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

Differential operators on CC^{*}-algebras and applications to smooth functional calculus and Schwartz functions on the tangent groupoid

Omar Mohsen
Abstract

We introduce the notion of a differential operator on CC^{*}-algebras. This is a noncommutative analogue of a differential operator on a smooth manifold. We show that the common closed domain of all differential operators is closed under smooth functional calculus. As a corollary, we show that Schwartz functions on Connes tangent groupoid are closed under smooth functional calculus.

Introduction

Let MM be a compact smooth manifold. In [Con94a, Chapter 2.5], Connes introduced the tangent groupoid

𝕋M:=M×M×+×TM×{0}.\mathbb{T}M:=M\times M\times\mathbb{R}_{+}^{\times}\sqcup TM\times\{0\}.

The space 𝕋M\mathbb{T}M is equipped with a smooth manifold structure which is a special case of the deformation to the normal cone construction, see [Ful84a, Chapter 5]. The space Cc(𝕋M)C^{\infty}_{c}(\mathbb{T}M) is equipped with a convolution law making it a *-algebra. The convolution law is a deformation of the usual convolution law on M×MM\times M which is used in Schwartz kernels of operators acting on L2ML^{2}M and the convolution law on TxMT_{x}M for xMx\in M which comes from the commutative group structure. As Connes shows, the space 𝕋M\mathbb{T}M very naturally gives a canonical deformation from a differential operator to its principal symbol. Connes used this deformation to give a conceptual proof of the Atiyah-Singer index theorem [AS68a].

The tangent groupoid has been studied and generalised by many authors. In this article, we will be interested in the space of Schwartz functions 𝒮(𝕋M)\mathcal{S}(\mathbb{T}M) on 𝕋M\mathbb{T}M introduced by Carrillo-Rouse [Car08a]. The *-algebra Cc(𝕋M)C^{\infty}_{c}(\mathbb{T}M) can be naturally completed into a CC^{*}-algebra C𝕋MC^{*}\mathbb{T}M. In [Car08a], Carrillo-Rouse proves that 𝒮(𝕋M)\mathcal{S}(\mathbb{T}M) is a *-subalgebra of C𝕋MC^{*}\mathbb{T}M. In this article, we improve his result by showing that 𝒮(𝕋M)\mathcal{S}(\mathbb{T}M) is closed under smooth functional calculus.

Theorem A.
  1. 1.

    If a𝒮(𝕋M)a\in\mathcal{S}(\mathbb{T}M) is a Schwartz function and aa is normal (aa=aaa^{*}a=aa^{*}) and ff is a smooth function on some open neighbourhood of spec(a)\mathrm{spec}(a) with f(0)=0f(0)=0, then f(a)𝒮(𝕋M)f(a)\in\mathcal{S}(\mathbb{T}M).

  2. 2.

    If a𝒮(𝕋M)a\in\mathcal{S}(\mathbb{T}M) is a Schwartz function and ff is a holomorphic function on some open neighbourhood of spec(a)\mathrm{spec}(a) with f(0)=0f(0)=0, then f(a)𝒮(𝕋M)f(a)\in\mathcal{S}(\mathbb{T}M).

Of course f(a)f(a) in Theorem A means the functional calculus in C𝕋MC^{*}\mathbb{T}M applied to aa. The main difficulty in Theorem A is the derivability of f(a)f(a) in the +\mathbb{R}_{+} direction of 𝕋M\mathbb{T}M because as t0+t\to 0^{+}, the structure of C𝕋MC^{*}\mathbb{T}M changes from compact operators on L2ML^{2}M to C0(TM)C_{0}(T^{*}M). We remark that we suppose f(0)=0f(0)=0 because C𝕋MC^{*}\mathbb{T}M isn’t unital.

To prove Theorem A we introduce differential operators on an arbitrary CC^{*}-algebra AA. Recall that a derivation on AA is a map δ:𝒜A\delta:\mathcal{A}\to A defined on a *-subalgebra 𝒜A\mathcal{A}\subseteq A such that

δ(ab)=δ(a)b+aδ(b),a,b𝒜.\delta(ab)=\delta(a)b+a\delta(b),\quad\forall a,b\in\mathcal{A}.

Derivations usually come from the derivative of an \mathbb{R} action on AA. If A=C(N)A=C(N) and 𝒜=C(N)\mathcal{A}=C^{\infty}(N) where NN is a smooth manifold NN, then a derivation is a vector field on NN. Therefore, one can think of derivations as differential operators of order 11 on AA with domain 𝒜\mathcal{A}. We introduce the notion of a differential operator of higher order on an arbitrary CC^{*}-algebra AA. Our differential operators coincide with differential operators on NN if A=C(N)A=C(N) and 𝒜=C(N)\mathcal{A}=C^{\infty}(N). We show that the derivative in the direction of +\mathbb{R}_{+} is a differential operator on C𝕋MC^{*}\mathbb{T}M of order 22. The main theorem of this article is the following, see Definition 1.8 for the precise definition of 𝒟(A)A\mathcal{D}(A)\subseteq A. We remark that 𝒟(A)\mathcal{D}(A) is a *-subalgebra of AA which contains 𝒜\mathcal{A}.

Theorem B.

Let 𝒜A\mathcal{A}\subseteq A be a dense *-subalgebra and 𝒟(𝒜)\mathcal{D}(\mathcal{A}) be the intersection of the domains of the closures of all differential operators defined on 𝒜\mathcal{A}. Then

  1. 1.

    If nn\in\mathbb{N}, aMn(𝒟(𝒜))a\in M_{n}(\mathcal{D}(\mathcal{A})) is normal (aa=aaa^{*}a=aa^{*}) and ff is a smooth function defined on some open neighbourhood of spec(a)\mathrm{spec}(a) with f(0)=0f(0)=0, then f(a)Mn(𝒟(𝒜))f(a)\in M_{n}(\mathcal{D}(\mathcal{A})).

  2. 2.

