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

On evolution equations for Lie groupoids

Jean-Marie Lescure Université Paris Est Créteil, CNRS, LAMA, F-94000 CRETEIL, FRANCE [email protected]  and  Stéphane Vassout Université de Paris, CNRS, IMJ-PRG, F-75205 PARIS CEDEX 13, FRANCE [email protected]
(Date: April 5, 2025)
Abstract.

Using the calculus of Fourier integral operators on Lie groupoids developped in [18], we study the fundamental solution of the evolution equation (t+iP)u=0(\frac{\partial}{\partial t}+iP)u=0 where PP is a self adjoint elliptic order one GG-pseudodifferential operator on the Lie groupoid GG. Along the way, we continue the study of distributions on Lie groupoids done in [17] by adding the reduced CC^{*}-algebra of GG in the picture and we investigate the local nature of the regularizing operators of [32].

1. Introduction

The main motivation of this paper is the construction of an approximate solution to the problem

(1) {(t+iP)u=fu(0)=g\begin{cases}(\dfrac{\partial}{\partial t}+iP)u=f\\ u(0)=g\end{cases}

in the framework of a Lie groupoid GMG\rightrightarrows M. This means that PP here is a suitable order 11 pseudodifferential GG-operator, that f,gf,g live in suitable spaces of distributions and that the approximate solution will be seeked among Fourier integral GG-operators. The present article can be considered as a continuation of [17], where properties of distributions on Lie groupoids, and convolution of them, are studied in a certain generality, and of [18], where Hörmander’s notion and calculus of Fourier integral operators on manifolds [11, 12] are exported to the framework of Lie groupoids. We will frequently refer to the results of these papers, and one of their cornerstones, namely the symplectic groupoid structure of TGT^{*}G [6]: sΓ,rΓ:Γ=TGAGs_{{}_{\Gamma}},r_{{}_{\Gamma}}:\Gamma=T^{*}G\rightrightarrows A^{*}G, will be of great importance here again.

The Cauchy problem (1) has been and can be of course investigated in many situations under many different assumptions. We refer more precisely to [12, Theorem 29.1.1] to illustrate the kind of results that we want to achieve on Lie groupoids. This can be summarized by the following problem:


(𝒫)(\mathcal{P}) Under an ellipticity assumption on PP, the fundamental solution of (1) should have, up to suitable regularizing error terms, an explicit approximation by Fourier integral GG-operators that describes in a simple and geometric way how the singularities of the initial data gg propagate at time tt under the action of the principal symbol of PP.


To set the problem on firm foundations, we first study in Section 3 existence and unicity conditions for (1) in the general framework of CC^{*}-algebras and Hilbertian modules, and we require there that PP is an unbounded self-adjoint regular operator on a CC^{*}-algebra AA [2, 3, 33, 13, 32, 30]. Then the fundamental solution of (1) denoted by E(t)=eitPE(t)=e^{-itP} is obtained by continuous functional calculus, which yields the existence of solutions, while easy computations identical to those for Hilbert spaces show the uniqueness. We get in particular:

Theorem 1.

Let AA be a CC^{*}-algebra, let HH be a Hilbertian AA-module and PP be a selfadjoint regular operator on AA. Let H=kdomPkH^{\infty}=\cap_{k}\mathop{\mathrm{dom}}\nolimits P^{k}. Then for any fC(,H)f\in C^{\infty}({\mathbb{R}},H^{\infty}) and gHg\in H^{\infty}, the Cauchy problem (1) has a unique solution in C(,H)C^{\infty}({\mathbb{R}},H^{\infty}), given by

(2) u(t)=eitPg+0tei(st)Pf(s)𝑑s.u(t)=e^{-itP}g+\int_{0}^{t}e^{i(s-t)P}f(s)ds.

This preliminary result allows us to speak about the fundamental solution of (1) in the case of a Lie groupoid GG with compact units space MM and of a first order elliptic symmetric and compactly supported, polyhomogeneous pseudodifferential GG-operator PP. Indeed, we then know by [32] that (the closure of) PP is selfadjoint and regular on, for instance, the reduced CC^{*}-algebra of GG, denoted by Cr(G)C^{*}_{r}(G). In particular, the theorem above applies and the task to find a nice approximation to E(t)E(t) among Fourier integral GG-operators is meaningful. Note that, because of (2), the error term will automatically belong to the space =H(H)\mathcal{H}^{\infty}=H^{\infty}\cap(H^{\infty})^{*}. Our answer to the problem (𝒫)(\mathcal{P}) is the main result of the paper:

Theorem 2.

There exists a CC^{\infty} family ΛtTG\Lambda_{t}\subset T^{*}G of GG-relations and a CC^{\infty} family of compactly supported Fourier integral GG-operators U(t)I(n(1)n(0))/4(G,Λt;Ω1/2)U(t)\in I^{(n^{(1)}-n^{(0)})/4}(G,\Lambda_{t};\Omega^{1/2}) such that :

(3) (t+iP)U(t)Cc(G,Ω1/2),(\frac{\partial}{\partial t}+iP)U(t)\in C^{\infty}_{c}(G,\Omega^{1/2}),

and for any tt, we have: E(t)U(t)\displaystyle E(t)-U(t)\in{\mathcal{H}^{\infty}}.

Let us now explain in some details the ingredients and the intermediate results, some of them being interesting on their owns, required in the proof of the main theorem.

First of all, Theorem 2 immediately rises the preliminary question of the regularity of elements in {\mathcal{H}^{\infty}}. Strictly speaking, elements of {\mathcal{H}^{\infty}} live in a noncommutative CC^{*}-algebra so dealing with their local properties makes a priori no sense. We manage on the one hand to prove that elements of the reduced CC^{*}-algebra Cr(G)C^{*}_{r}(G) of a Lie groupoid GG are distributions on GG, in a canonical way, and on the other hand, to precise the regularity of elements in {\mathcal{H}^{\infty}}. These intermediate tasks are the subject of Sections 4 and 5 and the details can be summarized as follows.

The space of distributions we deal with, denoted by 𝒟(G)\mathcal{D}^{\prime}(G), is the one of distributions on GG with values in the density bundle Ω1/2:=Ω1/2(rAG)Ω1/2(sAG)\Omega^{1/2}:=\Omega^{1/2}(r^{*}AG)\otimes\Omega^{1/2}(s^{*}AG) and thus the space of test functions we use, denoted by 𝒟(G)\mathcal{D}(G), is the one of compactly supported CC^{\infty} sections of the density bundle Ω01/2:=Ω1/2ΩG1\Omega^{1/2}_{0}:=\Omega^{-1/2}\otimes\Omega^{1}_{G}. Thus 𝒟(G)=(𝒟(G))\mathcal{D}^{\prime}(G)=(\mathcal{D}(G))^{\prime} and the choice of Ω1/2\Omega^{1/2} is relevant because Cc(G):=Cc(G,Ω1/2)𝒟(G)C^{\infty}_{c}(G):=C^{\infty}_{c}(G,\Omega^{1/2})\subset\mathcal{D}^{\prime}(G) is in a canonical way an involutive algebra, whose a certain completion is precisely the algebra Cr(G)C^{*}_{r}(G). Also, we have proved in [17] that the product \star in Cc(G)C^{\infty}_{c}(G) (called the convolution product for obvious reasons) widely generalizes, by continuity, to distributions in 𝒟(G)\mathcal{D}^{\prime}(G). For instance the space r,s(G)\mathcal{E}^{\prime}_{r,s}(G) of compactly supported distributions on GG whose pushforwards by the source and range maps are CC^{\infty} on MM (transversal distributions) forms a unital involutive algebra for the convolution product. Another justification for these choices of densities comes from the present work, indeed we prove that transversal distributions also act by convolution on 𝒟(G)\mathcal{D}(G) in a nice way, and that weak factorizations in the sense of [10] are available:

Theorem 3.

Let ,\langle\cdot,\cdot\rangle denote the pairing 𝒟(G)×𝒟(G)\mathcal{D}^{\prime}(G)\times\mathcal{D}(G)\to{\mathbb{C}} and ι\iota the inversion of the groupoid.

  1. (1)

    (u,ω)𝒟r,s(G)×𝒟(G),u,ω=ιu,ιω=δM,ιuω\displaystyle\forall(u,\omega)\in\mathcal{D}^{\prime}_{r,s}(G)\times\mathcal{D}(G),\quad\langle u,\omega\rangle=\langle\iota^{*}u,\iota^{*}\omega\rangle=\langle\delta_{M},\iota^{*}u\star\omega\rangle.

  2. (2)

    The space 𝒟(G)\mathcal{D}(G) is a bimodule over r,s(G)\mathcal{E}^{\prime}_{r,s}(G) and:

    u,vr,s(G),ω𝒟(G),uv,ω=v,ιuω=u,ωιv\displaystyle\forall u,v\in\mathcal{E}^{\prime}_{r,s}(G),\forall\omega\in\mathcal{D}(G),\quad\langle u\star v,\omega\rangle=\langle v,\iota^{*}u\star\omega\rangle=\langle u,\omega\star\iota^{*}v\rangle
  3. (3)

    Let ω𝒟(G)\omega\in\mathcal{D}(G). For any neighborhood VV of MM into GG , on can write ω\omega as a finite sum of elements ξχ\xi\star\chi where ξCc(G)\xi\in C^{\infty}_{c}(G), supp(ξ)V{\mathrm{supp}(\xi)}\subset V and χ𝒟(G)\chi\in\mathcal{D}(G), supp(χ)supp(ω){\mathrm{supp}(\chi)}\subset{\mathrm{supp}(\omega)}.

This material allows us in Section 5 to answer to the question about the local nature of elements in {\mathcal{H}^{\infty}}, and along the way, that of elements in Cr(G)C^{*}_{r}(G):

Theorem 4.
  1. (1)

    There is a continuous embedding Cr(G)𝒟(G)\displaystyle C^{*}_{r}(G)\hookrightarrow\mathcal{D}^{\prime}(G). This embedding extends the pairing ,\langle\cdot,\cdot\rangle between Cc(G)C^{\infty}_{c}(G) and 𝒟(G)\mathcal{D}(G).

  2. (2)

    The inclusions Corb,0(G)Cc(G)Corb,0(G)Cr(G)\displaystyle C^{\infty,0}_{\mathop{\mathrm{orb}}\nolimits}(G)\cap C_{c}(G)\subset{\mathcal{H}^{\infty}}\subset C^{\infty,0}_{\mathop{\mathrm{orb}}\nolimits}(G)\cap C^{*}_{r}(G) hold true.
    Here Corb,0C^{\infty,0}_{\mathop{\mathrm{orb}}\nolimits} refers to the space of continuous functions on GG that are CC^{\infty} on the subgroupoids G𝒪=s1(𝒪)G_{\mathcal{O}}=s^{-1}(\mathcal{O}), as well as all their derivatives along the fibers of ss and rr, for every orbits 𝒪=r(s1({x}))\mathcal{O}=r(s^{-1}(\{x\})) in MM.

Next, we explain how the principal symbol pp of PP gives rise to the family of Lagrangian submanifolds Λt\Lambda_{t}, tt\in{\mathbb{R}}, that will describe the propagation of singularities as expected in Problem (𝒫)(\mathcal{P}). By definition, PI1+(n(1)n(0))/4(G,M,Ω1/2)P\in I^{1+(n^{(1)}-n^{(0)})/4}(G,M,\Omega^{1/2}) is a polyhomogeneous conormal distribution, thus posseses a homogeneous principal symbol p00C(AG0)p^{0}_{0}\in C^{\infty}(A^{*}G\setminus 0). If one considers the family (Px)xM(P_{x})_{x\in M} of ordinary pseudodifferential operators in the fibers of ss and collects their principal symbols into a homogeneous function p0C(TsG0)p^{0}\in C^{\infty}(T_{s}^{*}G\setminus 0), TsG=(kerds)T^{*}_{s}G=(\ker ds)^{*}, that will be called the principal GG-symbol of PP. After lifting p0p^{0} to a function on TGkerrΓT^{*}G\setminus\ker r_{{}_{\Gamma}}, one gets the following identity:

(γ,ξ)TGkerrΓ,p(γ,ξ)=p0(rΓ(γ,ξ)).\forall(\gamma,\xi)\in T^{*}G\setminus\ker r_{{}_{\Gamma}},\quad p(\gamma,\xi)=p_{0}(r_{{}_{\Gamma}}(\gamma,\xi)).

Here rΓr_{{}_{\Gamma}} is the range map of the symplectic groupoid Γ=TG\Gamma=T^{*}G. The computations also give a local expression for the sub-principal GG-symbol of PP, that is, for the collection of the sub-principal symbols of the operators PxP_{x}. Now it turns out that the Hamiltonian flow χ\chi of the principal GG-symbol pp is complete and right invariant, and we get the required Lagrangian submanifolds that will describe the evolution of singularities:

t,Λt=χt(AG0).\forall t\in{\mathbb{R}},\qquad\Lambda_{t}=\chi_{t}(A^{*}G\setminus 0).

This already produces a CC^{\infty} family of homogeneous Lagrangian submanifolds of TG0T^{*}G\setminus 0 that satisfies the group relation, with respect to the product in TGT^{*}G:

Λt.Λs=Λt+s.\Lambda_{t}.\Lambda_{s}=\Lambda_{t+s}.

Moreover ΛtT.G:=TG(kerrΓkersΓ)\Lambda_{t}\subset\accentset{\mbox{\large.}}{T}^{*}G:=T^{*}G\setminus(\ker r_{{}_{\Gamma}}\cup\ker s_{{}_{\Gamma}}), that is, in the vocabulary of [18], every Λt\Lambda_{t} is a GG-relation, while the global object coming with the family (Λt)t(\Lambda_{t})_{t}:

Λ={(t,τ,γ,ξ);τ+p(γ,ξ)=0,(γ,ξ)Λt}T×TG\Lambda=\{(t,\tau,\gamma,\xi)\ ;\ \tau+p(\gamma,\xi)=0,\ (\gamma,\xi)\in\Lambda_{t}\}\subset T^{*}{\mathbb{R}}\times T^{*}G

is a family ×G{\mathbb{R}}\times G-relation. As in [18], this construction highlights the important role of the symplectic groupoid structure of TGT^{*}G in analysis.

There is a last result, of technical nature, that intervenes in the proof of Theorem 2. Indeed, assuming that the Lagrangian submanifolds Λt\Lambda_{t} provide the good candidate for Theorem 2, we are led to search a first order parametrix U0U_{0} for t+iP\partial_{t}+iP among the Fourier integral GG-operators associated with (Λt)t(\Lambda_{t})_{t}. This amounts to solve the transport equation for principal symbols:

tσpr(U0)+iσpr(PU0)=0\frac{\partial}{\partial t}\sigma_{\mathop{\mathrm{pr}}\nolimits}(U_{0})+i\sigma_{\mathop{\mathrm{pr}}\nolimits}(PU_{0})=0

and thus it requires to express the principal symbol of the convolution product PU0PU_{0} of the lagrangian distributions PP and U0U_{0}. Since by construction and on purpose, the principal symbol p0p_{0} vanishes on rΓΛr_{{}_{\Gamma}}\Lambda, we need to look for the next term in the asymptotic expansion of the total symbol of PU0PU_{0}. This is what is achieved, modulo some technical details, by using the following result:

Theorem 5.

Let QΨcm(G)Q\in\Psi_{c}^{m}(G), with principal GG-symbol qq, sub-principal symbol q1sq^{1s}, and let CC be a GG-relation such that qq vanishes on CC. Let AIm(G,C;Ω1/2)A\in I^{m^{\prime}}(G,C;\Omega^{1/2}) and aa be a principal symbol of AA.
Then

QAIm+m1(G,C;Ω1/2) and σpr(QA)=iqa+q1sa.QA\in I^{m+m^{\prime}-1}(G,C;\Omega^{1/2})\text{ and }\sigma_{\mathop{\mathrm{pr}}\nolimits}(QA)=-i\mathcal{L}_{q}a+q^{1s}a.

Here q\mathcal{L}_{q} is the Lie derivative along the Hamiltonian vector field HqH_{q} of qq.

Many interesting situations produce non compactly supported operators PP: for instance, if Δ\Delta is a Laplacian on GG then Δ=P+S\sqrt{\Delta}=P+S with PP as above and SS\in\mathcal{H}^{\infty} [32]. The main theorem trivially extends to such non compactly supported operators: one just needs to replace Cc(G)C^{\infty}_{c}(G) by {\mathcal{H}^{\infty}} in (3). We describe at the end of the paper several situations where Theorem 2 applies:

  1. (1)

    The usual pseudodifferential calculus on a compact manifold without boundary XX. We use the pair groupoid G=X×XXG=X\times X\rightrightarrows X. Since XX itself is an orbit, we have Corb,0(G,E)=C(X×X,E)C^{\infty,0}_{\mathop{\mathrm{orb}}\nolimits}(G,E)=C^{\infty}(X\times X,E) and we just recover the classical result on manifolds.

  2. (2)

    The longitudinal calculus on foliations [4]. We use the holonomy groupoid. We recover a construction in [15] by a quite different approach.

  3. (3)

    Right invariant calculus on a Lie group GG. We use GG as a groupoid with units space {e}\{e\}.

  4. (4)

    The bb-calculus on manifolds with corners [23]. We use the bb-groupoid [25].

  5. (5)

    The calculus on manifolds with fibred boundary or with iterated fibred corners [20, 8]. We use the groupoid of [8].

As far as we know, the results obtained for cases (3), (4), (5) above are new.

The next section contains the basic definitions and notation necessary for the sequel and can be considered as an extension of the introduction for the unfamiliar reader.


Acknowledgments The authors are grateful to Claire Debord, Omar Mohsen, Victor Nistor and Georges Skandalis for helpful and stimulating discussions or remarks. The first author is thankful for the hospitality of the IMJ-PRG, Paris University, where part of this project was achieved. Most part of this work has been realized with the support of the Grant ANR-14-CE25-0012-01 SINGSTAR.

2. Notation and reminders

Densities on manifolds. If EE is a real vector space of dimension nn and α\alpha\in{\mathbb{R}}, we denote by Ωα(E)\Omega^{\alpha}(E) the vector space of maps ω:ΛnE0\omega:\Lambda^{n}E\setminus 0\to{\mathbb{C}}, called α\alpha-densities, such that ω(tV)=|t|αω(V)\omega(tV)=|t|^{\alpha}\omega(V) for any t0t\not=0 and VΛnE0V\in\Lambda^{n}E\setminus 0. For any CC^{\infty} real vector bundle EXE\to X, the vector bundle Ωα(E)=xΩα(Ex)X\Omega^{\alpha}(E)=\cup_{x}\Omega^{\alpha}(E_{x})\to X is a CC^{\infty} line bundle, with transition functions given by |det(gij)|α|\det(g_{ij})|^{\alpha} if (gij)(g_{ij}) is a set of transition functions for EE. Sections of Ωα(E)\Omega^{\alpha}(E) are called α\alpha-densities on EE and sections of ΩXα:=Ωα(TX)\Omega^{\alpha}_{X}:=\Omega^{\alpha}(TX) are called α\alpha-densities on XX. Densities bundles are always trivialisable, but not canonically in general: one can construct an everywhere positive section using local trivializations.

A fundamental point is that compactly supported one densities on XX can be integrated over XX. More precisely, there is a unique linear form X:Cc(X,ΩX1)\int_{X}:C^{\infty}_{c}(X,\Omega^{1}_{X})\longrightarrow{\mathbb{R}} such that if f=𝖿(x)|dx|f=\mathsf{f}(x)|dx| is compactly supported in a local chart UU with local coordinates x=(x1,,xn)x=(x_{1},\ldots,x_{n}), then

Xf=n𝖿(x)𝑑x.\int_{X}f=\int_{{\mathbb{R}}^{n}}\mathsf{f}(x)dx.

Above, |dx||dx| is the one density defined by |dx|=|dx1dxn||dx|=|dx_{1}\wedge\cdots\wedge dx_{n}|. Diffeomorphisms ϕ:XY\phi:X\to Y provide isomorphisms ϕ:ΩYαΩXα\phi^{*}:\Omega^{\alpha}_{Y}\to\Omega^{\alpha}_{X} given by ϕω(V)=ω(ϕV)\phi^{*}\omega(V)=\omega(\phi_{*}V). By construction, the integral of one densities is invariant under the action of diffeomorphisms. Densities are usually handled with the following canonical isomorphisms:

  1. -

    Ωα(E)Ωβ(E)Ωα+β(E)\Omega^{\alpha}(E)\otimes\Omega^{\beta}(E)\simeq\Omega^{\alpha+\beta}(E)

  2. -

    Ωα(EF)Ωα(E)Ωα(F)\Omega^{\alpha}(E\oplus F)\simeq\Omega^{\alpha}(E)\otimes\Omega^{\alpha}(F),

  3. -

    Ωα(E)Ωα(E)\Omega^{\alpha}(E^{*})\simeq\Omega^{-\alpha}(E)

  4. -

    if 0FEG00\to F\to E\to G\to 0 is exact, then Ωα(E)Ωα(F)Ωα(G)\Omega^{\alpha}(E)\simeq\Omega^{\alpha}(F)\otimes\Omega^{\alpha}(G).

Lie groupoids. A Lie groupoid GMG\rightrightarrows M is a pair of manifolds (G,M)(G,M) of respective dimensions generally denoted by n=n(1)+n(0)n=n^{(1)}+n^{(0)} and n(0)n^{(0)}, together with the following data and required properties. The data are:

  1. (a)

    two surjective submersions r,s:GMr,s:G\to M, called range and source,

  2. (b)

    a CC^{\infty} section υ:MG\upsilon:M\to G of both rr and ss, assimilated to an inclusion,

  3. (c)

    a CC^{\infty} map ι:GG\iota:G\to G called inversion, noted: γ1:=ι(γ)\gamma^{-1}:=\iota(\gamma),

  4. (d)

    a CC^{\infty} map G(2)={(γ1,γ2)G2;s(γ1)=r(γ2)}GG^{(2)}=\{(\gamma_{1},\gamma_{2})\in G^{2}\ ;\ s(\gamma_{1})=r(\gamma_{2})\}\to G called multiplication: γ1γ2:=m(γ1,γ2)\gamma_{1}\gamma_{2}:=m(\gamma_{1},\gamma_{2}).

The required properties are those giving a sense to the following intuition: a groupoid is the algebraic structure obtained from a group GG after spreading out its unit into a whole subset MM, that is

  1. (i)

    r(γ1γ2)=r(γ1)r(\gamma_{1}\gamma_{2})=r(\gamma_{1}), s(γ1γ2)=s(γ2)s(\gamma_{1}\gamma_{2})=s(\gamma_{2}) whenever it makes sense,

  2. (ii)

    υ(r(γ))γ=γ\upsilon(r(\gamma))\gamma=\gamma, γυ(s(γ))=γ\gamma\upsilon(s(\gamma))=\gamma for all γ\gamma,

  3. (iii)

    r(γ1)=s(γ)r(\gamma^{-1})=s(\gamma), s(γ1)=r(γ)s(\gamma^{-1})=r(\gamma) for all γ\gamma,

  4. (iv)

    γγ1=υ(r(γ))\gamma\gamma^{-1}=\upsilon(r(\gamma)), γ1γ=υ(s(γ))\gamma^{-1}\gamma=\upsilon(s(\gamma)) for all γ\gamma,

  5. (v)

    (γ1γ2)γ3=γ1(γ2γ3)(\gamma_{1}\gamma_{2})\gamma_{3}=\gamma_{1}(\gamma_{2}\gamma_{3}) whenever it makes sense.

It follows that υ\upsilon is an embedding (often omitted in the notation), that ι\iota is an involutive diffeomorphism and mm a surjective submersion. We note GxG_{x}, the ss-fiber at xMx\in M, GxG^{x} its rr-fiber, and we set TsG=kerdsT_{s}G=\ker ds, TrG=kerdrT_{r}G=\ker dr. We note LγL_{\gamma}, RγR_{\gamma} the left and right multiplication by γ\gamma. The Lie algebroid AGAG of GMG\rightrightarrows M is by definition here the vector bundle kerds|MM\ker ds|_{M}\to M. The differential map dr:AGTMdr:AG\to TM is denoted by 𝔞\mathfrak{a} and called the anchor map. To any CC^{\infty} section XΓ(AG)X\in\Gamma(AG) corresponds a right invariant vector field X~Γ(TG)\widetilde{X}\in\Gamma(TG), defined by X~γ:=dRγ(Xr(γ))\widetilde{X}_{\gamma}:=dR_{\gamma}(X_{r(\gamma)}), and conversely. The right invariance means X~γη=dRη(X~γ)\widetilde{X}_{\gamma\eta}=dR_{\eta}(\widetilde{X}_{\gamma}). This allows to define a Lie algebra structure on Γ(AG)\Gamma(AG) that satisfies

X,YΓ(AG),fC(M),[X,fY]=f[X,Y]+(𝔞(X)f)Y.\forall X,Y\in\Gamma(AG),\ \forall f\in C^{\infty}(M),\quad[X,fY]=f[X,Y]+(\mathfrak{a}(X)f)Y.

