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

\contact

[carrillo@math.jussieu.fr]Paulo Carrillo Rouse
Projet d’algèbres d’opérateurs
Université de Paris 7
175, rue de Chevaleret
Paris, France

Compactly supported analytic indices for Lie groupoids

Paulo Carrillo Rouse
Abstract

For any Lie groupoid we construct an analytic index morphism taking values in a modified KtheoryK-theory group which involves the convolution algebra of compactly supported smooth functions over the groupoid. The construction is performed by using the deformation algebra of smooth functions over the tangent groupoid constructed in [CR06]. This allows in particular to prove a more primitive version of the Connes-Skandalis Longitudinal index Theorem for foliations, that is, an index theorem taking values in a group which pairs with Cyclic cocycles. As other application, for DD a 𝒢\mathscr{G}-PDO elliptic operator with associated index indDK0(Cc(𝒢))ind\,D\in K_{0}(C^{\infty}_{c}(\mathscr{G})), we prove that the pairing

<indD,τ>,<ind\,D,\tau>,

with τ\tau a bounded continuous cyclic cocycle, only depends on the principal symbol class [σ(D)]K0(A𝒢)[\sigma(D)]\in K^{0}(A^{*}\mathscr{G}). The result is completely general for Étale groupoids. We discuss some potential applications to the Novikov’s conjecture.

keywords:
Lie groupoids, Tangent groupoid, K-theory, Cyclic cohomology, Index theory.

Primary 19K56; Secondary 58J42,53C10.

1 Introduction

Index theory has as a starting point the Atiyah-Singer index theorem [AS68]: Let DD be an elliptic operator over a compact smooth manifold MM, in particular DD has finite dimensional Kernel and Cokernel and the Fredholm index can be defined as

indD:=dimKerDdimCokerD.ind\,D:=dim\,Ker\,D-dim\,Coker\,D\in\mathbb{Z}.

The map DindDD\mapsto ind\,D is completely codified by a group morphism K0(TM),K^{0}(T^{*}M)\longrightarrow\mathbb{Z}, called the analytic index of MM. That is, if Ell(M)Ell(M) denotes the set of elliptic pseudodifferential operators over MM, then the following diagram is commutative:

Ell(M)\textstyle{Ell(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ind\scriptstyle{ind}σ\scriptstyle{\sigma}\textstyle{\mathbb{Z}}K0(TM)\textstyle{K^{0}(T^{*}M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,M\scriptstyle{ind_{a,M}}, (1)

where Ell(M)σK0(TM)Ell(M)\stackrel{{\scriptstyle\sigma}}{{\longrightarrow}}K^{0}(T^{*}M) is the surjective map that associates the class of the principal symbol of the operator in K0(TM)K^{0}(T^{*}M). This is a fundamental property because it allows to use the (cohomological) properties of KK-theory and gives stability to the index. The topological formula given by Atiyah-Singer for the analytic index is the following:

inda,M([σP])=TMch([σP])Td(M).ind_{a,M}([\sigma_{P}])=\int_{T^{*}M}ch([\sigma_{P}])Td(M). (2)

This kind of formulas give very interesting invariants of the manifold [LM89, Mel93, Sha78].

We discuss now the Lie groupoids case. This concept is central in non commutative geometry. Groupoids generalize the concepts of spaces, groups and equivalence relations. In the late 70’s, mainly with the work of Alain Connes, it became clear that groupoids appeared naturally as substitutes of singular spaces [Con79, Mac87, Ren80, Pat99]. Furthermore, Connes showed that many groupoids and algebras associated to them appeared as ‘non commutative analogues‘ of smooth manifolds to which many tools of geometry such as K-theory and Characteristic classes could be applied [Con79, Con94]. Lie groupoids became a very natural place where to perform pseudodifferential calculus and index theory, [Con79, MP97, NWX99].

The study of the indices in the groupoid case is, as we will see, more delicate than the classical case. There are new phenomena appearing. If 𝒢\mathscr{G} is a Lie groupoid, a 𝒢\mathscr{G}-pseudodifferential operator is a differentiable family (see [MP97, NWX99]) of operators. Let PP be such an operator, the index of PP, indPind\,P, is an element of K0(Cc(𝒢))K_{0}(C_{c}^{\infty}(\mathscr{G})). There is a diagram similar to (1) for the Lie groupoid setting:

Ell(𝒢)\textstyle{Ell(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ind\scriptstyle{ind}symb\scriptstyle{symb}K0(Cc(𝒢))\textstyle{K_{0}(C_{c}^{\infty}(\mathscr{G}))}K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\nexists}. (3)

We would like, as above, that the map DindDK0(Cc(𝒢))D\mapsto ind\,D\in K_{0}(C_{c}^{\infty}(\mathscr{G})) factor through K0(A𝒢)K^{0}(A^{*}\mathscr{G}): this group codifies the principal symbols of all the elliptic 𝒢\mathscr{G}-operators and carries the topological invariants. Unfortunately, the map DindDK0(Cc(𝒢))D\mapsto ind\,D\in K_{0}(C_{c}^{\infty}(\mathscr{G})) does not factor in general through the principal symbol [Con94]. Nevertheless, if we consider the KK-theory morphism K0(Cc(𝒢))jK0(Cr(𝒢))K_{0}(C_{c}^{\infty}(\mathscr{G}))\stackrel{{\scriptstyle j}}{{\longrightarrow}}K_{0}(C^{*}_{r}(\mathscr{G})) induced by the inclusion Cc(𝒢)Cr(𝒢)C_{c}^{\infty}(\mathscr{G})\subset C^{*}_{r}(\mathscr{G}), then the composition

Ell(𝒢)indK0(Cc(𝒢))jK0(Cr(𝒢))Ell(\mathscr{G})\stackrel{{\scriptstyle ind}}{{\longrightarrow}}K_{0}(C_{c}^{\infty}(\mathscr{G}))\stackrel{{\scriptstyle j}}{{\longrightarrow}}K_{0}(C^{*}_{r}(\mathscr{G}))

do factor through the principal symbol by a morphism indaind_{a}, called the analytic index of 𝒢\mathscr{G}:

Ell(𝒢)\textstyle{Ell(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ind\scriptstyle{ind}symb\scriptstyle{symb}K0(Cc(𝒢))\textstyle{K_{0}(C_{c}^{\infty}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j\scriptstyle{j}K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda\scriptstyle{ind_{a}}K0(Cr(𝒢)).\textstyle{K_{0}(C^{*}_{r}(\mathscr{G})).} (4)

In fact, the CC^{*}-analytic index indaind_{a} can be completely described using the Connes tangent groupoid, [Con94, MP97, MN05, DLN06, ANS06, HS87]. We briefly recall this fact.

The tangent groupoid associated to a Lie groupoid 𝒢𝒢(0)\mathscr{G}\rightrightarrows\mathscr{G}^{(0)} is a Lie groupoid

𝒢T:=A𝒢×{0}𝒢×(0,1]𝒢(0)×[0,1],\mathscr{G}^{T}:=A\mathscr{G}\times\{0\}\bigsqcup\mathscr{G}\times(0,1]\rightrightarrows\mathscr{G}^{(0)}\times[0,1],

that is compatible with A𝒢A\mathscr{G} and 𝒢\mathscr{G}. Its CC^{*}algebra is a continuous field of CC^{*}-algebras over [0,1][0,1] whose fiber at zero is Cr(A𝒢)C0(A𝒢)C_{r}^{*}(A\mathscr{G})\cong C_{0}(A^{*}\mathscr{G}) and Cr(𝒢)C_{r}^{*}(\mathscr{G}) elsewhere, it is a CC^{*}-algebraic deformation in the sense of [Lan03]. In particular it gives a short exact sequence of CC^{*}-algebras

0Cr(𝒢×(0,1])Cr(𝒢T)ev0Cr(A𝒢)0,0\rightarrow C_{r}^{*}(\mathscr{G}\times(0,1])\longrightarrow C_{r}^{*}(\mathscr{G}^{T})\stackrel{{\scriptstyle ev_{0}}}{{\longrightarrow}}C_{r}^{*}(A\mathscr{G})\rightarrow 0, (5)

Now, thanks to the good properties of the KK-theory for CC^{*}-algebras, Monthubert and Pierrot show in [MP97] that

inda=(ev1)(ev0)1,ind_{a}=(ev_{1})_{*}\circ(ev_{0})_{*}^{-1}, (6)

where Cr(𝒢T)ev1Cr(𝒢)C_{r}^{*}(\mathscr{G}^{T})\stackrel{{\scriptstyle ev_{1}}}{{\longrightarrow}}C_{r}^{*}(\mathscr{G}) is the evaluation at 11.

If we try to adapt the above arguments for the index indDK0(Cc(𝒢))ind\,D\in K_{0}(C_{c}^{\infty}(\mathscr{G})), we find several difficulties: For instance, the map K0(Cc(A𝒢))K0(Cr(A𝒢))=K0(A𝒢)K_{0}(C^{\infty}_{c}(A\mathscr{G}))\longrightarrow K_{0}(C^{*}_{r}(A\mathscr{G}))=K^{0}(A^{*}\mathscr{G}) is not an isomorphism, i.e., K0(Cc(A𝒢))K_{0}(C^{\infty}_{c}(A\mathscr{G})) is not the group that we would like. On the other hand, the kernel of the evaluation map Cc(𝒢T)ev0Cc(A𝒢)C^{\infty}_{c}(\mathscr{G}^{T})\stackrel{{\scriptstyle ev_{0}}}{{\longrightarrow}}C^{\infty}_{c}(A\mathscr{G}) is not easy to describe, and we cannot conclude its K0K_{0} group vanishes since the KK-theory for general topological algebras is not necessarily homotopy invariant. In general, the morphism K0(Cc(𝒢))jK0(Cr(𝒢))K_{0}(C_{c}^{\infty}(\mathscr{G}))\stackrel{{\scriptstyle j}}{{\rightarrow}}K_{0}(C^{*}_{r}(\mathscr{G})) is not an isomorphism. A very simple example of this situation is {}\mathbb{R}\rightrightarrows\{*\} with the vectorial group structure ([Con94], p.142). When the groupoid is proper we do not perceive this nuance: it is an isomorphism. This is because Cc(𝒢)C_{c}^{\infty}(\mathscr{G}) is stable under holomorphic calculus ([CMR07, Kar78]). In general, the KK-theory groups as K0(Cc(𝒢))K_{0}(C_{c}^{\infty}(\mathscr{G})) do not satisfy in general some fundamental properties as homotopy invariance and Bott periodicity.

Our first step is to consider a ”good quotient” of K0(Cc(𝒢))K_{0}(C_{c}^{\infty}(\mathscr{G})) where we are going to be able to factorize [indD]K0(Cc(𝒢))/[ind\,D]\in K_{0}(C_{c}^{\infty}(\mathscr{G}))/\sim through the principal symbol, using the Connes tangent groupoid.

Let us note, for each kk\in\mathbb{N}, K0h,k(𝒢)K_{0}^{h,k}(\mathscr{G}) the quotient group of K0(Cck(𝒢))K_{0}(C_{c}^{k}(\mathscr{G})) by the equivalence relation induced by K0(Cck(𝒢×[0,1]))e1e0K0(Cck(𝒢))\tiny{K_{0}(C_{c}^{k}(\mathscr{G}\times[0,1]))\overset{e_{0}}{\underset{e_{1}}{\rightrightarrows}}K_{0}(C_{c}^{k}(\mathscr{G}))}. Let K0F(𝒢)=limkK0h,k(𝒢)K_{0}^{F}(\mathscr{G})=\varprojlim_{k}K_{0}^{h,k}(\mathscr{G}) be the projective limit relative to the inclusions Cck(𝒢)Cck1(𝒢)C_{c}^{k}(\mathscr{G})\subset C_{c}^{k-1}(\mathscr{G}). The main result of this paper is the following.

Theorem 1

There exists a group morphism

indaF:K0(A𝒢)K0F(𝒢)ind_{a}^{F}:K^{0}(A^{*}\mathscr{G})\rightarrow K_{0}^{F}(\mathscr{G})

such that the following diagram is commutative

Ell(𝒢)\textstyle{Ell(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}symb\scriptstyle{symb}ind\scriptstyle{ind}K0(Cc(𝒢))\textstyle{K_{0}(C^{\infty}_{c}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K0(A𝒢)\textstyle{K_{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{id}indaF\scriptstyle{ind_{a}^{F}}K0F(𝒢)\textstyle{K_{0}^{F}(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K0(A𝒢)\textstyle{K_{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda\scriptstyle{ind_{a}}K0(Cr(𝒢))\textstyle{K_{0}(C_{r}^{*}(\mathscr{G}))} (7)

The index of the above theorem is called “the compactly supported analytic index of 𝒢\mathscr{G}. The construction of these indices is also based on the Connes tangent groupoid, as in case of CC^{*}-algebras. In fact, in [CR06] we constructed an algebra of smooth functions over the tangent groupoid that will allow in this paper to perform the compactly supported indices as ”deformations”. We recall the main result of [CR06].

Theorem

There exists an intermediate algebra 𝒮c(𝒢T)\mathscr{S}_{c}(\mathscr{G}^{T}) consisting of smooth functions over the tangent groupoid

Cc(𝒢T)𝒮c(𝒢T)Cr(𝒢T),C_{c}^{\infty}(\mathscr{G}^{T})\subset\mathscr{S}_{c}(\mathscr{G}^{T})\subset C_{r}^{*}(\mathscr{G}^{T}),

such that it is a field of algebras over [0,1][0,1], whose fibers are

𝒮(A𝒢) at t=0, and \mathscr{S}(A\mathscr{G})\text{ at }t=0,\text{ and }
Cc(𝒢) for t0.C_{c}^{\infty}(\mathscr{G})\text{ for }t\neq 0.

The compactly supported index fits in a commutative diagram of the following type:

K0(𝒮c(𝒢T))\textstyle{K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e1F\scriptstyle{e_{1}^{F}}e0\scriptstyle{e_{0}}K0(𝒮(A𝒢))\textstyle{K_{0}(\mathscr{S}(A\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}indaF\scriptstyle{ind_{a}^{F}}K0F(𝒢)\textstyle{K_{0}^{F}(\mathscr{G})}.

Applications

Longitudinal index theorem

First, we slightly modify the indices defined above. We take K0B(𝒢)=limmK0F(𝒢×2m)K_{0}^{B}(\mathscr{G})=\varinjlim_{m}K_{0}^{F}(\mathscr{G}\times\mathbb{R}^{2m}), the inductive limit induced by the Bott morphisms K0F(𝒢×2m)BottK0F(𝒢×2(m+1))K_{0}^{F}(\mathscr{G}\times\mathbb{R}^{2m})\stackrel{{\scriptstyle Bott}}{{\longrightarrow}}K_{0}^{F}(\mathscr{G}\times\mathbb{R}^{2(m+1)}). Let

inda,𝒢B:K0(A𝒢)K0B(𝒢)ind_{a,\mathscr{G}}^{B}:K^{0}(A^{*}\mathscr{G})\rightarrow K_{0}^{B}(\mathscr{G})

be the morphism given by the composition of inda,𝒢F:K0(A𝒢)K0F(𝒢)ind_{a,\mathscr{G}}^{F}:K^{0}(A^{*}\mathscr{G})\rightarrow K_{0}^{F}(\mathscr{G}) followed by K0F(𝒢)BottK0B(𝒢)K_{0}^{F}(\mathscr{G})\stackrel{{\scriptstyle Bott}}{{\longrightarrow}}K_{0}^{B}(\mathscr{G}). This morphism satisfies also theorem 1, and so it is called the Periodic compactly supported index of 𝒢\mathscr{G}.

The periodic index can be calculated by topological methods in the case of foliations. That is, we have a longitudinal index theorem that reinforce the Connes-Skandalis one, [CS84]. Indeed, we will show that the equality between the analytic and the topological indices takes already place in K0B(𝒢)K_{0}^{B}(\mathscr{G}).

Theorem 2

Let (M,F)(M,F) be foliated manifold with holonomy groupoid 𝒢\mathscr{G}. Then, the topological index (à la Connes-Skandalis) can be defined with values in K0B(𝒢)K_{0}^{B}(\mathscr{G}) and it coincides with the Periodic compactly supported index of 𝒢\mathscr{G}:

inda,𝒢B=indt,𝒢B.ind_{a,\mathscr{G}}^{B}=ind_{t,\mathscr{G}}^{B}.

The reinforced longitudinal index theorem allows us to define an assembly map in our setting:

μF:K,τ(B𝒢)K0B(𝒢),\mu_{F}:K_{*,\tau}(B\mathscr{G})\rightarrow K_{0}^{B}(\mathscr{G}), (8)

given by μF(δD)=indaB(σD)\mu_{F}(\delta_{D})=ind_{a}^{B}(\sigma_{D}). It fits in the following commutative diagram

K,τ(B𝒢)\textstyle{K_{*,\tau}(B\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ\scriptstyle{\mu}μF\scriptstyle{\mu_{F}}K0(Cr(𝒢))\textstyle{K_{0}(C_{r}^{*}(\mathscr{G}))}K0B(𝒢),\textstyle{K_{0}^{B}(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces,}i\scriptstyle{i} (9)

where μ\mu is the classical Baum-Connes map, [BCH94, Tu00].

Pairings with cyclic cohomology

The interest to keep track on the CcC^{\infty}_{c}-indices is because at this level we can make explicit calculations via the Chern-Weil-Connes theory. In fact there is a pairing [Con85, Con94, Kar87]

_,_:K0(Cc(𝒢))×HP(Cc(𝒢))\langle\_\,,\_\rangle:K_{0}(C_{c}^{\infty}(\mathscr{G}))\times HP^{*}(C_{c}^{\infty}(\mathscr{G}))\rightarrow\mathbb{C} (10)

There are several known cocycles over Cc(𝒢)C_{c}^{\infty}(\mathscr{G}). An important problem in Noncommutative Geometry is to calculate the above pairing in order to obtain numerical invariants from the indices in K0(Cc(𝒢))K_{0}(C_{c}^{\infty}(\mathscr{G})), [Con94, CM90, GL06].