    If nn\in\mathbb{N}, aMn(𝒟(𝒜))a\in M_{n}(\mathcal{D}(\mathcal{A})) and ff is a holomorphic function defined on some open neighbourhood of spec(a)\mathrm{spec}(a) with f(0)=0f(0)=0, then f(a)Mn(𝒟(𝒜))f(a)\in M_{n}(\mathcal{D}(\mathcal{A})).

Here spec(a)\mathrm{spec}(a) is the spectrum of aa as an element of Mn(A)M_{n}(A), the CC^{*}-algebra of n×nn\times n matrices with coefficients in AA.

The methods we use to prove B are based on arguments by Bratteli and Robinson [BR75a, BR76a] and Pedersen [Ped00a]. We obtain Theorem A as a corollary of Theorem B because 𝒟(𝒮(TM))=𝒮(TM)\mathcal{D}(\mathcal{S}(TM))=\mathcal{S}(TM).

The main interest in subalgebras closed under holomorphic calculus in the theory of operator algebras is that they give isomorphisms in KK-theory, thus allowing tools from cyclic cohomology to be used. Cyclic cocycles are rarely defined on the whole CC^{*}-algebra and only defined on a subalgebra, see for example [Con86a]

We remark that our approach to Schwartz functions is different from that of Carrillo-Rouse, and our algebra differs slightly from his algebra, see Remark 2.10. His approach is based on using local coordinates for 𝕋M\mathbb{T}M and demanding growth conditions in such coordinates. He then shows that this is independent of local coordinates. Our approach is more global in nature. Recall that Schwartz functions on a vector space are smooth functions ff such that D(f)D(f) is bounded whenever DD is a differential operator with polynomial coefficients. Analogously, on the space 𝕋M\mathbb{T}M, we define differential operators which play the role of differential operators with polynomial coefficients. We then define Schwartz functions to be functions whose derivative by such operators is bounded.

Acknowledgments

We thank the referee for his valuable remarks which helped immensely with the exposition.

Organization of the article.

The article is organized as follows.

  • In Section 1, we define differential operators on CC^{*}-algebras and prove Theorem B.

  • In Section 2, we give a quick introduction to Connes’s tangent groupoid. We then prove Theorem A.

1 Differential operators on CC^{*}-algebras

Throughout the article \mathbb{N} stands for {1,2,}\{1,2,\cdots\}, +×\mathbb{R}_{+}^{\times} for {x:x>0}\{x\in\mathbb{R}:x>0\} and +\mathbb{R}_{+} for {x:x0}\{x\in\mathbb{R}:x\geq 0\}.

Example 1.1.

Let

:=3,(x,y,z)(x,y,z)=(x+x,y+y,z+z+xyyx)\mathbb{H}:=\mathbb{R}^{3},\quad(x,y,z)\cdot(x^{\prime},y^{\prime},z^{\prime})=(x+x^{\prime},y+y^{\prime},z+z^{\prime}+xy^{\prime}-yx^{\prime})

be the Heisenberg group and

δx,δy,δz:Cc()Cc(),δx(f)=xf,δy(f)=yf,δz(f)=zf.\displaystyle\delta_{x},\delta_{y},\delta_{z}:C^{\infty}_{c}(\mathbb{H})\to C^{\infty}_{c}(\mathbb{H}),\quad\delta_{x}(f)=xf,\quad\delta_{y}(f)=yf,\quad\delta_{z}(f)=zf. (1)

The maps δx\delta_{x} and δy\delta_{y} satisfy the Leibniz rule

δx(fg)=δx(f)g+fδx(g),δy(fg)=δy(f)g+fδy(g)\displaystyle\delta_{x}(f\star g)=\delta_{x}(f)\star g+f\star\delta_{x}(g),\quad\delta_{y}(f\star g)=\delta_{y}(f)\star g+f\star\delta_{y}(g)

but δz\delta_{z} satisfies

δz(fg)=δz(f)g+fδz(g)+δx(f)δy(g)δy(f)δx(g),\displaystyle\delta_{z}(f\star g)=\delta_{z}(f)\star g+f\star\delta_{z}(g)+\delta_{x}(f)\star\delta_{y}(g)-\delta_{y}(f)\star\delta_{x}(g), (2)

where \star denotes the convolution. We define differential operators on noncommutative algebras to be maps which satisfy a Leibniz rule as in (2).

Definition 1.2.

Let 𝒜\mathcal{A} be a \mathbb{C}-algebra equipped with an involution, 𝒟End(𝒜)\mathcal{D}\subseteq\operatorname{End}(\mathcal{A}) a \mathbb{C}-subalgebra of \mathbb{C}-linear endomorphisms of 𝒜\mathcal{A} which contains the identity.

  1. 1.

    We say that δ𝒟\delta\in\mathcal{D} is a 𝒟\mathcal{D}-differential operator of order 0 if δ(ab)=δ(a)b\delta(ab)=\delta(a)b for all a,b𝒜a,b\in\mathcal{A}.

  2. 2.

    We define 𝒟\mathcal{D}-differential operators of higher order by recurrence. A map δ𝒟\delta\in\mathcal{D} is called a 𝒟\mathcal{D}-differential operator of order nn\in\mathbb{N}, if there exists a finite family of linear maps

    δ10,,δ1r,δ20,,δ2r𝒟\delta_{10},\cdots,\delta_{1r},\delta_{20},\cdots,\delta_{2r}\in\mathcal{D}

    which are 𝒟\mathcal{D}-differential operators of order <n<n such that

    δ(ab)=δ(a)δ20(b)+δ10(a)δ(b)+i=1rδ1i(a)δ2i(b).\displaystyle\delta(ab)=\delta(a)\delta_{20}(b)+\delta_{10}(a)\delta(b)+\sum_{i=1}^{r}\delta_{1i}(a)\delta_{2i}(b). (3)

    Furthermore, we suppose that

    δ20(x)=δ20(x),x𝒜\displaystyle\delta_{20}(x^{*})=\delta_{20}(x)^{*},\quad\forall x\in\mathcal{A} (4)

We denote by Diff𝒟(𝒜)𝒟\operatorname{\mathrm{Diff}_{\mathcal{D}}}(\mathcal{A})\subseteq\mathcal{D} the space of 𝒟\mathcal{D}-differential operators of any order.