We refer to [24, 19] for a detailed account on Lie groupoids and Lie algebroids.

We will use several α\alpha-densities bundles over GG, often for α=±1/2,±1\alpha=\pm 1/2,\pm 1:

  1. -

    The bundles Ωα(kerdπ)\Omega^{\alpha}(\ker d\pi) of densities along the fibers of π=s,r\pi=s,r. They are conveniently replaced for computations by the respective isomorphic bundles Ωsα=Ωα(rAG)\Omega^{\alpha}_{s}=\Omega^{\alpha}(r^{*}AG) and Ωrα=Ωα(sAG)\Omega^{\alpha}_{r}=\Omega^{\alpha}(s^{*}AG). The isomorphisms are induced by:

    (4) :rAGTsG,(γ,X)(γ,(dRγ)r(γ)(X)),\mathcal{R}:r^{*}AG\longrightarrow T_{s}G,\quad(\gamma,X)\longmapsto\big{(}\gamma,(dR_{\gamma})_{r(\gamma)}(X)\big{)},
    (5) 𝒮:sAGTrG,(γ,X)(γ,(dLγι)s(γ)(X)).\mathcal{S}:s^{*}AG\longrightarrow T_{r}G,\quad(\gamma,X)\longmapsto\big{(}\gamma,(dL_{\gamma}\circ\iota)_{s(\gamma)}(X)\big{)}.
  2. -

    The “symmetrisation” of the preceeding ones: Ωα=ΩsαΩrα\Omega^{\alpha}=\Omega^{\alpha}_{s}\otimes\Omega^{\alpha}_{r}, which is suitable for convolution on GG.

  3. -

    The bundle Ω01/2=Ω1/2ΩG1\Omega^{1/2}_{0}=\Omega^{-1/2}\otimes\Omega^{1}_{G} necessary for the pairing:

    fC(G,Ω1/2),ωCc(G,Ω01/2),f,ω=Gfg.f\in C^{\infty}(G,\Omega^{1/2}),\ \omega\in C^{\infty}_{c}(G,\Omega^{1/2}_{0}),\quad\langle f,\omega\rangle=\int_{G}fg.

    Actually, there is a natural isomorphism Ω01/2Ω1/2(rTM)Ω1/2(sTM)\Omega^{1/2}_{0}\simeq\Omega^{1/2}(r^{*}TM)\otimes\Omega^{1/2}(s^{*}TM).

The cotangent groupoid The cotangent space TGT^{*}G has a non trivial groupoid structure: Γ=TGAG\Gamma=T^{*}G\rightrightarrows A^{*}G, with structure maps rΓ,sΓ,mΓ,ιΓr_{{}_{\Gamma}},s_{{}_{\Gamma}},m_{{}_{\Gamma}},\iota_{{}_{\Gamma}} defined as follows:

  1. -

    rΓ(γ,ξ)=(r(γ),dtRγ(ξ|TγGs(γ)))r_{{}_{\Gamma}}(\gamma,\xi)=\big{(}r(\gamma),{}^{t}dR_{\gamma}(\xi|_{T_{\gamma}G_{s(\gamma)}})\big{)} and sΓ(γ,ξ)=(s(γ),dt(Lγι)(ξ|TγGr(γ)))s_{{}_{\Gamma}}(\gamma,\xi)=\big{(}s(\gamma),-{}^{t}d(L_{\gamma}\circ\iota)(\xi|_{T_{\gamma}G^{r(\gamma)}})\big{)},

  2. -

    (γ1,ξ1)(γ2,ξ2)=(γ1γ2,ξ)(\gamma_{1},\xi_{1})(\gamma_{2},\xi_{2})=(\gamma_{1}\gamma_{2},\xi) with ξ(dm(t1,t2))=ξ1(t1)+ξ2(t2)\xi(dm(t_{1},t_{2}))=\xi_{1}(t_{1})+\xi_{2}(t_{2}),

  3. -

    (γ,ξ)1=(γ1,dtι(ξ))(\gamma,\xi)^{-1}=(\gamma^{-1},-{}^{t}d\iota(\xi)).

This is a symplectic groupoid, which means that the graph of mΓm_{{}_{\Gamma}} is a Lagrangian submanifold of (TG)3(T^{*}G)^{3} provided with the symplectic form ωωω\omega\oplus\omega\oplus-\omega, with ω\omega the canonical symplectic form of TGT^{*}G. We refer to [6, 19] for a detailed account on symplectic groupoids and on the related notion of VBVB-groupoids, as well as to [17, 18] for the interest of this symplectic structure regarding the theory of distributions on groupoids. We will denote

T.G=TGkerrΓ and T.G=TG(kerrΓkersΓ).\quad T^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G=T^{*}G\setminus\ker r_{{}_{\Gamma}}\ \text{ and }\quad\accentset{\mbox{\large.}}{T}^{*}G=T^{*}G\setminus(\ker r_{{}_{\Gamma}}\cup\ker s_{{}_{\Gamma}}).

We will consider in this paper homogeneous lagrangian submanifods of TG0T^{*}G\setminus 0 that avoids the kernel of rΓr_{{}_{\Gamma}} and sΓs_{{}_{\Gamma}}. We call them GG-relations, in reference to the term canonical relations often employed for (product) manifolds. Under mild assumptions, GG-relations compose well in the groupoid TGT^{*}G [18]. GG-relations Λ\Lambda such that sΓ|Λs_{{}_{\Gamma}}|_{\Lambda} and rΓ|Λr_{{}_{\Gamma}}|_{\Lambda} are diffeomorphisms onto their ranges are called invertible. We will sometimes use densities along the sΓs_{{}_{\Gamma}} and rΓr_{{}_{\Gamma}}-fibers of the cotangent groupoid TGT^{*}G. Both are naturally isomorphic and:

ΩsΓαΩrΓαΩ^αΩ^Gα\Omega^{\alpha}_{s_{{}_{\Gamma}}}\simeq\Omega^{\alpha}_{r_{{}_{\Gamma}}}\simeq\hat{\Omega}^{\alpha}\otimes\hat{\Omega}^{-\alpha}_{G}

where E^\hat{E} denotes the pull back to TGT^{*}G of the bundle EGE\to G. Also, we note that ΩsΓ1|AG=Ω1(ATG)=(𝒟AGtr)1\Omega^{1}_{s_{{}_{\Gamma}}}|_{A^{*}G}=\Omega^{1}(AT^{*}G)=(\mathcal{D}_{AG}^{\mathop{\mathrm{tr}}\nolimits})^{-1} where 𝒟AGtr\mathcal{D}_{AG}^{\mathop{\mathrm{tr}}\nolimits} is the transverse density bundle of AGAG [7].

The convolution algebra Throughout this paper we make the convention:

C(G):=C(G,Ω1/2),C^{\infty}(G):=C^{\infty}(G,\Omega^{1/2}),

that is, we omit the ubiquitous density bundle Ω1/2\Omega^{1/2} in the notation. We apply the same convention for other sections of Ω1/2\Omega^{1/2} with various regularity and support conditions. When the sections of a different bundle are considered, this bundle will be always mentionned.

The convolution algebra structure on Cc(G)C^{\infty}_{c}(G) refers to the product \star canonically defined from any of the following three intuitive formulas:

(6) fg(γ)=γ2Gs(γ)f(γγ21)g(γ2)=γ1Gr(γ)f(γ1)g(γ11γ)=(γ1,γ2)m1(γ)f(γ1)g(γ2)f\star g(\gamma)=\int_{\gamma_{2}\in G_{s(\gamma)}}f(\gamma\gamma_{2}^{-1})g(\gamma_{2})=\int_{\gamma_{1}\in G^{r(\gamma)}}f(\gamma_{1})g(\gamma_{1}^{-1}\gamma)=\int_{(\gamma_{1},\gamma_{2})\in m^{-1}(\gamma)}f(\gamma_{1})g(\gamma_{2})

This is justified as follows. Write f=𝖿(μsμr)1/2f=\mathsf{f}(\mu_{s}\mu_{r})^{1/2}, g=𝗀(μsμr)1/2g=\mathsf{g}(\mu_{s}\mu_{r})^{1/2} with 𝖿,𝗀Cc(G,)\mathsf{f},\mathsf{g}\in C^{\infty}_{c}(G,{\mathbb{C}}), μs=rμC(G,Ωs1)\mu_{s}=r^{*}\mu\in C^{\infty}(G,\Omega^{1}_{s}), μr=sμC(G,Ωr1)\mu_{r}=s^{*}\mu\in C^{\infty}(G,\Omega^{1}_{r}) for some positive μC(M,Ω1(AG))\mu\in C^{\infty}(M,\Omega^{1}(AG)). Then, whenever γ1γ2=γ\gamma_{1}\gamma_{2}=\gamma:

f(γ1)g(γ2)=𝖿(γ1)𝗀(γ2)μr1/2(γ1)μs1/2(γ2)(μsμr)1/2(γ) and μr(γ1)=μs(γ2).f(\gamma_{1})g(\gamma_{2})=\mathsf{f}(\gamma_{1})\mathsf{g}(\gamma_{2})\mu_{r}^{1/2}(\gamma_{1})\mu_{s}^{1/2}(\gamma_{2})(\mu_{s}\mu_{r})^{1/2}(\gamma)\text{ and }\mu_{r}(\gamma_{1})=\mu_{s}(\gamma_{2}).

We now may set rigorously:

(7) fg(γ)=(γ2Gs(γ)𝖿(γγ21)𝗀(γ2)μs(γ2))(μsμr)1/2(γ).f\star g(\gamma)=\Big{(}\int_{\gamma_{2}\in G_{s(\gamma)}}\mathsf{f}(\gamma\gamma_{2}^{-1})\mathsf{g}(\gamma_{2})\mathcal{R}_{*}\mu_{s}(\gamma_{2})\Big{)}(\mu_{s}\mu_{r})^{1/2}(\gamma).

This gives consistance to the first formula in (6). The second and third formulas are obtained from the first one using the diffeormorphisms Lγι:Gs(γ)Gr(γ)L_{\gamma}\circ\iota:G_{s(\gamma)}\to G^{r(\gamma)} and (Lγι,Id):Gs(γ)m1(γ)(L_{\gamma}\circ\iota,\operatorname{Id}):G_{s(\gamma)}\to m^{-1}(\gamma). Equivalently, one can directly define them as we did for the first one using the suitable structural isomorphisms to create the appropriate one densities on Gr(γ)G^{r(\gamma)} and m1(γ)m^{-1}(\gamma). With the notation above, this concretely means:

(8) fg(γ)\displaystyle f\star g(\gamma) =(γ1Gr(γ)𝖿(γ1)𝗀(γ11γ)𝒮μr(γ1))(μsμr)1/2(γ)\displaystyle=\Big{(}\int_{\gamma_{1}\in G^{r(\gamma)}}\mathsf{f}(\gamma_{1})\mathsf{g}(\gamma_{1}^{-1}\gamma)\mathcal{S}_{*}\mu_{r}(\gamma_{1})\Big{)}(\mu_{s}\mu_{r})^{1/2}(\gamma)
(9) =((γ1,γ2)m1(γ)𝖿(γ1)𝗀(γ2)μs(γ2))(μsμr)1/2(γ).\displaystyle=\Big{(}\int_{(\gamma_{1},\gamma_{2})\in m^{-1}(\gamma)}\mathsf{f}(\gamma_{1})\mathsf{g}(\gamma_{2})\mathcal{M}_{*}\mu_{s}(\gamma_{2})\Big{)}(\mu_{s}\mu_{r})^{1/2}(\gamma).

with, for the last line:

(10) :rAG|Gs(γ)Tm1(γ),(γ2,X)(γγ21,γ2,d(Lγγ21ι)(X),dRγ2(X)).\mathcal{M}:r^{*}AG|_{G_{s(\gamma)}}\longrightarrow Tm^{-1}(\gamma),\quad(\gamma_{2},X)\longmapsto\big{(}\gamma\gamma_{2}^{-1},\gamma_{2},d(L_{\gamma\gamma_{2}^{-1}}\circ\iota)(X),dR_{\gamma_{2}}(X)\big{)}.

By Cπ,0(G)C^{\infty,0}_{\pi}(G), we denote the space of elements in Cc(G)C_{c}(G) that belong to C(U(0),C(U(1)))C(U_{(0)},C^{\infty}(U_{(1)})) over any local trivializations κ:UU(0)×U(1)\kappa:U\overset{\simeq}{\to}U_{(0)}\times U_{(1)} of π\pi (here π=pr1κ\pi=\mathop{\mathrm{pr}}\nolimits_{1}\circ\kappa). The topology is modeled on that of C(U(0),C(U(1)))C(U_{(0)},C^{\infty}(U_{(1)})) and is Fréchet. We write Cπ,c,0C^{\infty,0}_{\pi,c} for Cπ,0CcC^{\infty,0}_{\pi}\cap C_{c}, and equip it with the corresponding LF-topology.

The reduced CC^{*}-algebra of a groupoid. The space Cc(G,Ωs1/2)C^{\infty}_{c}(G,\Omega^{1/2}_{s}) comes with a natural prehilbertian C(M)C(M)-module structure:

(11) f,gCc(G,Ωs1/2),f|gs(x)=Gxf(γ)¯g(γ).f,g\in C^{\infty}_{c}(G,\Omega^{1/2}_{s}),\quad\langle f\ |\ g\rangle_{s}(x)=\int_{G_{x}}\overline{f(\gamma)}g(\gamma).

Its completion as a hilbertian C(M)C(M)-module is denoted by Ls2(G)L_{s}^{2}(G). The homomorphism λ:Cc(G)(Ls2(G))\lambda:C^{\infty}_{c}(G)\longrightarrow\mathcal{L}(L_{s}^{2}(G)) given by:

fCc(G),gCc(G,Ωs1/2),λ(f)(g)(γ)=fg(γ)=Gs(γ)f(γα1)g(α)\forall f\in C^{\infty}_{c}(G),\ g\in C^{\infty}_{c}(G,\Omega^{1/2}_{s}),\quad\lambda(f)(g)(\gamma)=f\star g(\gamma)=\int_{G_{s(\gamma)}}f(\gamma\alpha^{-1})g(\alpha)

is well defined, injective, and the reduced CC^{*}-algebra of GG, denoted by Cr(G)C^{*}_{r}(G), is the completion of Cc(G)C^{\infty}_{c}(G) with respect to the CC^{*}-norm f=λ(f)op\|f\|=\|\lambda(f)\|_{\mathrm{op}}. The extended homomorphism

λ:Cr(G)(Ls2(G))\lambda:C^{*}_{r}(G)\longrightarrow\mathcal{L}(L_{s}^{2}(G))

is called the left regular representation. Starting from Cc(G,Ωr1/2)C^{\infty}_{c}(G,\Omega^{1/2}_{r}), we get a Hilbert C(M)C(M)-module Lr2(G)L^{2}_{r}(G) and the right regular representation ρ:Cr(G)(Lr2(G))\rho:C^{*}_{r}(G)\longrightarrow\mathcal{L}(L_{r}^{2}(G)). The adjunction map :Ls2(G)Lr2(G)*:L_{s}^{2}(G)\longrightarrow L_{r}^{2}(G) provides a unitary anti-homomorphism. The unfamiliar reader may consult [29, 5, 14] for groupoids CC^{*}-algebras and [33, 13, 30] for Hilbertian modules.

Distributions. We consider in this article various spaces of distributions on GG, always valued in Ω1/2\Omega^{1/2}, which thus is safely omitted. We set:

𝒟(G):=𝒟(G,Ω1/2).\mathcal{D}^{\prime}(G):=\mathcal{D}^{\prime}(G,\Omega^{1/2}).

This is the topological dual of the space:

𝒟(G):=Cc(G,Ω01/2),\mathcal{D}(G):=C^{\infty}_{c}(G,\Omega^{1/2}_{0}),

where Ω01/2:=Ω1/2ΩG1\Omega^{1/2}_{0}:=\Omega^{-1/2}\otimes\Omega^{1}_{G}. The elements of 𝒟(G)\mathcal{D}(G) will be called test functions, with a slight abuse of vocabulary. We denote by (G)\mathcal{E}^{\prime}(G) the subspace of 𝒟(G)\mathcal{D}^{\prime}(G) consisting of compactly supported distributions. We set:

(12) 𝒟π(G)={u𝒟(G);fCc(G,Ωs1/2),π!(uf)C(M,Ω1/2(AG))}\mathcal{D}^{\prime}_{\pi}(G)=\{u\in\mathcal{D}^{\prime}(G)\ ;\ \forall f\in C^{\infty}_{c}(G,\Omega^{1/2}_{s}),\ \pi_{!}(uf)\in C^{\infty}(M,\Omega^{1/2}(AG))\}

where π!\pi_{!} denotes the pushforward of distributions and π=r,s\pi=r,s. Elements of 𝒟π(G)\mathcal{D}^{\prime}_{\pi}(G) are called CC^{\infty}-transversal distributions with respect to π\pi [1, 17, 31]. The convolution product \star extends by continuity to transversal distibutions, providing π(G)\mathcal{E}_{\pi}^{\prime}(G) with the structure of a unital algebra and r,s(G)=s(G)r(G)\mathcal{E}_{r,s}^{\prime}(G)=\mathcal{E}_{s}^{\prime}(G)\cap\mathcal{E}_{r}^{\prime}(G) with the structure of an involutive unital algebra. The unit is δM(f)=Mf\delta_{M}(f)=\int_{M}f and the involution is u=ιu¯u^{*}=\overline{\iota^{*}u}. Elements of 𝒟π(G)\mathcal{D}^{\prime}_{\pi}(G) can be restricted fiberwise, giving CC^{\infty} families over MM of distributions in the fibers, the space of whose families being denoted by Cπ(M,𝒟(G))C^{\infty}_{\pi}(M,\mathcal{D}^{\prime}(G)), or viewed as C(M)C^{\infty}(M)-linear continuous operators, the space of whose operators being denoted by C(M)(Cc(G),C(M,Ω1/2(AG)))\mathcal{L}_{C^{\infty}(M)}(C^{\infty}_{c}(G),C^{\infty}(M,\Omega^{1/2}(AG))), and there are canonical isomorphisms:

𝒟π(G)Cπ(M,𝒟(G))C(M)(Cc(G),C(M,Ω1/2(AG))).\mathcal{D}^{\prime}_{\pi}(G)\simeq C^{\infty}_{\pi}(M,\mathcal{D}^{\prime}(G))\simeq\mathcal{L}_{C^{\infty}(M)}(C^{\infty}_{c}(G),C^{\infty}(M,\Omega^{1/2}(AG))).

We will also consider continuously transversal distributions with respect to π=r,s\pi=r,s:

(13) 𝒟π,0(G)={u𝒟(G);fCc(G,Ωs1/2),π!(uf)C(M,Ω1/2(AG))}.\mathcal{D}^{\prime}_{\pi,0}(G)=\{u\in\mathcal{D}^{\prime}(G)\ ;\ \forall f\in C^{\infty}_{c}(G,\Omega^{1/2}_{s}),\ \pi_{!}(uf)\in C(M,\Omega^{1/2}(AG))\}.

By rephrazing the arguments in [17], one gets:

𝒟π,0(G)Cπ(M,𝒟(G))C(M)(Cπ,c,0(G),C(M,Ω1/2(AG))).\mathcal{D}^{\prime}_{\pi,0}(G)\simeq C_{\pi}(M,\mathcal{D}^{\prime}(G))\simeq\mathcal{L}_{C(M)}(C^{\infty,0}_{\pi,c}(G),C(M,\Omega^{1/2}(AG))).

GG-operators: they are the continuous linear maps Cc(G)C(G)C^{\infty}_{c}(G)\to C^{\infty}(G) given by right invariant families of (linear continuous) operators acting in the ss-fibers. More precisely, PP is a GG-operator if there exists a family Px:Cc(Gx,ΩGx1/2)C(Gx,ΩGx1/2)P_{x}:C^{\infty}_{c}(G_{x},\Omega^{1/2}_{G_{x}})\longrightarrow C^{\infty}(G_{x},\Omega^{1/2}_{G_{x}}), xMx\in M, such that for all xMx\in M, γG\gamma\in G, fCc(G)f\in C^{\infty}_{c}(G):

(14) P(f)|Gx=Px(f|Gx) and Pr(γ)Rγ=RγPs(γ),γG.P(f)|_{G_{x}}=P_{x}(f|_{G_{x}})\text{ and }P_{r(\gamma)}\circ R_{\gamma}=R_{\gamma}\circ P_{s(\gamma)},\ \forall\gamma\in G.

This is equivalent to requiring that PP maps Cc(G)C(G)C^{\infty}_{c}(G)\to C^{\infty}(G) continuously and that:

P(f)g=P(fg) for any f,gCc(G).P(f)\star g=P(f\star g)\quad\text{ for any }f,g\in C^{\infty}_{c}(G).

A GG-operator PP as an adjoint if there exists a GG-operator QQ such that (Pf)g=f(Qg)(Pf)^{*}\star g=f^{*}\star(Qg) for any f,gf,g. We denote by OpG\mathop{\mathrm{Op}}\nolimits_{G} (resp. OpG\mathop{\mathrm{Op}}\nolimits^{*}_{G}, OpG,c\mathop{\mathrm{Op}}\nolimits^{*}_{G,c}) the space of (resp. adjointable, compactly supported and adjointable) GG-operators.

It is proved in [17] that the map

𝒟r(G)OpG,uu\mathcal{D}^{\prime}_{r}(G)\to\mathop{\mathrm{Op}}\nolimits_{G},\quad u\mapsto u\star\cdot

is an isomorphism, with inverse PkPP\mapsto k_{P}, given by kP(γ)=pr(γ)(r(γ),γ1)k_{P}(\gamma)=p_{r(\gamma)}(r(\gamma),\gamma^{-1}) where pxp_{x} denotes the Schwartz kernel of PxP_{x}. The same map induces an isomorphism:

OpG,cr,s(G).\mathrm{Op}_{G,c}^{*}\simeq\mathcal{E}^{\prime}_{r,s}(G).

Pseudodifferential GG-operators and regularizing operators. Among the class of GG-operators one finds the well known subclass of pseudodifferential GG-operators (GG-PDO) [4, 26, 28, 32], that is, of right invariant families of pseudodifferential operators in the ss-fibers: they coincide with left convolution by distributions in:

(15) ΨG=I+(n(1)n(0))/4(G,M;Ω1/2).\Psi^{*}_{G}=I^{*+(n^{(1)}-n^{(0)})/4}(G,M;\Omega^{1/2}).

where here II refers to the space of conormal distributions. One has a principal symbol map:

σ0:ΨGmS[m](AG)\sigma_{0}:\Psi^{m}_{G}\longrightarrow S^{[m]}(A^{*}G)

with kernel ΨGm1\Psi^{m-1}_{G}. Here S[m]=Sm/Sm1S^{[m]}=S^{m}/S^{m-1}. It is well known that (ΨG,c,)(\Psi^{*}_{G,c},\star) is an involutive unital algebra and σ0\sigma_{0} an algebra homomorphism. When PΨG,c1P\in\Psi^{1}_{G,c} is elliptic and symmetric, then its closure, as an unbounded operator on Cr(G)C^{*}_{r}(G) with domain Cc(G)C^{\infty}_{c}(G), is selfadjoint and regular [32, 2, 3, 30]. There is a canonical scale HtH^{t}, tt\in{\mathbb{R}}, of Hilbert Cr(G)C^{*}_{r}(G)-modules, that we call intrinsic Sobolev modules, which do not depend, up to isomorphism of Hilbertian structures, on the symmetric elliptic operator PΨG,c1P\in\Psi^{1}_{G,c} used to define them:

Ht=Cc(G)¯|t,f|gt=(1+P2)tf|gCr(G),H^{t}=\overline{C^{\infty}_{c}(G)}^{\langle\cdot|\cdot\rangle_{t}},\quad\langle f\ |\ g\rangle_{t}=\langle(1+P^{2})^{t}\star f\ |\ g\rangle\in C^{*}_{r}(G),

where a|b=ab\langle a|b\rangle=a^{*}b. Then any QΨG,cmQ\in\Psi^{m}_{G,c} gives a bounded homomorphism Q(Ht,Htm)Q\in\mathcal{L}(H^{t},H^{t-m}) and for any m>0m>0, the inclusion HtHt+mH^{t}\hookrightarrow H^{t+m} is a compact homomorphism of Hilbert modules. All of this material is developped in [32]. Although we call the spaces HtH^{t} Sobolev modules, we may think of them as modules of abstract pseudodifferential operators of order <t<-t. Indeed, HtH^{t} is also the completion of ΨG,c<t\Psi^{<-t}_{G,c} for the norm Qt=(1+P2)t/2QCr(G)\|Q\|_{t}=\|(1+P^{2})^{t/2}Q\|_{C^{*}_{r}(G)}. Scales of Hilbert modules closer to the usual notion of Sobolev regularity of order tt for functions or distributions will be obtained using the left regular representation of HtH^{t}.

