is a -homomorphism. For this, we define 𝐹𝑡(𝑥,𝑡)=𝑓(𝑥𝑡)𝑔1𝜋(𝑥)𝜉(𝑠(𝑥)),𝜂(𝑟(𝑥))𝑑1/2(𝑥).(5.11) Then Π𝐼(𝑓𝑔)𝜉,𝜂=𝐺𝑓𝑔(𝑥)𝜋(𝑥)𝜉(𝑠(𝑥)),𝜂(𝑟(𝑥))𝑑𝜈0(=𝑥)𝐺𝐺𝑡𝑓(𝑥𝑡)𝑔1𝑑𝜆𝑠(𝑥)(𝑡)𝜋(𝑥)𝜉𝑠(𝑥),𝜂𝑟(𝑥)𝑑1/2(𝑥)𝐷(𝑥)𝑑𝜈1=(𝑥)𝐺2𝐹(𝑥,𝑡)𝑑𝜈2=(𝑥,𝑡)𝐺2𝐹𝑥𝑡,𝑡1𝜌1𝑥𝑡,𝑡1𝑑𝜈2=(𝑥,𝑡)𝐺0𝐺𝑢𝑓(𝑥)𝜋(𝑥)𝐺𝑠(𝑥)𝑔(𝑡)𝜋(𝑡)𝜉𝑠(𝑡)𝐷1/2(𝑡)𝑑𝜆𝑠(𝑥)(𝑡),𝜂(𝑢)×𝐷1/2(𝑥)𝑑𝜆𝑢=(𝑥)𝑑𝜇(𝑢)𝐺𝑓(𝑥)𝜋(𝑥)Π(𝑔)𝜉(𝑠(𝑥)),𝜂(𝑟(𝑥))𝑑𝜈0(𝑥)=Π𝐼(𝑓)Π𝐼(𝑔)𝜉,𝜂.(5.12) Hence, Π𝐼(𝑓𝑔)=Π𝐼(𝑓)Π𝐼(𝑔). Also Π𝐼𝑓𝜉,𝜂=𝐺𝑓(𝑥)𝜋(𝑥)𝜉(𝑠(𝑥)),𝜂(𝑟(𝑥))𝑑𝜈0(=𝑥)𝐺𝑓𝑥1𝐷𝑥1𝜋(𝑥)𝜉(𝑠(𝑥)),𝜂(𝑟(𝑥))𝑑1/2=(𝑥)𝑑𝜈(𝑥)𝐺𝑥𝑓(𝑥)𝐷1𝜋(𝑥)𝜉(𝑟(𝑥)),𝜂(𝑠(𝑥))𝑑1/2𝑥(𝑥)𝐷1=𝑑𝜈(𝑥)𝐺𝑓(𝑥)𝜂(𝑠(𝑥)),𝜋(𝑥)𝜉(𝑟(𝑥))𝑑𝜈0=(𝑥)𝐺𝑓(𝑥)𝜋(𝑥)𝜂(𝑠(𝑥)),𝜉(𝑟(𝑥))𝑑𝜈0=(𝑥)Π𝐼(𝑓)𝜉,𝜂=𝜉,Π𝐼Π(𝑓)𝜂=𝐼.(𝑓)𝜉,𝜂(5.13) The nondegeneracy follows from Π𝐼(𝑓)𝜉𝑓𝐶𝑐Π(𝐺),𝜉𝐻𝐼(𝑓)𝜉𝑓𝐿1𝐺,𝐼,𝜉𝐻𝐻.(5.14)