Remarks 1.3.

Let us explain Definition 1.2 in more details.

  1. 1.

    We use a subspace 𝒟\mathcal{D} of linear maps instead of the space of all linear endomorphisms of 𝒜\mathcal{A}, because in applications, it seems unhelpful to consider all linear endomorphisms. Usually, one has a given subspace of linear endomorphisms of interest.

  2. 2.

    Condition (4) is included because our main theorem fails without it. It plays an absolutely necessary role in the proof of Lemma 1.15. In all our applications (Theorem A and Corollary 1.18), the 𝒟\mathcal{D}-differential operators we consider have the property δ20=cId\delta_{20}=c\mathrm{Id} with c{0,1}c\in\{0,1\} and thus trivially satisfy (4). We include the general case of arbitrary δ20\delta_{20} satisfying (4) because the arguments in this section don’t change with its inclusion.

  3. 3.

    It is worth nothing that if δ\delta is a 𝒟\mathcal{D}-differential operator of order nn, then it is also a 𝒟\mathcal{D}-differential operator of order mm for any m>nm>n. It might be tempting to define the order of a 𝒟\mathcal{D}-differential operator δ\delta to be the minimal nn such that δ\delta is of order nn. We won’t do so because the order behaves in a rather unintuitive way. For example, the sum of a 𝒟\mathcal{D}-differential operator of order nn and another of order mm isn’t necessarily of order max(n,m)\max(n,m). It is always of order max(n,m)+1\max(n,m)+1. See the proof of Proposition 1.5.

Remark 1.4.

The following observation will be used in the proof of Proposition 1.5. If δ𝒟\delta\in\mathcal{D} such that there exists 𝒟\mathcal{D}-differential operators δ10,,δ1r,δ20,,δ2r𝒟\delta_{10},\cdots,\delta_{1r},\delta_{20},\cdots,\delta_{2r}\in\mathcal{D} such that (3) and (4) hold, then δ\delta is a 𝒟\mathcal{D}-differential operator of order one plus the maximum of the orders of δ10,,δ1r,δ20,,δ2r\delta_{10},\cdots,\delta_{1r},\delta_{20},\cdots,\delta_{2r}. Whenever we apply this remark, to make it clearer how it is applied, we colored the term δ(a)δ20(b)\delta(a)\delta_{20}(b) by red, the term δ10(a)δ(b)\delta_{10}(a)\delta(b) by green, and the rest by blue.

Proposition 1.5.

The space Diff𝒟(𝒜)\operatorname{\mathrm{Diff}_{\mathcal{D}}}(\mathcal{A}) is a unital subalgebra of 𝒟\mathcal{D} and thus of End(𝒜)\operatorname{End}(\mathcal{A}) as well.

Proof.

It is clear that Id\operatorname{Id} is a 𝒟\mathcal{D}-differential operator of order 0. Let δ\delta and δ\delta^{\prime} be 𝒟\mathcal{D}-differential operators of order nn and mm respectively with δ10,,δ1r,δ20,,δ2r\delta_{10},\cdots,\delta_{1r},\delta_{20},\cdots,\delta_{2r} and δ10,,δ1r,δ20,,δ2r\delta_{10}^{\prime},\cdots,\delta_{1r}^{\prime},\delta_{20}^{\prime},\cdots,\delta_{2r}^{\prime} as in Definition 1.2. Then,

δ(ab)+δ(ab)\displaystyle\delta(ab)+\delta^{\prime}(ab) =δ(a)δ20(b)+δ10(a)δ(b)+i=1rδ1i(a)δ2i(b)\displaystyle=\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}{\delta(a)\delta_{20}(b)}\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\delta_{10}(a)\delta(b)\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\sum_{i=1}^{r}\delta_{1i}(a)\delta_{2i}(b)
+δ(a)δ20(b)+δ10(a)δ(b)+i=1rδ1i(a)δ2i(b).\displaystyle\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\delta^{\prime}(a)\delta_{20}^{\prime}(b)\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\delta_{10}^{\prime}(a)\delta^{\prime}(b)\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\sum_{i=1}^{r^{\prime}}\delta^{\prime}_{1i}(a)\delta^{\prime}_{2i}(b).

Hence, δ+δ\delta+\delta^{\prime} is a 𝒟\mathcal{D}-differential operator of order max(n,m)+1\max(n,m)+1 by Remark 1.4. Notice here that there are only blue terms. This is possible because, one already supposes that δ\delta and δ\delta^{\prime} are 𝒟\mathcal{D}-differential operators.

We prove that δδ\delta\circ\delta^{\prime} is a 𝒟\mathcal{D}-differential operator by induction on the order. The case n=m=0n=m=0 is clear. We prove the case m=0m=0 by induction on nn. We have

δ(δ(ab))=δ(δ(a)b)=δ(δ(a))δ20(b)+δ10(δ(a))δ(b)+i=1rδ1i(δ(a))δ2i(b)\displaystyle\delta(\delta^{\prime}(ab))=\delta(\delta^{\prime}(a)b)=\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}\delta(\delta^{\prime}(a))\delta_{20}(b)\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\delta_{10}(\delta^{\prime}(a))\delta(b)\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\sum_{i=1}^{r}\delta_{1i}(\delta^{\prime}(a))\delta_{2i}(b) (5)

All the terms δ10δ,,δ1rδ\delta_{10}\circ\delta^{\prime},\cdots,\delta_{1r}\circ\delta^{\prime} are 𝒟\mathcal{D}-differential operators by recurrence. Hence, by Remark 1.4, δδ\delta\circ\delta^{\prime} is also a 𝒟\mathcal{D}-differential operator.