The algebra ΨG,c\Psi^{*}_{G,c} is too small for practical purposes. For instance, the inverse of an elliptic element in ΨG,c\Psi^{*}_{G,c} which is invertible as an operator between Sobolev modules, has no reason to be compactly supported. This phenomenom propagates to operators obtained by holomorphic functional calculus and we will eventually face it also when building an approximation of E(t)=eitPE(t)=e^{itP} by Fourier integral GG-operators. A suitable enlargement of ΨG,c\Psi^{*}_{G,c} is provided by:

(16) ΨG:=ΨG,c+, where =H(H)\Psi^{*}_{G}:=\Psi^{*}_{G,c}+\mathcal{H}^{\infty},\quad\text{ where }\mathcal{H}^{\infty}=H^{\infty}\cap(H^{\infty})^{*}

and H=tHtCr(G)H^{\infty}=\cap_{t}H^{t}\subset C^{*}_{r}(G). Actually, \mathcal{H}^{\infty} coincides with the ideal of regularizing operators introduced in [32].

Fourier integral GG-operators. Another remarkable subclass of GG-operators is given by that of lagrangian distributions on GG with respect to arbitrary GG-relations. We call them Fourier integral GG-operators (GG-FIO) and we set for a given GG-relation Λ\Lambda:

(17) ΦG(Λ)=I+(n(1)n(0))/4(G,Λ;Ω1/2).\Phi^{*}_{G}(\Lambda)=I^{*+(n^{(1)}-n^{(0)})/4}(G,\Lambda;\Omega^{1/2}).

where here II refers to the space of Lagrangian distributions. The convolution product gives a map

(18) ΦG,c(Λ1)×ΦG(Λ2)ΦG(Λ1Λ2)\Phi^{*}_{G,c}(\Lambda_{1})\times\Phi^{*}_{G}(\Lambda_{2})\to\Phi^{*}_{G}(\Lambda_{1}\Lambda_{2})

as soon as Λ1×Λ2\Lambda_{1}\times\Lambda_{2} has a clean intersection with TG(2)T^{*}G^{(2)}. This proves in particular that ΦG(Λ)\Phi^{*}_{G}(\Lambda) is a bimodule over ΨG,c\Psi^{*}_{G,c} and that F1PFΨG,cF^{-1}PF\in\Psi^{*}_{G,c} if PΨG,cP\in\Psi^{*}_{G,c} and FΦG,c(Λ)F\in\Phi^{*}_{G,c}(\Lambda) is invertible. Also, when Λ\Lambda is invertible, one gets ΦG,c0(Λ)(Cr(G))\Phi^{0}_{G,c}(\Lambda)\subset\mathcal{M}(C^{*}_{r}(G)) and ΦG,c<0(Λ)Cr(G)\Phi^{<0}_{G,c}(\Lambda)\subset C^{*}_{r}(G). In general, if AΦG(Λ)A\in\Phi^{*}_{G}(\Lambda), the corresponding family (Ax)xM(A_{x})_{x\in M} consists of operators AxA_{x} given by locally finite sums of oscillatory integrals and when Λ\Lambda is transversal to TLGT_{L}^{*}G, for any L=s1(𝒪)L=s^{-1}(\mathcal{O}) and 𝒪M/G\mathcal{O}\in M/G, (this is for instance the case if Λ\Lambda is invertible), then each AxA_{x} is a genuine Fourier integral operator on the manifold GxG_{x}. All the statements here about GG-FIOs are proved in [18].

3. The one parameter group eitP,te^{-itP},\ t\in{\mathbb{R}}

Before analyzing evolution equations on groupoids, we study the functional analytic aspects of them in a reasonably general and simple framework. So, let us consider the Cauchy problem:

(19) {(t+iP)u=fu(0)=g\begin{cases}(\frac{\partial}{\partial t}+iP)u=f\\ u(0)=g\end{cases}

in the following situation: PP is a regular self-adjoint operator on HH [2, 3, 30, 32] where HH is a Hilbert module over some CC^{*}-algebra AA. It turns out that under natural assumptions on ff and gg, this problem has a unique solution given in term of the operator eitPe^{-itP}. This operator is first defined in term of the unbounded continuous functional calculus for regular operators [30, Paragraph 14.3.3]. We recall that any nondegenerate representation

π:C0()(H)\pi:C_{0}({\mathbb{R}})\longrightarrow\mathscr{L}(H)

extends into a map π~\widetilde{\pi} from C()C({\mathbb{R}}) (viewed as regular operators on C0()C_{0}({\mathbb{R}})) to the set of regular operators on HH. The map π~\widetilde{\pi} is defined through the identification C0()πHHC_{0}({\mathbb{R}})\otimes_{\pi}H\simeq H and the formula:

π~(f)=fπId.\widetilde{\pi}(f)=f\otimes_{\pi}\operatorname{Id}.

Moreover, there exists a unique such representation π\pi such that π~(Id)=P\widetilde{\pi}(\operatorname{Id}_{{\mathbb{R}}})=P and we fix this particular one from now on. Introducing ftC()f_{t}\in C({\mathbb{R}}), ft(λ)=eitλf_{t}(\lambda)=e^{-it\lambda}, we set:

eitP=π~(ft).e^{-itP}=\widetilde{\pi}(f_{t}).

Actually, the restriction of π~\widetilde{\pi} to Cb()C_{b}({\mathbb{R}}) is a strictly continuous homomorphism [30, Proposition 5.19] :

π¯=π~|Cb():Cb()(H).\overline{\pi}=\widetilde{\pi}|_{C_{b}({\mathbb{R}})}:C_{b}({\mathbb{R}})\longrightarrow\mathcal{L}(H).

Here, strict continuity refers to the topologies of Cb()C_{b}({\mathbb{R}}) and (H)\mathscr{L}(H) as multiplier algebras of C0()C_{0}({\mathbb{R}}) and 𝒦(H)\mathcal{K}(H) respectively. The map tftCb(){\mathbb{R}}\ni t\mapsto f_{t}\in C_{b}({\mathbb{R}}) being strictly continuous, the map teitP(H){\mathbb{R}}\ni t\to e^{-itP}\in\mathcal{L}(H) is thus strictly continuous too. Specializing the semi-norms giving the strict topology to rank one operators, this means that teitPxC(,H)t\longmapsto e^{-itP}x\in C({\mathbb{R}},H). The following properties are valid:

(20) ei(t+s)P=eitPeisPe^{-i(t+s)P}=e^{-itP}e^{-isP}

and defining (Ef)(t)=eitPf(t)(Ef)(t)=e^{-itP}f(t), tt\in{\mathbb{R}}, for fCb(,H)f\in C_{b}({\mathbb{R}},H) we also get:

(21) E(Cb(,H)).E\in\mathcal{L}(C_{b}({\mathbb{R}},H)).

To further analyse eitPe^{-itP}, we introduce the sequence of Hilbert AA-modules associated to PP:

s+,Hs=dom(1+P2)s/2 and H0=H\forall s\in{\mathbb{R}}_{+},\ H^{s}=\mathop{\mathrm{dom}}\nolimits(1+P^{2})^{s/2}\text{ and }H^{0}=H

Note that Hk=domPkH^{k}=\mathop{\mathrm{dom}}\nolimits P^{k} for kk\in{\mathbb{N}}. The Hilbertian structure of HsH^{s} is given by:

u,vs=(1+P2)su,v.\langle u,v\rangle_{s}=\langle(1+P^{2})^{s}u,v\rangle.

For negative order ss, we define HsH^{s} to be the completion of domP\mathop{\mathrm{dom}}\nolimits P with respect to the prehilbertian structure given by the scalar product above. We refer to this family of Hilbert AA-modules as the intrinsic scale of Sobolev modules of PP. It was introduced in [32] in the framework of groupoid CC^{*}-algebras.

We recall that π~(ft(λ)λk)=π~(ft(λ))π~(λk)\widetilde{\pi}(f_{t}(\lambda)\lambda^{k})=\widetilde{\pi}(f_{t}(\lambda))\widetilde{\pi}(\lambda^{k}) and that π~(λk)=Pk\widetilde{\pi}(\lambda^{k})=P^{k} for any k0k\geq 0, therefore:

eitP(Hk)=Hk and eitPPk=PkeitP.e^{-itP}(H^{k})=H^{k}\text{ and }e^{-itP}P^{k}=P^{k}e^{-itP}.

In particular, we get eitP(Hk)e^{itP}\in\mathcal{L}(H^{k}) and teitPxC(,Hk)t\longmapsto e^{-itP}x\in C({\mathbb{R}},H^{k}) for any xHkx\in H^{k}. Since 1t(eitλ1)t0iλ\frac{1}{t}(e^{-it\lambda}-1)\overset{t\to 0}{\longrightarrow}i\lambda uniformly on compact subsets of {\mathbb{R}}, we get using [9, Appendix] that 1t(eitP1)\frac{1}{t}(e^{itP}-1) converges to iPiP strongly, that is,

1t(eitPxx)iPxHt00, for all xH1.\|\frac{1}{t}(e^{itP}x-x)-iPx\|_{H}\overset{t\to 0}{\longrightarrow}0,\qquad\text{ for all }x\in H^{1}.

Therefore

(22) xH1,(teitPx)C1(,H)C0(,H1)\forall x\in H^{1},\quad(t\longmapsto e^{-itP}x)\in C^{1}({\mathbb{R}},H)\cap C^{0}({\mathbb{R}},H^{1})

and

(23) xH1,t,ddteitPx=iPeitPx.\forall x\in H^{1},\ \forall t\in{\mathbb{R}},\quad\frac{d}{dt}e^{-itP}x=-iPe^{-itP}x.

Repeating the previous arguments gives for any natural number kk:

(24) xHk,(teitPx)0jkCj(,Hkj).\forall x\in H^{k},\quad(t\longmapsto e^{-itP}x)\in\bigcap_{0\leq j\leq k}C^{j}({\mathbb{R}},H^{k-j}).

This eventually implies:

(25) xH,(teitPx)kC(,Hk)=:C(,H),\forall x\in H^{\infty},\quad(t\longmapsto e^{-itP}x)\in\bigcap_{k}C^{\infty}({\mathbb{R}},H^{k})=:C^{\infty}({\mathbb{R}},H^{\infty}),

where H=kHkH^{\infty}=\cap_{k}H^{k} has Frechet space structure given by the seminorms Hk\|\cdot\|_{H^{k}}, k0k\geq 0. We can now state the result:

Theorem 1.

Let kk be a positive integer. For any fCk1(,Hk)f\in C^{k-1}({\mathbb{R}},H^{k}) and gHkg\in H^{k}, the Cauchy problem:

(26) {(t+iP)u=fu(0)=g\begin{cases}(\frac{\partial}{\partial t}+iP)u=f\\ u(0)=g\end{cases}

has a unique solution in 0jkCj(,Hkj)\bigcap_{0\leq j\leq k}C^{j}({\mathbb{R}},H^{k-j}), given by

(27) u(t)=eitPg+0tei(st)Pf(s)𝑑s.u(t)=e^{-itP}g+\int_{0}^{t}e^{i(s-t)P}f(s)ds.
Proof.

That eitPge^{-itP}g is in the required space and satisfies the equation when f=0f=0 is done before the statement of the theorem. Straightforward arguments prove that the second term in the expression of u(t)u(t) in (27) is in the required space too, and it is then obvious that uu solves (19). For unicity, consider the case f=g=0f=g=0 and let uu be a solution. Pairing the equation with uu on both sides gives the relations

itu,u+Pu,u\displaystyle-i\langle\frac{\partial}{\partial t}u,u\rangle+\langle Pu,u\rangle =0\displaystyle=0
iu,tu+u,Pu\displaystyle i\langle u,\frac{\partial}{\partial t}u\rangle+\langle u,Pu\rangle =0.\displaystyle=0.

Since PP is selfadjoint, substracting both relations gives

tu,u=tu,u+u,tu=0.\displaystyle\frac{\partial}{\partial t}\langle u,u\rangle=\langle\frac{\partial}{\partial t}u,u\rangle+\langle u,\frac{\partial}{\partial t}u\rangle=0.

Therefore, u(t)H2=u(t),u(t)A=u(0),u(0)A=0\|u(t)\|_{H}^{2}=\|\langle u(t),u(t)\rangle\|_{A}=\|\langle u(0),u(0)\rangle\|_{A}=0 for any tt. ∎

Keeping the previous setting, let BB be a CC^{*}-algebra, LL be a Hilbert BB-module and λ:A(L)\lambda:A\longrightarrow\mathcal{L}(L) be a representation. Then Hλ=HλLH_{\lambda}=H\otimes_{\lambda}L is a Hilbert BB-module and Pλ=PλIdP_{\lambda}=P\otimes_{\lambda}\operatorname{Id} is a selfadjoint regular operator acting on it. Then, Proposition 1 applies to PλP_{\lambda} and we get the following corollary, using [30, 14.3.2].

Corollary 2.
(28)  for k>0,domPλk=Hλk\text{ for }k>0,\quad\mathop{\mathrm{dom}}\nolimits P_{\lambda}^{k}=H^{k}_{\lambda}

and we have the equality:

(29) eitPλ=eitPλId.e^{itP_{\lambda}}=e^{itP}\otimes_{\lambda}\operatorname{Id}.

4. Distributions, test functions and weak factorizations for a Lie groupoid

From now on, and in the remaining parts of this article, we fix a Lie groupoid GG of dimension n=n(1)+n(0)n=n^{(1)}+n^{(0)} with compact basis G(0)=MG^{(0)}=M of dimension n(0)n^{(0)}. We recall that:

Ω1/2:=Ωs1/2Ωr1/2=Ω1/2(rAG)Ω1/2(sAG)Ω1/2(kerds)Ω1/2(kerdr).\Omega^{1/2}:=\Omega^{1/2}_{s}\otimes\Omega^{1/2}_{r}=\Omega^{1/2}(r^{*}AG)\otimes\Omega^{1/2}(s^{*}AG)\simeq\Omega^{1/2}(\ker ds)\otimes\Omega^{1/2}(\ker dr).

and that the bundle Ω01/2\Omega^{1/2}_{0} used in the space of test functions 𝒟(G)=Cc(G,Ω01/2)\mathcal{D}(G)=C^{\infty}_{c}(G,\Omega^{1/2}_{0}) satisfies:

(30) Ω01/2:=Ω1/2ΩG1Ωs1/2Ωr1/2sΩMΩr1/2Ωs1/2rΩMrΩM1/2sΩM1/2.\Omega^{1/2}_{0}:=\Omega^{-1/2}\otimes\Omega^{1}_{G}\simeq\Omega^{1/2}_{s}\otimes\Omega^{-1/2}_{r}\otimes s^{*}\Omega_{M}\simeq\Omega^{1/2}_{r}\otimes\Omega^{-1/2}_{s}\otimes r^{*}\Omega_{M}\simeq r^{*}\Omega^{1/2}_{M}\otimes s^{*}\Omega^{1/2}_{M}.

All the isomorphisms above are easily checked using the isomorphisms

ΩGαΩsαsΩMαΩrαrΩMα\Omega^{\alpha}_{G}\simeq\Omega^{\alpha}_{s}\otimes s^{*}\Omega^{\alpha}_{M}\simeq\Omega^{\alpha}_{r}\otimes r^{*}\Omega^{\alpha}_{M}

that result from the exact sequences:

0kerdσTGdσσTM0,σ=s,r0\longrightarrow\ker d\sigma\longrightarrow TG\overset{d\sigma}{\longrightarrow}\sigma^{*}TM\longrightarrow 0,\qquad\sigma=s,r

as well as straight properties of the calculus of densities. To finish with this description, we mention that Ω01/2\Omega^{1/2}_{0} is related, but distinct, to the transverse density bundle 𝒟AGtr\mathcal{D}^{\mathop{\mathrm{tr}}\nolimits}_{AG} of [7]. The latter is GG-invariant and serves to produce geometric transverse measures useful for the geometry of groupoids and stacks, while our choice of “transverse” bundle is required for the pairing with densities in Ω1/2\Omega^{1/2}, but only equivariant with respect to {\mathbb{R}}-actions provided by invariant vectors fields.

Moreover, besides its pairing with distributions, the space 𝒟(G)\mathcal{D}(G) appears to be a bimodule over Cc(G)C^{\infty}_{c}(G), with left and right multiplication given by the canonically defined integrals:

(31) fξ(γ)=Gs(γ)f(γα1)g(α) and ξf(γ)=Gs(γ)ξ(γα1)f(α).f\star\xi(\gamma)=\int_{G_{s(\gamma)}}f(\gamma\alpha^{-1})g(\alpha)\text{ and }\xi\star f(\gamma)=\int_{G_{s(\gamma)}}\xi(\gamma\alpha^{-1})f(\alpha).

Finally, we recall that the embedding Cc(G)𝒟(G)C^{\infty}_{c}(G)\subset\mathcal{D}^{\prime}(G) is given by:

uCc(G),ω𝒟(G),u,ω=Gu(γ)ω(γ)𝑑γ.\forall u\in C^{\infty}_{c}(G),\omega\in\mathcal{D}(G),\quad\langle u,\omega\rangle=\int_{G}u(\gamma)\omega(\gamma)d\gamma.

The inversion map ι:GG\iota:G\to G acts on sections of Ω1/2\Omega^{1/2} and Ω01/2\Omega^{1/2}_{0} in the natural way. This gives involutive isomorphisms:

(32) ι:𝒟(G)𝒟(G) and ι:Cc(G)Cc(G).\ \iota^{*}:\mathcal{D}(G)\longrightarrow\mathcal{D}(G)\text{ and }\iota^{*}:C^{\infty}_{c}(G)\longrightarrow C^{\infty}_{c}(G).

The second one extends to an involutive isomorphism ι:𝒟(G)𝒟(G)\iota^{*}:\mathcal{D}^{\prime}(G)\longrightarrow\mathcal{D}^{\prime}(G).

Proposition 3.
  1. (1)

    For any (u,ω)𝒟r,s(G)×𝒟(G)(u,\omega)\in\mathcal{D}^{\prime}_{r,s}(G)\times\mathcal{D}(G), we have

    (33) u,ω=δM,ιuω=δM,ωιu=ιu,ιω (trace property). \langle u,\omega\rangle=\langle\delta_{M},\iota^{*}u\star\omega\rangle=\langle\delta_{M},\omega\star\iota^{*}u\rangle=\langle\iota^{*}u,\iota^{*}\omega\rangle\text{ (trace property). }

    The trace property u,ω=ιu,ιω\langle u,\omega\rangle=\langle\iota^{*}u,\iota^{*}\omega\rangle is still valid with u𝒟(G)u\in\mathcal{D}^{\prime}(G).

  2. (2)

    The map ι\iota^{*} is an anti-isomorphism of the algebra r,s(G)\mathcal{E}^{\prime}_{r,s}(G):

    (34) u,vr,s(G),ι(uv)=ιvιu,\forall u,v\in\mathcal{E}^{\prime}_{r,s}(G),\quad\iota^{*}(u\star v)=\iota^{*}v\star\iota^{*}u,
  3. (3)

    The space 𝒟(G)\mathcal{D}(G) is a bimodule over r,s(G)\mathcal{E}^{\prime}_{r,s}(G) and ι\iota^{*} is a bimodule antisomorphism:

    u,vr,s(G),ω𝒟(G),ι(uωv)=ιvιωιu,\forall u,v\in\mathcal{E}^{\prime}_{r,s}(G),\forall\omega\in\mathcal{D}(G),\quad\iota^{*}(u\star\omega\star v)=\iota^{*}v\star\iota^{*}\omega\star\iota^{*}u,
  4. (4)

    For any u,vr,s(G)u,v\in\mathcal{E}^{\prime}_{r,s}(G) and ω𝒟(G)\omega\in\mathcal{D}(G), we have:

    uv,ω\displaystyle\langle u\star v,\omega\rangle =v,ιuω=u,ωιv=δM,ιuωιv\displaystyle=\langle v,\iota^{*}u\star\omega\rangle=\langle u,\omega\star\iota^{*}v\rangle=\langle\delta_{M},\iota^{*}u\star\omega\star\iota^{*}v\rangle
Proof.

That 𝒟(G)\mathcal{D}(G) is a bimodule over r,s(G)\mathcal{E}^{\prime}_{r,s}(G) follows directly from [1, 17]. If uu is CC^{\infty}, the quantity u,ω\langle u,\omega\rangle is the integral of the one density on GG defined by the product uωu\omega, whose integral is then invariant by action of diffeomorphisms. In particular, u,ω=Gu(γ1)ω(γ1)\langle u,\omega\rangle=\int_{G}u(\gamma^{-1})\omega(\gamma^{-1}). On the other hand, one is allowed to write

u,ω=M(Gxιu(γ1)ω(γ))𝑑x=Mιuω(x)𝑑x.\langle u,\omega\rangle=\int_{M}\big{(}\int_{G_{x}}\iota^{*}u(\gamma^{-1})\omega(\gamma)\big{)}dx=\int_{M}\iota^{*}u\star\omega(x)dx.

Both identities together give (1) when uu is CC^{\infty}, and the general case follows by density and continuity. The identities given in (2) and (3) are then checked easily. ∎

Let XΓ(AG)X\in\Gamma(AG). Since AGTGAG\subset TG, the vector field XX provides at any xMx\in M a local derivation Xx:𝒟(G)Ω1(TxM)X_{x}:\mathcal{D}(G)\to\Omega^{1}(T_{x}M) and xXxωx\mapsto X_{x}\omega is CC^{\infty} for any ω𝒟(G)\omega\in\mathcal{D}(G). Therefore XΓ(AG)X\in\Gamma(AG) provides a distribution

τXDiff(G)={uΨG;supp(u)M}Ψc(G),\tau_{X}\in\mathop{\mathrm{Diff}}\nolimits(G)=\{u\in\Psi^{*}_{G}\ ;\ {\mathrm{supp}(u)}\subset M\}\subset\Psi^{*}_{c}(G),

via the formula:

ωC(G),τX,ω=MXω.\forall\omega\in C^{\infty}(G),\quad\langle\tau_{X},\omega\rangle=\int_{M}X\omega.

We recall that the algebra isomorphism

(35) op:\displaystyle\mathop{\mathrm{op}}\nolimits: (r,s(G),)\displaystyle(\mathcal{E}^{\prime}_{r,s}(G),\star) (OpG,c,)\displaystyle\longrightarrow(\mathop{\mathrm{Op}}\nolimits_{G}^{*,c},\circ)
u\displaystyle u u\displaystyle\longmapsto u\star\cdot

maps ΨG,c\Psi^{*}_{G,c} (resp. DiffG\mathop{\mathrm{Diff}}\nolimits^{*}_{G}) to the algebra of uniformly supported and equivariant CC^{\infty} family of pseudodifferential (resp. equivariant CC^{\infty} family of differential) operators on the fibers of ss [28, 25, 17].

Note that the action of τX\tau_{X} as a differential GG-operator is given, up to inversion, by the right invariant vector field X~\widetilde{X} associated with XX:

uCc(G),ιτXu=X~u.\forall u\in C^{\infty}_{c}(G),\quad\iota^{*}\tau_{X}\star u=\widetilde{X}u.

Let φ\varphi be the flow of the vector field X~\widetilde{X}. By compacity of MM, there exists ε>0\varepsilon>0 and a neighborhood UU of MM into GG such that φ\varphi is defined on ]ε,ε[×U]-\varepsilon,\varepsilon[\times U. Since X~γ=dRγ(Xr(γ))\widetilde{X}_{\gamma}=dR_{\gamma}(X_{r(\gamma)}) for any γ\gamma we get the relation φ(t,γη)=φ(t,γ)η\varphi(t,\gamma\eta)=\varphi(t,\gamma)\eta whenever both terms are well defined. Therefore the flow φ\varphi is well defined on ]ε,ε[×G]-\varepsilon,\varepsilon[\times G, and then on ×G{\mathbb{R}}\times G using the one parameter group property. This proves that the flow of X~\widetilde{X} is complete and commutes with right multiplication in GG:

(36) t,γ,ηG(2),φ(t,γ)Gs(γ) and φ(t,γη)=φ(t,γ)η.\forall t\in{\mathbb{R}},\forall\gamma,\eta\in G^{(2)},\quad\varphi(t,\gamma)\in G_{s(\gamma)}\text{ and }\varphi(t,\gamma\eta)=\varphi(t,\gamma)\eta.