Now let us turn to the problem that a continuous representation of 𝐼(𝐺,𝜆,𝜇) is not necessarily continuous in 𝐿1-norm. Let 𝐻=𝐿2(𝐺0,{𝐻𝑢}𝑢,𝜇) and put 𝐻Π1=𝜉𝐻Themap𝑓Π𝐼(𝑓)𝜉,𝜉iscontinuousin𝐿1-norm.(5.15) Observe that 𝐻Π1 is a nontrivial subspace, as 𝜇 is a probability measure and if 𝜉(𝑢)=1 in 𝐻𝑢 for each 𝑢𝐺0, then 𝜉=1 in 𝐻 and the calculations in the proof of Lemma 5.5 shows that |Π𝐼(𝑓)𝜉,𝜉|𝑓1, hence 𝜉𝐻1. On the other hand, Π𝐼(𝑓)𝐻Π1𝐻Π1 for each 𝑓𝐼(𝐺,𝜆,𝜇), because the map 𝑓Π𝐼(𝑓)Π𝐼(𝑔)𝜉,Π𝐼Π(𝑔)𝜉=𝐼𝑔𝑓𝑔𝜉,𝜉(5.16) is continuous. Therefore, Π𝐼(𝑓)𝐻Π1𝐻Π1. Hence, we have Π𝐼(𝑓)=Π𝐼(𝑓)|𝐻Π1Π𝐼(𝑓)|𝐻Π1 for each 𝑓𝐼(𝐺,𝜆,𝜇).

Put 𝐻1=𝐻Π1 and define Π1(𝑓)=Π𝐼(𝑓)|𝐻1, for 𝑓𝐼(𝐺,𝜆,𝜇). Then, it follows from continuity of 𝑓Π𝐼(𝑓)𝜉,𝜉 and polarization identity that Π1 extends to a continuous representation of 𝐿1(𝐺,𝜆,𝜇) on 𝐻1, still denoted by Π1.

Next we focus on the notion of irreducibility which plays an important role in the theory of representations. We show that if 𝜋 is an irreducible representation of 𝐺, the integrated representation Π1 of 𝐿1(𝐺,𝜆,𝜇) on 𝐻1 is irreducible. Basic materials come from [6].

Definition 5.6. Let (𝐺0,{𝐻𝑢}𝑢𝐺0,𝜇) be a Hilbert bundle. A family 𝑀={𝑀𝑢}𝑢𝐺0, where 𝑀𝑢 is a closed subspace of 𝐻𝑢 for each 𝑢𝐺0, is called a subbundle. A subbundle {𝑀𝑢}𝑢𝐺0 is called nontrivial if 0𝑀𝑢𝐻𝑢 for some 𝑢𝐺0. For a representation 𝜋 of a locally compact groupoid 𝐺 associated with the Hilbert bundle (𝐺0,{𝐻𝑢}𝑢𝐺0,𝜇), a subbundle {𝑀𝑢}𝑢𝐺0 is called invariant if 𝜋(𝑥)𝑀𝑠(𝑥)𝑀𝑟(𝑥) for each 𝑥𝐺. Note that if 𝑀 is an invariant subbundle, and 0𝑀𝑢𝐻𝑢 for some 𝑢𝐺0, then 0𝑀𝑤𝐻𝑤 for every 𝑤[𝑢].

The following lemma is proved in [6, Lemma  3.4].

Lemma 5.7. Let 𝜋 be a representation of a locally compact groupoid 𝐺 associated with the Hilbert bundle (𝐺0,{𝐻𝑢}𝑢𝐺0,𝜇). If 𝑀={𝑀𝑢}𝑢𝐺0 is an invariant subbundle, then so is 𝑀={𝑀𝑢}𝑢𝐺0.

Definition 5.8. A representation 𝜋 of a locally compact groupoid 𝐺 is called reducible if 𝜋 admits a nontrivial invariant subbundle 𝑀={𝑀𝑢}𝑢𝐺0, otherwise 𝜋 is called irreducible. In this case, it is easy to check that 𝜋𝑀 with 𝜋𝑀(𝑥)=𝜋(𝑥)|𝑀𝑠(𝑥)𝑀𝑠(𝑥)𝑀𝑟(𝑥) is called a subrepresentation of 𝜋. If 𝜋 and 𝜋 are two representations of a locally compact groupoid 𝐺 associated with two Hilbert bundles (𝐺0,{𝐻𝑢}𝑢𝐺0,𝜇) and (𝐺0,{𝐻𝑢}𝑢𝐺0,𝜇), respectively. Then, we put 𝒞𝜋,𝜋=𝑇𝑢𝑢𝑢𝐺0𝐻𝑢,𝐻𝑢𝜋(𝑥)𝑇𝑠(𝑥)=𝑇𝑟(𝑥)𝜋(𝑥)(𝑥𝐺)(5.17) and write 𝒞(𝜋,𝜋)=𝒞(𝜋).