The case n=0n=0 is also proved by induction on mm. We have

δ(δ(ab))=δ(δ(a))δ20(b)+δ(δ10(a))δ(b)+i=1rδ(δ1i(a))δ2i(b).\displaystyle\delta(\delta^{\prime}(ab))=\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}\delta(\delta^{\prime}(a))\delta^{\prime}_{20}(b)\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\delta(\delta^{\prime}_{10}(a))\delta^{\prime}(b)\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\sum_{i=1}^{r^{\prime}}\delta(\delta^{\prime}_{1i}(a))\delta^{\prime}_{2i}(b). (6)

All the terms δδ10,,δδ1r\delta\circ\delta^{\prime}_{10},\cdots,\delta\circ\delta^{\prime}_{1r^{\prime}} are 𝒟\mathcal{D}-differential operators by recurrence. Hence, by Remark 1.4, δδ\delta\circ\delta^{\prime} is again a 𝒟\mathcal{D}-differential operator.

We can now prove the general case of the composition of 𝒟\mathcal{D}-differential operators by induction on (n,m)(n,m). We suppose that it is true for any n,m{0}n^{\prime},m^{\prime}\in\mathbb{N}\cup\{0\} such that nnn^{\prime}\leq n and mmm^{\prime}\leq m and n+m<n+mn^{\prime}+m^{\prime}<n+m. One has

δ(δ(ab))\displaystyle\delta(\delta^{\prime}(ab)) =δ(δ(a))δ20(δ20(b))+δ10(δ(a))δ(δ20(b))+i=1rδ1i(δ(a))δ2i(δ20(b)).\displaystyle=\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}\delta(\delta^{\prime}(a))\delta_{20}(\delta_{20}^{\prime}(b))\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\delta_{10}(\delta^{\prime}(a))\delta(\delta_{20}^{\prime}(b))\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\sum_{i=1}^{r}\delta_{1i}(\delta^{\prime}(a))\delta_{2i}(\delta_{20}^{\prime}(b)). (7)
+δ(δ10(a))δ20(δ(b))+δ10(δ10(a))δ(δ(b))+i=1rδ1i(δ10(a))δ2i(δ(b)).\displaystyle\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\delta(\delta_{10}^{\prime}(a))\delta_{20}(\delta^{\prime}(b))\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,1,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,1,0}\delta_{10}(\delta_{10}^{\prime}(a))\delta(\delta^{\prime}(b))\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\sum_{i=1}^{r}\delta_{1i}(\delta_{10}^{\prime}(a))\delta_{2i}(\delta^{\prime}(b)).
+j=1r(δ(δ1j(a))δ20(δ2j(b))+δ10(δ1j(a))δ(δ2j(b))+i=1rδ1i(δ1j(a))δ2i(δ2j(b))).\displaystyle\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\sum_{j=1}^{r^{\prime}}\Big{(}\delta(\delta^{\prime}_{1j}(a))\delta_{20}(\delta^{\prime}_{2j}(b))\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\delta_{10}(\delta^{\prime}_{1j}(a))\delta(\delta^{\prime}_{2j}(b))\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\sum_{i=1}^{r}\delta_{1i}(\delta^{\prime}_{1j}(a))\delta_{2i}(\delta^{\prime}_{2j}(b))\Big{)}.

Again by Remark 1.4, δδ\delta\circ\delta^{\prime} is a 𝒟\mathcal{D}-differential operator.

Finally, since the map aλaa\mapsto\lambda a is a differential operator of order 0 for any λ\lambda\in\mathbb{C}, it follows that λδ\lambda\delta, which is a composition of two 𝒟\mathcal{D}-differential operators, is also a 𝒟\mathcal{D}-differential operator. This finishes the proof that Diff𝒟(𝒜)\operatorname{\mathrm{Diff}_{\mathcal{D}}}(\mathcal{A}) is a unital subalgebra of 𝒟\mathcal{D}.: ∎

Definition 1.6.

We say that a unital subalgebra 𝒟\mathcal{D} of End(𝒜)\operatorname{End}(\mathcal{A}) is stable if 𝒟=Diff𝒟(𝒜)\mathcal{D}=\operatorname{\mathrm{Diff}_{\mathcal{D}}}(\mathcal{A}).

It is clear from Definition 1.2 that for any unital subalgebra 𝒟\mathcal{D} of End(𝒜)\operatorname{End}(\mathcal{A}), Diff𝒟(𝒜)\operatorname{\mathrm{Diff}_{\mathcal{D}}}(\mathcal{A}) is stable.

We now give natural examples of stable subalgebras. Let 𝔤\mathfrak{g} be a nilpotent Lie algebra. The Baker-Campbell-Hausdorff formula is a finite sum because 𝔤\mathfrak{g} is nilpotent. Hence, we can view 𝔤\mathfrak{g} as a nilpotent Lie group. Let 𝒜=Cc(𝔤)\mathcal{A}=C^{\infty}_{c}(\mathfrak{g}). We equip 𝒜\mathcal{A} with the convolution product after fixing some Haar measure on 𝔤\mathfrak{g}. We remark that a Haar measure is equivalently a Lebesgue measure because 𝔤\mathfrak{g} is nilpotent.

Proposition 1.7.

For DD a differential operator on 𝔤\mathfrak{g} with polynomial coefficients, let

δD:𝒜𝒜,δD(f)=D(f).\delta_{D}:\mathcal{A}\to\mathcal{A},\quad\delta_{D}(f)=D(f).

The space 𝒟\mathcal{D} of all maps δD\delta_{D} as DD varies is a stable unital subalgebra of End(𝒜)\operatorname{End}(\mathcal{A}), i.e., δD\delta_{D} is a 𝒟\mathcal{D}-differential operator for every DD.

Proof.