In other words, XX provides an action of {\mathbb{R}} on the manifold GG, which is equivariant with respect to right multiplication. Also, the map ψ:=rφ:×MM\psi:=r\circ\varphi:{\mathbb{R}}\times M\longrightarrow M is the flow of the vector field 𝔞(X)Γ(TM)\mathfrak{a}(X)\in\Gamma(TM) where 𝔞=dr|TM\mathfrak{a}=dr|_{TM} is the anchor map of GG [19] and the map:

φ:ψMG,(t,x)φ(t,x)\varphi:{\mathbb{R}}\ltimes_{\psi}M\longrightarrow G,\quad(t,x)\longmapsto\varphi(t,x)

is a (CC^{\infty}) homomorphism of groupoids over MM. We recall that a groupoid homorphism h:G1G2h:G_{1}\to G_{2} over (the identity map of) X=G1(0)=G2(0)X=G_{1}^{(0)}=G_{2}^{(0)} is a map satisfying h(αβ)=h(α)h(β)h(\alpha\beta)=h(\alpha)h(\beta) whenever it makes sense and rh=rr\circ h=r, sh=ss\circ h=s.

We record the following simple fact:

Proposition 4.

Let G,HG,H be two Lie groupoids with same units space MM.

  1. (1)

    Let h:GHh:G\longrightarrow H a CC^{\infty} be a homomorphism over MM. Then the pushforward map h!h_{!} gives rise to a (unital, involutive) algebra homomorphism:

    (37) h!:r,s(G)r,s(H).h_{!}:\mathcal{E}^{\prime}_{r,s}(G)\longrightarrow\mathcal{E}^{\prime}_{r,s}(H).
  2. (2)

    Let h1,h2:GHh_{1},h_{2}:G\longrightarrow H be two CC^{\infty} homomorphisms over MM and set h12:=m(h1h2):G(2)Hh_{12}:=m\circ(h_{1}\otimes h_{2}):G^{(2)}\longrightarrow H. Then for any u,vr,s(G)u,v\in\mathcal{E}^{\prime}_{r,s}(G), we have

    (38) h1!uh2!v=h12!(uv|G(2)).h_{1!}u\star h_{2!}v=h_{12!}(u\otimes v|_{G^{(2)}}).
Proof.

First of all, h!:(G)(H)h_{!}:\mathcal{E}^{\prime}(G)\longrightarrow\mathcal{E}^{\prime}(H) is well defined. Indeed, if ω𝒟(H)\omega\in\mathcal{D}(H) and uCc(G)u\in C^{\infty}_{c}(G), then ω(h(γ))ΩM,r(h(γ))1/2ΩM,s(h(γ))1/2=ΩM,r(γ)1/2ΩM,s(γ)1/2\omega(h(\gamma))\in\Omega^{1/2}_{M,r(h(\gamma))}\otimes\Omega^{1/2}_{M,s(h(\gamma))}=\Omega^{1/2}_{M,r(\gamma)}\otimes\Omega^{1/2}_{M,s(\gamma)} and therefore:

h!u,ω:=u,ωh=Gu(γ)ω(h(γ))\langle h_{!}u,\omega\rangle:=\langle u,\omega\circ h\rangle=\int_{G}u(\gamma)\omega(h(\gamma))

is canonically defined. The algebraic remaining assertions come from the identities: m(hh)=hmm\circ(h\otimes h)=h\circ m on G(2)G^{(2)}, hι=ιhh\iota=\iota h on GG, from the functoriality of pushforwards: f!g!=(fg)!f_{!}g_{!}=(fg)_{!}, and from the definition of the convolution product of distributions: uv=m!(uv|G(2))u\star v=m_{!}(u\otimes v|_{G^{(2)}}). ∎

The goal now is to export to Lie groupoids (with compact unit spaces) a classic result by Dixmier and Malliavin about Lie groups [10]. This will be the main technical tool used to embed reduced CC^{*}-algebras into distributions.

Theorem 5.

Let VV be an open neighborhood of MM into GG and ω𝒟(G)\omega\in\mathcal{D}(G). Then ω\omega is a finite sum of elements:

(39) ξχ\xi\star\chi

where ξCc(G)\xi\in C^{\infty}_{c}(G), supp(ξ)V{\mathrm{supp}(\xi)}\subset V and χ𝒟(G)\chi\in\mathcal{D}(G), supp(χ)supp(ω){\mathrm{supp}(\chi)}\subset{\mathrm{supp}(\omega)}. The result is still valid with the factors flipped in the convolution above.

We adapt the proof of [10, Theorem 3.1] to groupoids. Firstly, [10, Lemma 2.5 and Remark 2.6] gives rise to:

Lemma 6.

Let XΓ(AG)X\in\Gamma(AG) and φ\varphi, ψ\psi be the associated actions of {\mathbb{R}} on GG and MM. Let ε>0\varepsilon>0. For any test function ω𝒟(G)\omega\in\mathcal{D}(G), there exists a1,b1Cc(]ε,ε[)r,s(ψM)a_{1},b_{1}\in C^{\infty}_{c}(]-\varepsilon,\varepsilon[)\subset\mathcal{E}^{\prime}_{r,s}({\mathbb{R}}\ltimes_{\psi}M) and ω1𝒟(G)\omega_{1}\in\mathcal{D}(G) with supp(ω1)supp(ω){\mathrm{supp}(\omega_{1})}\subset{\mathrm{supp}(\omega)} such that

(40) ω=φ!a1ω1+φ!b1ω.\omega=\varphi_{!}a_{1}\star\omega_{1}+\varphi_{!}b_{1}\star\omega.
Proof of the Lemma.

First of all, we pick up a sequence (pj)(p_{j}) of semi-norms characterizing the topology of 𝒟(supp(ω))\mathcal{D}({\mathrm{supp}(\omega)}), and set βk=k2inf{(pj(DX2iω)+1)1;i,jk}\beta_{k}=k^{-2}\inf\{(p_{j}(D_{X}^{2i}\omega)+1)^{-1}\ ;\ i,j\leq k\}. Then the series (1)kαkDX2kω\sum(-1)^{k}\alpha_{k}D_{X}^{2k}\omega converges in 𝒟(G)\mathcal{D}(G) for any sequence 0αkβk0\leq\alpha_{k}\leq\beta_{k}. Next, we choose by [10, Lemma 2.5 and Remark 2.6], two functions a1,b1Cc(]ε,ε[)a_{1},b_{1}\in C^{\infty}_{c}(]-\varepsilon,\varepsilon[) and a sequence 0αkβk0\leq\alpha_{k}\leq\beta_{k} such that

(41) δ=a1k=0(1)kαkδ(2k)+b1=k=0(1)kαka1(2k)+b1 in ()r,s(ψM)\delta=a_{1}\star\sum_{k=0}^{\infty}(-1)^{k}\alpha_{k}\delta^{(2k)}+b_{1}=\sum_{k=0}^{\infty}(-1)^{k}\alpha_{k}a_{1}^{(2k)}+b_{1}\text{ in }\mathcal{E}^{\prime}({\mathbb{R}})\subset\mathcal{E}^{\prime}_{r,s}({\mathbb{R}}\ltimes_{\psi}M)

Now (1) of Proposition 4 gives the identity (40), with ω1=k=0(1)kαkDX2kω\omega_{1}=\sum_{k=0}^{\infty}(-1)^{k}\alpha_{k}D_{X}^{2k}\omega. ∎

Proof of the theorem.

Let X1,,XX_{1},\ldots,X_{\ell} be a family generating the C(M)C^{\infty}(M)-module Γ(AG)\Gamma(AG), and

φi:Gi:=ψiMG\varphi_{i}:G_{i}:={\mathbb{R}}\ltimes_{\psi_{i}}M\to G

be the associated homomorphisms. Applying the lemma to ω\omega with φ=φ1\varphi=\varphi_{1}, we get

(42) ω=λ1ω1+μ1ω in r,s(G),\omega=\lambda_{1}\star\omega_{1}+\mu_{1}\star\omega\text{ in }\mathcal{E}^{\prime}_{r,s}(G),

with λ1,μ1\lambda_{1},\mu_{1} in φ1!(Cc(]ε,ε[))\varphi_{1!}(C^{\infty}_{c}(]-\varepsilon,\varepsilon[)). Applying the lemma to ω\omega and ω1\omega_{1} with φ2\varphi_{2} we get, with intuitive notation:

(43) ω1=λ2,1ω2,1+μ2,1ω1;ω=λ2ω2+μ2ω in r,s(G).\omega_{1}=\lambda_{2,1}\star\omega_{2,1}+\mu_{2,1}\star\omega_{1}\quad;\quad\omega=\lambda_{2}\star\omega_{2}+\mu_{2}\star\omega\text{ in }\mathcal{E}^{\prime}_{r,s}(G).

Inserting (43) into (42), we get:

(44) ω=λ1λ2,1ω2,1+λ1μ2,1ω1+μ1μ2ω2+μ1μ2ω in r,s(G),\omega=\lambda_{1}\star\lambda_{2,1}\star\omega_{2,1}+\lambda_{1}\star\mu_{2,1}\star\omega_{1}+\mu_{1}\star\mu_{2}\star\omega_{2}+\mu_{1}\star\mu_{2}\star\omega\text{ in }\mathcal{E}^{\prime}_{r,s}(G),

where all the λj,μj\lambda_{j\bullet},\mu_{j\bullet} are in the range of Cc(]ε,ε[)C^{\infty}_{c}(]-\varepsilon,\varepsilon[) by φj!\varphi_{j!}, j=1,2j=1,2, and all ω\omega_{\bullet} are test functions with support in supp(ω){\mathrm{supp}(\omega)}.

Repeating the argument with φ3!\varphi_{3!}, …, φ!\varphi_{\ell!} we get that ω\omega is equal to a sum of 22^{\ell} distributions of the form:

(45) ξ1ξ2ξχ\xi_{1}\star\xi_{2}\star\cdots\star\xi_{\ell}\star\chi

where ξj=φj!(kj)𝒟(G)\xi_{j}=\varphi_{j!}(k_{j})\in\mathcal{D}^{\prime}(G) for some kjCc(]ε,ε[)k_{j}\in C^{\infty}_{c}(]-\varepsilon,\varepsilon[) and χ𝒟(G)\chi\in\mathcal{D}(G) with supp(χ)supp(ω){\mathrm{supp}(\chi)}\subset{\mathrm{supp}(\omega)}. Setting as in Proposition (4):

(46) φ=φ1:×MG1×𝑀×𝑀GG\varphi=\varphi_{1\cdots\ell}:{\mathbb{R}}^{\ell}\times M\simeq G_{1}\underset{M}{\times}\cdots\underset{M}{\times}G_{\ell}\longrightarrow G

and after an obvious induction, we get

(47) ξ1ξ2ξ=φ!k with k=k1kCc(]ε,ε[).\xi_{1}\star\xi_{2}\star\cdots\star\xi_{\ell}=\varphi_{!}k\quad\text{ with }\quad k=k_{1}\otimes\cdots\otimes k_{\ell}\in C^{\infty}_{c}(]-\varepsilon,\varepsilon[^{\ell}).

Since tjφ(t,x)|t=0=Xj(x)\partial_{t_{j}}\varphi(t,x)|_{t=0}=X_{j}(x) and φ(0,x)=x\varphi(0,x)=x for any 1j1\leq j\leq\ell and xMx\in M, we get that φ\varphi is a submersion on ]ε,ε[×M]-\varepsilon,\varepsilon[^{\ell}\times M if ε>0\varepsilon>0 is small enough. Since the push forward of a CC^{\infty} distribution by a submersion is CC^{\infty}, we get that φ!k\varphi_{!}k:

(48) η𝒟(G),φ!k,ω=×Mk(t)η(φ(t,x))𝑑t𝑑x,\forall\eta\in\mathcal{D}(G),\quad\langle\varphi_{!}k,\omega\rangle=\int_{{\mathbb{R}}^{\ell}\times M}k(t)\eta(\varphi(t,x))dtdx,

is CC^{\infty} and supported in φ(]ε,ε[×M)\varphi(]-\varepsilon,\varepsilon[^{\ell}\times M). Taking ε>0\varepsilon>0 small enough ensures that this last set is contained in VV. ∎

5. Embedding Cr(G)C^{*}_{r}(G) into 𝒟(G)\mathcal{D}^{\prime}(G) and regularizing operators

From now on and in the remaining parts of this article, we fix a compactly supported, first order elliptic pseudodifferential GG-operator PΨG,c1P\in\Psi^{1}_{G,c} and we denote by Cr(G)C^{*}_{r}(G) the reduced CC^{*}-algebra of GG.

Theorem 7.

There is a continuous embedding:

(49) Cr(G)𝒟(G)C^{*}_{r}(G)\hookrightarrow\mathcal{D}^{\prime}(G)

that extends the pairing:

(50) uCc(G),ω𝒟(G),u,ω=Mιuω\forall u\in C^{\infty}_{c}(G),\ \forall\omega\in\mathcal{D}(G),\quad\langle u\ ,\ \omega\rangle=\int_{M}\iota^{*}u\star\omega
Proof.

Let v,ξCc(G)v,\xi\in C^{\infty}_{c}(G) and χ𝒟(G)\chi\in\mathcal{D}(G). We have:

v,ξχ\displaystyle\langle v,\xi\star\chi\rangle =ιξv,χ\displaystyle=\langle\iota^{*}\xi\star v,\chi\rangle
=ιvξ,ιχ (trace property)\displaystyle=\langle\iota^{*}v\star\xi,\iota^{*}\chi\rangle\text{ (trace property)}
(51) =Gιvξ(α)ιχ(α).\displaystyle=\int_{G}\iota^{*}v\star\xi(\alpha)\iota^{*}\chi(\alpha).

Let μ\mu and μ0\mu_{0} be positive sections of, respectively, the degree 11 densities bundles of AGAG and TMTM. We define μrC(G,Ωr1)\mu_{r}\in C^{\infty}(G,\Omega^{1}_{r}) and μs,0C(G,sΩM1)\mu_{s,0}\in C^{\infty}(G,s^{*}\Omega^{1}_{M}) by

(52) μr(γ)=μ(s(γ)) and μs,0(γ)=μ0(s(γ)).\mu_{r}(\gamma)=\mu(s(\gamma))\quad\text{ and }\quad\mu_{s,0}(\gamma)=\mu_{0}(s(\gamma)).

We observe:

(ιvξ).μr1/2=ιvξCc(G,Ωs1/2) with ξ=ξμr1/2Cc(G,Ωs1/2)(\iota^{*}v\star\xi).\mu^{-1/2}_{r}=\iota^{*}v\star\xi^{\prime}\in C^{\infty}_{c}(G,\Omega^{1/2}_{s})\text{ with }\xi^{\prime}=\xi\mu^{-1/2}_{r}\in C^{\infty}_{c}(G,\Omega^{1/2}_{s})

and

χ=χ.μr1/2.μs,01Cc(G,Ωs1/2).\chi^{\prime}=\chi.\mu^{1/2}_{r}.\mu^{-1}_{s,0}\in C^{\infty}_{c}(G,\Omega^{1/2}_{s}).

This allows us to write

(53) v,ξχ=Gιvξ(α)ιχ(α)\displaystyle\langle v,\xi\star\chi\rangle=\int_{G}\iota^{*}v\star\xi(\alpha)\iota^{*}\chi(\alpha) =M(Gxιvξ(α)ιχ(α))𝑑μ0\displaystyle=\int_{M}\Big{(}\int_{G_{x}}\iota^{*}v\star\xi^{\prime}(\alpha)\iota^{*}\chi^{\prime}(\alpha)\Big{)}d\mu_{0}

and to use the Cauchy Schwarz inequalities for the Hilbert spaces (L2(Gx,ΩGx1/2),x)(L^{2}(G_{x},\Omega^{1/2}_{G_{x}}),\|\cdot\|_{x}) in the following computations:

|v,ξχ|\displaystyle|\langle v,\xi\star\chi\rangle| M𝑑μ0.supxM|Gxιvξ(α)ιχ(α)|\displaystyle\leq\int_{M}d\mu_{0}.\sup_{x\in M}|\int_{G_{x}}\iota^{*}v\star\xi^{\prime}(\alpha)\iota^{*}\chi^{\prime}(\alpha)|
cMsupxMιvξxιχx\displaystyle\leq c_{M}\sup_{x\in M}\|\iota^{*}v\star\xi^{\prime}\|_{x}\|\iota^{*}\chi^{\prime}\|_{x}
cMιvCr(G)ξLs2(G)ιχLs2(G)\displaystyle\leq c_{M}\|\iota^{*}v\|_{C^{*}_{r}(G)}\|\xi^{\prime}\|_{L^{2}_{s}(G)}\|\iota^{*}\chi^{\prime}\|_{L^{2}_{s}(G)}
(54) cvCr(G)\displaystyle\leq c\|v\|_{C^{*}_{r}(G)}

Now let ω𝒟(G)\omega\in\mathcal{D}(G) and pick up a weak factorisation ω=jξjχj\omega=\sum_{j}\xi_{j}\star\chi_{j}. Let uCr(G)u\in C^{*}_{r}(G) and choose a sequence (uk)(u_{k}) with ukCc(G)u_{k}\in C^{\infty}_{c}(G) and ukuu_{k}\to u in Cr(G)C^{*}_{r}(G). Using the previous estimates, we see that the sequence uk,ω\langle u_{k},\omega\rangle\in{\mathbb{C}} satisfies the Cauchy criterium and thus converges. Setting u(ω)=limk+uk,ωu(\omega)=\lim_{k\to+\infty}\langle u_{k},\omega\rangle with get that u𝒟(G)u\in\mathcal{D}^{\prime}(G) and that ukuu_{k}\to u in 𝒟(G)\mathcal{D}^{\prime}(G). ∎

We now give some complements to the properties of the regularizing operators:

(55) ΨG\displaystyle\Psi^{-\infty}_{G} :={R(Cr(G));R(Hs,Ht) for all s,t}\displaystyle:=\{R\in\mathcal{L}(C^{*}_{r}(G))\ ;\ R\in\mathcal{L}(H^{s},H^{t})\text{ for all }s,t\in{\mathbb{N}}\}
(56) ={R(Cr(G));P1RP2(Cr(G)) for all PjΨG,csj,sj,j=1,2}.\displaystyle=\{R\in\mathcal{L}(C^{*}_{r}(G))\ ;\ P_{1}RP_{2}\in\mathcal{L}(C^{*}_{r}(G))\text{ for all }P_{j}\in\Psi^{s_{j}}_{G,c},\ s_{j}\in{\mathbb{N}},\ j=1,2\}.

introduced exaclty in this form in [32] and in an equivalent form in [16]. In both references, this ideal of the CC^{*}-closure of ΨG,c\Psi_{G,c} is proved to be stable under holomorphic functional calculus. Here HsH^{s} denotes the scale of intrinsic Sobolev Cr(G)C^{*}_{r}(G)-modules.

Proposition 8.

Operators in ΨG\Psi^{-\infty}_{G} are exaclty convolution operators by elements of \mathcal{H}^{\infty}. In other words, as subsets of the multipliers algebra (Cr(G))\mathcal{M}(C^{*}_{r}(G)), these sets coincide:

ΨG=(Cr(G)).\Psi^{-\infty}_{G}=\mathcal{H}^{\infty}\subset\mathcal{M}(C^{*}_{r}(G)).
Proof.

We know that ΨG𝒦(Cr(G))=Cr(G)\Psi^{-\infty}_{G}\subset\mathcal{K}(C^{*}_{r}(G))=C^{*}_{r}(G). Let TΨGT\in\Psi^{-\infty}_{G}. For any kk\in{\mathbb{N}}, we have:

(1+P2)kT=SkCr(G) and T(1+P2)k=SkCr(G)(1+P^{2})^{k}T=S_{k}\in C^{*}_{r}(G)\text{ and }T(1+P^{2})^{k}=S^{\prime}_{k}\in C^{*}_{r}(G)

Then T=(1+P2)kSk=Sk(1+P2)kH2k(H2k)T=(1+P^{2})^{-k}S_{k}=S^{\prime}_{k}(1+P^{2})^{-k}\in H^{2k}\cap(H^{2k})^{*} for any kk, which proves the first inclusion. The second one is obvious. ∎

All the previous statements hold true for the maximal CC^{*}-algebra of GG but we stay in the framework of the reduced CC^{*}-algebra, because the embedding Cr(G)𝒟(G)C^{*}_{r}(G)\hookrightarrow\mathcal{D}^{\prime}(G) and the regular representation allow us to precise in what extent elements of ΨG=\Psi^{-\infty}_{G}=\mathcal{H}^{\infty} are regularizing. For that purpose, we let ΨG\Psi^{*}_{G} act not on the scale of intrinsic Sobolev modules HsH^{s}, but on their representation via the left regular representation. These C(M)C(M)-modules are concretely given as follows, for kk\in{\mathbb{Z}}:

(57) Hsk=Cc(G,Ωs1/2)¯|k,s,ω|ηk,s=(1+P2)kω|ηsC(M).H^{k}_{s}=\overline{C^{\infty}_{c}(G,\Omega^{1/2}_{s})}^{\langle\cdot|\cdot\rangle_{k,s}},\quad\langle\omega\ |\ \eta\rangle_{k,s}=\langle(1+P^{2})^{k}\star\omega\ |\ \eta\rangle_{s}\in C(M).
Lemma 9.

We have:

(58) Hs:=kHskCs,0(G,Ωs1/2) and Hs:=kHsks,0(G,Ωs1/2).H^{\infty}_{s}:=\bigcap_{k\in{\mathbb{Z}}}H^{k}_{s}\subset C^{\infty,0}_{s}(G,\Omega^{1/2}_{s})\text{ and }H^{-\infty}_{s}:=\bigcup_{k\in{\mathbb{Z}}}H^{k}_{s}\supset\mathcal{E}^{\prime}_{s,0}(G,\Omega^{1/2}_{s}).
Proof of the lemma.

Let ωHs\omega\in H^{\infty}_{s}. Since pointwise multiplication operators by compactly supported CC^{\infty} functions are in (Hsk)\mathscr{L}(H^{k}_{s}) for any kk, we can assume that ω\omega is compactly supported in the domain UU of a local trivialization κ:Un(1)×n(0)\kappa:U\to{\mathbb{R}}^{n^{(1)}}\times{\mathbb{R}}^{n^{(0)}}, κ(x)=(x,x′′)\kappa(x)=(x^{\prime},x^{\prime\prime}) of the submersion ss. By assumption, we have

(59) k0,ΔGkωCc0(n(0),L2(n(1))).\forall k\geq 0,\quad\Delta^{k}_{G}\omega\in C^{0}_{c}({\mathbb{R}}^{n^{(0)}},L^{2}({\mathbb{R}}^{n^{(1)}})).

Here ΔG=ddDiff2(G)\Delta_{G}=d^{*}d\in\mathop{\mathrm{Diff}}\nolimits^{2}(G) is the Laplacian associated with a given euclidean structure on AGAG. The ellipticity of each term of the CC^{\infty} family (ΔG,x′′)x′′n(0)(\Delta_{G,x^{\prime\prime}})_{x^{\prime\prime}\in{\mathbb{R}}^{n^{(0)}}} and the compactness of supp(ω){\mathrm{supp}(\omega)} imply using usual Garding inequality that ωC0(n(0),H2k(n(1)))\omega\in C^{0}({\mathbb{R}}^{n^{(0)}},H^{2k}({\mathbb{R}}^{n^{(1)}})) for any kk, where HH^{*} denotes here the usual Sobolev spaces of euclidean spaces. We then conclude that ωC0(n0,C(n(1)))=Cpr2,0(n(1)×n(0))\omega\in C^{0}({\mathbb{R}}^{n_{0}},C^{\infty}({\mathbb{R}}^{n^{(1)}}))=C^{\infty,0}_{\mathop{\mathrm{pr}}\nolimits_{2}}({\mathbb{R}}^{n^{(1)}}\times{\mathbb{R}}^{n^{(0)}}). This proves HsCs,0(G,Ωs1/2)H^{\infty}_{s}\subset C^{\infty,0}_{s}(G,\Omega^{1/2}_{s}).
Let us,0(G,Ωs1/2)u\in\mathcal{E}^{\prime}_{s,0}(G,\Omega^{1/2}_{s}). The result [12, Theorem 4.4.7] extends immediately to continuous family of distributions so there exists kk\in{\mathbb{N}} and finite collections: uICc(G,Ωs1/2)u_{I}\in C_{c}(G,\Omega^{1/2}_{s}), DIDiffGkD_{I}\in\mathop{\mathrm{Diff}}\nolimits^{k}_{G} such that

(60) u=IDIuI.u=\sum_{I}D_{I}u_{I}.