Two representations 𝜋 and 𝜋 are called (unitarily) equivalent if 𝜇𝜇, and there is (𝑇𝑢)𝑢𝐺0𝒞(𝜋,𝜋) such that 𝑇𝑢 is a unitary operator for every 𝑢𝐺0. Note that if (𝑇𝑢)𝑢𝐺0𝒞(𝜋) and 𝑇𝑢 denotes the adjoint operator to 𝑇𝑢, then 𝜋(𝑥)𝑇𝑠(𝑥)=𝑇𝑠(𝑥)𝜋𝑥1=𝜋𝑥1𝑇𝑟(𝑥)=𝑇𝑟(𝑥)𝜋(𝑥),(5.18) hence, (𝑇𝑢)𝑢𝐺0𝒞(𝜋). We observe that 𝒞(𝜋) is a unital -algebra, where the operations are defined pointwise.

Following [6], for a representation 𝜋 of a locally compact groupoid 𝐺 associated with the Hilbert bundle (𝐺0,{𝐻𝑢}𝑢𝐺0,𝜇), we set 𝜆Λ=𝑢𝜋(𝑢)𝑢𝑢𝐺0𝐻𝑢𝜆𝑢,𝜆𝑢=𝜆𝑤whenever.𝑢𝑤(5.19) If (𝜆𝑢𝜋(𝑢))𝑢𝒞(𝜋), then 𝜆𝑢=𝜆𝑤 whenever 𝑢𝑤 that is (𝜆𝑢𝜋(𝑢))𝑢Λ. Therefore, if (𝑇𝑢)𝑢𝐺0𝒞(𝜋)Λ, then there exists 𝑢𝐺0 with 𝑇𝑢 not in 𝜋(𝑢). It is also obvious that Λ𝒞(𝜋).

We need the following version of Schur's lemma [6, Lemma  3.11].

Lemma 5.9. A representation 𝜋 of a locally compact groupoid 𝐺 is irreducible if and only if Λ=𝒞(𝜋). In particular, in the case where 𝐺 is transitive, then 𝜋 is irreducible if and only if 𝒞(𝜋)=(𝜋(𝑢))𝑢𝐺0.

Lemma 5.10. If =(𝐺0,{𝑀𝑢}𝑢,𝜇) is closed subbundle of =(𝐺0,{𝐻𝑢}𝑢,𝜇) and 𝑃𝑢 is an orthogonal projection 𝐻𝑢 onto 𝑀𝑢, then is invariant under 𝜋 if and only if (𝑃𝑢)𝑢𝒞(𝜋).

Proof. If (𝑃𝑢)𝑢𝒞(𝜋) and 𝑚𝑢𝑀𝑢, then 𝜋(𝑥)𝑚𝑠(𝑥)=𝜋(𝑥)𝑃𝑠(𝑥)𝑚𝑠(𝑥)=𝑃𝑟(𝑥)𝜋(𝑥)𝑚𝑠(𝑥)𝑀𝑟(𝑥).(5.20) Thus, 𝜋(𝑥)𝑀𝑠(𝑥)𝑀𝑟(𝑥) so is invariant. Conversely, if is invariant, then for 𝑚𝑀𝑠(𝑥)𝜋(𝑥)𝑃𝑠(𝑥)𝑚=𝜋(𝑥)𝑚=𝑃𝑟(𝑥)𝜋(𝑥)𝑚(5.21) and for 𝑚𝑀𝑠(𝑥)𝜋(𝑥)𝑃𝑠(𝑥)𝑚=0=𝑃𝑟(𝑥)𝜋(𝑥)𝑚.(5.22) Hence, 𝜋(𝑥)𝑃𝑠(𝑥)=𝑃𝑟(𝑥)𝜋(𝑥).

We show that if a representation 𝜋 of a locally compact groupoid 𝐺 associated with the Hilbert bundle (𝐺0,{𝐻𝑢}𝑢𝐺0,𝜇) is irreducible and Π1 is the corresponding integrated representation of 𝐿1(𝐺,𝜆,𝜇) on 𝐻1𝐿2(𝐺0,{𝐻𝑢}𝑢,𝜇), then Π1 is irreducible. If 𝐻=𝐿2(𝐺0,{𝐻𝑢}𝑢,𝜇), then for each subbundle {𝑀𝑢}𝑢, 𝑀=𝐿2(𝐺0,{𝑀𝑢}𝑢,𝜇) is closed subspace of 𝐻 with orthogonal complement 𝑀=𝐿2(𝐺0,{𝑀𝑢}𝑢,𝜇). A map Θ𝐿2(𝐺0,{𝐻𝑢}𝑢,𝜇)𝐿2(𝐺0,{𝑀𝑢}𝑢,𝜇);𝜉Θ𝜉 with Θ𝜉(𝑠(𝑥))𝑀𝑟(𝑥) is called an orthogonal projection of 𝐻 onto 𝑀.