It is clear that 𝒟\mathcal{D} is a unital subalgebra. We now show that it is stable. For X𝔤X\in\mathfrak{g}, let XRX_{R} be the associated right invariant vector field. One easily sees that δXR\delta_{X_{R}} is a 𝒟\mathcal{D}-differential operator of order 0. We can write DD as a sum of terms of the form PX1RXkRPX_{1R}\cdots X_{kR}, where PP is a polynomial and X1R,,XkRX_{1R},\cdots,X_{kR} are right-invariant vector fields. Hence, by Proposition 1.5, it suffices to prove that if D(f)=PfD(f)=Pf, then δD\delta_{D} is a 𝒟\mathcal{D}-differential operator. Again, by Proposition 1.5, we can suppose that PP is linear. Let 𝔤1=𝔤\mathfrak{g}^{1}=\mathfrak{g} and 𝔤n+1=[𝔤n,𝔤]\mathfrak{g}^{n+1}=[\mathfrak{g}^{n},\mathfrak{g}] be the central series of 𝔤\mathfrak{g}. Since 𝔤\mathfrak{g} is nilpotent, there exists kk\in\mathbb{N} such that 𝔤k+1=0\mathfrak{g}^{k+1}=0. We choose a linear decomposition 𝔤=V1Vk\mathfrak{g}=V^{1}\oplus\cdots\oplus V^{k} such that 𝔤n=VnVk\mathfrak{g}^{n}=V^{n}\oplus\cdots\oplus V^{k}. It is enough to consider PViP\in V^{i*} for some ii. We proceed by induction on ii. For i=1i=1, δP\delta_{P} is a 𝒟\mathcal{D}-differential operator of order 11. In fact

δP(fg)=δP(f)g+fδP(g).\displaystyle\delta_{P}(f\star g)=\delta_{P}(f)\star g+f\star\delta_{P}(g). (8)

The induction case follows by using the Baker-Campbell-Hausdorff formula like in (2). ∎

From now on, we suppose that 𝒜\mathcal{A} is a dense *-subalgebra of a CC^{*}-algebra AA.

Definition 1.8.

Let 𝒟\mathcal{D} be a stable unital subalgebra of End(𝒜)\operatorname{End}(\mathcal{A}). We denote by 𝒟(𝒜)\mathcal{D}(\mathcal{A}) the set of aAa\in A such that there exists a sequence (xn)n𝒜(x_{n})_{n\in\mathbb{N}}\in\mathcal{A} such that xnax_{n}\to a and for every δ𝒟\delta\in\mathcal{D}, δ(xn)\delta(x_{n}) and δ(xn)\delta(x_{n}^{*}) converge in AA.

By taking constant sequences, we see that 𝒜𝒟(𝒜)\mathcal{A}\subseteq\mathcal{D}(\mathcal{A}).

Remark 1.9.

Let δ𝒟\delta\in\mathcal{D}. Since δ\delta is generally an unbounded operator, it is desirable to take the closure of the graph of δ\delta. Some care should be taken, as there exists derivations which aren’t closable. See [BR75a, BR76a] for examples and further discussion of this issue. This issue doesn’t concern us as we only care about the domain of the closure and not the linear map on the closure.

In the next theorem, Mn(𝒜)M_{n}(\mathcal{A}), Mn(𝒟(𝒜))M_{n}(\mathcal{D}(\mathcal{A})) and Mn(A)M_{n}(A) denotes the space of n×nn\times n matrices with coefficients in 𝒜\mathcal{A}, 𝒟(𝒜)\mathcal{D}(\mathcal{A}) and AA respectively.

Theorem 1.10.

Let 𝒟\mathcal{D} be a stable unital subalgebra of End(𝒜)\operatorname{End}(\mathcal{A}). The set 𝒟(𝒜)\mathcal{D}(\mathcal{A}) is a dense *-subalgebra of AA. It is also closed under smooth functional calculus, i.e,

  1. 1.

    If nn\in\mathbb{N}, aMn(𝒟(𝒜))a\in M_{n}(\mathcal{D}(\mathcal{A})) is normal (aa=aaa^{*}a=aa^{*}) and ff is a smooth function defined on some neighborhood of spec(a)\mathrm{spec}(a) with f(0)=0f(0)=0, then f(a)Mn(𝒟(𝒜))f(a)\in M_{n}(\mathcal{D}(\mathcal{A})).

  2. 2.

    If nn\in\mathbb{N}, aMn(𝒟(𝒜))a\in M_{n}(\mathcal{D}(\mathcal{A})) and ff is a holomorphic function defined on some neighborhood of spec(a)\mathrm{spec}(a) with f(0)=0f(0)=0, then f(a)Mn(𝒟(𝒜))f(a)\in M_{n}(\mathcal{D}(\mathcal{A})).

Here spec(a)\mathrm{spec}(a) is the spectrum of aa as an element of Mn(A)M_{n}(A).

For clarity of the exposition, the proof of Theorem 1.10 will be divided into several lemmas

Lemma 1.11.

The space 𝒟(𝒜)\mathcal{D}(\mathcal{A}) is a dense *-subalgebra of 𝒜\mathcal{A}.

Proof.

The space 𝒟(𝒜)\mathcal{D}(\mathcal{A}) is dense because 𝒜𝒟(𝒜)\mathcal{A}\subseteq\mathcal{D}(\mathcal{A}). Proving that 𝒟(𝒜)\mathcal{D}(\mathcal{A}) is closed under addition, involution, multiplication by scalars is straightforward. Let a,b𝒟(𝒜)a,b\in\mathcal{D}(\mathcal{A}), (xn)n,(yn)n𝒜(x_{n})_{n\in\mathbb{N}},(y_{n})_{n\in\mathbb{N}}\in\mathcal{A} such that xnax_{n}\to a and ynby_{n}\to b and δ(xn)\delta(x_{n}), δ(xn)\delta(x_{n}^{*}), δ(yn)\delta(y_{n}), δ(yn)\delta(y_{n}^{*}) converge for every δ𝒟\delta\in\mathcal{D}.  For any δ𝒟\delta\in\mathcal{D}, we can find δ10,,δ1r\delta_{10},\cdots,\delta_{1r}, δ20,,δ2r𝒟\delta_{20},\cdots,\delta_{2r}\in\mathcal{D} such that