Since Cc(G,Ωs1/2)Ls2(G)C_{c}(G,\Omega^{1/2}_{s})\subset L^{2}_{s}(G) and (1+P2)k/2DI(Ls2(G))(1+P^{2})^{-k/2}D_{I}\in\mathcal{L}(L^{2}_{s}(G)), we then conclude that uHsku\in H^{-k}_{s}. ∎

We recall [24, 19] that for any xMx\in M, the orbit 𝒪=r(s1({x}))M\mathcal{O}=r(s^{-1}(\{x\}))\subset M is an immersed submanifold, the map r:Gx𝒪r:G_{x}\longrightarrow\mathcal{O} is a submersion (actually a GxxG_{x}^{x} principal bundle) and that G𝒪𝒪G_{\mathcal{O}}\rightrightarrows\mathcal{O} is an immersed subgroupoid. We set:

(61) Corb,0(G,E)={uC(G,E);xM,DDiff(G),DuC(G𝒪,E),𝒪=r(s1({x})}.C^{\infty,0}_{\mathop{\mathrm{orb}}\nolimits}(G,E)=\{u\in C(G,E)\ ;\ \forall x\in M,\forall D\in\mathop{\mathrm{Diff}}\nolimits(G),Du\in C^{\infty}(G_{\mathcal{O}},E),\ \mathcal{O}=r(s^{-1}(\{x\})\}.
Theorem 10.

The following inclusions hold true:

(62) Corb,0(G)Cc(G)Corb,0(G)Cr(G).C^{\infty,0}_{\mathop{\mathrm{orb}}\nolimits}(G)\cap C_{c}(G)\subset\mathcal{H}^{\infty}\subset C^{\infty,0}_{\mathop{\mathrm{orb}}\nolimits}(G)\cap C^{*}_{r}(G).

In particular, since \mathcal{H}^{\infty} is an ideal in Cr(G)C^{*}_{r}(G):

Corollary 11.

Any hh\in\mathcal{H}^{\infty} provides continuous operators :

h:Cr(G)Corb,0(G)Cr(G) and h:Cr(G)Corb,0(G)Cr(G).h\star\cdot:C^{*}_{r}(G)\longrightarrow C^{\infty,0}_{\mathop{\mathrm{orb}}\nolimits}(G)\cap C^{*}_{r}(G)\text{ and }\cdot\star\,h:C^{*}_{r}(G)\longrightarrow C^{\infty,0}_{\mathop{\mathrm{orb}}\nolimits}(G)\cap C^{*}_{r}(G).
Proof of the theorem.

Let uu\in\mathcal{H}^{\infty}. By [32] and the left regular representation, uu maps HskHskH^{-k}_{s}\to H^{k}_{s} continuously for any kk\in{\mathbb{N}}. Therefore, the previous lemma implies that uu maps s,0(G,Ωs1/2)Cs,0(G,Ωs1/2)\mathcal{E}^{\prime}_{s,0}(G,\Omega^{1/2}_{s})\to C^{\infty,0}_{s}(G,\Omega^{1/2}_{s}) continuously. In particular for every xMx\in M, the distribution κx(γ1,γ2)=u(γ1γ21)𝒟(Gx×Gx,Ωx1/2)\kappa_{x}(\gamma_{1},\gamma_{2})=u(\gamma_{1}\gamma_{2}^{-1})\in\mathcal{D}^{\prime}(G_{x}\times G_{x},\Omega^{1/2}_{x}) extends to a continuous map:

(63) κx:(Gx,ΩGx1/2)C(Gx,ΩGx1/2)\kappa_{x}:\mathcal{E}^{\prime}(G_{x},\Omega^{1/2}_{G_{x}})\to C^{\infty}(G_{x},\Omega^{1/2}_{G_{x}})

which implies that κxC(Gx×Gx,Ωx1/2)\kappa_{x}\in C^{\infty}(G_{x}\times G_{x},\Omega^{1/2}_{x}) for fixed xx. Next, consider xMx\in M, 𝒪=r(s1({x}))\mathcal{O}=r(s^{-1}(\{x\})) the orbit of xx in MM and fix (γ1,γ2)Gx×GxG×𝑠G(\gamma_{1},\gamma_{2})\in G_{x}\times G_{x}\subset G\underset{s}{\times}G. We denote by π:G×𝑠GM\pi:G\underset{s}{\times}G\longrightarrow M the obvious submersion. Since r:Gx𝒪r:G_{x}\longrightarrow\mathcal{O} is a submersion, there exists a CC^{\infty} local section η:UyηyGx\eta:U\in y\mapsto\eta_{y}\in G_{x} of rr such that ηx=x\eta_{x}=x, defined on some open neighborhood UU of xx into OO. Then V=π𝒪1(U)={(η1,η2);s(η1)=s(η2)U}V=\pi_{\mathcal{O}}^{-1}(U)=\{(\eta_{1},\eta_{2})\ ;\ s(\eta_{1})=s(\eta_{2})\in U\} is an open neighborhood of (γ1,γ2)(\gamma_{1},\gamma_{2}) into G𝒪×𝑠G𝒪G_{\mathcal{O}}\underset{s}{\times}G_{\mathcal{O}} and we have:

(64) (η1,η2)V,κy(η1,η2)=κx(Rηyη1,Rηyη2)\forall(\eta_{1},\eta_{2})\in V,\quad\kappa_{y}(\eta_{1},\eta_{2})=\kappa_{x}(R_{\eta_{y}}\eta_{1},R_{\eta_{y}}\eta_{2})

which proves that κ\kappa is CC^{\infty} on G𝒪×𝑠G𝒪G_{\mathcal{O}}\underset{s}{\times}G_{\mathcal{O}}, and thus that uu is CC^{\infty} on G𝒪G_{\mathcal{O}}. It is clear that Corb,0(G)Cc(G)C^{\infty,0}_{\mathop{\mathrm{orb}}\nolimits}(G)\cap C_{c}(G) is contained in Cr(G)C^{*}_{r}(G) and is invariant under the left and right convolution by PP. The inclusion Corb,0(G)Cc(G)C^{\infty,0}_{\mathop{\mathrm{orb}}\nolimits}(G)\cap C_{c}(G)\subset\mathcal{H}^{\infty} follows. ∎

Summarizing the content above, we have proved that regularizing operators are actually convolution operators by distributions on GG lying in the class \mathcal{H}^{\infty}, the latter class being included in the class of functions that are continuous on GG and infinitely differentiable over any orbit, and thus in particular along the fibers of ss and rr. Closely related results were obtained in [16] under the assumption of bounded geometry for GG. In the following sections, we are going to prove that E(t)=eitPE(t)=e^{itP} is a family ×G{\mathbb{R}}\times G-FIO [18], modulo such regularizing operators.

6. Principal and Subprincipal symbols of GG-PDOs

As a conormal distribution, any element of ΨGm=Im+(n(1)n(0))/4(G,M;Ω1/2)\Psi^{m}_{G}=I^{m+(n^{(1)}-n^{(0)})/4}(G,M;\Omega^{1/2}) has a principal symbol [12, Theorem 18.2.11] in:

(65) S[m+(n(1)n(0))/4+n/4](AG,ΩAG1/2Ω^1/2Ω^G1/2).S^{[m+(n^{(1)}-n^{(0)})/4+n/4]}(A^{*}G,\Omega^{1/2}_{A^{*}G}\otimes\hat{\Omega}^{1/2}\otimes\hat{\Omega}^{-1/2}_{G}).

The density bundle above is canonically trivial:

(66) ΩAG1/2Ω^1/2Ω^G1/2=Ω1/2(TMAG)Ω1(AG)Ω1/2(TMAG)M×,\ \Omega^{1/2}_{A^{*}G}\otimes\hat{\Omega}^{1/2}\otimes\hat{\Omega}^{-1/2}_{G}=\Omega^{1/2}(TM\oplus A^{*}G)\oplus\Omega^{1}(AG)\otimes\Omega^{-1/2}(TM\oplus AG)\simeq M\times{\mathbb{C}},

and since half densities on AGA^{*}G contribute with a value of n(1)/2n^{(1)}/2 to the degree of symbols, the simplification above lowers the degree by the same value. In conclusion the principal symbol map is a well defined map:

(67) σ0:ΨGmS[m](AG).\sigma_{0}:\Psi^{m}_{G}\longrightarrow S^{[m]}(A^{*}G).

Alternatively, given PΨGmP\in\Psi^{m}_{G}, one may consider the family P~=(Px)xM\widetilde{P}=(P_{x})_{x\in M}, PxΨm(Gx,ΩGx1/2)P_{x}\in\Psi^{m}(G_{x},\Omega^{1/2}_{G_{x}}) associated with PP by the isomorphism (35) and collect the family of principal symbols σ(Px)S[m](TGx)\sigma(P_{x})\in S^{[m]}(T^{*}G_{x}) into the element σ(P)S[m](TsG)\sigma(P)\in S^{[m]}(T^{*}_{s}G), where TsG=(kerds)T^{*}_{s}G=(\ker ds)^{*}, defined by:

σ(P)(γ,ξ)=σ(Ps(γ))(γ,ξ).\sigma(P)(\gamma,\xi)=\sigma(P_{s(\gamma)})(\gamma,\xi).

In this point of view, the principal symbol is a map :

(68) σ:ΨGmS[m](TsG).\sigma:\Psi^{m}_{G}\longrightarrow S^{[m]}(T_{s}^{*}G).

Both notions are related by:

Proposition 12.

With the notation above, the following identity holds true:

(69) σ=σ0rΓ.\sigma=\sigma_{0}\circ r_{{}_{\Gamma}}.
Remark 13.

Strictly speaking, the target map rΓr_{{}_{\Gamma}} is defined on TGT^{*}G. It is by construction the composition of the natural restriction map TGTsGT^{*}G\to T^{*}_{s}G with the natural map TsGAGT^{*}_{s}G\to A^{*}G. It is understood in the Proposition above that rΓr_{{}_{\Gamma}} means the latter.

Proof.

Let PΨGmP\in\Psi^{m}_{G}. Without loss of generality, we can assume that PP is supported in a local chart UU around some point of MM and satisfying:

  1. -

    the local coordinates trivialize the source map, that is γ=(x′′,x)\gamma=(x^{\prime\prime},x^{\prime}) with s(γ)=x′′s(\gamma)=x^{\prime\prime} on UU,

  2. -

    the domain UU is invariant for the inversion map : U1=UU^{-1}=U.

We then pick up a positive one density μ\mu on AGAG such that:

xUG(0),μ(x)=|dx|,\forall x\in U\cap G^{(0)},\quad\mu(x)=|dx^{\prime}|,

and define μsC(G,Ωs1/2)\mu_{s}\in C^{\infty}(G,\Omega^{1/2}_{s}), μrC(G,Ωr1/2)\mu_{r}\in C^{\infty}(G,\Omega^{1/2}_{r}) by:

μs(γ)=μ(r(γ)) and μr(γ)=μ(s(γ)).\mu_{s}(\gamma)=\mu(r(\gamma))\text{ and }\mu_{r}(\gamma)=\mu(s(\gamma)).

We can set on UU:

(70) P(γ)=𝖯(γ).μs1/2μr1/2P(\gamma)=\mathsf{P}(\gamma).\mu^{1/2}_{s}\mu^{1/2}_{r}

where 𝖯\mathsf{P} is a scalar oscillatory integral conveniently given in the following form:

(71) 𝖯(γ)=eiγ1.ξp0(r(γ),ξ)𝑑ξ.\mathsf{P}(\gamma)=\int e^{-i\gamma^{-1}.\xi^{\prime}}p_{0}(r(\gamma),\xi^{\prime})d\xi^{\prime}.

Let us describe the various ingredients of this formula. First, p0Sm(AG)p_{0}\in S^{m}(A^{*}G) is a (classical) symbol, and the integral (in the distribution sense) is performed with respect to (0,ξ)Ar(γ)GTr(γ)G(0,\xi^{\prime})\in A_{r(\gamma)}^{*}G\subset T_{r(\gamma)}^{*}G. Secondly, it is understood that γ1=ι(γ)\gamma^{-1}=\iota(\gamma) stands for the nn-tuple of coordinates of the inverse of γ\gamma in GG, and then γ1.ξ\gamma^{-1}.\xi^{\prime} stands for its scalar product with (0,ξ)(0,\xi^{\prime}) in n{\mathbb{R}}^{n}. We could use the inverse of an exponential map to give an invariant meaning to γ1.ξ\gamma^{-1}.\xi with ξAr(γ)G\xi\in A_{r(\gamma)}^{*}G, but since we already work in local coordinates, this is pointless. Finally, we read from (71) that:

(72) σ0(P)=p0modSm1(AG),\sigma_{0}(P)=p_{0}\mod S^{m-1}(A^{*}G),

and since the symbols used here are classical, we may identify σ0(P)\sigma_{0}(P) with the leading homogeneous part p00p_{0}^{0} of p0p_{0}. Now let uCc(G)u\in C^{\infty}_{c}(G) with support in a local chart VV of GG, and set:

(73) u(γ)=𝗎(γ).μs1/2μr1/2u(\gamma)=\mathsf{u}(\gamma).\mu^{1/2}_{s}\mu^{1/2}_{r}

with 𝗎Cc(V,)\mathsf{u}\in C^{\infty}_{c}(V,{\mathbb{C}}). To express PuP{u} in local coordinates in terms of 𝖯\mathsf{P} and 𝗎\mathsf{u}, we need to recall the necessary identifications of densities allowing the convolution product:

(74) P(u)(γ)=αGs(γ)P(γα1)u(α).P({u})(\gamma)=\int_{\alpha\in G_{s(\gamma)}}P(\gamma\alpha^{-1}){u}(\alpha).

For that purpose, note that for any γ,α\gamma,\alpha with same source point:

μs(γα1)=μ(r(γ))=μs(γ),μr(γα1)=μ(r(α))=μs(α),μr(α)=μr(γ).\mu_{s}(\gamma\alpha^{-1})=\mu(r(\gamma))=\mu_{s}(\gamma),\ \mu_{r}(\gamma\alpha^{-1})=\mu(r(\alpha))=\mu_{s}(\alpha),\ \mu_{r}(\alpha)=\mu_{r}(\gamma).

Hence:

P(γα1)u(α)=𝖯(γα1)𝗎(α)μs(α)μs1/2(γ)μr1/2(γ).P(\gamma\alpha^{-1}){u}(\alpha)=\mathsf{P}(\gamma\alpha^{-1})\mathsf{u}(\alpha)\mu_{s}(\alpha)\mu^{1/2}_{s}(\gamma)\mu^{1/2}_{r}(\gamma).

It remains to express μs(α)\mu_{s}(\alpha) in term of a one density on GxVG_{x}\cap V. We also assume that the coordinates fixed on VV trivializes the source map ss:

Vα=(α′′,α) with s(α)=α′′V\ni\alpha=(\alpha^{\prime\prime},\alpha^{\prime})\text{ with }s(\alpha)=\alpha^{\prime\prime}

In the coordinates fixed on UU and VV, we get using (dRα)r(α):(rAG)α(kerds)α(dR_{\alpha})_{r(\alpha)}:(r^{*}AG)_{\alpha}\overset{\simeq}{\longrightarrow}(\ker ds)_{\alpha}:

μs(α)=|dx|=|(dRα)r(α)|1|dα|.\mu_{s}(\alpha)=|dx^{\prime}|=|(dR_{\alpha})_{r(\alpha)}|^{-1}|d\alpha^{\prime}|.

It follows that, setting P~(u)=𝗏μs1/2μr1/2\widetilde{P}({u})=\mathsf{v}\mu^{1/2}_{s}\mu_{r}^{1/2} on WW:

(75) 𝗏(γ)\displaystyle\mathsf{v}(\gamma) =P(γα1)𝗎(α)|(dRα)r(α)|1𝑑α\displaystyle=\int P(\gamma\alpha^{-1})\mathsf{u}(\alpha)|(dR_{\alpha})_{r(\alpha)}|^{-1}d\alpha^{\prime}
(76) =eiαγ1.ξp0(r(γ),ξ)|(dRα)r(α)|1𝗎(α)𝑑α𝑑ξ.\displaystyle=\int e^{-i\alpha\gamma^{-1}.\xi^{\prime}}p_{0}(r(\gamma),\xi^{\prime})|(dR_{\alpha})_{r(\alpha)}|^{-1}\mathsf{u}(\alpha)d\alpha^{\prime}d\xi^{\prime}.

Actually, the action of the induced family of operators PxΨm(Gx,ΩGx1/2)P_{x}\in\Psi^{m}(G_{x},\Omega^{1/2}_{G_{x}}) on half-densities fCc(Gx,ΩGx1/2){f}\in C^{\infty}_{c}(G_{x},\Omega^{1/2}_{G_{x}}) is given by the same formula:

(77)  if f=𝖿μs1/2, then Px(f)=𝗏μs1/2 with 𝗏(γ)=eiαγ1.ξp0(r(γ),ξ)𝖿(α)|(dRα)r(α)|1𝑑α𝑑ξ.\text{ if }{f}=\mathsf{f}\mu_{s}^{1/2},\text{ then }P_{x}({f})=\mathsf{v}\mu_{s}^{1/2}\text{ with }\mathsf{v}(\gamma)=\int e^{-i\alpha\gamma^{-1}.\xi}p_{0}(r(\gamma),\xi)\mathsf{f}(\alpha)|(dR_{\alpha})_{r(\alpha)}|^{-1}d\alpha^{\prime}d\xi.

Let us set

(78) φ(α)=|(dRα)r(α)|1.\varphi(\alpha)=|(dR_{\alpha})_{r(\alpha)}|^{-1}.

Since αγ1\alpha\gamma^{-1} vanishes at α=γGs(γ)\alpha=\gamma\in G_{s(\gamma)}, there exists a linear map

(79) ψ(α,γ):n(1)n(1)\psi(\alpha,\gamma):{\mathbb{R}}^{n^{(1)}}\longrightarrow{\mathbb{R}}^{n^{(1)}}

which is CC^{\infty} in (α,γ)(\alpha,\gamma), bijective for α\alpha in a neighborhood of γ\gamma and satisfies:

(80) αγ1=ψ(α,γ)(αγ).\alpha\gamma^{-1}=\psi(\alpha,\gamma)(\alpha-\gamma).

By construction we have:

(81) ψ(γ,γ)=(dRγ1)γ=(dRγ)r(γ)1.\psi(\gamma,\gamma)=(dR_{\gamma^{-1}})_{\gamma}=(dR_{\gamma})_{r(\gamma)}^{-1}.

Now we work on (77) to find the amplitude and symbol of PxP_{x} in local coordinates on VV:

𝗏(γ)\displaystyle\mathsf{v}(\gamma) =eiαγ,ψt(α,γ)ξp0(r(γ),ξ)𝗎(α)φ(α)𝑑α𝑑ξ\displaystyle=\int e^{-i\langle\alpha-\gamma,{}^{t}\psi(\alpha,\gamma)\xi^{\prime}\rangle}p_{0}(r(\gamma),\xi^{\prime})\mathsf{u}(\alpha)\varphi(\alpha)d\alpha^{\prime}d\xi^{\prime}
=eiγα,ξ(p0(r(γ),ψt(α,γ)1ξ)φ(α)|ψt(α,γ)|1𝗎(α)dαdξ\displaystyle=\int e^{i\langle\gamma-\alpha,\xi^{\prime}\rangle}\big{(}p_{0}(r(\gamma),{}^{t}\psi(\alpha,\gamma)^{-1}\xi^{\prime})\varphi(\alpha)|{}^{t}\psi(\alpha,\gamma)|^{-1}\ \mathsf{u}(\alpha)d\alpha^{\prime}d\xi^{\prime}
=eiγα,ξp~(γ,α,ξ)𝗎(α)𝑑α𝑑ξ\displaystyle=\int e^{i\langle\gamma-\alpha,\xi^{\prime}\rangle}\widetilde{p}(\gamma,\alpha,\xi^{\prime})\mathsf{u}(\alpha)d\alpha^{\prime}d\xi^{\prime}
(82) =eiγα,ξp(γ,ξ)𝗎(α)𝑑α𝑑ξ\displaystyle=\int e^{i\langle\gamma-\alpha,\xi^{\prime}\rangle}p(\gamma,\xi^{\prime})\mathsf{u}(\alpha)d\alpha^{\prime}d\xi^{\prime}

where we have set

(83) p~(γ,α,ξ)=p0(r(γ),ψt(α,γ)1ξ)φ(α)|ψt(α,γ)|1\widetilde{p}(\gamma,\alpha,\xi^{\prime})=p_{0}(r(\gamma),{}^{t}\psi(\alpha,\gamma)^{-1}\xi^{\prime})\varphi(\alpha)|{}^{t}\psi(\alpha,\gamma)|^{-1}

and

(84) p(γ,ξ)=eiDα,Dξp~(γ,α,ξ)|α=γp(\gamma,\xi^{\prime})=e^{i\langle D_{\alpha^{\prime}},D_{\xi^{\prime}}\rangle}\widetilde{p}(\gamma,\alpha,\xi^{\prime})|_{\alpha=\gamma}

which gives the asymptotic expansion:

(85) p(γ,ξ)1k!iDα,Dξkp~(γ,α,ξ)|α=γ.p(\gamma,\xi^{\prime})\sim\sum\frac{1}{k!}\langle iD_{\alpha^{\prime}},D_{\xi^{\prime}}\rangle^{k}\widetilde{p}(\gamma,\alpha,\xi^{\prime})\,|_{\alpha=\gamma}.

Since (r(γ),ψt(γ,γ)1ξ)=(r(γ),(dRγ)r(γ)tξ)=rΓ(γ,ξ)(r(\gamma),{}^{t}\psi(\gamma,\gamma)^{-1}\xi^{\prime})=(r(\gamma),{}^{t}(dR_{\gamma})_{{}_{r(\gamma)}}\xi^{\prime})=r_{{}_{\Gamma}}(\gamma,\xi^{\prime}), the expression of the principal symbol of PxP_{x} over VV is the first term in the sum (85):

(86) σ(Px)(γ,ξ)=p0(rΓ(γ,ξ))modSm1(TsG),\sigma(P_{x})(\gamma,\xi^{\prime})=p_{0}(r_{{}_{\Gamma}}(\gamma,\xi^{\prime}))\mod S^{m-1}(T^{*}_{s}G),

or equivalently using homogeneous expansions: σ(P)=p0=p00rΓ\sigma(P)=p^{0}=p_{0}^{0}\circ r_{{}_{\Gamma}}. ∎

Remark 14.

We will often consider CC^{\infty} functions on TsGT^{*}_{s}G as CC^{\infty} functions on TGT^{*}G, thanks to the convention a(γ,ξ)=a(γ,ξ|Ts(γ)G)a(\gamma,\xi)=a(\gamma,\xi|_{T_{s(\gamma)}G}).

We now turn our attention to the sub-principal symbols. It is not obvious to us how to define the sub-principal symbol for general conormal distributions, but in the case of ΨG=I(G,M,Ω1/2)\Psi^{*}_{G}=I(G,M,\Omega^{1/2}), we may again consider the family of usual sub-principal symbols of the operators PxΨm(Gx,ΩGx1/2)P_{x}\in\Psi^{m}(G_{x},\Omega^{1/2}_{G_{x}}) and set:

(87) (γ,ξ)TsG,σ1s(P)(γ,ξ):=σ1s(Ps(γ))(γ,ξ)S[m1](TsG).(\gamma,\xi)\in T_{s}^{*}G,\quad\sigma^{1s}(P)(\gamma,\xi):=\sigma^{1s}(P_{s(\gamma)})(\gamma,\xi)\in S^{[m-1]}(T_{s}^{*}G).

This gives a well defined map:

(88) σ1s:ΨGmS[m1](TsG),\sigma^{1s}:\Psi^{m}_{G}\longrightarrow S^{[m-1]}(T_{s}^{*}G),

When PP is given by (71), we recall that the sub-principal symbol above is given in terms of the homogeneous expansion of total symbol pp, expressed in the last proof (see formula (83), (84) and (85)), by:

(89) p1s(γ,ξ):=σ1s(P)(γ,ξ)=p1(γ,ξ)i2Dγ,Dξp0(γ,ξ),p^{1s}(\gamma,\xi^{\prime}):=\sigma^{1s}(P)(\gamma,\xi^{\prime})=p^{1}(\gamma,\xi^{\prime})-\frac{i}{2}\langle D_{\gamma^{\prime}},D_{\xi^{\prime}}\rangle p^{0}(\gamma,\xi^{\prime}),

We now consider p0=σ(P)p^{0}=\sigma(P) as a CC^{\infty} homogeneous function on T.G=TGkerrΓT^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G=T^{*}G\setminus\ker r_{{}_{\Gamma}} (see Remark 14) and we denote Hp0Γ(TT.G)H_{p^{0}}\in\Gamma(TT^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G) the hamiltonian vector field of p0p^{0}. We recall that the latter is defined by dp0()=ωG(Hp0,)dp^{0}(\cdot)=\omega_{G}(H_{p^{0}},\cdot), and in local coordinates (γ,ξ)(\gamma,\xi) we get:

Hp0=j=1np0ξjγjp0γjξj.\quad H_{p^{0}}=\sum_{j=1}^{n}\frac{\partial p^{0}}{\partial\xi_{j}}\frac{\partial}{\partial\gamma_{j}}-\frac{\partial p^{0}}{\partial\gamma_{j}}\frac{\partial}{\partial\xi_{j}}.