Now, we can expect an easy (topological) calculation only if the map DD,τD\mapsto\langle D\,,\tau\rangle (τHP(Cc(𝒢))\tau\in HP^{*}(C_{c}^{\infty}(\mathscr{G})) fix) factors through the symbol class of DD, [σ(D)]K0(A𝒢)[\sigma(D)]\in K^{0}(A^{*}\mathscr{G}): we want to have a diagram of the following kind:

Ell(𝒢)\textstyle{Ell(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ind\scriptstyle{ind}symb.\scriptstyle{symb.}K0(Cc(𝒢))\textstyle{K_{0}(C_{c}^{\infty}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}_,τ\scriptstyle{\langle\_,\tau\rangle}\textstyle{\mathbb{C}}K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ\scriptstyle{\tau}.

The next step in this work will consist of solving this factorization problem, using the compactly supported indices. In fact, the Chern-Connes theory applies naturally to algebras as Cck(𝒢)C_{c}^{k}(\mathscr{G}). Moreover, the pairing with Periodic cyclic cohomology preserves the relation h\sim_{h} and is compatible with Bott morphism. The result is the following:

Theorem 3 (Factorization theorem)

Let τ\tau be a bounded continuous cocycle. Then τ\tau defines a morphism Ψτ:K0(A𝒢)\Psi_{\tau}:K^{0}(A^{*}\mathscr{G})\rightarrow\mathbb{C} such that the following diagram is commutative

Ell(𝒢)\textstyle{Ell(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ind\scriptstyle{ind}symb.\scriptstyle{symb.}K0(Cc(𝒢))\textstyle{K_{0}(C_{c}^{\infty}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}_,τ\scriptstyle{\langle\_,\tau\rangle}\textstyle{\mathbb{C}}K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψτ\scriptstyle{\Psi_{\tau}}.

The restriction of taking bounded continuous cyclic cocycles in the last theorem is not at all restrictive. In fact, all the geometrical cocycles are of this kind (Group cocycles, The transverse fundamental class, Godbillon-Vey and all the Gelfand-Fuchs cocycles for instance). Moreover, for the case of étale groupoids, the explicit calculations of the Periodic cohomologies spaces developed in [BN94, Cra99] allow us to conclude that the above result is completely general in this setting (see Theorem 6.9). In the last part of this work we will discuss how to recover some known index formulas for foliations.

Acknowledgments The present work is part of my PHD thesis at the University of Paris 7, [CR07]. I would like to thank my advisor, Professor Georges Skandalis, for proposing me this subject and for his guide and support.

2 Lie groupoids

Let us recall what a groupoid is:

Definition 2.1.

A 𝑔𝑟𝑜𝑢𝑝𝑜𝑖𝑑\it{groupoid} consists of the following data: two sets 𝒢\mathscr{G} and 𝒢(0)\mathscr{G}^{(0)}, and maps

  • \cdot

    s,r:𝒢𝒢(0)s,r:\mathscr{G}\rightarrow\mathscr{G}^{(0)} called the source and target map respectively,

  • \cdot

    m:𝒢(2)𝒢m:\mathscr{G}^{(2)}\rightarrow\mathscr{G} called the product map (where 𝒢(2)={(γ,η)𝒢×𝒢:s(γ)=r(η)}\mathscr{G}^{(2)}=\{(\gamma,\eta)\in\mathscr{G}\times\mathscr{G}:s(\gamma)=r(\eta)\}),

such that there exist two maps, u:𝒢(0)𝒢u:\mathscr{G}^{(0)}\rightarrow\mathscr{G} (the unit map) and i:𝒢𝒢i:\mathscr{G}\rightarrow\mathscr{G} (the inverse map), such that, if we note m(γ,η)=γηm(\gamma,\eta)=\gamma\cdot\eta, u(x)=xu(x)=x and i(γ)=γ1i(\gamma)=\gamma^{-1}, we have

  • 1.

    r(γη)=r(γ)r(\gamma\cdot\eta)=r(\gamma) and s(γη)=s(η)s(\gamma\cdot\eta)=s(\eta).

  • 2.

    γ(ηδ)=(γη)δ\gamma\cdot(\eta\cdot\delta)=(\gamma\cdot\eta)\cdot\delta, γ,η,δ𝒢\forall\gamma,\eta,\delta\in\mathscr{G} when this is possible.

  • 3.

    γx=γ\gamma\cdot x=\gamma and xη=ηx\cdot\eta=\eta, γ,η𝒢\forall\gamma,\eta\in\mathscr{G} with s(γ)=xs(\gamma)=x and r(η)=xr(\eta)=x.

  • 4.

    γγ1=u(r(γ))\gamma\cdot\gamma^{-1}=u(r(\gamma)) and γ1γ=u(s(γ))\gamma^{-1}\cdot\gamma=u(s(\gamma)), γ𝒢\forall\gamma\in\mathscr{G}.

Generally, we denote a groupoid by 𝒢𝒢(0)\mathscr{G}\rightrightarrows\mathscr{G}^{(0)}.

Along this paper we will only deal with Lie groupoids, that is, a groupoid in which 𝒢\mathscr{G} and 𝒢(0)\mathscr{G}^{(0)} are smooth manifolds (possibly with boundary), and s,r,m,us,r,m,u are smooth maps (with s and r submersions, see [Mac87, Pat99]). For A,BA,B subsets of 𝒢(0)\mathscr{G}^{(0)} we use the notation 𝒢AB\mathscr{G}_{A}^{B} for the subset {γ𝒢:s(γ)A,r(γ)B}\{\gamma\in\mathscr{G}:s(\gamma)\in A,\,r(\gamma)\in B\}.

We recall how to define an algebra structure in Cc(𝒢)C_{c}^{\infty}(\mathscr{G}) using smooth Haar systems.

Definition 2.2.

A 𝑠𝑚𝑜𝑜𝑡ℎ𝐻𝑎𝑎𝑟𝑠𝑦𝑠𝑡𝑒𝑚\it{smooth\,Haar\,system} over a Lie groupoid is a family of measures μx\mu_{x} in 𝒢x\mathscr{G}_{x} for each x𝒢(0)x\in\mathscr{G}^{(0)} such that,

  • for η𝒢xy\eta\in\mathscr{G}_{x}^{y} we have the following compatibility condition:

    𝒢xf(γ)𝑑μx(γ)=𝒢yf(γη)𝑑μy(γ)\int_{\mathscr{G}_{x}}f(\gamma)d\mu_{x}(\gamma)=\int_{\mathscr{G}_{y}}f(\gamma\circ\eta)d\mu_{y}(\gamma)
  • for each fCc(𝒢)f\in C_{c}^{\infty}(\mathscr{G}) the map

    x𝒢xf(γ)𝑑μx(γ)x\mapsto\int_{\mathscr{G}_{x}}f(\gamma)d\mu_{x}(\gamma)

    belongs to Cc(𝒢(0))C_{c}^{\infty}(\mathscr{G}^{(0)})

A Lie groupoid always posses a smooth Haar system. In fact, if we fix a smooth (positive) section of the 1-density bundle associated to the Lie algebroid we obtain a smooth Haar system in a canonical way. We suppose for the rest of the paper a given smooth Haar system given by 1-densities (for complete details see [Pat99]). We can now define a convolution product on Cc(𝒢)C_{c}^{\infty}(\mathscr{G}): Let f,gCc(𝒢)f,g\in C_{c}^{\infty}(\mathscr{G}), we set

(fg)(γ)=𝒢s(γ)f(γη1)g(η)𝑑μs(γ)(η)(f*g)(\gamma)=\int_{\mathscr{G}_{s(\gamma)}}f(\gamma\cdot\eta^{-1})g(\eta)d\mu_{s(\gamma)}(\eta)

This gives a well defined associative product.

Remark 2.3.

There is a way to define the convolution algebra using half densities (see Connes book [Con94]).

As we mentioned in the introduction, we are going to consider some elements in the KK-theory group K0(Cc(𝒢))K_{0}(C_{c}^{\infty}(\mathscr{G})). We recall how these elements are usually defined (See [NWX99] for complete details): First we recall a few facts about 𝒢\mathscr{G}-Pseudodifferential calculus:

A 𝒢\mathscr{G}-𝑃𝑠𝑒𝑢𝑑𝑜𝑑𝑖𝑓𝑓𝑒𝑟𝑒𝑛𝑡𝑖𝑎𝑙\it{Pseudodifferential} 𝑜𝑝𝑒𝑟𝑎𝑡𝑜𝑟\it{operator} is a family of pseudodifferential operators {Px}x𝒢(0)\{P_{x}\}_{x\in\mathscr{G}^{(0)}} acting in Cc(𝒢x)C^{\infty}_{c}(\mathscr{G}_{x}) such that if γ𝒢\gamma\in\mathscr{G} and

Uγ:Cc(𝒢s(γ))Cc(𝒢r(γ))U_{\gamma}:C^{\infty}_{c}(\mathscr{G}_{s(\gamma)})\rightarrow C^{\infty}_{c}(\mathscr{G}_{r(\gamma)})

the induced operator, then we have the following compatibility condition

Pr(γ)Uγ=UγPs(γ)P_{r(\gamma)}\circ U_{\gamma}=U_{\gamma}\circ P_{s(\gamma)}

There is also a differentiability condition with respect to xx that can be found in [NWX99].

For PP a 𝒢\mathscr{G}-Pseudodifferential elliptic operator there is a parametrix, i.e.\it{i.e.}, a 𝒢\mathscr{G}-Pseudodifferential operator QQ such that PQ1PQ-1 and QP1QP-1 belong to Cc(𝒢)C_{c}^{\infty}(\mathscr{G}) (where we are identifying Cc(𝒢)C_{c}^{\infty}(\mathscr{G}) with Ψ(𝒢)\Psi^{-\infty}(\mathscr{G}) as in [NWX99]). In other words, PP defines a quasi-isomorphism in (Ψ+,Cc(𝒢))(\Psi^{+\infty},C_{c}^{\infty}(\mathscr{G})) and thus an element in K0(Cc(𝒢))K_{0}(C_{c}^{\infty}(\mathscr{G})) that we call the index ind(P)ind(P). Similarly to the classical case, a 𝒢\mathscr{G}-PDO operator has a principal symbol that defines an element in the KK-theory group K0(A𝒢)K^{0}(A^{*}\mathscr{G}).

3 Deformation to the normal cone

The tangent groupoid is a particular case of a geometric construction that we describe here.

Let MM be a CC^{\infty} manifold and XMX\subset M be a CC^{\infty} submanifold. We denote by 𝒩XM\mathscr{N}_{X}^{M} the normal bundle to XX in MM, i.e.\it{i.e.}, 𝒩XM:=TXM/TX\mathscr{N}_{X}^{M}:=T_{X}M/TX.

We define the following set

𝒟XM:=𝒩XM×0M×\displaystyle\mathscr{D}_{X}^{M}:=\mathscr{N}_{X}^{M}\times{0}\bigsqcup M\times\mathbb{R}^{*} (11)

The purpose of this section is to recall how to define a CC^{\infty}-structure in 𝒟XM\mathscr{D}_{X}^{M}. This is more or less classical, for example it was extensively used in [HS87].

Let us first consider the case where M=p×qM=\mathbb{R}^{p}\times\mathbb{R}^{q} and X=p×{0}X=\mathbb{R}^{p}\times\{0\} (where we identify canonically X=pX=\mathbb{R}^{p}). We denote by q=npq=n-p and by 𝒟pn\mathscr{D}_{p}^{n} for 𝒟pn\mathscr{D}_{\mathbb{R}^{p}}^{\mathbb{R}^{n}} as above. In this case we clearly have that 𝒟pn=p×q×\mathscr{D}_{p}^{n}=\mathbb{R}^{p}\times\mathbb{R}^{q}\times\mathbb{R} (as a set). Consider the bijection ψ:p×q×𝒟pn\psi:\mathbb{R}^{p}\times\mathbb{R}^{q}\times\mathbb{R}\rightarrow\mathscr{D}_{p}^{n} given by

ψ(x,ξ,t)={(x,ξ,0) if t=0(x,tξ,t) if t0\psi(x,\xi,t)=\left\{\begin{array}[]{cc}(x,\xi,0)&\mbox{ if }t=0\\ (x,t\xi,t)&\mbox{ if }t\neq 0\end{array}\right. (12)

which inverse is given explicitly by

ψ1(x,ξ,t)={(x,ξ,0) if t=0(x,1tξ,t) if t0\psi^{-1}(x,\xi,t)=\left\{\begin{array}[]{cc}(x,\xi,0)&\mbox{ if }t=0\\ (x,\frac{1}{t}\xi,t)&\mbox{ if }t\neq 0\end{array}\right.

We can consider the CC^{\infty}-structure on 𝒟pn\mathscr{D}_{p}^{n} induced by this bijection.

We pass now to the general case. A local chart (𝒰,ϕ)(\mathscr{U},\phi) in MM is said to be a XX-slice if

  • 1)

    ϕ:𝒰Up×q\phi:\mathscr{U}\stackrel{{\scriptstyle\cong}}{{\rightarrow}}U\subset\mathbb{R}^{p}\times\mathbb{R}^{q}

  • 2)

    If 𝒰X=𝒱\mathscr{U}\cap X=\mathscr{V}, 𝒱=ϕ1(Up×{0})\mathscr{V}=\phi^{-1}(U\cap\mathbb{R}^{p}\times\{0\}) (we note V=Up×{0}V=U\cap\mathbb{R}^{p}\times\{0\})

With this notation, 𝒟VU𝒟pn\mathscr{D}_{V}^{U}\subset\mathscr{D}_{p}^{n} as an open subset. We may define a function

ϕ~:𝒟𝒱𝒰𝒟VU\tilde{\phi}:\mathscr{D}_{\mathscr{V}}^{\mathscr{U}}\rightarrow\mathscr{D}_{V}^{U} (13)

in the following way: For x𝒱x\in\mathscr{V} we have ϕ(x)p×{0}\phi(x)\in\mathbb{R}^{p}\times\{0\}. If we write ϕ(x)=(ϕ1(x),0)\phi(x)=(\phi_{1}(x),0), then

ϕ1:𝒱Vp\phi_{1}:\mathscr{V}\rightarrow V\subset\mathbb{R}^{p}

is a diffeomorphism. We set ϕ~(v,ξ,0)=(ϕ1(v),dNϕv(ξ),0)\tilde{\phi}(v,\xi,0)=(\phi_{1}(v),d_{N}\phi_{v}(\xi),0) and ϕ~(u,t)=(ϕ(u),t)\tilde{\phi}(u,t)=(\phi(u),t) for t0t\neq 0. Here dNϕv:Nvqd_{N}\phi_{v}:N_{v}\rightarrow\mathbb{R}^{q} is the normal component of the derivate dϕvd\phi_{v} for v𝒱v\in\mathscr{V}. It is clear that ϕ~\tilde{\phi} is also a bijection (in particular it induces a CC^{\infty} structure on 𝒟𝒱𝒰\mathscr{D}_{\mathscr{V}}^{\mathscr{U}}). Now, let us consider an atlas {(𝒰α,ϕα)}αΔ\{(\mathscr{U}_{\alpha},\phi_{\alpha})\}_{\alpha\in\Delta} of MM consisting of XX-slices. Then the collection {(𝒟𝒱α𝒰α,ϕα)~}αΔ\{(\mathscr{D}_{\mathscr{V}_{\alpha}}^{\mathscr{U}_{\alpha}},\tilde{\phi_{\alpha})}\}_{\alpha\in\Delta} is a CC^{\infty}-atlas of 𝒟XM\mathscr{D}_{X}^{M} (proposition 3.1 in [CR06]).

Definition 3.1 (Deformation to the normal cone).

Let XMX\subset M be as above. The set 𝒟XM\mathscr{D}_{X}^{M} equipped with the CC^{\infty} structure induced by the atlas described in the last proposition is called 𝑇ℎ𝑒𝑑𝑒𝑓𝑜𝑟𝑚𝑎𝑡𝑖𝑜𝑛𝑡𝑜𝑛𝑜𝑟𝑚𝑎𝑙𝑐𝑜𝑛𝑒𝑎𝑠𝑠𝑜𝑐𝑖𝑎𝑡𝑒𝑑𝑡𝑜\it{"The\,deformation\,to\,normal\,cone\,associated\,to\,} XMX\subset M”.

Remark 3.2.

Following the same steps, we can define a deformation to the normal cone associated to an injective immersion XMX\hookrightarrow M.

One important feature about this construction is that it is in some sense functorial. More explicitly, let (M,X)(M,X) and (M,X)(M^{\prime},X^{\prime}) be CC^{\infty}-pairs as above and let F:(M,X)(M,X)F:(M,X)\rightarrow(M^{\prime},X^{\prime}) be a pair morphism, i.e., a CC^{\infty} map F:MMF:M\rightarrow M^{\prime}, with F(X)XF(X)\subset X^{\prime}. We define 𝒟(F):𝒟XM𝒟XM\mathscr{D}(F):\mathscr{D}_{X}^{M}\rightarrow\mathscr{D}_{X^{\prime}}^{M^{\prime}} by the following formulas:

𝒟(F)(x,ξ,0)=(F(x),dNFx(ξ),0)\mathscr{D}(F)(x,\xi,0)=(F(x),d_{N}F_{x}(\xi),0) and

𝒟(F)(m,t)=(F(m),t)\mathscr{D}(F)(m,t)=(F(m),t) for t0t\neq 0, where dNFxd_{N}F_{x} is by definition the map

(𝒩XM)xdNFx(𝒩XM)F(x)(\mathscr{N}_{X}^{M})_{x}\stackrel{{\scriptstyle d_{N}F_{x}}}{{\longrightarrow}}(\mathscr{N}_{X^{\prime}}^{M^{\prime}})_{F(x)}

induced by TxMdFxTF(x)MT_{x}M\stackrel{{\scriptstyle dF_{x}}}{{\longrightarrow}}T_{F(x)}M^{\prime}.

Then 𝒟(F):𝒟XM𝒟XM\mathscr{D}(F):\mathscr{D}_{X}^{M}\rightarrow\mathscr{D}_{X^{\prime}}^{M^{\prime}} is CC^{\infty}-map (proposition 3.4 in [CR06]).

3.1 The tangent groupoid

Definition 3.3 (Tangent groupoid).