δ(xnyn)=δ(xn)δ20(yn)+δ10(xn)δ(yn)+i=1rδ1i(xn)δ2i(yn).\displaystyle\delta(x_{n}y_{n})=\delta(x_{n})\delta_{20}(y_{n})+\delta_{10}(x_{n})\delta(y_{n})+\sum_{i=1}^{r}\delta_{1i}(x_{n})\delta_{2i}(y_{n}). (9)

Since δ1i(xn)\delta_{1i}(x_{n}) and δ2i(yn)\delta_{2i}(y_{n}) converge for every i{0,,r}i\in\{0,\cdots,r\}, we obtain that δ(xnyn)\delta(x_{n}y_{n}) converges. By a similar argument, we deduce that δ(ynxn)\delta(y_{n}^{*}x_{n}^{*}) converges. Hence, ab𝒟(𝒜)ab\in\mathcal{D}(\mathcal{A}). Therefore, 𝒟(𝒜)\mathcal{D}(\mathcal{A}) is a dense *-algebra. ∎

Let nn\in\mathbb{N},

Mn:End(𝒜)End(Mn(𝒜)),Mn(δ)((a)1i,jn)=(δ(aij))1i,jn.\displaystyle M_{n}:\operatorname{End}(\mathcal{A})\to\operatorname{End}(M_{n}(\mathcal{A})),\quad M_{n}(\delta)((a)_{1\leq i,j\leq n})=(\delta(a_{ij}))_{1\leq i,j\leq n}. (10)

Clearly MnM_{n} is an algebra homomorphism.

Lemma 1.12.

For any nn\in\mathbb{N}, Mn(𝒟)M_{n}(\mathcal{D}) is a stable subalgebra of End(Mn(𝒜))\operatorname{End}(M_{n}(\mathcal{A})). Furthermore, Mn(𝒟)(Mn(𝒜))=Mn(𝒟(𝒜))M_{n}(\mathcal{D})(M_{n}(\mathcal{A}))=M_{n}(\mathcal{D}(\mathcal{A})).

Proof.

A straightforward computation shows that if δ\delta is a left multiplier, then Mn(δ)M_{n}(\delta) is also a left multiplier. Similar computation shows that if (3) holds, then

Mn(δ)(ab)=Mn(δ)(a)Mn(δ20)(b)+Mn(δ10)(a)Mn(δ)(b)+i=1rMn(δ1i)(a)Mn(δ2i)(b).\displaystyle M_{n}(\delta)(ab)=M_{n}(\delta)(a)M_{n}(\delta_{20})(b)+M_{n}(\delta_{10})(a)M_{n}(\delta)(b)+\sum_{i=1}^{r}M_{n}(\delta_{1i})(a)M_{n}(\delta_{2i})(b). (11)

It is also clear that if δ20\delta_{20} satisfies (4), then Mn(δ20)M_{n}(\delta_{20}) satisfies (4). The lemma follows. ∎

By Lemma 1.12, it follows that to prove Theorem 1.10, it suffices to deal with the case n=1n=1.

Lemma 1.13.

Let a𝒟(𝒜)a\in\mathcal{D}(\mathcal{A}) be self-adjoint and ξ\xi\in\mathbb{R}. Then, aeiξa𝒟(𝒜)ae^{i\xi a}\in\mathcal{D}(\mathcal{A}). Furthermore, we can find a family (xn,ξ)n,ξ𝒜(x_{n,\xi})_{n\in\mathbb{N},\xi\in\mathbb{R}}\in\mathcal{A} such that

  1. 1.

    For every compact subset KK of \mathbb{R}, the sequence (xn,ξ)n(x_{n,\xi})_{n\in\mathbb{N}} converges uniformly in ξK\xi\in K to aeiξaae^{i\xi a} as n+n\to+\infty. Moreover,

    sup{xn,ξ:n,ξ}<+\displaystyle\sup\{\left\lVert x_{n,\xi}\right\rVert:n\in\mathbb{N},\xi\in\mathbb{R}\}<+\infty (12)
  2. 2.

    For every compact subset KK of \mathbb{R} and for every δ𝒟\delta\in\mathcal{D}, the sequence δ(xn,ξ)\delta(x_{n,\xi}) converges uniformly in ξK\xi\in K to an element in AA as n+n\to+\infty. We denote the limit by δ(aeiξa)\delta(ae^{i\xi a}).

  3. 3.

    For every δ𝒟\delta\in\mathcal{D}, the function

    A,ξδ(aeiξa)\displaystyle\mathbb{R}\to A,\quad\xi\mapsto\delta(ae^{i\xi a}) (13)

    is continuous. Furthermore, if δ\delta is of order ll, then

    sup{δ(xn,ξ)|ξ|l+1:n,ξ}<+\displaystyle\sup\left\{\frac{\left\lVert\delta(x_{n,\xi})\right\rVert}{|\xi|^{l}+1}:n\in\mathbb{N},\xi\in\mathbb{R}\right\}<+\infty (14)
  4. 4.

    For any ξ\xi\in\mathbb{R}, nn\in\mathbb{N}, xn,ξ=xn,ξx_{n,\xi}^{*}=x_{n,-\xi}.

We remark that the notation δ(aeiξa)\delta(ae^{i\xi a}) is an abuse of notation. In general aeiξaae^{i\xi a} doesn’t belong to 𝒜\mathcal{A}, therefore the value of δ(aeiξa)\delta(ae^{i\xi a}) usually depends on the choice of a sequence xnx_{n} converging to aa as in Definition 1.8, see Remark 1.9.

We will delay the proof of Lemma 1.13 for the moment. Instead, we will finish the proof of Theorem 1.10.1.

Proof of Theorem 1.10.1.

To prove Theorem 1.10.1, we first deal with the self-adjoint case, then with the general case of a normal element.

  1. 1.