Now we shall compute the principal symbol of a product PAPA where PP is a GG-PDO and AA a GG-FIO in the situation later encountered in the construction of the parametrix of eitPe^{itP}. To that purpose, we recall that the principal symbol of GG-FIO is a homomorphism [12]:

(90) Im(G,Λ;Ω1/2)S[m+n/4](Λ,MΛΩΛ1/2Ω^1/2Ω^G1/2)I^{m}(G,\Lambda;\Omega^{1/2})\longrightarrow S^{[m+n/4]}(\Lambda,M_{\Lambda}\otimes\Omega^{1/2}_{\Lambda}\otimes\hat{\Omega}^{1/2}\otimes\hat{\Omega}^{-1/2}_{G})

where MΛM_{\Lambda} is the Maslov bundle and E^\hat{E} denotes the pull back of the vector bundle EGE\to G over Λ\Lambda. By [18], we know that there is canonical isomorphism:

(91) Ω^1/2Ω^G1/2ΩrΓ1/2=sΓΩ1/2(ATG)\hat{\Omega}^{1/2}\otimes\hat{\Omega}^{-1/2}_{G}\simeq\Omega^{1/2}_{r_{{}_{\Gamma}}}=s_{{}_{\Gamma}}^{*}\Omega^{1/2}(AT^{*}G)

This isomorphism uses the product and inversion map of GG but their contributions cancel and thus, elements in Ω^1/2Ω^G1/2\hat{\Omega}^{1/2}\otimes\hat{\Omega}^{-1/2}_{G} do define, without any other data, pull back of half densities on the vector bundle ATGAGAT^{*}G\longrightarrow A^{*}G. We thus may consider the principal symbol of Fourier integral GG-operators as a homomorphism:

(92) Im(G,Λ;Ω1/2)S[m+n/4](Λ,MΛΩΛ1/2ΩrΓ1/2)I^{m}(G,\Lambda;\Omega^{1/2})\longrightarrow S^{[m+n/4]}(\Lambda,M_{\Lambda}\otimes\Omega^{1/2}_{\Lambda}\otimes\Omega^{1/2}_{r_{{}_{\Gamma}}})

We recall that for a manifold XX and a vector field VV on XX with flow ϕt\phi_{t}, the Lie derivative of a α\alpha-density aa is the α\alpha-density given by, in local coordinates a=𝖺|dx|αa=\mathsf{a}|dx|^{\alpha}:

(93) V(𝖺|dx|α)=ddtϕt𝖺|dx|α|t=0=(V𝖺+αdiv(V)𝖺)|dx|α.\mathcal{L}_{V}(\mathsf{a}|dx|^{\alpha})=\frac{d}{dt}\phi^{*}_{t}\mathsf{a}|dx|^{\alpha}|_{t=0}=\big{(}V\cdot\mathsf{a}+\alpha\,\mathrm{div}(V)\mathsf{a}\big{)}|dx|^{\alpha}.

This is the same for sections aC(Λ,MΛΩΛ1/2)a\in C^{\infty}(\Lambda,M_{\Lambda}\otimes\Omega^{1/2}_{\Lambda}) and vector fields VΓ(TΛ)V\in\Gamma(T\Lambda). Indeed, the transition functions of MΛM_{\Lambda} are locally constant, so the bundle MΛM_{\Lambda} can be factorized out of (93).

On the other hand, we are mainly interested in Hamiltonian vector fields V=HfV=H_{f} that are also right invariant, which happens if and only if f=f0rΓf=f_{0}\circ r_{{}_{\Gamma}} [6], and such that f|Λ=0f|_{\Lambda}=0, which implies that VV is tangent to Λ\Lambda. For such vector fields, we can extend the Lie derivative above to a map f=Hf\mathcal{L}_{f}=\mathcal{L}_{H_{f}} acting on sections of the line bundle appearing in the symbols space in (92). To do that, consider νrΓ=νsΓC(TG,ΩrΓ1/2)\nu_{r_{{}_{\Gamma}}}=\nu\circ s_{{}_{\Gamma}}\in C^{\infty}(T^{*}G,\Omega^{1/2}_{r_{{}_{\Gamma}}}) with ν\nu a positive density on ATGAT^{*}G. Since by assumption ssΓϕt=sΓs_{s_{{}_{\Gamma}}}\circ\phi_{t}=s_{{}_{\Gamma}}, we get:

f(νrΓ):=ddtϕt(νrΓ)|t=0=ddtν(sΓϕt)|t=0=ddtν1/2(sΓ)|t=0=0.\displaystyle\mathcal{L}_{f}(\nu_{r_{{}_{\Gamma}}}):=\frac{d}{dt}\phi^{*}_{t}(\nu_{r_{{}_{\Gamma}}})|_{t=0}=\frac{d}{dt}\nu(s_{{}_{\Gamma}}\circ\phi_{t})|_{t=0}=\frac{d}{dt}\nu^{1/2}(s_{{}_{\Gamma}})|_{t=0}=0.

Combining the usual action of V|ΛV|_{\Lambda} recalled in (93) with the above trivial one, we obtain that V=HfV=H_{f} acts on C(Λ,MΛΩΛ1/2ΩrΓ1/2)C^{\infty}(\Lambda,M_{\Lambda}\otimes\Omega^{1/2}_{\Lambda}\otimes\Omega^{1/2}_{r_{{}_{\Gamma}}}) by the formula:

(94) f(𝖺μΛ1/2νrΓ1/2)=(Hf𝖺+12div(Hf)𝖺)μΛ1/2νrΓ1/2.\displaystyle\mathcal{L}_{f}(\mathsf{a}\mu_{\Lambda}^{1/2}\nu_{r_{{}_{\Gamma}}}^{1/2})=\big{(}H_{f}\cdot\mathsf{a}+\frac{1}{2}\,\mathrm{div}(H_{f})\mathsf{a}\big{)}\mu_{\Lambda}^{1/2}\nu_{r_{{}_{\Gamma}}}^{1/2}.

In the important particular case where the GG-relation Λ\Lambda is a bissection, that is, when sΓ,rΓs_{{}_{\Gamma}},r_{{}_{\Gamma}} are diffeomorphisms from Λ\Lambda to open subsets of AGA^{*}G, then

(95) TΛkerdsΓ|Λ=TΛTG.T\Lambda\oplus\ker ds_{{}_{\Gamma}}|_{\Lambda}=T_{\Lambda}T^{*}G.

Since v=Hf0rΓv=H_{f_{0}\circ r_{{}_{\Gamma}}} is tangent to both Λ\Lambda and to the sΓs_{{}_{\Gamma}}-fibers, we conclude that vv vanishes on Λ\Lambda, which implies that f=0\mathcal{L}_{f}=0 in (94).

Theorem 15.

Let Λ\Lambda be a GG-relation and QΨG,cmQ\in\Psi_{G,c}^{m} with principal and sub-principal GG-symbols q0q^{0} and q1sq^{1s}. Assume that q0q^{0} vanishes on Λ\Lambda. Let AIm(G,Λ;Ω1/2)A\in I^{m^{\prime}}(G,\Lambda;\Omega^{1/2}) and let aSm+n/4(Λ,MΛΩΛ1/2Ω^1/2Ω^G1/2)a\in S^{m^{\prime}+n/4}(\Lambda,M_{\Lambda}\otimes\Omega^{1/2}_{\Lambda}\otimes\hat{\Omega}^{1/2}\otimes\hat{\Omega}^{-1/2}_{G}) be a principal symbol of AA.
Then

(96) QAIm+m1(G,Λ;Ω1/2)QA\in I^{m+m^{\prime}-1}(G,\Lambda;\Omega^{1/2})

and PAPA has a principal symbol represented by

(97) iq0a+q1sa.-i\mathcal{L}_{q^{0}}a+q^{1s}a.

We could consider the distribution QAQA as the family of operators QxAxQ_{x}\circ A_{x} and apply [12, Theorem 25.2.4]. However, we are going to consider QAQA as a single lagrangian distribution on GG given by the convolution in GG of two distributions, and then make the minor necessary adaptations of the proof of [12, Theorem 25.2.4]. This yields more conceptual and self-contained explanations for the assertion to be proved.

Proof.

We keep the assumptions and notation introduced for QQ in the proof of Proposition 12. Using a partition of unity and [18], we can assume that AA is supported in the domain VV of local coordinates trivialising ss such that there exists a conic open set CC in n{\mathbb{R}}^{n} and a homogeneous CC^{\infty} function hh such that:

(98) ΛTV={(h(ξ),ξ);ξC}.\Lambda\cap T^{*}V=\{(h^{\prime}(\xi),\xi)\ ;\ \xi\in C\}.

The existence of such coordinates follows from [12, Lemma 25.2.5 and Theorem 21.2.16]. We can write in these local coordinates above:

(99) A=𝖠.μs1/2μr1/2 with 𝖠(γ)=ei(<γ,η>h(η))𝖺(η)𝑑η,A=\mathsf{A}.\mu^{1/2}_{s}\mu^{1/2}_{r}\text{ with }\mathsf{A}(\gamma)=\int e^{i(<\gamma,\eta>-h(\eta))}\mathsf{a}(\eta)d\eta,

where 𝖺Smn/4(n)\mathsf{a}\in S^{m^{\prime}-n/4}({\mathbb{R}}^{n}) has support in a conic neighborhood of CC. Then on VV:

(100) QA=𝖡μs1/2μr1/2 with 𝖡(γ)=ei(<αγ1,ξ>+<α,η>h(η))q0(r(γ),ξ)𝖺(η)φ(α)𝑑α𝑑ξ𝑑η.QA=\mathsf{B}\mu^{1/2}_{s}\mu^{1/2}_{r}\text{ with }\mathsf{B}(\gamma)=\int e^{i(-<\alpha\gamma^{-1},\xi>+<\alpha,\eta>-h(\eta))}q_{0}(r(\gamma),\xi)\mathsf{a}(\eta)\varphi(\alpha)d\alpha d\xi d\eta.

Remember that, according to the decomposition η=(η′′,η)\eta=(\eta^{\prime\prime},\eta^{\prime}) provided by the local trivialisation of ss, the symbol q(γ,η)q(\gamma,\eta^{\prime}) of Qs(γ)Q_{s(\gamma)} (see (6) and above) is given by

(101) q(γ,η)=eiγ,ηQs(γ)(eiα,η)=ei(<αγ1,ξ>+<αγ,η>)q0(r(γ),ξ)φ(α)𝑑α𝑑ξ.q(\gamma,\eta^{\prime})=e^{-i\langle\gamma,\eta^{\prime}\rangle}Q_{s(\gamma)}(e^{i\langle\alpha,\eta^{\prime}\rangle})=\int e^{i(-<\alpha\gamma^{-1},\xi>+<\alpha-\gamma,\eta^{\prime}>)}q_{0}(r(\gamma),\xi)\varphi(\alpha)d\alpha d\xi.

Since γ,αGs(γ)\gamma,\alpha\in G_{s(\gamma)}, the same identity is licit for qq considered as a function of (γ,η)(\gamma,\eta), but does not define anymore a symbol in general (it satisfies symbolic estimates of order mm in η\eta^{\prime} but is independent of η′′\eta^{\prime\prime}). However, the assumption on the wave front set of AA implies that the symbol aa is or order -\infty in some open cone around (η′′,0)(\eta^{\prime\prime},0). Indeed, (η′′,0)ker(sΓ)(\eta^{\prime\prime},0)\in\ker(s_{\Gamma}) and by assumption Λ\Lambda is a GG relation, hence WFAker(sΓ)=\operatorname{WF}{A}\cap\ker(s_{\Gamma})=\emptyset. Therefore, the product 𝖻(γ,η)=q(γ,η)𝖺(η)\mathsf{b}(\gamma,\eta)=q(\gamma,\eta)\mathsf{a}(\eta) is a symbol of order m+mn/4m+m^{\prime}-n/4. We get from (100) and (101):

(102) 𝖡(γ)=ei(<γ,η>h(η))q(γ,η)𝖺(η)𝑑η,\mathsf{B}(\gamma)=\int e^{i(<\gamma,\eta>-h(\eta))}q(\gamma,\eta)\mathsf{a}(\eta)d\eta,

It is a lagrangian distribution of order m+mm+m^{\prime} a priori, but since the leading part of 𝖻\mathsf{b}, which is represented 𝖻0(γ,η)=q0(γ,η)𝖺(η)\mathsf{b}^{0}(\gamma,\eta)=q^{0}(\gamma,\eta)\mathsf{a}(\eta), vanishes on Λ\Lambda, it is actually of order m+m1m+m^{\prime}-1 and we need to work out more the expression (102) to get its principal symbol. To this end, we set:

(103) q(γ,ξ)=q0(γ,ξ)+e(γ,ξ),q(\gamma,\xi)=q^{0}(\gamma,\xi)+e(\gamma,\xi),

and using the assumption: q0(γ,ξ)=0q^{0}(\gamma,\xi)=0 whenever γ=h(ξ)\gamma=h^{\prime}(\xi); we make the factorisation:

(104) q0(γ,ξ)=jqj(γ,ξ)(γjhξj).q^{0}(\gamma,\xi)=\sum_{j}q_{j}(\gamma,\xi)(\gamma_{j}-\frac{\partial h}{\partial\xi_{j}}).

Now, after an integration by parts in (102), we get:

(105) 𝖡(γ)=ei(<γ,ξ>h(ξ))(e𝖺jDξj(qj𝖺))𝑑ξ.\mathsf{B}(\gamma)=\int e^{i(<\gamma,\xi>-h(\xi))}(e\mathsf{a}-\sum_{j}D_{\xi_{j}}(q_{j}\mathsf{a}))d\xi.

It follows that QAQA has principal symbol represented by

(106) (e𝖺jDξj(qj𝖺))(h(ξ),ξ)|dξ|1/2(μrμs|dγ|1)1/2(e\mathsf{a}-\sum_{j}D_{\xi_{j}}(q_{j}\mathsf{a}))(h^{\prime}(\xi),\xi)|d\xi|^{1/2}(\mu_{r}\mu_{s}|d\gamma|^{-1})^{1/2}

while AA has principal symbol represented by

(107) 𝖺(ξ)|dξ|1/2(μrμs|dγ|1)1/2.\mathsf{a}(\xi)|d\xi|^{1/2}(\mu_{r}\mu_{s}|d\gamma|^{-1})^{1/2}.

Since Hq0H_{q^{0}} is tangent to Λ\Lambda we have on Λ\Lambda parametrized by ξ\xi:

(108) Hq0(ξ)=jq0γj(h(ξ),ξ)ξj=jqj(h(ξ),ξ)ξj.H_{q^{0}}(\xi)=-\sum_{j}\frac{\partial q^{0}}{\partial\gamma_{j}}(h^{\prime}(\xi),\xi)\frac{\partial}{\partial\xi_{j}}=-\sum_{j}q_{j}(h^{\prime}(\xi),\xi)\frac{\partial}{\partial\xi_{j}}.

Then, as a vector field on Λ\Lambda, the divergence of Hq0Γ(TΛ)H_{q^{0}}\in\Gamma(T\Lambda) is given by:

(109) div(Hq0)=jξj[qj(h(ξ),ξ)]=j,k2hξkξjqjγkjqjξj.\mathrm{div}(H_{q^{0}})=-\sum_{j}\frac{\partial}{\partial\xi_{j}}\Big{[}q_{j}(h^{\prime}(\xi),\xi)\Big{]}=-\sum_{j,k}\frac{\partial^{2}h}{\partial\xi_{k}\partial\xi_{j}}\frac{\partial q_{j}}{\partial\gamma_{k}}-\sum_{j}\frac{\partial q_{j}}{\partial\xi_{j}}.

On the other hand we have:

2q0ξkγk(γ,ξ)\displaystyle\frac{\partial^{2}{q^{0}}}{\partial\xi_{k}\partial\gamma_{k}}(\gamma,\xi) =ξkj(γkqj).(γjhj(ξ))+qk\displaystyle=\frac{\partial}{\partial\xi_{k}}\sum_{j}(\frac{\partial}{\partial\gamma_{k}}q_{j}).(\gamma_{j}-h^{\prime}_{j}(\xi))+q_{k}
=j2qjγkξk.(γjhj(ξ))jqjγk2hξkξj+qkξk,\displaystyle=\sum_{j}\frac{\partial^{2}q_{j}}{\partial\gamma_{k}\partial\xi_{k}}.(\gamma_{j}-h^{\prime}_{j}(\xi))-\sum_{j}\frac{\partial q_{j}}{\partial\gamma_{k}}\frac{\partial^{2}h}{\partial\xi_{k}\partial\xi_{j}}+\frac{\partial q_{k}}{\partial\xi_{k}},

which gives after evaluation at γ=h(ξ)\gamma=h^{\prime}(\xi):

(110) k2q0ξkγk(h(ξ),ξ)=kqkξkj,kqjγk2hξkξj=2kqkξk+div(Hq0).\sum_{k}\frac{\partial^{2}{q^{0}}}{\partial\xi_{k}\partial\gamma_{k}}(h^{\prime}(\xi),\xi)=\sum_{k}\frac{\partial q_{k}}{\partial\xi_{k}}-\sum_{j,k}\frac{\partial q_{j}}{\partial\gamma_{k}}\frac{\partial^{2}h}{\partial\xi_{k}\partial\xi_{j}}=2\sum_{k}\frac{\partial q_{k}}{\partial\xi_{k}}+\mathrm{div}(H_{q^{0}}).

Setting

𝖫q0(a):=q0(a)(μrμs/|dγ|)1/2=jqj𝖺ξj+12div(Hq0)𝖺,\mathsf{L}_{q^{0}}(a):=\mathcal{L}_{q^{0}}(a)(\mu_{r}\mu_{s}/|d\gamma|)^{-1/2}=-\sum_{j}q_{j}\frac{\partial\mathsf{a}}{\partial\xi_{j}}+\frac{1}{2}\mathrm{div}(H_{q^{0}})\mathsf{a},

we get, still for γ=h(ξ)\gamma=h^{\prime}(\xi):

(e𝖺jDξj(qj𝖺))\displaystyle(e\mathsf{a}-\sum_{j}D_{\xi_{j}}(q_{j}\mathsf{a})) =(e+ijqjξj)𝖺+ijqj𝖺ξj\displaystyle=\Big{(}e+i\sum_{j}\frac{\partial q_{j}}{\partial\xi_{j}}\Big{)}\mathsf{a}+i\sum_{j}q_{j}\frac{\partial\mathsf{a}}{\partial\xi_{j}}
=(e+ijqjξj+i2div(Hq0))𝖺i𝖫q0(a)\displaystyle=\Big{(}e+i\sum_{j}\frac{\partial q_{j}}{\partial\xi_{j}}+\frac{i}{2}\mathrm{div}(H_{q^{0}})\Big{)}\mathsf{a}-i\mathsf{L}_{q^{0}}(a)
=(e+i2k2q0ξkγk)𝖺i𝖫q0(a)\displaystyle=\Big{(}e+\frac{i}{2}\sum_{k}\frac{\partial^{2}{q^{0}}}{\partial\xi_{k}\partial\gamma_{k}}\Big{)}\mathsf{a}-i\mathsf{L}_{q^{0}}(a)
(111) =i𝖫q0(a)+(e+i2k2q0ξkγk)𝖺.\displaystyle=-i\mathsf{L}_{q^{0}}(a)+\left(e+\frac{i}{2}\sum_{k}\frac{\partial^{2}{q^{0}}}{\partial\xi^{\prime}_{k}\partial\gamma^{\prime}_{k}}\right)\mathsf{a}.

In the last line, we have decomposed ξ=(ξ,ξ′′)\xi=(\xi^{\prime},\xi^{\prime\prime}) where ξ\xi^{\prime} is cotangent to the ss-fibers and used the fact that q0q^{0} does not depend on the ξ′′\xi^{\prime\prime} variables. This proves, looking at formula (89), that iq0(a)+q1s.a-i\mathcal{L}_{q^{0}}(a)+q^{1s}.a is a principal symbol of QAQA. ∎

7. The Hamiltonian flow of the principal symbol and the associated GG-relations

We set:

Λ0=AG0,T.G=TGkerrΓ and T.G=TG(kerrΓkersΓ).\Lambda_{0}=A^{*}G\setminus 0,\quad T^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G=T^{*}G\setminus\ker r_{{}_{\Gamma}}\ \text{ and }\quad\accentset{\mbox{\large.}}{T}^{*}G=T^{*}G\setminus(\ker r_{{}_{\Gamma}}\cup\ker s_{{}_{\Gamma}}).

The operator PP being elliptic, we have WFP=Λ0\operatorname{WF}{P}=\Lambda_{0}. Recall that p00C(Λ0)p^{0}_{0}\in C^{\infty}(\Lambda_{0}) denotes the homogeneous representative of σ0(P)\sigma_{0}(P) and that p0=p00rΓp^{0}=p^{0}_{0}\circ r_{{}_{\Gamma}} is then the homogeneous representative of σ(P)C(TsG0)\sigma(P)\in C^{\infty}(T^{*}_{s}G\setminus 0).

For every aa\in{\mathbb{R}} and (γ,ξ)TG(\gamma,\xi)\in T^{*}G, we let a.(γ,ξ)=𝜌a(γ,ξ)=(γ,aξ)a.(\gamma,\xi)=\mathop{\rho}\nolimits_{a}(\gamma,\xi)=(\gamma,a\xi).

Proposition 16.

The flow χ\chi of Hp0H_{p^{0}} satisfies the following properties:

  1. (1)

    It is complete.

  2. (2)

    It commutes with dilations in TGT^{*}G:

    a+,t,(γ,ξ)T.G,a.χ(t,γ,ξ)=χ(t,γ,aξ).\forall a\in{\mathbb{R}}^{*}_{+},\ t\in{\mathbb{R}},\ (\gamma,\xi)\in T^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G,\quad a.\chi(t,\gamma,\xi)=\chi(t,\gamma,a\xi).
  3. (3)

    It provides at each time tt a section of sΓs_{\Gamma} and commutes with right multiplication:

    t,(δ1,δ2)Γ(2),sΓ(χ(t,δ1))=sΓ(δ1) and χ(t,δ1δ2)=χ(t,δ1)δ2.\forall t\in{\mathbb{R}},\ (\delta_{1},\delta_{2})\in\Gamma^{(2)},\quad s_{{}_{\Gamma}}(\chi(t,\delta_{1}))=s_{{}_{\Gamma}}(\delta_{1})\text{ and }\chi(t,\delta_{1}\delta_{2})=\chi(t,\delta_{1})\delta_{2}.

    In particular, the integral curves of Hp0H_{p^{0}} go along the fibers of sΓs_{\Gamma}.

Proof.

This is essentially contained in [6]. More precisely:

  1. (2)

    The homogeneity of p0p^{0} implies: (𝜌a)(Hp0)=Hp0(\mathop{\rho}\nolimits_{a})_{*}(H_{p^{0}})=H_{p^{0}} and therefore 𝜌aχt𝜌a1=χt\mathop{\rho}\nolimits_{a}\circ\chi_{t}\circ\mathop{\rho}\nolimits_{a}^{-1}=\chi_{t}, which gives the result.

  2. (3)

    By definition, we have ω(Hp0,X)=dp0(X)=dp00(drΓ(X))\omega(H_{p^{0}},X)=dp^{0}(X)=dp_{0}^{0}(dr_{{}_{\Gamma}}(X)) which yields Hp0(dkerrΓ)ω=dkersΓH_{p^{0}}\in(d\ker r_{{}_{\Gamma}})^{\omega}=d\ker s_{{}_{\Gamma}}. Using the last part of the proof of [6, Lemma, p.22], we get that Hp0H_{p^{0}} is a right invariant vector field, which proves that the flow goes along the ss-fibers and is right invariant.