Let 𝒢𝒢(0)\mathscr{G}\rightrightarrows\mathscr{G}^{(0)} be a Lie groupoid. 𝑇ℎ𝑒𝑡𝑎𝑛𝑔𝑒𝑛𝑡𝑔𝑟𝑜𝑢𝑝𝑜𝑖𝑑\it{The\,tangent\,groupoid} associated to 𝒢\mathscr{G} is the groupoid that has 𝒟𝒢(0)𝒢\mathscr{D}_{\mathscr{G}^{(0)}}^{\mathscr{G}} as the set of arrows and 𝒢(0)×\mathscr{G}^{(0)}\times\mathbb{R} as the units, with:

  • \cdot

    sT(x,η,0)=(x,0)s^{T}(x,\eta,0)=(x,0) and rT(x,η,0)=(x,0)r^{T}(x,\eta,0)=(x,0) at t=0t=0.

  • \cdot

    sT(γ,t)=(s(γ),t)s^{T}(\gamma,t)=(s(\gamma),t) and rT(γ,t)=(r(γ),t)r^{T}(\gamma,t)=(r(\gamma),t) at t0t\neq 0.

  • \cdot

    The product is given by mT((x,η,0),(x,ξ,0))=(x,η+ξ,0)m^{T}((x,\eta,0),(x,\xi,0))=(x,\eta+\xi,0) et mT((γ,t),(β,t))=(m(γ,β),t)m^{T}((\gamma,t),(\beta,t))=(m(\gamma,\beta),t) if t0t\neq 0 and if r(β)=s(γ)r(\beta)=s(\gamma).

  • \cdot

    The unit map uT:𝒢(0)𝒢Tu^{T}:\mathscr{G}^{(0)}\rightarrow\mathscr{G}^{T} is given by uT(x,0)=(x,0)u^{T}(x,0)=(x,0) and uT(x,t)=(u(x),t)u^{T}(x,t)=(u(x),t) for t0t\neq 0.

We denote 𝒢T:=𝒟𝒢(0)𝒢\mathscr{G}^{T}:=\mathscr{D}_{\mathscr{G}^{(0)}}^{\mathscr{G}} and A𝒢:=𝒩𝒢(0)𝒢A\mathscr{G}:=\mathscr{N}_{\mathscr{G}^{(0)}}^{\mathscr{G}}.

As we have seen above 𝒢T\mathscr{G}^{T} can be considered as a CC^{\infty} manifold with border. As a consequence of the functoriality of the Deformation to the normal cone, one can show that the tangent groupoid is in fact a Lie groupoid. Indeed, it is easy to check that if we identify in a canonical way 𝒟𝒢(0)𝒢(2)\mathscr{D}_{\mathscr{G}^{(0)}}^{\mathscr{G}^{(2)}} with (𝒢T)(2)(\mathscr{G}^{T})^{(2)}, then

mT=𝒟(m),sT=𝒟(s),rT=𝒟(r),uT=𝒟(u)m^{T}=\mathscr{D}(m),\,s^{T}=\mathscr{D}(s),\,r^{T}=\mathscr{D}(r),\,u^{T}=\mathscr{D}(u)

where we are considering the following pair morphisms:

m:((𝒢)(2),𝒢(0))(𝒢,𝒢(0)),\displaystyle m:((\mathscr{G})^{(2)},\mathscr{G}^{(0)})\rightarrow(\mathscr{G},\mathscr{G}^{(0)}),
s,r:(𝒢,𝒢(0))(𝒢(0),𝒢(0)),\displaystyle s,r:(\mathscr{G},\mathscr{G}^{(0)})\rightarrow(\mathscr{G}^{(0)},\mathscr{G}^{(0)}),
u:(𝒢(0),𝒢(0))(𝒢,𝒢(0)).\displaystyle u:(\mathscr{G}^{(0)},\mathscr{G}^{(0)})\rightarrow(\mathscr{G},\mathscr{G}^{(0)}).
Remark 3.4.

Finally, let {μx}\{\mu_{x}\} be a smooth Haar system on 𝒢\mathscr{G}, i.e., a choice of 𝒢\mathscr{G}-invariant Lebesgue measures. In particular we have an associated smooth Haar system on A𝒢A\mathscr{G} (groupoid given by the vector bundle structure), which we note again by {μx}\{\mu_{x}\}. Then the following family {μ(x,t)}\{\mu_{(x,t)}\} is a smooth Haar system for the tangent groupoid of 𝒢\mathscr{G} (details may be found in [Pat99]):

  • μ(x,0):=μx\mu_{(x,0)}:=\mu_{x} at (𝒢T)(x,0)=Ax𝒢(\mathscr{G}^{T})_{(x,0)}=A_{x}\mathscr{G} and

  • μ(x,t):=tqμx\mu_{(x,t)}:=t^{-q}\cdot\mu_{x} at (𝒢T)(x,t)=𝒢x(\mathscr{G}^{T})_{(x,t)}=\mathscr{G}_{x} for t0t\neq 0, where q=dim𝒢xq=dim\,\mathscr{G}_{x}.

In this article, we are only going to consider this Haar systems for the tangent groupoids.

4 A Schwartz type algebra for the Tangent groupoid

In this section we will recall how to construct the deformation algebra mentioned at the introduction. For complete details, we refer the reader to [CR06].

The Schwartz algebra for the Tangent groupoid will be a particular case of a construction associated to any deformation to the normal cone.

Definition 4.1.

Let p,qp,q\in\mathbb{N} and Up×qU\subset\mathbb{R}^{p}\times\mathbb{R}^{q} an open subset, and let V=U(p×{0})V=U\cap(\mathbb{R}^{p}\times\{0\}).

  • (1)

    Let KU×K\subset U\times\mathbb{R} be a compact subset. We say that KK is a conic compact subset of U×U\times\mathbb{R} relative to VV if

    K0=K(U×{0})VK_{0}=K\cap(U\times\{0\})\subset V
  • (2)

    Let ΩVU={(x,ξ,t)p×q×:(x,tξ)U},\Omega_{V}^{U}=\{(x,\xi,t)\in\mathbb{R}^{p}\times\mathbb{R}^{q}\times\mathbb{R}:(x,t\cdot\xi)\in U\}, which is an open subset of p×q×\mathbb{R}^{p}\times\mathbb{R}^{q}\times\mathbb{R} and thus a CC^{\infty} manifold. Let gC(ΩVU)g\in C^{\infty}(\Omega_{V}^{U}). We say that gg has compact conic support, if there exists a conic compact KK of U×U\times\mathbb{R} relative to VV such that if (x,tξ,t)K(x,t\xi,t)\notin K then g(x,ξ,t)=0g(x,\xi,t)=0.

  • (3)

    We denote by 𝒮c(ΩVU)\mathscr{S}_{c}(\Omega_{V}^{U}) the set of functions gC(ΩVU)g\in C^{\infty}(\Omega_{V}^{U}) that have compact conic support and that satisfy the following condition:

    • (s1(s_{1})

      \forall k,mk,m\in\mathbb{N}, lpl\in\mathbb{N}^{p} and αq\alpha\in\mathbb{N}^{q} it exists C(k,m,l,α)>0C_{(k,m,l,\alpha)}>0 such that

      (1+ξ2)kxlξαtmg(x,ξ,t)C(k,m,l,α)(1+\|\xi\|^{2})^{k}\|\partial_{x}^{l}\partial_{\xi}^{\alpha}\partial_{t}^{m}g(x,\xi,t)\|\leq C_{(k,m,l,\alpha)}

Now, the spaces 𝒮c(ΩVU)\mathscr{S}_{c}(\Omega_{V}^{U}) are invariant under diffeomorphisms. More precisely: Let F:UUF:U\rightarrow U^{\prime} be a CC^{\infty}-diffeomorphism such that F(V)=VF(V)=V^{\prime}; let F~:ΩVUΩVU\tilde{F}:\Omega_{V}^{U}\rightarrow\Omega_{V^{\prime}}^{U^{\prime}} the induced map. Then, for every g𝒮c(ΩVU)g\in\mathscr{S}_{c}(\Omega_{V^{\prime}}^{U^{\prime}}), we have that g~:=gF~𝒮c(ΩVU)\tilde{g}:=g\circ\tilde{F}\in\mathscr{S}_{c}(\Omega_{V}^{U}) (proposition 4.2 in [CR06]).

This compatibility result allows to give the following definition.

Definition 4.2.

Let gC(𝒟XM)g\in C^{\infty}(\mathscr{D}_{X}^{M}).

  • (a)

    We say that gg has conic compact support KK, if there exists a compact subset KM×K\subset M\times\mathbb{R} with K0:=K(M×{0})XK_{0}:=K\cap(M\times\{0\})\subset X (conic compact relative to XX) such that if t0t\neq 0 and (m,t)K(m,t)\notin K then g(m,t)=0g(m,t)=0.

  • (b)

    We say that gg is rapidly decaying at zero if for every (𝒰,ϕ)(\mathscr{U},\phi) XX-slice chart and for every χCc(𝒰×)\chi\in C^{\infty}_{c}(\mathscr{U}\times\mathbb{R}), the map gχC(ΩVU)g_{\chi}\in C^{\infty}(\Omega_{V}^{U}) (ΩVU\Omega_{V}^{U} as in definition 4.1.) given by

    gχ(x,ξ,t)=(gφ1)(x,ξ,t)(χpφ1)(x,ξ,t)g_{\chi}(x,\xi,t)=(g\circ\varphi^{-1})(x,\xi,t)\cdot(\chi\circ p\circ\varphi^{-1})(x,\xi,t)

    is in 𝒮c(ΩVU)\mathscr{S}_{c}(\Omega_{V}^{U}), where

    • \cdot

      pp is the projection is the deformation of the pair map (M,X)Id(M,M)(M,X)\stackrel{{\scriptstyle Id}}{{\longrightarrow}}(M,M), i.e., p:𝒟XMM×p:\mathscr{D}_{X}^{M}\rightarrow M\times\mathbb{R} is given by (x,ξ,0)(x,0)(x,\xi,0)\mapsto(x,0), and (m,t)(m,t)(m,t)\mapsto(m,t) for t0t\neq 0, and

    • \cdot

      φ:=ϕ~1ψ:ΩVU𝒟𝒱𝒰\varphi:=\tilde{\phi}^{-1}\circ\psi:\Omega_{V}^{U}\rightarrow\mathscr{D}_{\mathscr{V}}^{\mathscr{U}}, where ψ\psi and ϕ~\tilde{\phi} are defined at (12) and (13) above.

Finally, we denote by 𝒮c(𝒟XM)\mathscr{S}_{c}(\mathscr{D}_{X}^{M}) the set of functions gC(𝒟XM)g\in C^{\infty}(\mathscr{D}_{X}^{M}) that are rapidly decaying at zero with conic compact support.

Remark 4.3.
  • (a)

    Obviously Cc(𝒟XM)C^{\infty}_{c}(\mathscr{D}_{X}^{M}) is a subspace of 𝒮c(𝒟XM)\mathscr{S}_{c}(\mathscr{D}_{X}^{M}).

  • (b)

    Let {(𝒰α,ϕα)}αΔ\{(\mathscr{U}_{\alpha},\phi_{\alpha})\}_{\alpha\in\Delta} a family of XX-slices covering XX. We have a decomposition of 𝒮c(𝒟XM)\mathscr{S}_{c}(\mathscr{D}_{X}^{M}) as follows (see remark 4.5 in [CR06] and discussion below it):

    𝒮c(𝒟XM)=αΛ𝒮c(𝒟𝒱α𝒰α)+Cc(M×).\displaystyle\mathscr{S}_{c}(\mathscr{D}_{X}^{M})=\sum_{\alpha\in\Lambda}\mathscr{S}_{c}(\mathscr{D}_{\mathscr{V}_{\alpha}}^{\mathscr{U}_{\alpha}})+C^{\infty}_{c}(M\times\mathbb{R}^{*}). (14)

The main theorem in [CR06] (Theorem 4.10) is the following

Theorem 4.4

The space 𝒮c(𝒢T)\mathscr{S}_{c}(\mathscr{G}^{T}) is stable under convolution. In particular, we have the following inclusions of algebras

Cc(𝒢T)𝒮c(𝒢T)Cr(𝒢T)C_{c}^{\infty}(\mathscr{G}^{T})\subset\mathscr{S}_{c}(\mathscr{G}^{T})\subset C_{r}^{*}(\mathscr{G}^{T})

Moreover, 𝒮c(𝒢T)\mathscr{S}_{c}(\mathscr{G}^{T}) is a field of algebras over \mathbb{R}, whose fibers are

𝒮(A𝒢) at t=0, and \mathscr{S}(A\mathscr{G})\text{ at }t=0,\text{ and }
Cc(𝒢) for t0.C_{c}^{\infty}(\mathscr{G})\text{ for }t\neq 0.

In the statement of this theorem, 𝒮(A𝒢)\mathscr{S}(A\mathscr{G}) denotes the Schwartz algebra over the Lie algebroid. Let us briefly recall the notion of Schwartz space associated to a vector bundle: For a trivial bundle X×qXX\times\mathbb{R}^{q}\rightarrow X, 𝒮(X×q):=Cc(X,𝒮(q))\mathscr{S}(X\times\mathbb{R}^{q}):=C^{\infty}_{c}(X,\mathscr{S}(\mathbb{R}^{q})) (see [Trè06]). In general, 𝒮(E)\mathscr{S}(E) is defined using local charts. More precisely, a partition of the unity argument, allows to see that if we take a covering of XX, {(𝒱α,τα)}αΔ\{(\mathscr{V}_{\alpha},\tau_{\alpha})\}_{\alpha\in\Delta}, consisting on trivializing charts, then we have a decomposition of the following kind:

𝒮(E)=α𝒮(𝒱α×q).\mathscr{S}(E)=\sum_{\alpha}\mathscr{S}(\mathscr{V}_{\alpha}\times\mathbb{R}^{q}). (15)

The ”Schwartz algebras” have in general the good KK-theory groups. As we said in te introduction, we are interested in the group K0(A𝒢)=K0(C0(A𝒢))K_{0}(A^{*}\mathscr{G})=K_{0}(C_{0}(A^{*}\mathscr{G})). It is not enough to take the KK-theory of Cc(A𝒢)C^{\infty}_{c}(A\mathscr{G}) (see for example [Con94]). We are going to see that 𝒮(A𝒢)\mathscr{S}(A^{*}\mathscr{G}) has the wanted KK-theory. In particular, our deformation algebra restricts at zero to the right algebra. The result is the following.

Proposition 4.5

Let EXE\rightarrow X be a smooth vector bundle over a smooth manifold XX.

  • (i)

    The algebra (𝒮(E),)(\mathscr{S}(E),*) with the convolution product is isomorphic to (𝒮(E),)(\mathscr{S}(E^{*}),\cdot), where \cdot denotes the punctual product.

  • (ii)

    The 𝒮(E)\mathscr{S}(E) is stable under holomorphic calculus on C0(E)C_{0}(E^{*}). In particular, K0(E)K0(𝒮(E))K^{0}(E^{*})\cong K_{0}(\mathscr{S}(E)).

Proof.
  • (i)

    Suppose first that E=X×qE=X\times\mathbb{R}^{q} is trivial. In this case 𝒮(E)=Cc(X,𝒮(q))\mathscr{S}(E)=C^{\infty}_{c}(X,\mathscr{S}(\mathbb{R}^{q})). Let gCc(X,𝒮(q))g\in C^{\infty}_{c}(X,\mathscr{S}(\mathbb{R}^{q})) and XqX\in\mathbb{R}^{q}. We let

    (g)(x)(X)=qeiXηg(x,η)𝑑η\mathscr{F}(g)(x)(X)=\int_{\mathbb{R}^{q}}e^{-iX\cdot\eta}g(x,\eta)d\eta

    which is the Fourier transform of g(x)g(x) evaluated at XX. It defines an element of Cc(X,𝒮(q))C^{\infty}_{c}(X,\mathscr{S}(\mathbb{R}^{q})). Now, since the product on 𝒮(E)\mathscr{S}(E) is given by

    (fg)(x)=f(x)g(x),(f*g)(x)=f(x)*g(x),

    we have, thanks to the continuity of the Fourier transform, that \mathscr{F} is an isomorphism (Cc(X,𝒮(q)),)(Cc(X,𝒮(q)),)(C^{\infty}_{c}(X,\mathscr{S}(\mathbb{R}^{q})),*)\cong(C^{\infty}_{c}(X,\mathscr{S}(\mathbb{R}^{q})),\cdot).

    For the general case, we look at the decomposition of 𝒮(E)\mathscr{S}(E) as in (15) and we have the isomorphism given by Fourier

    :(𝒮(E),)(𝒮(E),).\mathscr{F}:(\mathscr{S}(E),*)\rightarrow(\mathscr{S}(E^{*}),\cdot).
  • (ii)

    Let g𝒮(E)g\in\mathscr{S}(E) and f:f:\mathbb{C}\rightarrow\mathbb{C} an holomorphic map with f(0)=0f(0)=0. We have to verify that fg𝒮(E)f\circ g\in\mathscr{S}(E): This condition is local, in particular we can suppose E=V×qE=V\times\mathbb{R}^{q} with VpV\subset\mathbb{R}^{p} an open subset of p\mathbb{R}^{p}. Let αp×q\alpha\in\mathbb{N}^{p}\times\mathbb{N}^{q} and KVK\subset V be a compact subset. We et z=(x,ξ)V×qz=(x,\xi)\in V\times\mathbb{R}^{q} and u=g(x,ξ)u=g(x,\xi). We can show by induction that the derivative z|α|(fg)\partial_{z}^{|\alpha|}(f\circ g) can be written in the following way:

    z|α|(fg)=|β||α|u|β|f(u)Pβ(z),\partial_{z}^{|\alpha|}(f\circ g)=\sum_{|\beta|\leq|\alpha|}\partial_{u}^{|\beta|}f(u)P_{\beta}(z),

    where each Pβ(z)P_{\beta}(z) is a finite sum of products of the form

    zγ1g(z)zγrg(z),\partial_{z}^{\gamma_{1}}g(z)\cdots\partial_{z}^{\gamma_{r}}g(z),

    with |γ1|++|γr|=|β||\gamma_{1}|+...+|\gamma_{r}|=|\beta|. Since g𝒮(E)g\in\mathscr{S}(E), u|β|f(u)\partial_{u}^{|\beta|}f(u) is bounded for xKx\in K. The conclusion is immediate by using the Schwartz condition for gg.