    Let a𝒟(𝒜)a\in\mathcal{D}(\mathcal{A}) be self-adjoint, ff a CC^{\infty} function on a neighbourhood of Spec(a)\operatorname{Spec}(a)\subseteq\mathbb{R} with f(0)=0f(0)=0. Since we are only interested in f(a)f(a), without loss of generality, we can suppose fCc()f\in C^{\infty}_{c}(\mathbb{R}). Since f(0)=0f(0)=0, we can write f(x)=xg(x)f(x)=xg(x) where gCc()g\in C^{\infty}_{c}(\mathbb{R}). The Fourier transform g^\hat{g} of gg is a Schwartz function. By the Fourier inversion formula

    f(x)=12πxeixξg^(ξ)𝑑ξ.f(x)=\frac{1}{2\pi}\int_{-\infty}^{\infty}xe^{ix\xi}\hat{g}(\xi)d\xi.

    Hence,

    f(a)=12πaeiaξg^(ξ)𝑑ξ\displaystyle f(a)=\frac{1}{2\pi}\int_{-\infty}^{\infty}ae^{ia\xi}\hat{g}(\xi)d\xi (15)

    and the integral is absolutely convergent. We approximate the integral 12πaeiaξg^(ξ)𝑑ξ\frac{1}{2\pi}\int_{-\infty}^{\infty}ae^{ia\xi}\hat{g}(\xi)d\xi by the finite Riemann sums

    un:=12πj=n2n21naeiajng^(jn)u_{n}:=\frac{1}{2\pi}\sum_{j=-n^{2}}^{n^{2}}\frac{1}{n}ae^{ia\frac{j}{n}}\hat{g}\left(\frac{j}{n}\right)

    It is clear that unf(a)u_{n}\to f(a) as n+n\to+\infty. Let (xn,ξ)n,ξ(x_{n,\xi})_{n\in\mathbb{N},\xi\in\mathbb{R}} be as in Lemma 1.13. We define

    sn=12πj=n2n21nxn,jng^(jn).s_{n}=\frac{1}{2\pi}\sum_{j=-n^{2}}^{n^{2}}\frac{1}{n}x_{n,\frac{j}{n}}\hat{g}\left(\frac{j}{n}\right).

    It is clear that sn𝒜s_{n}\in\mathcal{A}. Since xn,ξx_{n,\xi} converges to aeiξaae^{i\xi a} uniformly in ξ\xi on compact sets, and g^\hat{g} is Schwartz, and by (12), we deduce that limn+snun=0\lim_{n\to+\infty}\left\lVert s_{n}-u_{n}\right\rVert=0. Hence, snf(a)s_{n}\to f(a). Let δ𝒟\delta\in\mathcal{D}. By Lemma 1.13.3, the integral

    12πδ(aeiξa)g^(ξ)𝑑ξ\displaystyle\frac{1}{2\pi}\int_{-\infty}^{\infty}\delta(ae^{i\xi a})\hat{g}(\xi)d\xi (16)

    is absolutely convergent. By using Riemann sums like we did with Integral (15), we deduce that

    limn+δ(sn)=12πδ(aeiξa)g^(ξ)𝑑ξ.\lim_{n\to+\infty}\delta(s_{n})=\frac{1}{2\pi}\int_{-\infty}^{\infty}\delta(ae^{i\xi a})\hat{g}(\xi)d\xi.

    Since xn,ξ=xn,ξx_{n,\xi}^{*}=x_{n,-\xi}, we can reuse our argument to deduce that

    limn+δ(sn)=12πδ(aeiξa)g^(ξ)¯𝑑ξ.\lim_{n\to+\infty}\delta(s_{n}^{*})=\frac{1}{2\pi}\int_{-\infty}^{\infty}\delta(ae^{i\xi a})\overline{\hat{g}(-\xi)}d\xi.

    Hence, f(a)𝒟(𝒜)f(a)\in\mathcal{D}(\mathcal{A}).

  2. 2.

    Let a𝒟(𝒜)a\in\mathcal{D}(\mathcal{A}) be a normal element, ff a CC^{\infty} function on a neighbourhood of Spec(a)\operatorname{Spec}(a)\subseteq\mathbb{C} with f(0)=0f(0)=0. Like before, we can suppose that fCc()f\in C^{\infty}_{c}(\mathbb{C}). We write z=x+iyz=x+iy for the coordinates in \mathbb{C}. Consider the function g(z)=f(x)g(z)=f(x). Then, g(a)=f|(a+a2)g(a)=f_{|\mathbb{R}}(\frac{a+a^{*}}{2}). It follows from the self-adjoint case that g(a)𝒟(𝒜)g(a)\in\mathcal{D}(\mathcal{A}). Hence, by replacing ff with fgf-g, we can suppose that f=0f=0 on \mathbb{R}. By a similar argument, we can suppose that f=0f=0 on both \mathbb{R} and ii\mathbb{R}. Hence, there exists gCc()g\in C^{\infty}_{c}(\mathbb{C}) such that f(z)=xyg(z)f(z)=xyg(z). Since gg is smooth compactly supported, it follows that g^\hat{g} is a Schwartz function. Let (a)=a+a2\Re(a)=\frac{a+a^{*}}{2} and (a)=aa2i\Im(a)=\frac{a-a^{*}}{2i}. By the Fourier inversion formula

    f(a)=1(2π)2(a)eiξ(a)(a)eiη(a)g^(ξ+iη)𝑑ξ𝑑η.\displaystyle f(a)=\frac{1}{(2\pi)^{2}}\int_{-\infty}^{\infty}\int_{-\infty}^{\infty}\Re(a)e^{i\xi\Re(a)}\Im(a)e^{i\eta\Im(a)}\hat{g}(\xi+i\eta)d\xi d\eta. (17)

    Theorem 1.10.1 follows from Lemma 1.13 applied to (a)\Re(a) and (a)\Im(a) by writing Integral (17) as a Riemann sum like we did in the self adjoint case. This finishes the proof of Theorem 1.10.1 under the assumption of Lemma 1.13.∎

The proof of Lemma 1.13 requires a few lemmas. Let a𝒟(𝒜)a\in\mathcal{D}(\mathcal{A}) be a self-adjoint element. We fix a sequence (xn)n(x_{n})_{n\in\mathbb{N}} such that xnax_{n}\to a and δ(xn)\delta(x_{n}) and δ(xn)\delta(x_{n}^{*}) converge for every δ𝒟\delta\in\mathcal{D}. By replacing xnx_{n} with xn+xn2\frac{x_{n}+x_{n}^{*}}{2}, we can suppose that xn=xnx_{n}=x_{n}^{*}.