  3. (1)

    Now, by compacity of M=G(0)M=G^{(0)} and the homogeneity of Hp0H_{p^{0}} in the fibers of Λ0\Lambda_{0}, we get the existence of ϵ>0\epsilon>0 such that χ:]ϵ,ϵ[×Λ0T.G\chi:]-\epsilon,\epsilon[\times\Lambda_{0}\to T^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G is well defined. By right invariance, we extend χp\chi_{p} onto ]ϵ,ϵ[×(T.G)]-\epsilon,\epsilon[\times(T^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G) and using the group property of flows we can choose for any tt an integer NN such that |t|/N<ϵ|t|/N<\epsilon and set

    χ(t,α)=χt/Nχt/N(α)\chi(t,\alpha)=\chi_{t/N}\circ\ldots\circ\chi_{t/N}(\alpha)

    which proves completeness of the flow.

Remark 17.
  1. (1)

    The compacity of M=G(0)M=G^{(0)} is only needed here for the completeness of Hp0H_{p^{0}}.

  2. (2)

    By construction, χt\chi_{t} is a diffeomorphism of T.GT^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G and since sΓχt=sΓs_{\Gamma}\circ\chi_{t}=s_{\Gamma}, we have:

    (112) χt(T.G)=T.G.\chi_{t}(\accentset{\mbox{\large.}}{T}^{*}G)=\accentset{\mbox{\large.}}{T}^{*}G.

We now set:

(113) t,Λt=χt(Λ0)T.G.\forall t\in{\mathbb{R}},\quad\Lambda_{t}=\chi_{t}(\Lambda_{0})\subset\accentset{\mbox{\large.}}{T}^{*}G.
Proposition 18.

For any real number tt, the set Λt\Lambda_{t} is an invertible GG-relation and:

(114) t1,t2,Λt1.Λt2=Λt1+t2.\forall t_{1},t_{2}\in{\mathbb{R}},\quad\Lambda_{t_{1}}.\Lambda_{t_{2}}=\Lambda_{t_{1}+t_{2}}.
Proof.

Since χ\chi is the Hamilton flow of a homogeneous function, each χt\chi_{t} is a homogeneous symplectomorphism. Since Λ0\Lambda_{0} is a homogeneous Lagrangian, its image Λt\Lambda_{t} by χt\chi_{t} is then a Lagrangian homogeneous submanifold, contained by construction in T.G\accentset{\mbox{\large.}}{T}^{*}G. Thus, Λt\Lambda_{t} is a GG-relation. Since sΓχt|Λ0=IdΛ0s_{\Gamma}\circ\chi_{t}|_{\Lambda_{0}}=\operatorname{Id}_{\Lambda_{0}}, we get that sΓ|Λts_{\Gamma}|_{\Lambda_{t}} is a diffeomorphism. The same conclusion is true for rΓ|Λtr_{\Gamma}|_{\Lambda_{t}} because the vector field Hp0H_{p^{0}} is right invariant and therefore rΓχ|×Λ0r_{\Gamma}\circ\chi|_{{\mathbb{R}}\times\Lambda_{0}} is the (complete) flow of the vector field (rΓ)(Hp0)C(Λ0,TΛ0)(r_{\Gamma})_{*}(H_{p^{0}})\in C^{\infty}(\Lambda_{0},T\Lambda_{0}) defined by: (rΓ)(Hp0)(δ)=(drΓ)δ(Hp0(δ)),δΛ0(r_{\Gamma})_{*}(H_{p^{0}})(\delta)=(dr_{\Gamma})_{\delta}(H_{p^{0}}(\delta)),\delta\in\Lambda_{0}. That sΓ|Λts_{\Gamma}|_{\Lambda_{t}} and rΓ|Λtr_{\Gamma}|_{\Lambda_{t}} are diffeomorphisms mean precisely that Λt\Lambda_{t} is an invertible GG-relation [18] (or a lagrangian bissection, if one accepts as bissections submanifolds of Γ\Gamma onto which rΓr_{\Gamma} and sΓs_{\Gamma} are diffeomorphisms onto open subsets of AG=(TG)(0)A^{*}G=(T^{*}G)^{(0)}).

Let us proceed to the proof of the one parameter group relation. Let δj=χtj(uj)Λtj\delta_{j}=\chi_{t_{j}}(u_{j})\in\Lambda_{t_{j}}, j=1,2j=1,2, be two composable elements. Then u1=sΓ(δ1)=rΓ(δ2)u_{1}=s_{\Gamma}(\delta_{1})=r_{\Gamma}(\delta_{2}) and by commutation of χ\chi with right multiplication:

δ1.δ2=χt1(rΓ(δ2))δ2=χt1(δ2)=χt1(χt2(u2))=χt1+t2(u2)Λt1+t2.\delta_{1}.\delta_{2}=\chi_{t_{1}}(r_{\Gamma}(\delta_{2}))\delta_{2}=\chi_{t_{1}}(\delta_{2})=\chi_{t_{1}}(\chi_{t_{2}}(u_{2}))=\chi_{t_{1}+t_{2}}(u_{2})\in\Lambda_{t_{1}+t_{2}}.

The converse inclusion is then obvious. ∎

8. Global aspects of the family (Λt)t(\Lambda_{t})_{t}

In the vocabulary of [18], the family (Λt)t(\Lambda_{t})_{t\in{\mathbb{R}}} admits a gluing into a single Lagrangian submanifold ΛT(×G)\Lambda\subset T^{*}({\mathbb{R}}\times G). The expression of Λ\Lambda is actually straightforward and we shall study it in relation with the both groupoid structures on ×G{\mathbb{R}}\times G.

Proposition 19.

Let it:G×Gi_{t}:G\to{\mathbb{R}}\times G be the inclusion γ(t,γ)\gamma\mapsto(t,\gamma). The set

(115) Λ={(t,p0(χt(x,ξ)),χt(x,ξ))T(×G);t,(x,ξ)AG0}\Lambda=\left\{(t,-p^{0}(\chi_{t}(x,\xi)),\chi_{t}(x,\xi))\in T^{*}({\mathbb{R}}\times G)\ ;\ t\in{\mathbb{R}},(x,\xi)\in A^{*}G\setminus 0\right\}

is a conic Lagrangian submanifold of T(×G)T^{*}({\mathbb{R}}\times G) satisfying:

t,itΛ=Λt.\forall t\in{\mathbb{R}},\quad i_{t}^{*}\Lambda=\Lambda_{t}.
Proof.

the map F(t,δ)=(t,χt(δ))F(t,\delta)=(t,\chi_{t}(\delta)) being a diffeomorphism of ×(T.G){\mathbb{R}}\times(T^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G), the set F(×Λ0)F({\mathbb{R}}\times\Lambda_{0}) is a submanifold of ×TG{\mathbb{R}}\times T^{*}G and therefore Λ\Lambda, as a graph, is a submanifold of T(×G)T^{*}({\mathbb{R}}\times G), obviously homogeneous and ϕ(t,x,ξ)=(t,p0(χt(x,ξ)),χt(x,ξ))\phi(t,x,\xi)=(t,-p^{0}(\chi_{t}(x,\xi)),\chi_{t}(x,\xi)) is a parametrization.

We check that Λ\Lambda is lagrangian, which is equivalent by homogeneity of Λ\Lambda to the vanishing on it of the canonical one form α×G=τdt+αG\alpha_{{\mathbb{R}}\times G}=\tau dt+\alpha_{G}, that is to the vanishing of the one form on ×Λ0{\mathbb{R}}\times\Lambda_{0} defined by ϕα×G\phi^{*}\alpha_{{\mathbb{R}}\times G}. We have:

ϕα×G\displaystyle\phi^{*}\alpha_{{\mathbb{R}}\times G} =p0(χ)dt+χαG\displaystyle=-p^{0}(\chi)dt+\chi^{*}\alpha_{G}
=p0(χ)dt+αG(χt)dt+(χt)αG\displaystyle=-p^{0}(\chi)dt+\alpha_{G}(\frac{\partial\chi}{\partial t})dt+(\chi_{t})^{*}\alpha_{G}

Since χt\chi_{t} is a homogeneous symplectomorphism of T.GT^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G, the one form (χt)αG(\chi_{t})^{*}\alpha_{G} vanishes on Λ0\Lambda_{0}. On the other hand, by homogeneity of p0p^{0} and Euler formula:

(116) αG(χt)=αG(Hp0(χt))=jξjp0ξj(χ)=p0(χ),\alpha_{G}(\frac{\partial\chi}{\partial t})=\alpha_{G}(H_{p^{0}}(\chi_{t}))=\sum_{j}\xi_{j}\frac{\partial p^{0}}{\partial\xi_{j}}(\chi)=p^{0}(\chi),

which proves the required assertion. ∎

There are two natural structures of groupoid on ×G{\mathbb{R}}\times G, with different unit space:

G~=×G×G(0) and G+=×GG(0)\widetilde{G}={\mathbb{R}}\times G\rightrightarrows{\mathbb{R}}\times G^{(0)}\text{ and }G_{+}={\mathbb{R}}\times G\rightrightarrows G^{(0)}

The first is the (constant) family of groupoids Gt=GG_{t}=G parametrized by the space {\mathbb{R}}, and the second one is the cartesian product of GG with the additive group {\mathbb{R}}. The corresponding symplectic groupoid structures on T(×G)T^{*}({\mathbb{R}}\times G) will be denoted by:

(117) Γ~=T(×G)pr2AG and Γ+=T(×G)×AG\widetilde{\Gamma}=T^{*}({\mathbb{R}}\times G)\rightrightarrows\mathop{\mathrm{pr}}\nolimits_{2}^{*}A^{*}G\ \text{ and }\ \Gamma_{+}=T^{*}({\mathbb{R}}\times G)\rightrightarrows{\mathbb{R}}\times A^{*}G

where pr2:×MM\mathop{\mathrm{pr}}\nolimits_{2}:{\mathbb{R}}\times M\to M is the second projection and ×AG{\mathbb{R}}\times A^{*}G denotes the bundle over MM with fibers ×AxG{\mathbb{R}}\times A_{x}^{*}G.

We will say that a subset A×XA\subset{\mathbb{R}}\times X is {\mathbb{R}}-proper if pr1:A\mathop{\mathrm{pr}}\nolimits_{1}:A\to{\mathbb{R}} is proper, that is

A[a,b]×X is compact for any a,b.A\cap[a,b]\times X\text{ is compact for any }a,b\in{\mathbb{R}}.

We will call support of AEA\subset E the set supp(A)=π(A)X{\mathrm{supp}(A)}=\pi(A)\subset X for any bundle map π:EX\pi:E\to X.

Proposition 20.

The submanifold Λ\Lambda of T(×G)T^{*}({\mathbb{R}}\times G) satisfies the following:

  1. (1)

    It is contained in (T0)×T.G(T^{*}{\mathbb{R}}\setminus 0)\times\accentset{\mbox{\large.}}{T}^{*}G and closed in T×(T.G)T^{*}{\mathbb{R}}\times(T^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G).

  2. (2)

    It is both an invertible G~\widetilde{G}-relation and a family G+G_{+}-relation.

  3. (3)

    The support of Λ\Lambda is {\mathbb{R}}-proper.

Proof.
  1. (1)

    We first check that Λ\Lambda is closed in T×T.GT^{*}{\mathbb{R}}\times T^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G. The map

    ϕ:×T.G×T.G,(t,λ)(t,χ(t,λ))\phi:{\mathbb{R}}\times T^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G\longrightarrow{\mathbb{R}}\times T^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G,\quad(t,\lambda)\longmapsto(t,\chi(t,\lambda))

    is a diffeomorphism and ×Λ0=×T.G×AG{\mathbb{R}}\times\Lambda_{0}={\mathbb{R}}\times T^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G\cap{\mathbb{R}}\times A^{*}G is closed in ×T.G{\mathbb{R}}\times T^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G since ×AG{\mathbb{R}}\times A^{*}G is closed in T(×G)T^{*}({\mathbb{R}}\times G). Thus ϕ(×Λ0)\phi({\mathbb{R}}\times\Lambda_{0}) is closed in ×T.G{\mathbb{R}}\times T^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G. It follows that

    Λ={(t,p0(λ),λ)T×T.G;(t,λ)ϕ(×Λ0)}\Lambda=\{(t,-p^{0}(\lambda),\lambda)\in T^{*}{\mathbb{R}}\times T^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G\ ;\ (t,\lambda)\in\phi({\mathbb{R}}\times\Lambda_{0})\}

    is closed in T×T.GT^{*}{\mathbb{R}}\times T^{*}_{\raisebox{1.42262pt}{\mbox{\large.}}}G.

  2. (2)

    By remark 17, the inclusion ΛT×T.G\Lambda\subset T^{*}{\mathbb{R}}\times\accentset{\mbox{\large.}}{T}^{*}G holds true and by ellipticity of PP, the function p0=p0rΓp^{0}=p^{0}\circ r_{{}_{\Gamma}} does not vanish on T.G\accentset{\mbox{\large.}}{T}^{*}G, hence

    Λ(T0)×T.G.\Lambda\subset(T^{*}{\mathbb{R}}\setminus 0)\times\accentset{\mbox{\large.}}{T}^{*}G.
  3. (3)

    Since rΓ~(t,τ,λ)=(t,rΓ(λ))r_{\widetilde{{}_{\Gamma}}}(t,\tau,\lambda)=(t,r_{{}_{\Gamma}}(\lambda)) and sΓ~(t,τ,λ)=(t,sΓ(λ))s_{\widetilde{{}_{\Gamma}}}(t,\tau,\lambda)=(t,s_{{}_{\Gamma}}(\lambda)), we immediately deduce the invertibility of Λ\Lambda from the invertibility of the GG-relations Λt\Lambda_{t} for all tt.

  4. (4)

    Since rΓ+(t,τ,λ)=(τ,rΓ(λ))r_{{}_{\Gamma_{+}}}(t,\tau,\lambda)=(\tau,r_{{}_{\Gamma}}(\lambda)), sΓ+(t,τ,λ)=(τ,sΓ(λ))s_{{}_{\Gamma_{+}}}(t,\tau,\lambda)=(\tau,s_{{}_{\Gamma}}(\lambda)) and Λ(T0)×T.G\Lambda\subset(T^{*}{\mathbb{R}}\setminus 0)\times\accentset{\mbox{\large.}}{T}^{*}G we get ΛkerrΓ+=\Lambda\cap\ker r_{{}_{\Gamma_{+}}}=\emptyset and the same for sΓ+s_{{}_{\Gamma_{+}}} so Λ\Lambda is a G+G_{+}-relation. Moreover, denoting by π,π0,π2\pi,\pi_{0},\pi_{2} the natural projection maps:

    π:T(×G)×G,π0:AGM,π2:T(×G)TG,\pi:T^{*}({\mathbb{R}}\times G)\to{\mathbb{R}}\times G,\quad\pi_{0}:A^{*}G\to M,\quad\pi_{2}:T^{*}({\mathbb{R}}\times G)\to T^{*}G,

    since (t,τ,λ)ΛsΓ(λ)AG(t,\tau,\lambda)\in\Lambda\mapsto s_{{}_{\Gamma}}(\lambda)\in A^{*}G is a submersion, the composition

    π0sΓπ2=sG+π|Λ:ΛAG+M,(t,τ,γ,ξ)s(γ)\pi_{0}\circ s_{{}_{\Gamma}}\circ\pi_{2}=s_{G_{+}}\circ\pi|_{\Lambda}:\Lambda\longrightarrow A^{*}G_{+}\longrightarrow M,(t,\tau,\gamma,\xi)\longmapsto s(\gamma)

    is a submersion. This proves that Λ\Lambda is a G+G_{+}-family by [18, Remark 15 and below]

  5. (5)

    This is a straightforward consequence of the compacity of M=G(0)M=G^{(0)}, of the homogeneity of χ\chi, and of standard continuity arguments.

9. Approximation of eitPe^{-itP} by GG-FIOs

The manifold ×G{\mathbb{R}}\times G will be provided by the pull back of the half density bundle used for GG, and it will still be denoted by Ω1/2\Omega^{1/2}.

Let Λ\Lambda be the G~\widetilde{G}-relation defined by PP as in (115). Since Λ\Lambda is a family G~\widetilde{G}-relation, any UIm(×G,Λ;Ω1/2)U\in I^{m}({\mathbb{R}}\times G,\Lambda;\Omega^{1/2}) is a Fourier integral G~\widetilde{G}-operator (see [18] for the details), also given as a distribution on G~\widetilde{G} by the CC^{\infty} family U(t)Im+1/4(G,Λt;Ω1/2)U(t)\in I^{m+1/4}(G,\Lambda_{t};\Omega^{1/2}) of GG-FIOs defined by U(t)=it(U)U(t)=i_{t}^{*}(U). Here it:G×Gi_{t}:G\to{\mathbb{R}}\times G is the inclusion it(γ)=(t,γ)i_{t}(\gamma)=(t,\gamma). The converse is true: any such family gives a single distribution in Im(×G,Λ;Ω1/2)I^{m}({\mathbb{R}}\times G,\Lambda;\Omega^{1/2}).

Theorem 21.

There exists a Fourier integral G~\widetilde{G}-operator UI14+(n(1)n(0))/4(×G,Λ;Ω1/2)U\in I^{-\frac{1}{4}+(n^{(1)}-n^{(0)})/4}({\mathbb{R}}\times G,\Lambda;\Omega^{1/2}) with {\mathbb{R}}-proper support such that:

(118) (t+iP)UC(×G,Ω1/2).(\frac{\partial}{\partial t}+iP)U\in C^{\infty}({\mathbb{R}}\times G,\Omega^{1/2}).

Moreover, if E=(eitP)tE=(e^{-itP})_{t\in{\mathbb{R}}} denotes the one parameter group defined in Section 3, we have:

(119) EUC(,).E-U\in C^{\infty}({\mathbb{R}},\mathcal{H}^{\infty}).
Remark 22.
  1. (1)

    It follows that (E(t))t(E(t))_{t\in{\mathbb{R}}} is a CC^{\infty} family of distributions, equivalently E𝒟pr1(×G,Ω1/2)E\in\mathcal{D}^{\prime}_{\mathop{\mathrm{pr}}\nolimits_{1}}({\mathbb{R}}\times G,\Omega^{1/2}).

  2. (2)

    Recall that Corb,0(G)\mathcal{H}^{\infty}\subset C^{\infty,0}_{\mathop{\mathrm{orb}}\nolimits}(G) by Section 5, in particular the error term (119) is CC^{\infty} on ×G𝒪{\mathbb{R}}\times G_{\mathcal{O}} for any orbit 𝒪M\mathcal{O}\subset M.

  3. (3)

    Theorem 21 also gives information about the operators eitPxe^{itP_{x}} on the (usually non compact, complete, with bounded geometry) manifolds GxG_{x}, xMx\in M. In the latter situation, we refer to [21] for related results.

Proof of the theorem.

Let UIm(×G,Λ;Ω1/2)U\in I^{m}({\mathbb{R}}\times G,\Lambda;\Omega^{1/2}). We first check that:

(120) tUIm+1(×G,Λ;Ω1/2) and PUIm+1(×G,Λ;Ω1/2).\frac{\partial}{\partial t}U\in I^{m+1}({\mathbb{R}}\times G,\Lambda;\Omega^{1/2})\text{ and }PU\in I^{m+1}({\mathbb{R}}\times G,\Lambda;\Omega^{1/2}).

The distribution PUPU is given by convolution product in G~\widetilde{G} of the G~\widetilde{G}-PDO PP with the G~\widetilde{G}-FIO UU. Therefore, the composition theorem of [18] applies and proves PUIm+1(×G,Λ,Ω1/2)PU\in I^{m+1}({\mathbb{R}}\times G,\Lambda,\Omega^{1/2}). Note that PUPU is also a convolution of distributions in G+G_{+}:

(121) PU=(δtsP)G+U,PU=(\delta_{t-s}\otimes P)*_{G_{+}}U,

but this time it is not a composition of G+G_{+}-FIO because δtsP\delta_{t-s}\otimes P fails to be in general a G+G_{+}-PDO. The other assertion in (120) can be checked either by directly differentiating with respect to tt the family (U(t))t(U(t))_{t} expressed in local coordinates with oscillatory integrals, or by composing the differential G+G_{+}-operator t\frac{\partial}{\partial t} with the G+G_{+}-FIO UU.

The next task is to prove that the sum (t+iP)U(\frac{\partial}{\partial t}+iP)U is actually of order mm and has principal symbol given by:

(122) τ+p0(u)+iσ1s(P)u.\mathcal{L}_{\tau+p^{0}}(u)+i\sigma^{1s}(P)u.

Since t+iP\frac{\partial}{\partial t}+iP is neither a G~\widetilde{G} nor G+G_{+} pseudodifferential operator, we can not directly apply Proposition 15 to extract the principal symbol of (120). We propose two ways to overcome this difficulty, both containing useful technics.

First approach. Both distributions tU\frac{\partial}{\partial t}U and PUPU are G~\widetilde{G}-FIO. Working as before in suitable local coordinates (t,γ,ξ)(t,\gamma,\xi), and using for instance [18, Theorems 5 and 6], there exists a CC^{\infty} function h(t,ξ)h(t,\xi), homogeneous of order 11 in ξ\xi, and a symbol u(t,ξ)u(t,\xi), such that:

(123) (t,τ,γ,ξ)Λτ=ht(t,ξ),γ=hξ(t,ξ),(t,\tau,\gamma,\xi)\in\Lambda\iff\tau=-h^{\prime}_{t}(t,\xi),\ \gamma=h^{\prime}_{\xi}(t,\xi),
(124) U(t,γ)=ei(γ,ξh(t,ξ))u(t,ξ)𝑑ξ.U(t,\gamma)=\int e^{i(\langle\gamma,\xi\rangle-h(t,\xi))}u(t,\xi)d\xi.

It immediately follows that

(125) (t+iP)U(t,γ)=ei(γ,ξh(t,ξ))ut(t,ξ)𝑑ξ+ei(γ,ξh(t,ξ))i(p(γ,ξ)ht(t,ξ))u(t,ξ)𝑑ξ.(\frac{\partial}{\partial t}+iP)U(t,\gamma)=\int e^{i(\langle\gamma,\xi\rangle-h(t,\xi))}\frac{\partial u}{\partial t}(t,\xi)d\xi+\int e^{i(\langle\gamma,\xi\rangle-h(t,\xi))}i(p(\gamma,\xi)-h^{\prime}_{t}(t,\xi))u(t,\xi)d\xi.

The right hand side is again a sum of Lagrangian distributions. The principal symbol of the first term in the right hand side of (125) is just the restriction to Λ\Lambda of:

(126) ut=τu\frac{\partial u}{\partial t}=\mathcal{L}_{\tau}u

In the second term, although phtp-h^{\prime}_{t} does not satisfy symbol estimates in ξ\xi, the product i(p(γ,ξ)ht(t,ξ))u(t,ξ)i(p(\gamma,\xi)-h^{\prime}_{t}(t,\xi))u(t,\xi) does and its leading part, which is represented by i(p0ht)ui(p^{0}-h^{\prime}_{t})u, vanishes on Λt\Lambda_{t} for any tt. We then reproduce the computations starting with (102), just replacing h(ξ)h(\xi) by h(t,ξ)h(t,\xi), q(γ,ξ)q(\gamma,\xi) by p(γ,ξ)ht(t,ξ)p(\gamma,\xi)-h^{\prime}_{t}(t,\xi) and a(ξ)a(\xi) by u(t,ξ)u(t,\xi), without omitting an extra factor ii. The reminder ee is unchanged e=(pht)(p0ht)=pp0e=(p-h^{\prime}_{t})-(p^{0}-h^{\prime}_{t})=p-p^{0}. The vector field Hp0htH_{p^{0}-h^{\prime}_{t}} being tangent to Λt\Lambda_{t} for any tt, we get, since hth^{\prime}_{t} is independent of γ\gamma:

Hp0ht=γj(p0ht)ξj=p0γjξj=Hp0.H_{p^{0}-h^{\prime}_{t}}=-\frac{\partial}{\partial\gamma_{j}}(p^{0}-h^{\prime}_{t})\frac{\partial}{\partial\xi_{j}}=-\frac{\partial p^{0}}{\partial\gamma_{j}}\frac{\partial}{\partial\xi_{j}}=H_{p^{0}}.

Now we can read the expression for the required principal symbol in (6):

(127) p0(u)+(ie12k2(p0ht)ξkγk)u.\mathcal{L}_{p^{0}}(u)+\left(ie-\frac{1}{2}\sum_{k}\frac{\partial^{2}{(p^{0}-h^{\prime}_{t})}}{\partial\xi_{k}\partial\gamma_{k}}\right)u.

Again, since hth^{\prime}_{t} is independent of γ\gamma and p0p^{0} independent of ξ′′\xi^{\prime\prime}, the last expression simplifies to:

(128) p0(u)+(ie12k2p0ξkγk)u=p0(u)+iσ1s(P)u.\mathcal{L}_{p^{0}}(u)+\left(ie-\frac{1}{2}\sum_{k}\frac{\partial^{2}{p^{0}}}{\partial\xi^{\prime}_{k}\partial\gamma^{\prime}_{k}}\right)u=\mathcal{L}_{p^{0}}(u)+i\sigma^{1s}(P)u.