From now on it will be important to restrict our functions on the tangent groupoid to the closed interval [0,1][0,1]. We keep the notation 𝒮c(𝒟XM)\mathscr{S}_{c}(\mathscr{D}_{X}^{M}) for the restricted space. All the results above remain true. So for instance 𝒮c(𝒢T)\mathscr{S}_{c}(\mathscr{G}^{T}) is an algebra which is a field of algebras over the closed interval [0,1][0,1] with 0-fiber 𝒮(A𝒢)\mathscr{S}(A\mathscr{G}) and Cc(𝒢)C^{\infty}_{c}(\mathscr{G}) otherwise. Before starting with the construction of the indices, we need to have an exact sequence analog as the one used in the construction of the CC^{*}-analytic indices (exact sequence (5)). The first step in this direction is the following proposition.

Proposition 4.6

The evaluation at zero, 𝒮c(𝒟XM)e0𝒮(𝒩XM)\mathscr{S}_{c}(\mathscr{D}_{X}^{M})\stackrel{{\scriptstyle e_{0}}}{{\longrightarrow}}\mathscr{S}(\mathscr{N}_{X}^{M}) is surjective.

Proof.

Thanks to the decomposition (14) discussed at the remark 4.3 above, it will be enough to prove that the evaluation map 𝒮c(ΩVU)e0𝒮(V×q)\mathscr{S}_{c}(\Omega_{V}^{U})\stackrel{{\scriptstyle e_{0}}}{{\longrightarrow}}\mathscr{S}(V\times\mathbb{R}^{q}) is surjective. Where we are using the same notations as in definition 4.1.

Let g𝒮(V×q)g\in\mathscr{S}(V\times\mathbb{R}^{q}). We consider

h(s)={e1s if s>00 if s0h(s)=\left\{\begin{array}[]{cc}e^{-\frac{1}{s}}&\mbox{ if }s>0\\ 0&\mbox{ if }s\leq 0\end{array}\right.

Let K0VK_{0}\subset V be the horizontal support of gg. We can assume without lost of generality that {(x,ξ)p×q:xK0 and ξ1}U\{(x,\xi)\in\mathbb{R}^{p}\times\mathbb{R}^{q}:x\in K_{0}\text{ and }\|\xi\|\leq 1\}\subset U. Let

K={(x,ξ,t)U×[0,1]:xK0,ξt},K=\{(x,\xi,t)\in U\times[0,1]:x\in K_{0},\|\xi\|\leq\sqrt{t}\},

then KU×[0,1]K\subset U\times[0,1] is a compact subset and KU×{0}=K0K\cap U\times\{0\}=K_{0}, i.e.\it{i.e.}, KK is a conic compact U×[0,1]U\times[0,1] relative to VV.

Let (x,ξ,t)ΩVU(x,\xi,t)\in\Omega_{V}^{U}, and

g~(x,ξ,t)={g(x,ξ)h(1tξ2) if xK00 otherwise \tilde{g}(x,\xi,t)=\left\{\begin{array}[]{cc}g(x,\xi)\cdot h(1-t\|\xi\|^{2})&\mbox{ if }x\in K_{0}\\ 0&\mbox{ otherwise }\end{array}\right.

then we have that

  • \cdot

    g~\tilde{g} has compact conic support KK

  • \cdot

    g~C(ΩVU)\tilde{g}\in C^{\infty}(\Omega_{V}^{U}) because gg and hh are CC^{\infty} and gg has horizontal compact support contained in K0K_{0}.

  • \cdot

    g~𝒮c(ΩVU)\tilde{g}\in\mathscr{S}_{c}(\Omega_{V}^{U}): Let h~(ξ,t)=h(1tξ2)\tilde{h}(\xi,t)=h(1-t\|\xi\|^{2}). By induction, an elementary computation shows that, for αq\alpha\in\mathbb{N}^{q}, ll\in\mathbb{N}, we have

    ξαtlh~(ξ,t)=|β||α|+laβPβ(ξ,t)h|β|(1tξ2)\partial_{\xi}^{\alpha}\partial_{t}^{l}\tilde{h}(\xi,t)=\sum_{|\beta|\leq|\alpha|+l}a_{\beta}P_{\beta}(\xi,t)h^{|\beta|}(1-t\|\xi\|^{2}),

    where βq+1\beta\in\mathbb{N}^{q+1}, aβa_{\beta} is a constant (depending on α\alpha and ll) and Pβ(ξ,t)P_{\beta}(\xi,t) is a finite sum of products

    tγ0ξ1γ1ξqγqt^{\gamma_{0}}\cdot\xi_{1}^{\gamma_{1}}\cdots\xi_{q}^{\gamma_{q}}, γi\gamma_{i}\in\mathbb{N}.

    Now, from the fact that hh and all its derivates are bounded, and using the Schwratz property for gg, we conclude that g~𝒮c(ΩVU)\tilde{g}\in\mathscr{S}_{c}(\Omega_{V}^{U}). Finally, e0(g~)=ge_{0}(\tilde{g})=g by construction.

We have a the short exact sequence of algebras:

0J𝒮c(𝒢T)e0𝒮(A𝒢)0,\displaystyle 0\longrightarrow J\longrightarrow\mathscr{S}_{c}(\mathscr{G}^{T})\stackrel{{\scriptstyle e_{0}}}{{\longrightarrow}}\mathscr{S}(A\mathscr{G})\longrightarrow 0, (16)

where J=Ker(e0)J=Ker(e_{0}) by definition.

5 Compactly supported analytic indices

This section is devoted to the construction of the indices announced in the introduction.

As a first step, we want to apply KK-theory to the exact sequence (16) above. But in principle there is no reason for obtaining an exact sequence of the same kind. We have the following proposition:

Proposition 5.1

The morphism in KK-theory,

K0(𝒮c(𝒢T))(e0)K0(𝒮(A𝒢))K0(A𝒢),K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T}))\stackrel{{\scriptstyle(e_{0})_{*}}}{{\longrightarrow}}K_{0}(\mathscr{S}(A\mathscr{G}))\approx K^{0}(A^{*}\mathscr{G}),

induced from the evaluation at zero is surjective.

Proof.

Let [σ]K0(𝒮(A𝒢))=K0(A𝒢)[\sigma]\in K_{0}(\mathscr{S}(A\mathscr{G}))=K^{0}(A^{*}\mathscr{G}). We know from the 𝒢\mathscr{G}-pseudodifferential calculus that [σ][\sigma] can be represented by a smooth homogeneus elliptic symbol. We can consider the symbol over A𝒢×[0,1]A^{*}\mathscr{G}\times[0,1] that coincides with σ\sigma for all tt, we note it by σ~\tilde{\sigma}. Now, since A𝒢T=A𝒢×[0,1]A\mathscr{G}^{T}=A\mathscr{G}\times[0,1], we can take P~=(Pt)t[0,1]\tilde{P}=(P_{t})_{t\in[0,1]} a 𝒢T\mathscr{G}^{T}-elliptic pseudodifferential operator associated to σ\sigma, that is, σP~=σ~\sigma_{\tilde{P}}=\tilde{\sigma}. Let i:Cc(𝒢T)𝒮c(𝒢T)i:C^{\infty}_{c}(\mathscr{G}^{T})\rightarrow\mathscr{S}_{c}(\mathscr{G}^{T}) be the inclusion (which is an algebra morphism), then i(indP~)K0(𝒮c(𝒢T))i_{*}(ind\,\tilde{P})\in K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T})) is such that e0,(i(indP~))=σe_{0,*}(i_{*}(ind\,\tilde{P}))=\sigma. ∎

By applying KK-theory to the exact sequence (16) we obtain

K0(J)K0(𝒮c(𝒢T))e0K0(A𝒢)0,\displaystyle K_{0}(J)\longrightarrow K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T}))\stackrel{{\scriptstyle e_{0}}}{{\longrightarrow}}K^{0}(A^{*}\mathscr{G})\longrightarrow 0, (17)

Let us now define the group where our indices take values.

Definition 5.2.

Let kk\in\mathbb{N}. We note by

K0h,k(𝒢):=Coeq[K0(Cck(𝒢×[0,1]))e1e0K0(Cck(𝒢))]K_{0}^{h,k}(\mathscr{G}):=Coeq\left[\tiny{K_{0}(C_{c}^{k}(\mathscr{G}\times[0,1]))\overset{e_{0}}{\underset{e_{1}}{\rightrightarrows}}K_{0}(C_{c}^{k}(\mathscr{G}))}\right]

the co-equalizer of the KK-theory morphisms induced by the evaluations at zero and at one. In other words, K0h,k(𝒢)K_{0}^{h,k}(\mathscr{G}) is just the quotient of K0(Cck(𝒢))K_{0}(C_{c}^{k}(\mathscr{G})) by the image of (e1e0):K0(Cck(𝒢×[0,1]))K0(Cck(𝒢))(e_{1}-e_{0}):K_{0}(C_{c}^{k}(\mathscr{G}\times[0,1]))\rightarrow K_{0}(C_{c}^{k}(\mathscr{G})). We note by π:K0(Cck(𝒢))K0h,k(𝒢)\pi:K_{0}(C_{c}^{k}(\mathscr{G}))\rightarrow K_{0}^{h,k}(\mathscr{G}) the quotient morphism. We define the Bounded KK-theory group of 𝒢\mathscr{G} as the projective limit

K0F(𝒢)=limkK0h,k(𝒢)K_{0}^{F}(\mathscr{G})=\varprojlim_{k}K_{0}^{h,k}(\mathscr{G})

induced by the inclusions Cck(𝒢)Cck+1(𝒢)...\hookrightarrow C_{c}^{k}(\mathscr{G})\hookrightarrow C_{c}^{k+1}(\mathscr{G})\hookrightarrow...

Remark 5.3.

In the above definition, the superscript hh in K0h,k(𝒢)K_{0}^{h,k}(\mathscr{G}) makes reference to homotopy. Indeed, we can see the co-equalizer K0h,k(𝒢)K_{0}^{h,k}(\mathscr{G}) otherwise: for x,yK0(Cck(𝒢))x,y\in K_{0}(C_{c}^{k}(\mathscr{G})), we say the they are homotopic, xhyx\sim_{h}y, if there exists zK0(Cck(𝒢×[0,1]))z\in K_{0}(C_{c}^{k}(\mathscr{G}\times[0,1])) such that

  • (0)

    e0(z)=xe_{0}(z)=x and

  • (11)

    e1(z)=ye_{1}(z)=y.

This gives an equivalence relation in K0(Cck(𝒢))K_{0}(C_{c}^{k}(\mathscr{G})) compatible with the group structure. The quotient K0(Cck(𝒢))/hK_{0}(C_{c}^{k}(\mathscr{G}))/\sim_{h} coincides with K0h,k(𝒢)K_{0}^{h,k}(\mathscr{G}).

Before stating our main theorem, let us see that there are natural morphisms

K0(Cc(𝒢))K0F(𝒢)K0(Cr(𝒢)).K_{0}(C_{c}^{\infty}(\mathscr{G}))\rightarrow K_{0}^{F}(\mathscr{G})\rightarrow K_{0}(C^{*}_{r}(\mathscr{G})).

The first is induced from the canonical morphisms K0(Cc(𝒢))K0h,k(𝒢)K_{0}(C_{c}^{\infty}(\mathscr{G}))\rightarrow K_{0}^{h,k}(\mathscr{G}) by using the universal property of the projective limits. For the second one, we see that

K0F(𝒢)Coeq[limkK0(Cck(𝒢×[0,1]))e1~e0~limkK0(Cck(𝒢))]K_{0}^{F}(\mathscr{G})\cong Coeq\left[\tiny{\varprojlim_{k}K_{0}(C_{c}^{k}(\mathscr{G}\times[0,1]))\overset{\tilde{e_{0}}}{\underset{\tilde{e_{1}}}{\rightrightarrows}}\varprojlim_{k}K_{0}(C_{c}^{k}(\mathscr{G}))}\right] (18)

That is, K0F(𝒢)K_{0}^{F}(\mathscr{G}) is also a coequalizer. Now, the morphism K0F(𝒢)K0(Cr(𝒢))K_{0}^{F}(\mathscr{G})\rightarrow K_{0}(C^{*}_{r}(\mathscr{G})) is induced from the canonical morphisms K0h,k(𝒢)K0(Cr(𝒢))K_{0}^{h,k}(\mathscr{G})\rightarrow K_{0}(C^{*}_{r}(\mathscr{G})) (the KK-theory for CC^{*}-algebras respects the homotopy relation).

Our main result is the following:

Theorem 5.4
  1. 1.

    There is a unique group morphism

    indaF:K0(A𝒢)K0F(𝒢)ind_{a}^{F}:K^{0}(A^{*}\mathscr{G})\rightarrow K_{0}^{F}(\mathscr{G})

    that fits in the following commutative diagram

    K0(𝒮c(𝒢T))\textstyle{K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e0\scriptstyle{e_{0}}e1F\scriptstyle{e_{1}^{F}}K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}indaF\scriptstyle{ind_{a}^{F}}K0F(𝒢).\textstyle{K_{0}^{F}(\mathscr{G}).} (19)
  2. 2.

    The morphism indaF:K0(A𝒢)K0F(𝒢)ind_{a}^{F}:K^{0}(A^{*}\mathscr{G})\rightarrow K_{0}^{F}(\mathscr{G}) fits in the following commutative diagram

    Ell(𝒢)\textstyle{Ell(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}symb\scriptstyle{symb}ind\scriptstyle{ind}K0(Cc(𝒢))\textstyle{K_{0}(C^{\infty}_{c}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K0(A𝒢)\textstyle{K_{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{id}indaF\scriptstyle{ind_{a}^{F}}K0F(𝒢)\textstyle{K_{0}^{F}(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K0(A𝒢)\textstyle{K_{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda\scriptstyle{ind_{a}}K0(Cr(𝒢))\textstyle{K_{0}(C_{r}^{*}(\mathscr{G}))}. (20)

The proof of the theorem will require several lemmas:

Lemma 5.5

Let λ:𝒢T[0,1]\lambda:\mathscr{G}^{T}\rightarrow[0,1] be the projection and J=Ker(e0)J=Ker(e_{0}), where 𝒮c(𝒢T)e0𝒮(A𝒢)\mathscr{S}_{c}(\mathscr{G}^{T})\stackrel{{\scriptstyle e_{0}}}{{\longrightarrow}}\mathscr{S}(A\mathscr{G}), as in (16). Then J=λ𝒮c(𝒢T)J=\lambda\cdot\mathscr{S}_{c}(\mathscr{G}^{T}).

Proof.

The inclusion λ𝒮c(𝒢T)J\lambda\cdot\mathscr{S}_{c}(\mathscr{G}^{T})\subset J is obvious. Let f𝒮c(𝒢T)f\in\mathscr{S}_{c}(\mathscr{G}^{T}) with e0(f)=0e_{0}(f)=0. As a CC^{\infty}-map, fC(𝒢T)f\in C^{\infty}(\mathscr{G}^{T}), we can look at its Taylor expansion with respect to tt. We have then that there is gC(𝒢T)g\in C^{\infty}(\mathscr{G}^{T}) such that f=λgf=\lambda\cdot g. Now, in the definition of 𝒮c(𝒢T)\mathscr{S}_{c}(\mathscr{G}^{T}) we imposed a condition on the partial derivates t\partial_{t} with respect to tt (condition (s1)(s_{1}) in definition 4.1). This condition implies g𝒮c(𝒢T)g\in\mathscr{S}_{c}(\mathscr{G}^{T}). That is, Jλ𝒮c(𝒢T)J\subset\lambda\cdot\mathscr{S}_{c}(\mathscr{G}^{T}). ∎

Lemma 5.6

Let (M,X)(M,X) be a CC^{\infty} pair. Let f𝒮c(𝒟XM)f\in\mathscr{S}_{c}(\mathscr{D}_{X}^{M}) and kk\in\mathbb{N}^{*}. We define, for (m,t)M×[0,1](m,t)\in M\times[0,1],

fk(m,t)={0 if t=0tkf(m,t) if t0f_{k}(m,t)=\left\{\begin{array}[]{cc}0&\mbox{ if }t=0\\ t^{k}f(m,t)&\mbox{ if }t\neq 0\end{array}\right.

Then fkCck(M×[0,1])f_{k}\in C_{c}^{k}(M\times[0,1]).

Proof.

Since CkC^{k} is a local property, we can assume M=Up×qM=U\subset\mathbb{R}^{p}\times\mathbb{R}^{q} open and X=V=U(p×{0})X=V=U\cap(\mathbb{R}^{p}\times\{0\}). We have then to show that if f𝒮c(Ω)f\in\mathscr{S}_{c}(\Omega) (we remind Ω={(x,ξ,t)p×q×[0,1]:(x,tξ)U}\Omega=\{(x,\xi,t)\in\mathbb{R}^{p}\times\mathbb{R}^{q}\times[0,1]:(x,t\xi)\in U\}) then

fk(x,ξ,t)={0 if t=0tkf(x,ξt,t) if t0f_{k}(x,\xi,t)=\left\{\begin{array}[]{cc}0&\mbox{ if }t=0\\ t^{k}f(x,\frac{\xi}{t},t)&\mbox{ if }t\neq 0\end{array}\right.

belongs to Cck(U×[0,1])C_{c}^{k}(U\times[0,1]). Let us show this point.

First, the only problem is at t=0t=0, out of zero fkf_{k} is CC^{\infty}. Let (x,ξ,0)Ω(x,\xi,0)\in\Omega, we are going to see that fkf_{k} is CkC^{k} at this point. We can suppose ξ0\xi\neq 0 because otherwise the result is trivial.

Let uξ(t)=ξtu_{\xi}(t)=\frac{\xi}{t}, we have uξC((0,1],q)u_{\xi}\in C^{\infty}((0,1],\mathbb{R}^{q}) and uξu_{\xi} satisfies

limt0uξ(t)=+lim_{t\rightarrow 0}\|u_{\xi}(t)\|=+\infty (21)

We recall 𝒮c(ΩVU)\mathscr{S}_{c}(\Omega_{V}^{U}) consists of compact conic supported maps gC(ΩVU)g\in C^{\infty}(\Omega_{V}^{U}) satisfying the following condition:

  • (r1(r_{1})

    \forall n,mn,m\in\mathbb{N}, lpl\in\mathbb{N}^{p} and αq\alpha\in\mathbb{N}^{q} it exists C(n,m,l,α)>0C_{(n,m,l,\alpha)}>0 such that

    (1+ξ2)nxlξαtmg(x,ξ,t)C(n,m,l,α)(1+\|\xi\|^{2})^{n}\|\partial_{x}^{l}\partial_{\xi}^{\alpha}\partial_{t}^{m}g(x,\xi,t)\|\leq C_{(n,m,l,\alpha)}

By hypothesis f𝒮c(Ω)f\in\mathscr{S}_{c}(\Omega), hence, thanks to equation (21) above, we have for that

limt0uξ(t)rzαf(x,uξ(t),t)=0,lim_{t\rightarrow 0}\|u_{\xi}(t)\|^{r}\|\partial_{z}^{\alpha}f(x,u_{\xi}(t),t)\|=0, (22)

for every rr\in\mathbb{N} and αp×q×\alpha\in\mathbb{N}^{p}\times\mathbb{N}^{q}\times\mathbb{N}, here z=(x,ξ,t)z=(x,\xi,t).