Let 𝜋 be a representation of 𝐺 on the Hilbert bundle (𝐺0,{𝐻𝑢}𝑢𝐺0,𝜇) and Π𝐼 be the bounded representation of 𝐼(𝐺,𝜆,𝜇) on (𝐿2(𝐺0,{𝐻𝑢}𝑢,𝜇)) constructed above.

Lemma 5.11. 𝑀 is invariant under Π𝐼 if and only if for each 𝑓𝐼(𝐺,𝜆,𝜇), Π𝐼(𝑓)Θ=ΘΠ𝐼(𝑓).

Proof. Suppose Π𝐼(𝑓)Θ=ΘΠ𝐼(𝑓) and 𝜉𝑀, then Π𝐼(𝑓)𝜉=Π𝐼(𝑓)Θ𝜉=ΘΠ𝐼(𝑓)𝜉𝑀.(5.23) Hence, 𝑀 is invariant. Conversely, if 𝑀 is invariant, then for 𝜉𝑀, Π𝐼(𝑓)Θ𝜉=Π𝐼(𝑓)𝜉=ΘΠ𝐼(𝑓)𝜉(5.24) and for 𝜂𝑀, Π𝐼(𝑓)Θ𝜂=0=ΘΠ𝐼(𝑓)𝜂.(5.25) Thus, Π𝐼(𝑓)Θ=ΘΠ𝐼(𝑓).

Proposition 5.12. If 𝜋 is an irreducible representation of 𝐺, then Π𝐼 is an irreducible representation of 𝐼(𝐺,𝜆,𝜇).

Proof. If Π𝐼 is reducible, then there exists a nontrivial invariant closed subspace 𝑀 of 𝐿2(𝐺0,{𝐻𝑢}𝑢,𝜇), and, hence, there is an orthogonal projection Θ𝒞(Π𝐼). It follows that 𝐺𝑓(𝑥)𝜋(𝑥)𝑃𝑠(𝑥)𝜉(𝑠(𝑥)),𝜂(𝑟(𝑥))𝑑𝜈0(𝑥)=Π𝐼(𝑓)Θ𝜉,𝜂=ΘΠ𝐼=(𝑓)𝜉,𝜂𝐺𝑃𝑓(𝑥)𝑟(𝑥)𝜋(𝑥)𝜉(𝑠(𝑥)),𝜂(𝑟(𝑥))𝑑𝜈0(𝑥),(5.26) for each 𝑓𝐶𝑐(𝐺) and 𝜉,𝜂𝐿2(𝐺0,{𝐻𝑢}𝑢,𝜇). Therefore, 𝜋(𝑥)𝑃𝑠(𝑥)=𝑃𝑟(𝑥)𝜋(𝑥), for each 𝑥𝐺, hence (𝑃𝑢)𝑢𝒞(𝜋). Now by the Schur's lemma, 𝜋 is reducible.

Theorem 5.13. If 𝜋 is an irreducible representation of 𝐺, then Π1 is an irreducible representation of 𝐿1(𝐺,𝜆,𝜇) on 𝐻1.

Proof. If Π1 is reducible, then there exists a nontrivial invariant closed subspace 𝑀1 of 𝐻1𝐻=𝐿2(𝐺0,{𝐻𝑢}𝑢,𝜇). By the calculations after Lemma 5.5, Π𝐼=Π1Π2, where Π1 and Π2 are the corresponding representations of 𝐼(𝐺,𝜆,𝜇) on 𝐻1 and 𝐻1. Therefore, 𝑀=𝑀1𝐻1 is a nontrivial invariant closed subspace 𝐻, and Π𝐼 is reducible. Hence, 𝜋 is reducible by the above proposition.