Summing up (126) and (128), we conclude that the principal symbol of (t+iP)U(\frac{\partial}{\partial t}+iP)U is (122).


Second approach. We wish to use Proposition 15 in the framework of the groupoid G+G_{+}. However, we need to have the convolution of a pseudodifferential G+G_{+}-operator with a G+G_{+}-FIO. The problem is that the distribution (t+iδtsP)(\frac{\partial}{\partial t}+i\delta_{t-s}\otimes P) is not a G+G_{+}-pseudodifferential operator, unless PP is differential. The trick (similar to the one used in the proof of [12, Theorem 25.2.4]), consists in finding a suitable microlocal approximation δtsP=P1+P2\delta_{t-s}\otimes P=P_{1}+P_{2} of δtsP\delta_{t-s}\otimes P by a G+G_{+}-PDO P1P_{1} such that P2UC(×G)P_{2}U\in C^{\infty}({\mathbb{R}}\times G). For that purpose, observe that we can deduce from (115) that there exists constants c1,c2>0c_{1},c_{2}>0 such for any (τ,x,ξ)rΓ+(Λ)(\tau,x,\xi)\in r_{\Gamma_{+}}(\Lambda), we have

(129) c1|τ||ξ|c2.c_{1}\leq\frac{|\tau|}{|\xi|}\leq c_{2}.

Indeed, we know that (τ,x,ξ)=rΓ+(t,p(λ),λ)=(p(λ),rΓ(λ))(\tau,x,\xi)=r_{\Gamma_{+}}(t,-p(\lambda),\lambda)=(-p(\lambda),r_{\Gamma}(\lambda)) for some tt\in{\mathbb{R}} and λΛt\lambda\in\Lambda_{t}. Denoting λ=(γ,η)TG0\lambda=(\gamma,\eta)\in T^{*}G\setminus 0 and rΓ(γ,η)=(x,ξ)AG0r_{\Gamma}(\gamma,\eta)=(x,\xi)\in A^{*}G\setminus 0, we then get by homegenity of p0p_{0},

|τ||ξ|=|p0rΓ(γ,η)||ξ|=|p0(x,ξ)||ξ|=|p0(x,ξ|ξ|)|\frac{|\tau|}{|\xi|}=\frac{|p_{0}\circ r_{\Gamma}(\gamma,\eta)|}{|\xi|}=\frac{|p_{0}(x,\xi)|}{|\xi|}=|p_{0}(x,\frac{\xi}{|\xi|})|

and the result follows by continuity of p0p_{0} and compacity of M=G(0)M=G^{(0)} (which implies that SG={(x,ξ)AG0,|ξ|=1}S^{*}G=\{(x,\xi)\in A^{*}G\setminus 0,|\xi|=1\} is compact.) We will use

Lemma 23.

The distribution δtsP\delta_{t-s}\otimes P on G+G_{+} can be written δtsP=P1+P2\delta_{t-s}\otimes P=P_{1}+P_{2} with P1P_{1} a G+G_{+}-pseudodifferential operator and P2P_{2} a distribution on G+G_{+} such that WFP2{(t,τ,λ)TG+0,sΓ(λ)=(x,ξ) with |ξ||τ|<ε}\operatorname{WF}{P_{2}}\subset\{(t,\tau,\lambda)\in T^{*}G_{+}\setminus 0,s_{{}_{\Gamma}}(\lambda)=(x,\xi)\mbox{ with }\frac{|\xi|}{|\tau|}<\varepsilon\}. In particular the total symbol of P1P_{1} and PP coincide in a neighborhood of rΓ+(Λ)r_{\Gamma_{+}}(\Lambda) and one has that P2UC(G+)P_{2}U\in C^{\infty}(G_{+}) if UIm(×G,Λ,Ω1/2).U\in I^{m}({\mathbb{R}}\times G,\Lambda,\Omega^{1/2}).

Proof.

Consider a map χ\chi on A(G+)=τ×AGA^{*}(G_{+})={\mathbb{R}}_{\tau}\times A^{*}G such that χ\chi is homogeneous of degree 0 in the cotangent variables outside a compact set, and such that for a chosen ε\varepsilon , one has

  1. (1)

    χ(x,ξ,τ)=0\chi(x,\xi,\tau)=0 unless 1<ε|τ|1<\varepsilon|\tau| and |ξ|<ε|τ||\xi|<\varepsilon|\tau| ;

  2. (2)

    χ(x,ξ,τ)=1\chi(x,\xi,\tau)=1 if 2<ε|τ|2<\varepsilon|\tau| and 2|ξ|<ε|τ|2|\xi|<\varepsilon|\tau| .

If P(x,ξ)P(x,\xi) is a total symbol for PP, then one can write

P(x,ξ)=p1(x,ξ,τ)+p2(x,ξ,τ)=(1χ(x,ξ,τ))P(x,ξ)+χ(x,ξ,τ)P(x,ξ).P(x,\xi)=p_{1}(x,\xi,\tau)+p_{2}(x,\xi,\tau)=(1-\chi(x,\xi,\tau))P(x,\xi)+\chi(x,\xi,\tau)P(x,\xi).

It is clear that p1(x,ξ,τ)S1(AG+)p_{1}(x,\xi,\tau)\in S^{1}(A^{*}G_{+}), so that the corresponding operator

P1(t,γ)=ei<κ(γ),ξ>+i<t,τ>p1(r(γ),ξ,τ)𝑑ξ𝑑τμs1/2(γ)μr1/2(γ)ΨG+1.P_{1}(t,\gamma)=\int e^{i<\kappa(\gamma),\xi>+i<t,\tau>}p_{1}(r(\gamma),\xi,\tau)d\xi d\tau\,\mu^{1/2}_{s}(\gamma)\mu^{1/2}_{r}(\gamma)\in\Psi^{1}_{G_{+}}.

Moreover, in the neighbourhood of rΓ+(Λ)r_{\Gamma_{+}}(\Lambda), one has that χ(x,ξ,τ)=0\chi(x,\xi,\tau)=0, because of (129) and hence the symbol of P1P_{1} is the symbol of PP.
Now the wave front of the distribution :

P2(t,γ)=ei<κ(γ),ξ>+i<t,τ>χ(r(γ),ξ,τ)P(r(γ),ξ)𝑑ξ𝑑τμs1/2(γ)μr1/2(γ)P_{2}(t,\gamma)=\int e^{i<\kappa(\gamma),\xi>+i<t,\tau>}\chi(r(\gamma),\xi,\tau)P(r(\gamma),\xi)d\xi d\tau\,\mu^{1/2}_{s}(\gamma)\mu^{1/2}_{r}(\gamma)

is such that if (t,τ,λ)TG+0(t,\tau,\lambda)\in T^{*}G_{+}\setminus 0 and sΓ(λ)=(x,ξ)AG0s_{\Gamma}(\lambda)=(x,\xi)\in A^{*}G\setminus 0 , rΓ(λ)=(y,η)AG0r_{\Gamma}(\lambda)=(y,\eta)\in A^{*}G\setminus 0, then (t,τ,λ)WFP2max(|ξ||τ|,|η||τ|)ε.(t,\tau,\lambda)\in\operatorname{WF}{P_{2}}\implies\max(\frac{|\xi|}{|\tau|},\frac{|\eta|}{|\tau|})\leq\varepsilon. This implies in particular that WFP2.Λ=.\operatorname{WF}{P_{2}}.\Lambda=\emptyset.

To conclude this second approach, note that the principal symbol of (t+iP1)(\frac{\partial}{\partial t}+iP_{1}) is equal to τ+p0\tau+p_{0} in a neighboorhood of rΓ+(Λ)r_{\Gamma_{+}}(\Lambda) and vanishes on rΓ+(Λ)r_{\Gamma_{+}}(\Lambda), because (τ+p0)rΓ+=τ+p0rΓ(\tau+p_{0})\circ r_{\Gamma_{+}}=\tau+p_{0}\circ r_{\Gamma} vanishes on Λ\Lambda. Thus we may apply Proposition 15 with G+G_{+} as underlying groupoid to the operators (t+iP1)(\frac{\partial}{\partial t}+iP_{1}) and UU, which allows to recover the formula (122) for the principal symbol of their product by remarking that the subprincipal symbol of P1P_{1} is also equal to the subprincipal of PP in a neighboorhood of rΓ+(Λ)r_{\Gamma_{+}}(\Lambda).

The rest of the proof is essentially identical to the proof of [12, Theorem 29.1.1]. Indeed the (transport) equation

(130) {(t+p0+iσ1s(P))u0=0u0(0,.)=1\begin{cases}(\frac{\partial}{\partial t}+\mathcal{L}_{p^{0}}+i\sigma^{1s}(P))u^{0}=0\\ u^{0}(0,.)=1\end{cases}

has a unique solution u0C(Λ)u^{0}\in C^{\infty}(\Lambda), and u0u^{0} homogeneous of degree 0 with respect to the +{\mathbb{R}}_{+} action on each Λt\Lambda_{t}. Let us fix a {\mathbb{R}}-proper set 𝒱×G\mathcal{V}\subset{\mathbb{R}}\times G such that supp(Λ)𝒱{\mathrm{supp}(\Lambda)}\subset\overset{\circ}{\mathcal{V}}. choose U0I(n(1)n(0)1)/4(×G,Λ,Ω1/2)U^{0}\in I^{(n^{(1)}-n^{(0)}-1)/4}({\mathbb{R}}\times G,\Lambda,\Omega^{1/2}) with principal symbol u0u^{0} and support in 𝒱\mathcal{V}. Note that U0(0)ΨG,c0U^{0}(0)\in\Psi^{0}_{G,c} because Λ0=AG0\Lambda_{0}=A^{*}G\setminus 0. It follows that:

(131) IU0(0)ΨG,c1 and (t+iP)U0=F1I1+(n(1)n(0)1)/4(×G,Λ,Ω1/2).I-U^{0}(0)\in\Psi^{-1}_{G,c}\text{ and }(\frac{\partial}{\partial t}+iP)U^{0}=F^{1}\in I^{-1+(n^{(1)}-n^{(0)}-1)/4}({\mathbb{R}}\times G,\Lambda,\Omega^{1/2}).

Next one chooses U1I1+(n(1)n(0)1)/4(×G,Λ,Ω1/2)U^{1}\in I^{-1+(n^{(1)}-n^{(0)}-1)/4}({\mathbb{R}}\times G,\Lambda,\Omega^{1/2}) with support in 𝒱\mathcal{V} and principal symbol u1u^{1} solving the transport equation

(132) {(t+p0+iσ1s(P))u1=f1u1(0,.)=σ(IU0(0))\begin{cases}(\frac{\partial}{\partial t}+\mathcal{L}_{p^{0}}+i\sigma^{1s}(P))u^{1}=-f^{1}\\ u^{1}(0,.)=\sigma(I-U^{0}(0))\end{cases}

and so on. We construct in this way a sequence UjIj+(n(1)n(0)1)/4(×G,Λ,Ω1/2)U^{j}\in I^{-j+(n^{(1)}-n^{(0)}-1)/4}({\mathbb{R}}\times G,\Lambda,\Omega^{1/2}). Finally we choose UI(n(1)n(0)1)/4(×G,Λ,Ω1/2)U\in I^{(n^{(1)}-n^{(0)}-1)/4}({\mathbb{R}}\times G,\Lambda,\Omega^{1/2}) with support in 𝒱\mathcal{V} such that:

UUj.U\sim\sum U^{j}.

By construction, we get

(133) R:=(t+iP)UC(×G) and S:=IdU(0)Cc(G).R:=(\frac{\partial}{\partial t}+iP)U\in C^{\infty}({\mathbb{R}}\times G)\quad\text{ and }\quad S:=\operatorname{Id}-U(0)\in C^{\infty}_{c}(G).

Modifying UU into U+φSU+\varphi S with φCc()\varphi\in C^{\infty}_{c}({\mathbb{R}}) and φ(0)=1\varphi(0)=1, we can directly assume that U(0)=IdU(0)=\operatorname{Id}. Also, the support of RR is contained in

𝒱=𝒱(×supp(P))G~𝒱=𝒱{(t,γ);γsupp(P)𝒱t}.\mathcal{V}^{\prime}=\mathcal{V}\cup({\mathbb{R}}\times{\mathrm{supp}(P)})\cdot_{\widetilde{G}}\mathcal{V}=\mathcal{V}\cup\{(t,\gamma)\ ;\ \gamma\in{\mathrm{supp}(P)}\cdot\mathcal{V}_{t}\}.

The set 𝒱\mathcal{V}^{\prime} is again {\mathbb{R}}-proper. This implies:

RC(,Cc(G))C(,)R\in C^{\infty}({\mathbb{R}},C^{\infty}_{c}(G))\subset C^{\infty}({\mathbb{R}},\mathcal{H}^{\infty})

We obtain, using (27) and following verbatim the proof of [12, Theorem 29.1.1]

(134) U(t)eitP=R~(t):=i0tei(ts)PR(s)𝑑sU(t)-e^{-itP}=\tilde{R}(t):=i\int_{0}^{t}e^{i(t-s)P}R(s)ds

Using the results of Section 3, we get R~C(,)\tilde{R}\in C^{\infty}({\mathbb{R}},\mathcal{H}^{\infty}), which ends the proof. ∎

The previous theorem is only stated for compactly supported operators, but it admits the following slight generalization:

Corollary 24.

Let T=Pc+SΨG1T=P_{c}+S\in\Psi^{1}_{G}, with PcΨG,c1P_{c}\in\Psi^{1}_{G,c} satisfying the assumption of Theorem 21 and SS\in\mathcal{H}^{\infty}. There exists UI14+(n(1)n(0))/4(×G,Λ;Ω1/2)U\in I^{-\frac{1}{4}+(n^{(1)}-n^{(0)})/4}({\mathbb{R}}\times G,\Lambda;\Omega^{1/2}) with {\mathbb{R}}-proper support such that

(135) (t+iT)UC(,).(\frac{\partial}{\partial t}+iT)U\in C^{\infty}({\mathbb{R}},\mathcal{H}^{\infty}).
Proof.

Apply Theorem 21 to PcP_{c} and let UI14+(n(1)n(0))/4(×G,Λ;Ω1/2)U\in I^{-\frac{1}{4}+(n^{(1)}-n^{(0)})/4}({\mathbb{R}}\times G,\Lambda;\Omega^{1/2}) be the corresponding parametrix. Then

(136) (t+iT)U=iSU+R,RC(,).(\frac{\partial}{\partial t}+iT)U=iSU+R,\qquad R\in C^{\infty}({\mathbb{R}},\mathcal{H}^{\infty}).

Using the continuity theorems for GG-FIO [18], one gets that for any tt, UtU_{t} acts continuously on the scale of Sobolev modules, which immediately implies that SUC(,)SU\in C^{\infty}({\mathbb{R}},\mathcal{H}^{\infty}). ∎

As examples of situations into which Theorem 21 and Corollary 24 apply, we mention:

  1. (1)

    The pair groupoid G=X×XXG=X\times X\rightrightarrows X of a compact manifold without boundary XX. Since XX itself is an orbit, we have Corb,0(G)=C(X×X,Ω1/2)C^{\infty,0}_{\mathop{\mathrm{orb}}\nolimits}(G)=C^{\infty}(X\times X,\Omega^{1/2}) and we just recover the classical result (see [12, Theorem 29.1.1] for instance), after the obvious identification between GG-operators and continuous linear operators C(X,ΩX1/2)C(X,ΩX1/2)C^{\infty}(X,\Omega^{1/2}_{X})\to C^{\infty}(X,\Omega^{1/2}_{X}).

  2. (2)

    The holonomy groupoid GG of a compact foliated manifold XX. We recover the construction of the leafwise geometrical optic approximation of eitPe^{itP} given in [15]. The latter is worked out for small time and by solving eikonal equations to find the required phases in local coordinates as well as by solving transport equations. Our construction here can be viewed as a complement, available for arbitrary time and regarding the evolution of singularities as well as the kind of Fourier integral operators involved in the problem.

  3. (3)

    G{e}G\rightrightarrows\{e\} a Lie group. Here again, there is only one orbit so Corb,0(G)=C(G)C^{\infty,0}_{\mathop{\mathrm{orb}}\nolimits}(G)=C^{\infty}(G). The result applies for instance to the square root Δ\sqrt{\Delta} of any elliptic laplacian Δ=Xj2DiffG2\Delta=-\sum X_{j}^{2}\in\mathop{\mathrm{Diff}}\nolimits^{2}_{G}, viewed as right invariant operators on GG. That Δ=Pc+SΨG1\sqrt{\Delta}=P_{c}+S\in\Psi^{1}_{G} with σG0(Pc)=ξj2\sigma_{G}^{0}(P_{c})=\sqrt{\sum\xi_{j}^{2}} follows from [32] and we get here the existence of a CC^{\infty} family UtU_{t} of right invariant FIOFIO on GG [27, 18] such that (t+iΔ)UtC(G)Cr(G)(\frac{\partial}{\partial t}+i\sqrt{\Delta})U_{t}\in C^{\infty}(G)\cap C^{*}_{r}(G) for any tt.

  4. (4)

    The groupoid GbXG_{b}\rightrightarrows X of the bb-calculus of a manifold with embedded corners XX [25]. We recall that GbG_{b} is the open submanifold with corners of the bb-stretched product of R. Melrose Xb2X^{2}_{b} in which all the lateral faces are removed. Identifying GbG_{b}-operators with pseudodifferential operators in the bb-calculus, and their restrictions at boundary hypersurfaces with indicial operators, we get for any elliptic symmetric PΨb1(X)P\in\Psi^{1}_{b}(X) in the small calculus the existence of a CC^{\infty} family UtU_{t} of bb-FIO on XX [22, 18] such that

    (137) (t+iP)Ut=RtC((XX)2)(Lb2(X)).(\frac{\partial}{\partial t}+iP)U_{t}=R_{t}\in C^{\infty}((X\setminus\partial X)^{2})\cap\mathcal{L}(L^{2}_{b}(X)).

    and for any boundary hypersurfaces HH (with normal bundle trivialized with a boundary defining function):

    IH(Rt)=(t+iIH(P))IH(Ut)C(H2×)(Lb2(H×)).I_{H}(R_{t})=(\frac{\partial}{\partial t}+iI_{H}(P))I_{H}(U_{t})\in C^{\infty}(H^{2}\times{\mathbb{R}})\cap\mathcal{L}(L^{2}_{b}(H\times{\mathbb{R}})).

    The error term RtR_{t} is C0C^{0} on GbG_{b} and there is no reason neither to expect that it is CC^{\infty} on GbG_{b}, nor that it extends continuously to Xb2X^{2}_{b}.

  5. (5)

    This discussion is similar to the previous one for the groupoid GπXG_{\pi}\rightrightarrows X [8] and its associated pseudodifferential calculus, where XX is a manifold with iterated fibred corners. In both cases, the regularity result that we reach for the error term RR is likely not optimal. This will be investigated, among other applications to singular spaces, in future works.

As far as we know, examples (3–5) above are new.

References

  • [1] I. Androulidakis and G. Skandalis. The analytic index of elliptic pseudodifferential operators on a singular foliation. J. K-Theory, 8(3):363–385, 2011.
  • [2] S. Baaj. Multiplicateurs non bornés. PhD thesis, Université Pierre et Marie Curie, Paris., 1980.
  • [3] S. Baaj and P. Julg. Théorie bivariante de Kasparov et opérateurs non bornés dans les CC^{\ast}-modules hilbertiens. C. R. Acad. Sci. Paris Sér. I Math., 296(21):875–878, 1983.
  • [4] A. Connes. Sur la théorie non commutative de l’intégration. In Algèbres d’opérateurs (Sém., Les Plans-sur-Bex, 1978), volume 725 of Lecture Notes in Math., pages 19–143. Springer, Berlin, 1979.
  • [5] A. Connes. Noncommutative Geometry. Academic Press, San Diego, CA, 1994.
  • [6] A. Coste, P. Dazord, and A. Weinstein. Groupoïdes symplectiques. Publications du Dep. de Maths. de l’Univ. de Lyon 1, 2/A, 1987.
  • [7] M. Crainic and J. Mestre. Measures on differentiable stacks. J. Noncommut. Geom., 13(4):1235–1294, 2019.
  • [8] C. Debord, J.-M. Lescure, and F. Rochon. Pseudodifferential operators on manifolds with fibred corners. Ann. Inst. Fourier, 65(3):1799–1880, 2015.
  • [9] C. Debord and G. Skandalis. Pseudodifferential extensions and adiabatic deformation of smooth groupoid actions. Bull. Sci. Math., 139(7):750–776, 2015.
  • [10] J. Dixmier and P. Malliavin. Factorisations de fonctions et de vecteurs indéfiniment différentiables. Bull. Sci. Math. (2), 102(4):307–330, 1978.
  • [11] L. Hörmander. Fourier integral operators. I. Acta Math., 127(1-2):79–183, 1971.
  • [12] L. Hörmander. The analysis of linear partial differential operators. I–IV. Classics in Mathematics. Springer, 2003–2009.
  • [13] K. Jensen and K. Thomsen. Elements of KKKK-theory. Mathematics: Theory & Applications. Birkhäuser Boston, Inc., Boston, MA, 1991.
  • [14] M. Khoshkam and G. Skandalis. Regular representation of groupoid CC^{*}-algebras and applications to inverse semigroups. J. Reine Angew. Math., 546:47–72, 2002.
  • [15] Yuri Kordyukov. Functional calculus for tangentially elliptic operators on foliated manifolds. In Analysis and geometry in foliated manifolds (Santiago de Compostela, 1994), pages 113–136. World Sci. Publ., River Edge, NJ, 1995.
  • [16] R. Lauter, B. Monthubert and V. Nistor Spectral invariance for certain algebras of pseudodifferential operators. J. Inst. Math. Jussieu, 3(4):405–442, 2005.
  • [17] J.-M. Lescure, D. Manchon, and S. Vassout. About the convolution of distributions on groupoids. J. Noncommut. Geom., 11(2):757–789, 2017.
  • [18] J.-M. Lescure and S. Vassout. Fourier integral operators on Lie groupoids. Adv. Math., 320:391–450, 2017. Reprinted in J. Math. Sci. (N.Y.) 236 (2019), no. 3, 367–375.
  • [19] K. Mackenzie. General theory of Lie groupoids and Lie algebroids, volume 213 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 2005.
  • [20] R. Mazzeo and R. Melrose. Pseudodifferential operators on manifolds with fibred boundaries. Asian J. Math., 2(4):833–866, 1999.
  • [21] A., Mazzucato and V. Nistor. Mapping properties of heat kernels, maximal regularity, and semi-linear parabolic equations on noncompact manifolds. J. Hyperbolic Differ. Equ., 3(4):599–629, 2006.
  • [22] R. Melrose. Transformation of boundary problems. Acta Math., 147(3-4):149–236, 1981.
  • [23] R. Melrose and P. Piazza. Analytic K-theory for manifolds with corners. Adv. in Math, 92:1–27, 1992.
  • [24] I. Moerdijk and J. Mrčun. Introduction to foliations and Lie groupoids, volume 91 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2003.
  • [25] B. Monthubert. Groupoids and pseudodifferential calculus on manifolds with corners. J. Funct. Anal., 199(1):243–286, 2003.
  • [26] B. Monthubert and F. Pierrot. Indice analytique et groupoïde de Lie. C.R.A.S Série 1, 325:193–198, 1997.
  • [27] B. Nielsen and H. Stetkaer. Invariant Fourier integral operators on Lie groups. Math. Scand., 35:193–210, 1974.
  • [28] V. Nistor, A. Weinstein, and P. Xu. Pseudodifferential operators on differential groupoids. Pacific J. of Math., 181(1):117–152, 1999.
  • [29] J. Renault. A groupoid approach to CC^{\ast}-algebras, volume 793 of Lecture Notes in Mathematics. Springer, Berlin, 1980.
  • [30] G. Skandalis. C-algèbres, Algèbres de Von Neumann, Exemples. Cours de M2, 2015. https://webusers.imj-prg.fr/ georges.skandalis/poly2015.pdf.
  • [31] E. van Erp and R. Yuncken. A groupoid approach to pseudodifferential calculi. J. Reine Angew. Math., 756:151–182, 2019.
  • [32] S. Vassout. Unbounded pseudodifferential calculus on Lie groupoids. J. Funct. Anal., 236(1):161–200, 2006.
  • [33] N. Wegge-Olsen. KK-theory and CC^{*}-algebras. Oxford Science Publications. The Clarendon Press Oxford University Press, New York, 1993. A friendly approach.