By definition fk(x,ξ,t)=tkf(x,uξ(t),t)f_{k}(x,\xi,t)=t^{k}\cdot f(x,u_{\xi}(t),t) for t0t\neq 0 and zero otherwise, in particular fkf_{k} is CC^{\infty} for t0t\neq 0. Now, out of zero, an induction argument shows that for α=(l,γ)p×(q×)\alpha=(l,\gamma)\in\mathbb{N}^{p}\times(\mathbb{N}^{q}\times\mathbb{N}) we have that

zαfk(x,ξ,t)=tk|γ||β||α|aβPβ(x,ξ,t),\partial_{z}^{\alpha}f_{k}(x,\xi,t)=t^{k-|\gamma|}\sum_{|\beta|\leq|\alpha|}a_{\beta}P_{\beta}(x,\xi,t),

where aβa_{\beta} are constants (depeding on α\alpha) and Pβ(x,ξ,t)P_{\beta}(x,\xi,t) is a finite sum of products of the following type

tm0(uξ1(t))m1(uξq(t))mqzδf(x,uξ(t),t)t^{m_{0}}\cdot(u_{\xi_{1}}(t))^{m_{1}}\cdots(u_{\xi_{q}}(t))^{m_{q}}\cdot\partial_{z}^{\delta}f(x,u_{\xi}(t),t), mim_{i}\in\mathbb{N},

where uξi(t)=ξitu_{\xi_{i}}(t)=\frac{\xi_{i}}{t}. Then, if |γ|k|\gamma|\leq k we can use limit (22) to obtain that

limt0zαfk(x,ξ,t)=0.lim_{t\rightarrow 0}\|\partial_{z}^{\alpha}f_{k}(x,\xi,t)\|=0.

That is, fkf_{k} is at least of class CkC^{k} at zero. Finally, the support of fkf_{k} is contained in the conic support of ff, which is compact by assumption. ∎

Lemma 5.7

Let kk\in\mathbb{N} and q:=dim𝒢xq:=dim\,\mathscr{G}_{x}, we define (for J=Ker(e0)J=Ker(e_{0}), where 𝒮c(𝒢T)e0𝒮(A𝒢)\mathscr{S}_{c}(\mathscr{G}^{T})\stackrel{{\scriptstyle e_{0}}}{{\longrightarrow}}\mathscr{S}(A\mathscr{G}))

φk:Jk+qCck(𝒢×[0,1])\varphi^{k}:J^{k+q}\rightarrow C_{c}^{k}(\mathscr{G}\times[0,1])

in the following way:

φk(f)(γ,t)={0 if t=0tqf(γ,t) if t0\varphi^{k}(f)(\gamma,t)=\left\{\begin{array}[]{cc}0&\mbox{ if }t=0\\ t^{-q}f(\gamma,t)&\mbox{ if }t\neq 0\end{array}\right.

Then, we have a well defined morphism of algebras

φk:Jk+qCck(𝒢×[0,1])\varphi^{k}:J^{k+q}\rightarrow C_{c}^{k}(\mathscr{G}\times[0,1])
Proof.

The fact that φk\varphi_{k} is well defined is an immediate consequence of lemmas 5.5 and 5.6 above. Let us check that it is an algebra morphism. We recall that we are only considering Haar systems of the type described at 3.4 above.

Let f,g𝒮c(𝒢T)f,g\in\mathscr{S}_{c}(\mathscr{G}^{T}), hence

φk(fg)(γ,t)={0 if t=0tq𝒢s(γ)f(γη1)g(η,t)tq𝑑μs(γ)(η) if t0\varphi^{k}(f*g)(\gamma,t)=\left\{\begin{array}[]{cc}0&\mbox{ if }t=0\\ \\ t^{-q}\displaystyle\int_{\mathscr{G}_{s(\gamma)}}f(\gamma\circ\eta^{-1})g(\eta,t)t^{-q}d\mu_{s(\gamma)}(\eta)&\mbox{ if }t\neq 0\end{array}\right.
={0 if t=0𝒢s(γ)tqf(γη1)tqg(η,t)𝑑μs(γ)(η) if t0=\left\{\begin{array}[]{cc}0&\mbox{ if }t=0\\ \\ \displaystyle\int_{\mathscr{G}_{s(\gamma)}}t^{-q}\cdot f(\gamma\circ\eta^{-1})t^{-q}\cdot g(\eta,t)d\mu_{s(\gamma)}(\eta)&\mbox{ if }t\neq 0\end{array}\right.
=(φk(f)φk(g))(γ,t).=(\varphi^{k}(f)*\varphi_{k}(g))(\gamma,t).

Remark 5.8.

Let φk:K0(Jk+q)K0(Cck(𝒢×[0,1]))\varphi_{k}:K_{0}(J^{k+q})\rightarrow K_{0}(C_{c}^{k}(\mathscr{G}\times[0,1])) the KK-theory morphism induced by φk\varphi^{k}. By construction we have the two following properties:

  • (a)

    (e0)φk=0(e_{0})_{*}\circ\varphi_{k}=0 où where e0e_{0} denotes the evaluation e0:Cck(𝒢×[0,1])Cck(𝒢)e_{0}:C_{c}^{k}(\mathscr{G}\times[0,1])\rightarrow C_{c}^{k}(\mathscr{G}).

  • (b)

    The following diagram is commutative

    0\textstyle{0}K0(J)\textstyle{K_{0}(J)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K0(𝒮c(𝒢T))\textstyle{K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e0\scriptstyle{e_{0}}e1k\scriptstyle{e_{1}^{k}}K0(𝒮(A𝒢))\textstyle{K_{0}(\mathscr{S}(A\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}K0(Jk+q)\textstyle{K_{0}(J^{k+q})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j\scriptstyle{j}φk\scriptstyle{\varphi_{k}}K0(Cck(𝒢×[0,1]))\textstyle{K_{0}(C_{c}^{k}(\mathscr{G}\times[0,1]))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e1\scriptstyle{e_{1}}K0(Cck(𝒢))\textstyle{K_{0}(C_{c}^{k}(\mathscr{G}))}

    where K0(Jk+q)jK0(J)K_{0}(J^{k+q})\stackrel{{\scriptstyle j}}{{\longrightarrow}}K_{0}(J) is induced by the inclusion Jk+qJJ^{k+q}\subset J. The fact that jj is surjective is immediate from the KK-theory exact sequence

    K0(JN)K0(J)K0(J/JN),K_{0}(J^{N})\rightarrow K_{0}(J)\rightarrow K_{0}(J/J^{N}),

    since K0(J/JN)=0K_{0}(J/J^{N})=0 because J/JNJ/J^{N} is a nilpotent algebra (see for instance [Ros94]).

Lemma 5.9

Let wK0(𝒮c(𝒢T))w\in K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T})) with wKer(e0)w\in Ker(e_{0})_{*}. Then e1h,k(w)=0e_{1}^{h,k}(w)=0, where e1h,ke_{1}^{h,k} is the composition

K0(𝒮c(𝒢T))e1K0(Cc(𝒢))ιkK0(Cck(𝒢))πK0h,k(𝒢).K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T}))\stackrel{{\scriptstyle e_{1}}}{{\rightarrow}}K_{0}(C^{\infty}_{c}(\mathscr{G}))\stackrel{{\scriptstyle\iota_{k}}}{{\rightarrow}}K_{0}(C_{c}^{k}(\mathscr{G}))\stackrel{{\scriptstyle\pi}}{{\rightarrow}}K_{0}^{h,k}(\mathscr{G}).
Proof.

By the exact sequence in KK-theory, (17), we can take xK0(J)x\in K_{0}(J) such that i(x)=wi_{*}(x)=w. We can choose yK0(Jk+q)y\in K_{0}(J^{k+q}) with j(y)=xj(y)=x, because K0(Jk+q)jK0(J)K_{0}(J^{k+q})\stackrel{{\scriptstyle j}}{{\longrightarrow}}K_{0}(J) is surjective (see discussion above). Now, the condition (b) in the precedent remark implies e1(φk(y))=e1k(w)e_{1}(\varphi_{k}(y))=e_{1}^{k}(w), and since e0(φk(y))=0e_{0}(\varphi_{k}(y))=0 it follows that e1k(w)=e1(φk(y))h0e_{1}^{k}(w)=e_{1}(\varphi_{k}(y))\sim_{h}0 in K0(Cck(𝒢))K_{0}(C_{c}^{k}(\mathscr{G})). ∎

We can now pass to the proof of the theorem:

Proof of theorem 5.4..
  1. 1.

    We have only to show the existence of the mentioned morphism since uniqueness will follow immediately from the surjectivity of the evaluation morphism K0(𝒮c(𝒢T))e0K0(A𝒢)K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T}))\stackrel{{\scriptstyle e_{0}}}{{\longrightarrow}}K_{0}(A^{*}\mathscr{G}), (proposition 5.1).

    Let kk\in\mathbb{N}. Let σK0(A𝒢)\sigma\in K_{0}(A^{*}\mathscr{G}) and take wσK0(𝒮c(𝒢T))w_{\sigma}\in K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T})) with e0(wσ)=σe_{0}(w_{\sigma})=\sigma. We put

    indah,k(σ):=e1h,k(wσ).ind_{a}^{h,k}(\sigma):=e_{1}^{h,k}(w_{\sigma}).

    From lemma above it follows that indah,k(σ)ind_{a}^{h,k}(\sigma) does not depend on the choice of wσw_{\sigma}. Finally, the fact that indah,kind_{a}^{h,k} is a group morphism follows immediately from the fact that e0:K0(𝒮c(𝒢T))K0(A𝒢)e_{0}:K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T}))\rightarrow K_{0}(A^{*}\mathscr{G}) and e1k:K0(𝒮c(𝒢T))K0h,k(𝒢)e_{1}^{k}:K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T}))\rightarrow K_{0}^{h,k}(\mathscr{G}) are group morphisms.

    Now, the morphism

    indaF:K0(A𝒢)K0F(𝒢)ind_{a}^{F}:K^{0}(A^{*}\mathscr{G})\rightarrow K_{0}^{F}(\mathscr{G})

    is just the induced by all the indices indah,kind_{a}^{h,k} by the universal property of the projective limit.

  2. 2.

    It is enough to show the commutativity of the diagram (20) for each indah,kind_{a}^{h,k}:

    First, let us consider the set of elliptic operators on the tangent groupoid, Ell(𝒢T)Ell(\mathscr{G}^{T}). From this set, we also have two evaluation maps

    Ell(A𝒢)e0Ell(𝒢T)e1Ell(𝒢).Ell(A\mathscr{G})\stackrel{{\scriptstyle e_{0}}}{{\longleftarrow}}Ell(\mathscr{G}^{T})\stackrel{{\scriptstyle e_{1}}}{{\longrightarrow}}Ell(\mathscr{G}).

    It is immediate that this evaluation maps commute with indices, i.e., the following diagrams are commutative

    Ell(A𝒢)\textstyle{Ell(A\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ind\scriptstyle{ind}Ell(𝒢T)\textstyle{Ell(\mathscr{G}^{T})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e0\scriptstyle{e_{0}}e1\scriptstyle{e_{1}}ind\scriptstyle{ind}Ell(𝒢)\textstyle{Ell(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ind\scriptstyle{ind}K0(Cc(A𝒢))\textstyle{K_{0}(C^{\infty}_{c}(A\mathscr{G}))}K0(Cc(𝒢T))\textstyle{K_{0}(C^{\infty}_{c}(\mathscr{G}^{T}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e0\scriptstyle{e_{0}}e1\scriptstyle{e_{1}}K0(Cc(𝒢)).\textstyle{K_{0}(C_{c}^{\infty}(\mathscr{G})).}

    Now, we know from the existence of a asymptotic pseudodifferential calculus that e1e_{1} is a surjective function (see [CH90, MP97, NWX99]). In fact, the symbol map Ell(𝒢)σK0(A𝒢)Ell(\mathscr{G})\stackrel{{\scriptstyle\sigma}}{{\rightarrow}}K^{0}(A^{*}\mathscr{G}) can (alternately) defined as follows: Let PEll(𝒢)P\in Ell(\mathscr{G}) and P~Ell(𝒢T)\tilde{P}\in Ell(\mathscr{G}^{T}) an e1e_{1}-lifting (e1(P~)=Pe_{1}(\tilde{P})=P). We set

    σ(P):=j(ind(e0(P~))),\sigma(P):=j(ind(e_{0}(\tilde{P}))),

    where Ell(A𝒢)indK0(Cc(A𝒢))jK0(A𝒢)Ell(A\mathscr{G})\stackrel{{\scriptstyle ind}}{{\rightarrow}}K_{0}(C^{\infty}_{c}(A\mathscr{G}))\stackrel{{\scriptstyle j}}{{\rightarrow}}K^{0}(A^{*}\mathscr{G}). That is, the symbol map fits in the following commutative diagram

    Ell(𝒢T)\textstyle{Ell(\mathscr{G}^{T})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e0\scriptstyle{e_{0}}e1\scriptstyle{e_{1}}Ell(𝒢)\textstyle{Ell(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}symb.\scriptstyle{symb.}Ell(A𝒢)\textstyle{Ell(A\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ind\scriptstyle{ind}inda,A𝒢\scriptstyle{ind_{a,A\mathscr{G}}}K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})}K0(Cc(A𝒢))\textstyle{K_{0}(C^{\infty}_{c}(A\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j\scriptstyle{j}

    Now, for proving the commutativity of

    Ell(𝒢)\textstyle{Ell(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}symb\scriptstyle{symb}ind\scriptstyle{ind}K0(Cc(𝒢))\textstyle{K_{0}(C^{\infty}_{c}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πι\scriptstyle{\pi\circ\iota}K0(A𝒢)\textstyle{K_{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}indah,k\scriptstyle{ind_{a}^{h,k}}K0h,k(𝒢)\textstyle{K_{0}^{h,k}(\mathscr{G})}

    we decompose it in commutative diagrams in the following way

    Ell(𝒢)\textstyle{Ell(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}symb\scriptstyle{symb}ind\scriptstyle{ind}Ell(𝒢T)\textstyle{Ell(\mathscr{G}^{T})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e1\scriptstyle{e_{1}}e0\scriptstyle{e_{0}}ind\scriptstyle{ind}K0(Cc(𝒢T))\textstyle{K_{0}(C^{\infty}_{c}(\mathscr{G}^{T}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e0\scriptstyle{e_{0}}j\scriptstyle{j}e1\scriptstyle{e_{1}}K0(Cc(𝒢))\textstyle{K_{0}(C^{\infty}_{c}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j\scriptstyle{j}πj\scriptstyle{\pi\circ j}Ell(A𝒢)\textstyle{Ell(A\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ind\scriptstyle{ind}K0(Cc(A𝒢))\textstyle{K_{0}(C^{\infty}_{c}(A\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j\scriptstyle{j}K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}indah,k\scriptstyle{ind_{a}^{h,k}}K0(𝒮c(𝒢T))\textstyle{K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e0\scriptstyle{e_{0}}e1k\scriptstyle{e_{1}^{k}}K0(Cck(𝒢))\textstyle{K_{0}(C_{c}^{k}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}K0h,k(𝒢)\textstyle{K_{0}^{h,k}(\mathscr{G})}

    Now, for the commutativity of

    K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{id}indah,k\scriptstyle{ind_{a}^{h,k}}K0h,k(𝒢)\textstyle{K_{0}^{h,k}(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι\scriptstyle{\iota}K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda\scriptstyle{ind_{a}}K0(Cr(𝒢))\textstyle{K_{0}(C_{r}^{*}(\mathscr{G}))}

    we proceed as above, i.e., we decompose it as

    K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{id}K0(𝒮c(𝒢T))\textstyle{K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e0\scriptstyle{e_{0}}ι\scriptstyle{\iota}e1k\scriptstyle{e_{1}^{k}}K0(Cck(𝒢))\textstyle{K_{0}(C_{c}^{k}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι\scriptstyle{\iota}π\scriptstyle{\pi}K0(Cr(𝒢T))\textstyle{K_{0}(C_{r}^{*}(\mathscr{G}^{T}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e1\scriptstyle{e_{1}}K0h,k(𝒢)\textstyle{K_{0}^{h,k}(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι~\scriptstyle{\tilde{\iota}}K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda\scriptstyle{ind_{a}}e01\scriptstyle{e_{0}^{-1}}K0(Cr(𝒢)),\textstyle{K_{0}(C_{r}^{*}(\mathscr{G})),}

    The fact that the index indaFind_{a}^{F} also satisfies diagram (20) comes from the universal properties of projective limits and co-equalizers (see (18)).

Definition 5.10 (Compactly supported analytic index).

The morphism given by the precedent theorem and its corollary is called The compactly supported analytic index of 𝒢\mathscr{G}.

In the next subsection we are going to slightly modify the indices we constructed, but before that, we give two important properties of these indices. The first one is related to the Bott morphism, so we first describe what the Bott morphism is in our setting.

Let 𝒢2:=𝒢×2𝒢(0)×2\mathscr{G}^{\mathbb{R}^{2}}:=\mathscr{G}\times\mathbb{R}^{2}\rightrightarrows\mathscr{G}^{(0)}\times\mathbb{R}^{2} be the product groupoid, where 22\mathbb{R}^{2}\rightrightarrows\mathbb{R}^{2} is the identity groupoid. Let us first recall that the algebra Cc(2)C^{\infty}_{c}(\mathbb{R}^{2}) (with the punctual product) is stable under holomorphic calculus in C0(2)C_{0}(\mathbb{R}^{2}), hence the inclusion induces an isomorphism in KK-theory. In particular, the Bott element can be seen as an element in K0(Cc(2))K0(C0(2))K_{0}(C^{\infty}_{c}(\mathbb{R}^{2}))\approx K_{0}(C_{0}(\mathbb{R}^{2})). Therefore, we can consider, for each 0k0\leq k\leq\infty, the Bott morphism

Bottk:K0(Cck(𝒢))K0(Cck(𝒢2)),Bott_{k}:K_{0}(C_{c}^{k}(\mathscr{G}))\rightarrow K_{0}(C_{c}^{k}(\mathscr{G}^{\mathbb{R}^{2}})),

that is just the product by the Bott element.

Now, from the fact that the product in KK-theory is natural it follows that the morphism BottkBott_{k} passes to the quotient K0h,k(𝒢)K_{0}^{h,k}(\mathscr{G}), i.e., we get a Bott morphism K0h,k(𝒢)BottK0h,k(𝒢2)K_{0}^{h,k}(\mathscr{G})\stackrel{{\scriptstyle Bott}}{{\longrightarrow}}K_{0}^{h,k}(\mathscr{G}^{\mathbb{R}^{2}}). Furthermore, by using universal properties we easily extend this morphism to K0F(𝒢)BottK0F(𝒢2)K_{0}^{F}(\mathscr{G})\stackrel{{\scriptstyle Bott}}{{\longrightarrow}}K_{0}^{F}(\mathscr{G}^{\mathbb{R}^{2}}). The following compatibility result will be very useful in te sequel.

Proposition 5.11

The compactly supported index inda,𝒢Find_{a,\mathscr{G}}^{F} is compatible with the Bott morphism, i.e.\it{i.e.}, the following diagram is commutative

K0(A𝒢)\textstyle{K_{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,𝒢F\scriptstyle{ind_{a,\mathscr{G}}^{F}}Bott\scriptstyle{Bott}K0F(𝒢)\textstyle{K_{0}^{F}(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}BottF\scriptstyle{Bott_{F}}K0(A𝒢2)\textstyle{K_{0}(A^{*}\mathscr{G}^{\mathbb{R}^{2}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,𝒢2F\scriptstyle{ind_{a,\mathscr{G}^{\mathbb{R}^{2}}}^{F}}K0F(𝒢2)\textstyle{K_{0}^{F}(\mathscr{G}^{\mathbb{R}^{2}})}
Proof.

It is enough to check that, for each kk\in\mathbb{N}, the following diagram is commutative:

K0(A𝒢)\textstyle{K_{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,𝒢h,k\scriptstyle{ind_{a,\mathscr{G}}^{h,k}}Bott\scriptstyle{Bott}K0h,k(𝒢)\textstyle{K_{0}^{h,k}(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Bottk\scriptstyle{Bott_{k}}K0(A𝒢2)\textstyle{K_{0}(A^{*}\mathscr{G}^{\mathbb{R}^{2}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,𝒢2h,k\scriptstyle{ind_{a,\mathscr{G}^{\mathbb{R}^{2}}}^{h,k}}K0h,k(𝒢2)\textstyle{K_{0}^{h,k}(\mathscr{G}^{\mathbb{R}^{2}})}

Now, by multiplying again by the Bott element (seen in K0(Cc(2))K_{0}(C^{\infty}_{c}(\mathbb{R}^{2})) as above) we have another Bott morphism

BottT:𝒮c(𝒢T)𝒮c((𝒢2)T).Bott_{T}:\mathscr{S}_{c}(\mathscr{G}^{T})\longrightarrow\mathscr{S}_{c}((\mathscr{G}^{\mathbb{R}^{2}})^{T}).

We use again the fact that the product in KK-theory is natural and in particular it commutes with the evaluation morphisms to see that we can decompose the last diagram in commutative diagrams in the following way:

K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,𝒢h,k\scriptstyle{ind_{a,\mathscr{G}}^{h,k}}Bott\scriptstyle{Bott}K0h,k(𝒢)\textstyle{K_{0}^{h,k}(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Bottk\scriptstyle{Bott_{k}}K0(𝒮c(𝒢T))\textstyle{K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e0\scriptstyle{e_{0}}e1k\scriptstyle{e_{1}^{k}}BottT\scriptstyle{Bott_{T}}K0(Cck(𝒢))\textstyle{K_{0}(C_{c}^{k}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}Bottk\scriptstyle{Bott_{k}}K0(𝒮c((𝒢2)T))\textstyle{K_{0}(\mathscr{S}_{c}((\mathscr{G}^{\mathbb{R}^{2}})^{T}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e0\scriptstyle{e_{0}}e1k\scriptstyle{e_{1}^{k}}K0(Cck(𝒢2))\textstyle{K_{0}(C_{c}^{k}(\mathscr{G}^{\mathbb{R}^{2}}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}K0(A(𝒢2))\textstyle{K^{0}(A^{*}(\mathscr{G}^{\mathbb{R}^{2}}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,𝒢2h,k\scriptstyle{ind_{a,\mathscr{G}^{\mathbb{R}^{2}}}^{h,k}}K0h,k(𝒢2).\textstyle{K_{0}^{h,k}(\mathscr{G}^{\mathbb{R}^{2}}).}

The second property is related with the inclusions of open subgroupoids. Let 𝒢𝒢(0)\mathscr{G}\rightrightarrows\mathscr{G}^{(0)} be a Lie groupoid and (0)\mathscr{H}\rightrightarrows\mathscr{H}^{(0)} be an open subgroupoid. We have the following compatibility result:

Proposition 5.12

The following diagram is commutative:

K0(A)\textstyle{K^{0}(A^{*}\mathscr{H})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,F\scriptstyle{ind_{a,\mathscr{H}}^{F}}K0F()\textstyle{K_{0}^{F}(\mathscr{H})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,𝒢F\scriptstyle{ind_{a,\mathscr{G}}^{F}}K0F(𝒢).\textstyle{K_{0}^{F}(\mathscr{G}).}

where the vertical maps are induced from the inclusions by open subgroupoids.

Proof.

It is enough to check the proposition for each index of order kk, indah,kind_{a}^{h,k}, for all kk\in\mathbb{N}.

First, note that T𝒢T\mathscr{H}^{T}\subset\mathscr{G}^{T} is an open subset. Even more, the algebra inclusion

𝒮c(T)𝒮c(𝒢T)\mathscr{S}_{c}(\mathscr{H}^{T})\hookrightarrow\mathscr{S}_{c}(\mathscr{G}^{T})

commutes with all evaluations. In particular, the following diagram is commutative

K0(A)\textstyle{K^{0}(A^{*}\mathscr{H})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j\scriptstyle{j}K0(𝒮c(T))\textstyle{K_{0}(\mathscr{S}_{c}(\mathscr{H}^{T}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e0\scriptstyle{e_{0}}j\scriptstyle{j}e1\scriptstyle{e_{1}}K0(Cc())\textstyle{K_{0}(C^{\infty}_{c}(\mathscr{H}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j\scriptstyle{j}K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})}K0(𝒮c(𝒢T))\textstyle{K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e0\scriptstyle{e_{0}}e1\scriptstyle{e_{1}}K0(Cc(𝒢))\textstyle{K_{0}(C_{c}^{\infty}(\mathscr{G}))}

where the morphisms noted by jj are induced by the extension by zero outside the open subsets. The conclusion is now immediate as we can decompose the diagram in the enouncement of the proposition in the following way

K0(A)\textstyle{K^{0}(A^{*}\mathscr{H})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,h,k\scriptstyle{ind_{a,\mathscr{H}}^{h,k}}j\scriptstyle{j}K0h,k()\textstyle{K_{0}^{h,k}(\mathscr{H})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j\scriptstyle{j}K0(𝒮c(T))\textstyle{K_{0}(\mathscr{S}_{c}(\mathscr{H}^{T}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e0\scriptstyle{e_{0}}e1\scriptstyle{e_{1}}j\scriptstyle{j}K0(Cc())\textstyle{K_{0}(C_{c}^{\infty}(\mathscr{H}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j\scriptstyle{j}π\scriptstyle{\pi}K0(𝒮c(𝒢T))\textstyle{K_{0}(\mathscr{S}_{c}(\mathscr{G}^{T}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e0\scriptstyle{e_{0}}e1\scriptstyle{e_{1}}K0(Cc(𝒢))\textstyle{K_{0}(C_{c}^{\infty}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,𝒢h,k\scriptstyle{ind_{a,\mathscr{G}}^{h,k}}K0h,k(𝒢),\textstyle{K_{0}^{h,k}(\mathscr{G}),}

Periodic compactly analytic index

Definition 5.13 (Periodic compactly analytic index).

As we saw above, we can consider the Bott morphism K0F(𝒢)BottK0F(𝒢2)K_{0}^{F}(\mathscr{G})\stackrel{{\scriptstyle Bott}}{{\longrightarrow}}K_{0}^{F}(\mathscr{G}^{\mathbb{R}^{2}}). We can take the inductive limit

limmK0F(𝒢×2m)\varinjlim_{m}K_{0}^{F}(\mathscr{G}\times\mathbb{R}^{2m})

induced by K0F(𝒢×2m)BottK0F(𝒢×2(m+1))K_{0}^{F}(\mathscr{G}\times\mathbb{R}^{2m})\stackrel{{\scriptstyle Bott}}{{\longrightarrow}}K_{0}^{F}(\mathscr{G}\times\mathbb{R}^{2(m+1)}). We note this group by

K0B(𝒢)=limmK0F(𝒢×2m),K_{0}^{B}(\mathscr{G})=\varinjlim_{m}K_{0}^{F}(\mathscr{G}\times\mathbb{R}^{2m}), (23)

and we call it the Periodic KK-theory of 𝒢\mathscr{G}.

Let

inda,𝒢B:K0(A𝒢)K0B(𝒢)ind_{a,\mathscr{G}}^{B}:K^{0}(A^{*}\mathscr{G})\rightarrow K_{0}^{B}(\mathscr{G})

be the morphism given by the composition of inda,𝒢F:K0(A𝒢)K0F(𝒢)ind_{a,\mathscr{G}}^{F}:K^{0}(A^{*}\mathscr{G})\rightarrow K_{0}^{F}(\mathscr{G}) followed by K0F(𝒢)BottK0B(𝒢)K_{0}^{F}(\mathscr{G})\stackrel{{\scriptstyle Bott}}{{\longrightarrow}}K_{0}^{B}(\mathscr{G}). We call this morphism the Periodic compactly analytic index of 𝒢\mathscr{G}.

Remark 5.14.
  1. 1.

    K0B(𝒢)K_{0}^{B}(\mathscr{G}) satisfy Bott periodicity by construction, i.e.,

    K0B(𝒢)BottK0B(𝒢2)K_{0}^{B}(\mathscr{G})\stackrel{{\scriptstyle Bott}}{{\longrightarrow}}K_{0}^{B}(\mathscr{G}^{\mathbb{R}^{2}})

    is an isomorphism.

  2. 2.

    The periodic analytic index is also intermediate between CC^{\infty} and CrC^{*}_{r}, i.e., it satisfy diagram (20) too:

    Ell(𝒢)\textstyle{Ell(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}symb\scriptstyle{symb}ind\scriptstyle{ind}K0(Cc(𝒢))\textstyle{K_{0}(C^{\infty}_{c}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K0(A𝒢)\textstyle{K_{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{id}indaB\scriptstyle{ind_{a}^{B}}K0B(𝒢)\textstyle{K_{0}^{B}(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K0(A𝒢)\textstyle{K_{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda\scriptstyle{ind_{a}}K0(Cr(𝒢))\textstyle{K_{0}(C_{r}^{*}(\mathscr{G}))}. (24)

    where K0B(𝒢)K0(Cr(𝒢))K_{0}^{B}(\mathscr{G})\rightarrow K_{0}(C_{r}^{*}(\mathscr{G})) is induced from

    K0F(𝒢×2m)K0(Cr(𝒢×2m))K0(Cr(𝒢))K_{0}^{F}(\mathscr{G}\times\mathbb{R}^{2m})\rightarrow K_{0}(C_{r}^{*}(\mathscr{G}\times\mathbb{R}^{2m}))\cong K_{0}(C_{r}^{*}(\mathscr{G}))

    using the Bott periodicity of KK-theory for CC^{*}-algebras.

6 Applications

6.1 Longitudinal index theorem

Let (M,F)(M,F) be a foliated manifold with holonomy groupoid 𝒢M\mathscr{G}\rightrightarrows M. In this case the Lie algebroid is given by the integrable subbundle FF.

In [CS84], Connes-Skandalis define a topological index indt:K0(F)K0(Cr(𝒢))ind_{t}:K^{0}(F^{*})\rightarrow K_{0}(C_{r}^{*}(\mathscr{G})) and they show the equality with the CC^{*}-analytic index of 𝒢\mathscr{G}.

We will establish a more primitive longitudinal index theorem. That is, we will see that the equality between the indices takes already place in the group K0B(𝒢)K_{0}^{B}(\mathscr{G}).

Before stating the longitudinal index theorem, we will need the next proposition (for more details see [Con94] II.5 or [DLN06] section 6.1).

Proposition 6.1

Let NpTN\stackrel{{\scriptstyle p}}{{\longrightarrow}}T be a vector bundle over TT. We consider the Thom groupoid associated to it, i.e., :=N×TNN\mathscr{H}:=N\times_{T}N\rightrightarrows N, which has Lie algebroid NNNN\oplus N^{*}\cong N\otimes\mathbb{C}. Then, the following diagram is commutative:

K0(A𝒢𝒯)\textstyle{K^{0}(A^{*}\mathscr{G}_{\mathscr{T}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,𝒢𝒯\scriptstyle{ind_{a,\mathscr{G}_{\mathscr{T}}}}Thom1\scriptstyle{Thom^{-1}}K0(T)\textstyle{K^{0}(T)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\mathscr{M}}K0F(𝒢𝒯)\textstyle{K_{0}^{F}(\mathscr{G}_{\mathscr{T}})}K0(Cc(𝒢𝒯)),\textstyle{K_{0}(C^{\infty}_{c}(\mathscr{G}_{\mathscr{T}}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces,}π\scriptstyle{\pi}

where \mathscr{M} is the morphism given by the Morita equivalence between 𝒢𝒯\mathscr{G}_{\mathscr{T}} and TT. In other words, modulo Fourier and Morita, the compactly supported analytic index of 𝒢𝒯\mathscr{G}_{\mathscr{T}} coincides with the Thom isomorphism’s.

Proof.

It is known (see for example [DLN06] theorem 6.2 or [Con94] II.5) that the following diagram is commutative:

K0(A)\textstyle{K^{0}(A\mathscr{H})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Thom1\scriptstyle{Thom^{-1}}K0(T)\textstyle{K^{0}(T)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\mathscr{M}}\scriptstyle{\approx}K0(A)\textstyle{K^{0}(A^{*}\mathscr{H})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,\scriptstyle{ind_{a,\mathscr{H}}}Fourier\scriptstyle{Fourier}\scriptstyle{\approx}K0(Cr()),\textstyle{K_{0}(C_{r}^{*}(\mathscr{H})),}

where \mathscr{M} is the morphism given by the Morita equivalence between 𝒢𝒯\mathscr{G}_{\mathscr{T}} and TT.

It will be then enough to prove that Cc()Cr()C^{\infty}_{c}(\mathscr{H})\subset C_{r}^{*}(\mathscr{H}) is stable under holomorphic calculus. Now, from the classical fact that the algebra of smooth kernel operators Cc(2m×2m)C^{\infty}_{c}(\mathbb{R}^{2m}\times\mathbb{R}^{2m}) is stable under holomorphic calculus on the compact operators algebra 𝒦\mathscr{K}, we easily get that Cc()C^{\infty}_{c}(\mathscr{H}) is stable under holomorphic calculus on Cr()C^{*}_{r}(\mathscr{H}) (locally it reduces to the case of smooth kernel operators, see [Trè80] for instance). In particular K0(Cc())K0(Cr())K_{0}(C^{\infty}_{c}(\mathscr{H}))\approx K_{0}(C^{*}_{r}(\mathscr{H})) and the proof is complete. ∎

Let us now define the periodic topological index of a foliation. The definition is analogue to the Connes-Skandalis definition of the CC^{*}-topological index.

Before stating the definition, let us recall the following fact: If TVT_{V} is an open transversal of a foliated manifold (V,FV)(V,F_{V}) with holonomy groupoid 𝒢V\mathscr{G}_{V}, then there a well defined morphism

K0(TV)=K0(Cc(TV))iK0(Cc(𝒢V)),K^{0}(T_{V})=K_{0}(C^{\infty}_{c}(T_{V}))\stackrel{{\scriptstyle i}}{{\longrightarrow}}K_{0}(C^{\infty}_{c}(\mathscr{G}_{V})), (25)

induced by the inclusion of the open subgroupoid resulting from the restriction to the transversal and a suitable Morita quivalence, [CS84, BH04].

Definition 6.2.

[Periodic topological index] Let g:M2mg:M\hookrightarrow\mathbb{R}^{2m} be an embedding, we consider the foliation M×2mM\times\mathbb{R}^{2m} given by the integrable vector bundle F×{0}F\times\{0\}. This foliation has 𝒢~=𝒢×2m\tilde{\mathscr{G}}=\mathscr{G}\times\mathbb{R}^{2m} as a holonomy groupoid. Let TT be the normal vector bundle to the foliation in 2m\mathbb{R}^{2m}, Tx:=(g(Fx))T_{x}:=(g_{*}(F_{x}))^{\bot}. Now, the map h:TM×2mh:T\rightarrow M\times\mathbb{R}^{2m} given by (x,ξ)(x,g(x)+ξ)(x,\xi)\mapsto(x,g(x)+\xi) allows to identify TT with an open transversal of (M×2m,F~)(M\times\mathbb{R}^{2m},\tilde{F}), that we still denote by TT. Let NN be the normal vector bundle to the inclusion TM×2mT\subset M\times\mathbb{R}^{2m}, we can take a neighborhood Ω\Omega of TT in M×2mM\times\mathbb{R}^{2m} in such a way that :=𝒢~|ΩN×TN\mathscr{H}:=\tilde{\mathscr{G}}|_{\Omega}\approx N\times_{T}N where N×TNNN\times_{T}N\rightrightarrows N is the pair groupoid over TT (we keep the notation from proposition 6.1). This last groupoid has Lie algebroid A=NNF2mA\mathscr{H}=N\oplus N\approx F\oplus\mathbb{R}^{2m}. We can then consider the Bott isomorphism

K0(F)BottK0(A).K^{0}(F)\stackrel{{\scriptstyle Bott}}{{\rightarrow}}K^{0}(A\mathscr{H}).

By the Periodic topological index of 𝒢\mathscr{G} we mean the morphism

indt,𝒢B:K0(F)K0B(𝒢)ind_{t,\mathscr{G}}^{B}:K^{0}(F)\rightarrow K_{0}^{B}(\mathscr{G})

given by the composition

K0(F)BottK0(A)Thom1K0(T)ιK0B(𝒢)K^{0}(F)\stackrel{{\scriptstyle Bott}}{{\longrightarrow}}K^{0}(A\mathscr{H})\stackrel{{\scriptstyle Thom^{-1}}}{{\longrightarrow}}K^{0}(T)\stackrel{{\scriptstyle\iota}}{{\longrightarrow}}K_{0}^{B}(\mathscr{G})

where ι\iota is given by the composition

K0(T)iK0(Cc(𝒢×2m))K0F(𝒢×2m)K0B(𝒢),K^{0}(T)\stackrel{{\scriptstyle i}}{{\longrightarrow}}K_{0}(C^{\infty}_{c}(\mathscr{G}\times\mathbb{R}^{2m}))\longrightarrow K_{0}^{F}(\mathscr{G}\times\mathbb{R}^{2m})\longrightarrow K_{0}^{B}(\mathscr{G}), (26)

The morphism

K0(T)iK0(Cc(𝒢×2m))K^{0}(T)\stackrel{{\scriptstyle i}}{{\longrightarrow}}K_{0}(C^{\infty}_{c}(\mathscr{G}\times\mathbb{R}^{2m}))

is the one described in (25) above.

Remark 6.3.

If K0B(𝒢)iBK0(Cr(𝒢))K_{0}^{B}(\mathscr{G})\stackrel{{\scriptstyle i_{B}}}{{\longrightarrow}}K_{0}(C_{r}^{*}(\mathscr{G})) is the morphism induced by the inclusion, then

iBindt,𝒢B=indt,𝒢,i_{B}\circ ind_{t,\mathscr{G}}^{B}=ind_{t,\mathscr{G}},

where indt,𝒢ind_{t,\mathscr{G}} is the CC^{*}-topological index of Connes-Skandalis.

Now, we state the theorem in our setting.

Theorem 6.4

Let (M,F)(M,F) be a foliated manifold. With the same notations as above we have that

inda,𝒢B=indt,𝒢B.ind_{a,\mathscr{G}}^{B}=ind_{t,\mathscr{G}}^{B}.

In particular indt,𝒢Bind_{t,\mathscr{G}}^{B} does not depend on the choices made for its definition.

Proof.

We use the same notations as in the definition of the periodic topological index. We are going to show that the following two morphisms coincide

Bottinda,𝒢F:K0(F)K0F(𝒢×2m),\displaystyle Bott\circ ind_{a,\mathscr{G}}^{F}:K^{0}(F)\rightarrow K_{0}^{F}(\mathscr{G}\times\mathbb{R}^{2m}), (27)

and

K0(F)BottK0(A)Thom1K0(T)ιK0F(𝒢×2m)\displaystyle K^{0}(F)\stackrel{{\scriptstyle Bott}}{{\longrightarrow}}K^{0}(A\mathscr{H})\stackrel{{\scriptstyle Thom^{-1}}}{{\longrightarrow}}K^{0}(T)\stackrel{{\scriptstyle\iota}}{{\longrightarrow}}K_{0}^{F}(\mathscr{G}\times\mathbb{R}^{2m}) (28)

where ι\iota denotes the composition

K0(T)K0(Cc())K0(Cc(𝒢×2m))K0F(𝒢×2m),K^{0}(T)\stackrel{{\scriptstyle\mathscr{M}}}{{\longrightarrow}}K_{0}(C^{\infty}_{c}(\mathscr{H}))\longrightarrow K_{0}(C^{\infty}_{c}(\mathscr{G}\times\mathbb{R}^{2m}))\longrightarrow K_{0}^{F}(\mathscr{G}\times\mathbb{R}^{2m}),

that is, we will see that the equality of the indices happens before taking the Bott limit.

Now, from the compatibility of the compactly supported index with open subgrupoids and with Bott morphism, seen in last section, we have that inda,Find_{a,\mathscr{H}}^{F} and inda,𝒢×2mFind_{a,\mathscr{G}\times\mathbb{R}^{2m}}^{F} coincide modulo the inclusions by open subgroupoids, and we have also that inda,𝒢×2mFind_{a,\mathscr{G}\times\mathbb{R}^{2m}}^{F} and inda,𝒢Find_{a,\mathscr{G}}^{F} coincide modulo Bott. Hence, all the following diagrams are commutative

K0(F)\textstyle{K^{0}(F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,𝒢F\scriptstyle{ind_{a,\mathscr{G}}^{F}}Bott\scriptstyle{Bott}Bott\scriptstyle{Bott}K0(A)\textstyle{K^{0}(A\mathscr{H})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,F\scriptstyle{ind_{a,\mathscr{H}}^{F}}\scriptstyle{\approx}i\scriptstyle{i}Thom1\scriptstyle{Thom^{-1}}K0(T)\textstyle{K^{0}(T)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\mathscr{M}}K0(A𝒢×2m)\textstyle{K^{0}(A\mathscr{G}\times\mathbb{R}^{2m})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inda,𝒢×2mF\scriptstyle{ind_{a,\mathscr{G}\times\mathbb{R}^{2m}}^{F}}K0F()\textstyle{K_{0}^{F}(\mathscr{H})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i\scriptstyle{i}K0(Cc())\textstyle{K_{0}(C^{\infty}_{c}(\mathscr{H}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i\scriptstyle{i}π\scriptstyle{\pi}K0F(𝒢)\textstyle{K_{0}^{F}(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Bott\scriptstyle{Bott}K0F(𝒢×2m)\textstyle{K_{0}^{F}(\mathscr{G}\times\mathbb{R}^{2m})}K0(Cc(𝒢×2m)),\textstyle{K_{0}(C^{\infty}_{c}(\mathscr{G}\times\mathbb{R}^{2m}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces,}π\scriptstyle{\pi}

where the top right square is commutative from proposition 6.1. To conclude, we have just to remark that in the precedent diagram we have precisely, on one side, the morphism (27), left and first below; and on the other side the morphism (28). ∎

6.1.1 Bounded assembly map

We recall that the map

Dinda(D)K0(Cr(𝒢))D\mapsto ind_{a}(D)\in K_{0}(C_{r}^{*}(\mathscr{G}))

allows to construct an assembly map

μ:K,τ(B𝒢)K0(Cr(𝒢))\mu:K_{*,\tau}(B\mathscr{G})\rightarrow K_{0}(C_{r}^{*}(\mathscr{G})) (29)

by putting μ(δD)=inda(σD)\mu(\delta_{D})=ind_{a}(\sigma_{D}). The fact that it is well defined can be deduced (at least in the case of foliations) from the Connes-Skandalis longitudinal index theorem, [CS84]. This morphism was first defined by Baum and Connes [BC00] for groups (see [Tu00] for the case of groupoids).

In our setting, the reinforced longitudinal index theorem allows us to define the corresponding assembly map,

μF:K,τ(B𝒢)K0B(𝒢),\mu_{F}:K_{*,\tau}(B\mathscr{G})\rightarrow K_{0}^{B}(\mathscr{G}), (30)

given as in (29) but with periodic analytic index: μF(δD)=indaB(σD)\mu_{F}(\delta_{D})=ind_{a}^{B}(\sigma_{D}). By definition we have a commutative diagram

K,τ(B𝒢)\textstyle{K_{*,\tau}(B\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ\scriptstyle{\mu}μF\scriptstyle{\mu_{F}}K0(Cr(𝒢))\textstyle{K_{0}(C_{r}^{*}(\mathscr{G}))}K0B(𝒢).\textstyle{K_{0}^{B}(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces.}i\scriptstyle{i} (31)
Remark 6.5.

We recall that the Baum-Connes conjecture establish that μ\mu is an isomorphism. Its importance stands in the several consequences it would have, [BCH94, KL05].

6.2 Pairings with Cyclic cohomology

As we mentioned in the introduction, the main motivation for constructing the compactly supported indices is that, at this level, we can extract numerical information from this KK-theory elements. There is indeed a pairing between KK-theory and Cyclic cohomology (see (10)), and an important problem in non commutative geometry is to give explicit (topological) formulae for this pairings.

Now, for fixed τHP(Cc(𝒢))\tau\in HP^{*}(C_{c}^{\infty}(\mathscr{G})), we can expect an easy (topological) calculation only if the map DD,τD\mapsto\langle D\,,\tau\rangle factors through the symbol class of DD, [σ(D)]K0(A𝒢)[\sigma(D)]\in K^{0}(A^{*}\mathscr{G}). We are going to solve the factorization problem for bounded cocycles. This restriction is not at all restrictive. The bounded cocycles are periodic cyclic cocycles whose formulas only use a finite number of derivates. This is of course the case of group (and group action) cocycles, the transverse fundamental class, Godbillon-Vey and all the cocycles coming from H(B𝒢)H^{*}(B\mathscr{G}) (as we see in the last section). In fact, as we will see below, we are going to completely solve the problem for étale groupoids. Before stating the factorization theorem, we give the precise definition of bounded cocycles.

Definition 6.6.

A multilinear map τ:Cc(𝒢)××Cc(𝒢)q+1times\tau:\underbrace{C_{c}^{\infty}(\mathscr{G})\times\cdots\times C_{c}^{\infty}(\mathscr{G})}_{q+1-times}\rightarrow\mathbb{C} is bounded if it extends to a continuous multilinear map Cck(𝒢)××Cck(𝒢)q+1timesτk\underbrace{C_{c}^{k}(\mathscr{G})\times\cdots\times C_{c}^{k}(\mathscr{G})}_{q+1-times}\stackrel{{\scriptstyle\tau_{k}}}{{\longrightarrow}}\mathbb{C}, for some kk\in\mathbb{N}. We can consider a sub-bicomplex of the Periodic bicomplex (Cn,m(Cc(𝒢)),b,B)(C^{n,m}(C_{c}^{\infty}(\mathscr{G})),b,B) consisting in bounded multilinear maps. We note this bicomplex by (CBn,m(Cc(𝒢)),b,B)(C_{B}^{n,m}(C_{c}^{\infty}(\mathscr{G})),b,B) and a cocycle for it will be called a bounded continuous cyclic cocycle.

In the following proposition we prove that for the pairing with a bounded cocycle extends to our groups K0B(𝒢)K_{0}^{B}(\mathscr{G}).

Proposition 6.7

Let τ\tau be a bounded cyclic cocycle. Then the pairing map K0(Cc(𝒢)),τK_{0}(C_{c}^{\infty}(\mathscr{G}))\stackrel{{\scriptstyle\langle\,,\tau\rangle}}{{\longrightarrow}} extends to K0B(𝒢)K_{0}^{B}(\mathscr{G}), i.e., we have a commutative diagram of the following type:

K0(Cc(𝒢))\textstyle{K_{0}(C_{c}^{\infty}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}<,τ>\scriptstyle{<,\tau>}ι\scriptstyle{\iota}𝐂\textstyle{\mathbf{C}}K0B(𝒢)\textstyle{K_{0}^{B}(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τB\scriptstyle{\tau_{B}} (32)
Proof.

We are going to see first that the pairing extends to K0F(𝒢)K_{0}^{F}(\mathscr{G}).

Let τk:Cck(𝒢)××Cck(𝒢)2n+1times\tau_{k}:\underbrace{C_{c}^{k}(\mathscr{G})\times\cdots\times C_{c}^{k}(\mathscr{G})}_{2n+1-times}\rightarrow\mathbb{C} be a continuous extension of τ\tau, for some kk\in\mathbb{N}. It is immediate by definition that K0(Cc(𝒢)),τK_{0}(C_{c}^{\infty}(\mathscr{G}))\stackrel{{\scriptstyle\langle\,,\tau\rangle}}{{\longrightarrow}}\mathbb{C} extends to K0(Cck(𝒢))K_{0}(C_{c}^{k}(\mathscr{G})), more explicitly, the following diagram is commutative

K0(Cc(𝒢))\textstyle{K_{0}(C_{c}^{\infty}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}<,τ>\scriptstyle{<,\tau>}(ιk)\scriptstyle{(\iota_{k})_{*}}𝐂\textstyle{\mathbf{C}}K0(Cck(𝒢))\textstyle{K_{0}(C_{c}^{k}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}<,τk>\scriptstyle{<,\tau_{k}>}

We are now going to see that the pairing passes to K0h,k(𝒢)K_{0}^{h,k}(\mathscr{G}). For that we just have to check that this pairing preserves the relation h\sim_{h} over K0(Cck(𝒢))K_{0}(C_{c}^{k}(\mathscr{G})). This can be done by adapting an argument already used by Connes [Con85] and Goodwillie [Goo85]: let eC([0,1],Cck(𝒢))e\in C^{\infty}([0,1],C_{c}^{k}(\mathscr{G}))\oplus\mathbb{C} be an idempotent. It defines a smooth family of idempotents ete_{t} in Cck(𝒢)~\widetilde{C_{c}^{k}(\mathscr{G})}. We set at:=detdt(2et1)a_{t}:=\frac{de_{t}}{dt}(2e_{t}-1). Hence, a simple calculation shows

ddtτ,et=i=02nτ(et,,[at,et],,et)=:Latτ(et,,et).\frac{d}{dt}\langle\tau,e_{t}\rangle=\sum_{i=0}^{2n}\tau(e_{t},...,[a_{t},e_{t}],...,e_{t})=:L_{a_{t}}\tau(e_{t},...,e_{t}).

Now, the Lie derivates LxtL_{x_{t}} act trivially on HP0(Cck(𝒢))HP^{0}(C_{c}^{k}(\mathscr{G})) (see [Con85, Goo85]), then τ,et\langle\tau,e_{t}\rangle is constant in tt. In particular, τ,e0=τ,e1\langle\tau,e_{0}\rangle=\langle\tau,e_{1}\rangle. It follows immediately that ,τ\langle\,,\tau\rangle extends to K0F(𝒢):=limkK0h,k(𝒢)K_{0}^{F}(\mathscr{G}):=\varinjlim_{k}K_{0}^{h,k}(\mathscr{G}).

Finally, the Periodic Cyclic cohomology HP(Cc(𝒢))HP^{*}(C_{c}^{\infty}(\mathscr{G})) satisfies Bott periodicity. Hence, the extension from K0F(𝒢)K_{0}^{F}(\mathscr{G}) to K0B(𝒢):=limmK0F(𝒢×2m)K_{0}^{B}(\mathscr{G}):=\varprojlim_{m}K_{0}^{F}(\mathscr{G}\times\mathbb{R}^{2m}) is now immediate. ∎

Remark 6.8.
  1. 1.

    The extension of diagram (32) is very explicit: let τCFn,m(𝒢)\tau\in C^{n,m}_{F}(\mathscr{G}) and x=(x1,x2,)K0F(𝒢)=limkK0h,k(𝒢)x=(x_{1},x_{2},...)\in K_{0}^{F}(\mathscr{G})=\varprojlim_{k}K_{0}^{h,k}(\mathscr{G}). Let τk\tau_{k} be an extension of τ\tau. Then

    τF(x)=τk,xk.\tau_{F}(x)=\langle\tau_{k},x_{k}\rangle.
  2. 2.

    If we note by HPB(Cc(𝒢))HP_{B}^{*}(C^{\infty}_{c}(\mathscr{G})) the cohomology of the bicomplex (CBn,m(Cc(𝒢)),b,B)(C_{B}^{n,m}(C_{c}^{\infty}(\mathscr{G})),b,B) described in the definition 6.6, then we have a pairing

    HPB(Cc(𝒢))×K0B(𝒢),HP_{B}^{*}(C^{\infty}_{c}(\mathscr{G}))\times K_{0}^{B}(\mathscr{G})\longrightarrow\mathbb{C}, (33)

    induced from the extension of the proposition above. More explicitly, if τ\tau is a bounded cocycle and xK0F(𝒢)x\in K^{F}_{0}(\mathscr{G}), then

    x,τ=Bott(x),Bott(τ).\langle x,\tau\rangle=\langle Bott(x),Bott(\tau)\rangle.
Theorem 6.9 (Factorization theorem)
  • (a)

    Let τ\tau be a bounded continuous cyclic cocycle. Then τ\tau defines a morphism φτ:K0B(𝒢)\varphi_{\tau}:K_{0}^{B}(\mathscr{G})\rightarrow\mathbb{C} such that the following diagram is commutative

    Ell(𝒢)\textstyle{Ell(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ind\scriptstyle{ind}symb.\scriptstyle{symb.}K0(Cc(𝒢))\textstyle{K_{0}(C_{c}^{\infty}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}_,τ\scriptstyle{\langle\_,\tau\rangle}\textstyle{\mathbb{C}}K0(A𝒢)\textstyle{K^{0}(A^{*}\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}indaB\scriptstyle{ind_{a}^{B}}K0B(𝒢)\textstyle{K_{0}^{B}(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φτ\scriptstyle{\varphi_{\tau}}.

    In particular, for a 𝒢\mathscr{G}-pseudodifferential elliptic operator DD, we have the following formula

    φτindaB([σ(D)])=<indD,τ>.\varphi_{\tau}\circ ind_{a}^{B}([\sigma(D)])=<ind\,D,\tau>. (34)
  • (b)

    Let 𝒢𝒢(0)\mathscr{G}\rightarrow\mathscr{G}^{(0)} be an étale groupoid. Then we have the result of precedent paragraph for every τHP(Cc(𝒢))\tau\in HP^{*}(C_{c}^{\infty}(\mathscr{G})). In particular, the map

    Ell(𝒢)indK0(Cc(𝒢))_,τEll(\mathscr{G})\stackrel{{\scriptstyle ind}}{{\longrightarrow}}K_{0}(C_{c}^{\infty}(\mathscr{G}))\stackrel{{\scriptstyle\langle\_,\tau\rangle}}{{\longrightarrow}}\mathbb{C}

    always factors through Ell(𝒢)K0(A𝒢)Ell(\mathscr{G})\rightarrow K^{0}(A^{*}\mathscr{G}).

Proof.
  • (a)

    It is immediate from proposition 6.7.

  • (b)

    Thanks to the works of Burghelea, Brylinski-Nistor and Crainic ([Bur85, BN94, Cra99]), we known a very explicit description of the Periodic cyclic cohomology for étale groupoids. For instance, we have a decomposition of the following kind (see for example [Cra99] theorems 4.1.2. and 4.2.5)

    HP(Cc(𝒢))=Π𝒪Hτ+r(B𝒩𝒪),HP^{*}(C_{c}^{\infty}(\mathscr{G}))=\Pi_{\mathscr{O}}H_{\tau}^{*+r}(B\mathscr{N}_{\mathscr{O}}), (35)

    where 𝒩𝒪\mathscr{N}_{\mathscr{O}} is an étale groupoid associated to 𝒪\mathscr{O} (the normalizer of 𝒪\mathscr{O}, see 3.4.8 in ref.cit.). For instance, when 𝒪=𝒢(0)\mathscr{O}=\mathscr{G}^{(0)}, 𝒩𝒪=𝒢\mathscr{N}_{\mathscr{O}}=\mathscr{G}.

    Now, all the cyclic cocycles coming from the cohomology of the classifying space are bounded. Indeed, we know that each factor of HP(Cc(𝒢))HP^{*}(C_{c}^{\infty}(\mathscr{G})) in the decomposition (35) consists of bounded cyclic cocycles (see the last section of this work). Now, the pairing

    HP(Cc(𝒢))×K0(Cc(𝒢))HP^{*}(C_{c}^{\infty}(\mathscr{G}))\times K_{0}(C_{c}^{\infty}(\mathscr{G}))\longrightarrow\mathbb{C}

    is well defined. In particular, the restriction to HP(Cc(𝒢))|𝒪HP^{*}(C_{c}^{\infty}(\mathscr{G}))|_{\mathscr{O}} vanishes for almost every 𝒪\mathscr{O}. The conclusion is now immediate from proposition 6.7.

6.2.1 Geometric Corollaries

The factorization problem we just have met is deeply related with the Novikov conjecture. Indeed, if the map

K0(Cc(𝒢))_,τK_{0}(C_{c}^{\infty}(\mathscr{G}))\stackrel{{\scriptstyle\langle\_,\tau\rangle}}{{\longrightarrow}}\mathbb{C}

where τHP(Cc(𝒢))\tau\in HP^{*}(C_{c}^{\infty}(\mathscr{G})), extends to K0(Cr(𝒢))K_{0}(C^{*}_{r}(\mathscr{G})), then the factorization through the principal symbol class is immediate. However, as the following example shows, it is far from being a trivial problem.

Example 1

[CM90, Con94] Let Γ\Gamma be a discrete group acting properly and freely on a smooth manifold M~\tilde{M} with compact quotient M~/Γ:=M\tilde{M}/\Gamma:=M. Let 𝒢𝒢(0)=M\mathscr{G}\rightrightarrows\mathscr{G}^{(0)}=M be the Lie groupoid quotient of M~×M~\tilde{M}\times\tilde{M} by the diagonal action of Γ\Gamma.

Let cH(Γ):=H(BΓ)c\in H^{*}(\Gamma):=H^{*}(B\Gamma). Connes-Moscovici showed in [CM90] that the higher Novikov signature, Signc(M)Sign_{c}(M), can be obtained with the paring of the signature operator DsignD_{sign} and a cyclic cocycle τc\tau_{c} associated to cc:

τc,indDsgn=Signc(M,ψ).\langle\tau_{c},ind\,D_{sgn}\rangle=Sign_{c}(M,\psi). (36)

The Novikov conjecture states that these higher signatures are oriented homotopy invariants of MM. Hence, if indDsignK0(Cc(𝒢))ind\,D_{sign}\in K_{0}(C_{c}^{\infty}(\mathscr{G})) is a homotopy invariant of (M,ψ)(M,\psi) then the Novikov conjecture would follow. We only known that j(indDsign)K0(Cr(𝒢))j(ind\,D_{sign})\in K_{0}(C^{*}_{r}(\mathscr{G})) is a homotopy invariant. But then we have to extend the action of τc\tau_{c} to K0(Cr(𝒢))K_{0}(C^{*}_{r}(\mathscr{G})). Connes-Moscovici show that this action extends for Hyperbolic groups.

Question Is indaB(D)ind_{a}^{B}(D) a homotopy invariant? An affirmative answer to this question would imply the Novikov conjecture because the pairing (36) extends to K0B(𝒢)K_{0}^{B}(\mathscr{G}). In fact, since we know that inda(D)K0(Cr(𝒢))ind_{a}(D)\in K_{0}(C^{*}_{r}(\mathscr{G})) is a homotopy invariant, another way to establish the Novikov conjecture would be to prove the injectivity of the map K0B(𝒢)iK0(Cr(𝒢))K_{0}^{B}(\mathscr{G})\stackrel{{\scriptstyle i}}{{\longrightarrow}}K_{0}(C^{*}_{r}(\mathscr{G})).

In [Con86], theorem 8.1, Connes solves the extension problem for some kind of cyclic cocycles over the holonomy groupoid of a foliation, and he gives a topological formula for the pairing. A main step in his proof is the Connes-Skandalis longitudinal index theorem.

Using our longitudinal index theorem we obtain, as a corollary of Connes theorem, the analog result in our setting. In this case we do not have to deal with the extension problem, and so the result applies to all classes cH(B𝒢)c\in H^{*}(B\mathscr{G}).

Corollary 6.10

Let (V,F)(V,F) be a foliated manifold (non necessarily compact) transversally oriented. Let 𝒢\mathscr{G} be its holonomy groupoid. For any cH(B𝒢)c\in H^{*}(B\mathscr{G}) there is an additive map

φc:K0B(𝒢)\varphi_{c}:K_{0}^{B}(\mathscr{G})\rightarrow\mathbb{C}

such that

φc(μF(x))=chτ(x),cxK0B(𝒢).\varphi_{c}(\mu_{F}(x))=\langle ch_{\tau}(x),c\rangle\,\,\forall x\in K_{0}^{B}(\mathscr{G}). (37)

Haefliger Cohomology

The reinforced longitudinal theorem can also be used for re-establish the index formulas in Haefliger cohomology found by Benameur-Heitsch in [BH04].

Benameur-Heitsch start by defining an algebraic Chern character

cha:K0(Cc(𝒢))Hc,bas(M/F),ch_{a}:K_{0}(C_{c}^{\infty}(\mathscr{G}))\rightarrow H_{c,bas}^{*}(M/F), (38)

where Hc,bas(M/F)H_{c,bas}^{*}(M/F) is the Haefliger cohomology (see [Hae84] or [CM04]). This character is compatible with the “shriek maps” in KK-theory and in Haefliger cohomology.

The main result in [BH04] requires the construction of a map cha(indt):K0(TF)Hc,bas(M/F)ch_{a}(ind_{t}):K^{0}(T^{*}F)\rightarrow H_{c,bas}^{*}(M/F) that is morally the topological index followed by the Chern character. Then, they prove the following formula (theorem 5.11 ref.cit.):

For any uK0(TF)u\in K^{0}(T^{*}F),

cha(indt)(u)=(1)pFπF!(ch(u))Td(F)Hc,bas(M/F),ch_{a}(ind_{t})(u)=(-1)^{p}\int_{F}\pi_{F}!(ch(u))Td(F\otimes\mathbb{C})\in H_{c,bas}^{*}(M/F), (39)

where πF!:Hc(F,)H(M,)\pi_{F}!:H^{*}_{c}(F,\mathbb{R})\rightarrow H^{*}(M,\mathbb{R}) is the fiberwise integration, and ch:K0(F)Hc(F)ch:K^{0}(F)\rightarrow H^{*}_{c}(F) is the usual Chern character.

Using the same kind of arguments as in proposition 6.7, one can show that the character chach_{a} extends to the group K0B(𝒢)K_{0}^{B}(\mathscr{G}):

K0(Cc(𝒢))\textstyle{K_{0}(C_{c}^{\infty}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}cha\scriptstyle{ch_{a}}K0B(𝒢)\textstyle{K_{0}^{B}(\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}cha\scriptstyle{ch_{a}}Hc,bas(M/F)\textstyle{H_{c,bas}^{*}(M/F)}.

The next result is an immediate consequence of the reinforced longitudinal theorem and the Benameur-Heitsch formula.

Corollary 6.11

For any longitudinal elliptic pseudodifferential operator DD, we have that

cha(indD)=cha(indt)[σD]Hc,bas(M/F).ch_{a}(ind\,D)=ch_{a}(ind_{t})[\sigma_{D}]\in H_{c,bas}^{*}(M/F). (40)

The fact that the actions of a large class of cyclic cocycles extend naturally to the periodic group K0B(𝒢)K_{0}^{B}(\mathscr{G}) allows to think that formulas as those of Benameur-Heitsch could be developed in cohomology spaces more complex than Hc,bas(M/F)H_{c,bas}^{*}(M/F).

6.2.2 Bounded Cyclic cohomology

Let HPB(Cc(𝒢))HP^{*}_{B}(C_{c}^{\infty}(\mathscr{G})) be the cohomology of the bicomplex (CBn,m(Cc(𝒢)),b,B)(C_{B}^{n,m}(C_{c}^{\infty}(\mathscr{G})),b,B) (see definition 6.6 for notations), we call it the ”Bounded (Periodic) Cyclic Cohomology of 𝒢\mathscr{G}.

The case of Étale groupoids

Let us consider the classifying space, B𝒢B\mathscr{G}, and its twisted cohomology, Hτ(B𝒢)H_{\tau}^{*}(B\mathscr{G}).

In [Con86], Connes constructed a group morphism (defined at the level of cocycles)

ϕ:Hτ(B𝒢)HP(Cc(𝒢)),\phi:H_{\tau}^{*}(B\mathscr{G})\rightarrow HP^{*}(C_{c}^{\infty}(\mathscr{G})), (41)

that is in fact an inclusion as a direct factor (see also [Con94], III.2.δIII.2.\delta).

By examinining Connes’ construction, it follows easily that the cocycles one gets are all bounded. In other words, one constructs a morphism ([CR07])

ϕB:Hτ(B𝒢)HPB(Cc(𝒢)),\phi_{B}:H_{\tau}^{*}(B\mathscr{G})\rightarrow HP^{*}_{B}(C_{c}^{\infty}(\mathscr{G})), (42)

fitting the diagram

Hτ(B𝒢)\textstyle{H_{\tau}^{*}(B\mathscr{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕ\scriptstyle{\phi}ϕB\scriptstyle{\phi_{B}}HP(Cc(𝒢))\textstyle{HP^{*}(C_{c}^{\infty}(\mathscr{G}))}HPB(Cc(𝒢)).\textstyle{HP^{*}_{B}(C_{c}^{\infty}(\mathscr{G}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces.}ι\scriptstyle{\iota} (43)

This means that the image of the morphism in (41) consists only on Bounded Cyclic cocycles.

References

  • [ANS06] Johannes Aastrup, Ryszard Nest, and Elmar Schrohe, A continuous field of CC^{*}-algebras and the tangent groupoid for manifolds with boundary, J. Funct. Anal. 237 (2006), no. 2, 482–506.
  • [AS68] M. F. Atiyah and I. M. Singer, The index of elliptic operators. I, Ann. of Math. (2) 87 (1968), 484–530.
  • [BC00] Paul Baum and Alain Connes, Geometric KK-theory for Lie groups and foliations, Enseign. Math. (2) 46 (2000), no. 1-2, 3–42.
  • [BCH94] Paul Baum, Alain Connes, and Nigel Higson, Classifying space for proper actions and KK-theory of group CC^{\ast}-algebras, CC^{\ast}-algebras: 1943–1993 (San Antonio, TX, 1993), Contemp. Math., vol. 167, Amer. Math. Soc., Providence, RI, 1994, pp. 240–291.
  • [BH04] Moulay-Tahar Benameur and James L. Heitsch, Index theory and non-commutative geometry. I. Higher families index theory, KK-Theory 33 (2004), no. 2, 151–183.
  • [BN94] J.-L. Brylinski and Victor Nistor, Cyclic cohomology of étale groupoids, KK-Theory 8 (1994), no. 4, 341–365.
  • [Bur85] Dan Burghelea, The cyclic homology of the group rings, Comment. Math. Helv. 60 (1985), no. 3, 354–365.
  • [CH90] Alain Connes and Nigel Higson, Déformations, morphismes asymptotiques et KK-théorie bivariante, C. R. Acad. Sci. Paris Sér. I Math. 311 (1990), no. 2, 101–106.
  • [CM90] Alain Connes and Henri Moscovici, Cyclic cohomology, the Novikov conjecture and hyperbolic groups, Topology 29 (1990), no. 3, 345–388.
  • [CM04] Marius Crainic and Ieke Moerdijk, Čech-De Rham theory for leaf spaces of foliations, Math. Ann. 328 (2004), no. 1-2, 59–85.
  • [CMR07] Joachim Cuntz, Ralf Meyer, and Jonathan Rosenberg, Topological and bivariant k-theory, Oberwolfach Seminars, vol. 36, Birkh user, Berlin, 2007.
  • [Con79] Alain Connes, Sur la théorie non commutative de l’intégration, Algèbres d’opérateurs (Sém., Les Plans-sur-Bex, 1978), Lecture Notes in Math., vol. 725, Springer, Berlin, 1979, pp. 19–143.
  • [Con85]   , Noncommutative differential geometry, Inst. Hautes Études Sci. Publ. Math. (1985), no. 62, 257–360.
  • [Con86]   , Cyclic cohomology and the transverse fundamental class of a foliation, Geometric methods in operator algebras (Kyoto, 1983), Pitman Res. Notes Math. Ser., vol. 123, Longman Sci. Tech., Harlow, 1986, pp. 52–144.
  • [Con94]   , Noncommutative geometry, Academic Press Inc., San Diego, CA, 1994.
  • [CR06] Paulo Carrillo-Rouse, A Schwartz type algebra for the tangent groupoid, to appear in the ”Proceedings of the International Conference on K-theory and Noncommutative Geometry” held in Valladolid, Spain. Preprint: arxiv:math.DG/08023596. (2006).
  • [CR07]   , Indices analytiques à support compact pour des groupoides de Lie, Thèse de Doctorat à l’Université de Paris 7 (2007).
  • [Cra99] Marius Crainic, Cyclic cohomology of étale groupoids: the general case, KK-Theory 17 (1999), no. 4, 319–362.
  • [CS84] Alain Connes and Georges Skandalis, The longitudinal index theorem for foliations, Publ. Res. Inst. Math. Sci. 20 (1984), no. 6, 1139–1183.
  • [DLN06] Claire Debord, Jean-Marie Lescure, and Victor Nistor, Groupoids and an index theorem for conical pseudomanifolds, arxiv:math.OA/0609438 (2006).
  • [GL06] Alexander Gorokhovsky and John Lott, Local index theory over foliation groupoids, Adv. Math. 204 (2006), no. 2, 413–447.
  • [Goo85] Thomas G. Goodwillie, Cyclic homology, derivations, and the free loopspace, Topology 24 (1985), no. 2, 187–215.
  • [Hae84] André Haefliger, Groupoïdes d’holonomie et classifiants, Astérisque (1984), no. 116, 70–97, Transversal structure of foliations (Toulouse, 1982).
  • [HS87] Michel Hilsum and Georges Skandalis, Morphismes KK-orientés d’espaces de feuilles et fonctorialité en théorie de Kasparov (d’après une conjecture d’A. Connes), Ann. Sci. École Norm. Sup. (4) 20 (1987), no. 3, 325–390.
  • [Kar78] Max Karoubi, KK-theory, Springer-Verlag, Berlin, 1978, An introduction, Grundlehren der Mathematischen Wissenschaften, Band 226.
  • [Kar87]   , Homologie cyclique et KK-théorie, Astérisque (1987), no. 149, 147.
  • [KL05] Matthias Kreck and Wolfgang Luck, The Novikov Conjecture, Oberwolfach Seminars, vol. 33, Birkh user, Berlin, 2005.
  • [Lan03] N. P. Landsman, Quantization and the tangent groupoid, Operator algebras and mathematical physics (Constanţa, 2001), Theta, Bucharest, 2003, pp. 251–265.
  • [LM89] H. Blaine Lawson, Jr. and Marie-Louise Michelsohn, Spin geometry, Princeton Mathematical Series, vol. 38, Princeton University Press, Princeton, NJ, 1989.
  • [Mac87] K. Mackenzie, Lie groupoids and Lie algebroids in differential geometry, London Mathematical Society Lecture Note Series, vol. 124, Cambridge University Press, Cambridge, 1987.
  • [Mel93] Richard B. Melrose, The Atiyah-Patodi-Singer index theorem, Research Notes in Mathematics, vol. 4, A K Peters Ltd., Wellesley, MA, 1993.
  • [MN05] Bertrand Monthubert and Victor Nistor, A topological index theorem for manifolds with corners, arxiv:math.KT/0507601 (2005).
  • [MP97] Bertrand Monthubert and François Pierrot, Indice analytique et groupoïdes de Lie, C. R. Acad. Sci. Paris Sér. I Math. 325 (1997), no. 2, 193–198.
  • [NWX99] Victor Nistor, Alan Weinstein, and Ping Xu, Pseudodifferential operators on differential groupoids, Pacific J. Math. 189 (1999), no. 1, 117–152.
  • [Pat99] Alan L. T. Paterson, Groupoids, inverse semigroups, and their operator algebras, Progress in Mathematics, vol. 170, Birkhäuser Boston Inc., Boston, MA, 1999.
  • [Ren80] Jean Renault, A groupoid approach to CC^{\ast}-algebras, Lecture Notes in Mathematics, vol. 793, Springer, Berlin, 1980.
  • [Ros94] Jonathan Rosenberg, Algebraic KK-theory and its applications, Graduate Texts in Mathematics, vol. 147, Springer-Verlag, New York, 1994.
  • [Sha78] Patrick Shanahan, The Atiyah-Singer index theorem, Lecture Notes in Mathematics, vol. 638, Springer, Berlin, 1978, An introduction.
  • [Trè80] François Trèves, Introduction to pseudodifferential and Fourier integral operators. Vol. 1, Plenum Press, New York, 1980, Pseudodifferential operators, The University Series in Mathematics.
  • [Trè06]   , Topological vector spaces, distributions and kernels, Dover Publications Inc., Mineola, NY, 2006, Unabridged republication of the 1967 original.
  • [Tu00] Jean-Louis Tu, The Baum-Connes conjecture for groupoids, CC^{*}-algebras (Münster, 1999), Springer, Berlin, 2000, pp. 227–242.