Lemma 1.14.

For every ll\in\mathbb{N} and δ𝒟\delta\in\mathcal{D} of order ll, there exists C1,C2>0C_{1},C_{2}>0 such that

δ(xnm)C1mlC2m,m,n.\left\lVert\delta(x_{n}^{m})\right\rVert\leq C_{1}m^{l}C_{2}^{m},\quad\forall m,n\in\mathbb{N}. (18)
Proof.

We prove the lemma by induction on ll. For l=0l=0 and mm\in\mathbb{N}, we have

δ(xnm+1)=δ(xn)xnm\displaystyle\delta(x_{n}^{m+1})=\delta(x_{n})x_{n}^{m} (19)

Since δ(xn)\left\lVert\delta(x_{n})\right\rVert and xn\left\lVert x_{n}\right\rVert converge, it follows that they are bounded. The case l=0l=0 is finished.

Suppose the lemma holds for δ𝒟\delta\in\mathcal{D} of order <l<l. Let δ𝒟\delta\in\mathcal{D} be of order ll. For each i{0,,r}i\in\{0,\cdots,r\}, the Lemma holds for δ1i\delta_{1i}. Therefore, there exists constants C1,C2C_{1},C_{2} such that

δ1i(xnm)C1morder of δ1iC2mC1ml1C2m,n,m,i{0,,r}.\displaystyle\left\lVert\delta_{1i}(x_{n}^{m})\right\rVert\leq C_{1}m^{\text{order of }\delta_{1i}}C_{2}^{m}\leq C_{1}m^{l-1}C_{2}^{m},\quad\forall n,m\in\mathbb{N},i\in\{0,\cdots,r\}. (20)

Let C1C_{1}^{\prime} be a constant such that

δ(xn)C1,andδ2i(xn)C1,andC1C1,n,i{1,,r}.\displaystyle\left\lVert\delta(x_{n})\right\rVert\leq C_{1}^{\prime},\quad\text{and}\quad\left\lVert\delta_{2i}(x_{n})\right\rVert\leq C_{1}^{\prime},\quad\text{and}\quad C_{1}\leq C_{1}^{\prime},\quad\forall n\in\mathbb{N},i\in\{1,\cdots,r\}. (21)

Let C2C_{2}^{\prime} be a constant such that

δ20(xn)C2,andC2C2,n.\displaystyle\left\lVert\delta_{20}(x_{n})\right\rVert\leq C_{2}^{\prime},\quad\text{and}\quad C_{2}\leq C_{2}^{\prime},\quad\forall n\in\mathbb{N}. (22)

Some constants C1,C2C_{1}^{\prime},C_{2}^{\prime} exist because δ(xn)\delta(x_{n}), δ20(xn),,δ2r(xn)\delta_{20}(x_{n}),\cdots,\delta_{2r}(x_{n}) converge in AA.

By (3), for mm\in\mathbb{N}, we have

δ(xnm+1)=δ(xnm)δ20(xn)+δ10(xnm)δ(xn)+k=1rδ1k(xnm)δ2k(xn),n,m.\displaystyle\delta(x_{n}^{m+1})=\delta(x_{n}^{m})\delta_{20}(x_{n})+\delta_{10}(x_{n}^{m})\delta(x_{n})+\sum_{k=1}^{r}\delta_{1k}(x_{n}^{m})\delta_{2k}(x_{n}),\quad\forall n,m\in\mathbb{N}. (23)

By recurrence, we deduce that

δ(xnm+1)=δ(xn)δ20(xn)m+j=1mδ10(xnj)δ(xn)δ20(xn)mj+j=1mk=1rδ1k(xnj)δ2k(xn)δ20(xn)mj\displaystyle\delta(x^{m+1}_{n})=\delta(x_{n})\delta_{20}(x_{n})^{m}+\sum_{j=1}^{m}\delta_{10}(x^{j}_{n})\delta(x_{n})\delta_{20}(x_{n})^{m-j}+\sum_{j=1}^{m}\sum_{k=1}^{r}\delta_{1k}(x^{j}_{n})\delta_{2k}(x_{n})\delta_{20}(x_{n})^{m-j} (24)

Hence

δ(xnm+1)\displaystyle\left\lVert\delta(x_{n}^{m+1})\right\rVert C1C2m+j=1mjl1C12C2m+j=1mjl1rC12C2m\displaystyle\leq C_{1}^{\prime}C_{2}^{\prime m}+\sum_{j=1}^{m}j^{l-1}C_{1}^{\prime 2}C_{2}^{\prime m}+\sum_{j=1}^{m}j^{l-1}rC_{1}^{\prime 2}C_{2}^{\prime m} (25)
C1C2m+ml+1C12C2m+ml+1rC12C2m\displaystyle\leq C_{1}^{\prime}C_{2}^{\prime m}+m^{l+1}C_{1}^{\prime 2}C_{2}^{\prime m}+m^{l+1}rC_{1}^{\prime 2}C_{2}^{\prime m}

The result follows. ∎

Let ξ\xi\in\mathbb{R}, δ𝒟\delta\in\mathcal{D}. By Lemma 1.14, the sum

m=0(iξ)mδ(xnm+1)m!\displaystyle\sum_{m=0}^{\infty}(i\xi)^{m}\frac{\delta(x_{n}^{m+1})}{m!} (26)

converges in norm uniformly in nn. By an abuse of notation, we denote the sum by δ(xneiξxn)\delta(x_{n}e^{i\xi x_{n}}). Since limn+δ(xnm+1)\lim_{n\to+\infty}\delta(x_{n}^{m+1}) exists for all mm\in\mathbb{N} and the convergence in (26) is uniform, it follows that

δ(aeiξa):=limn+δ(xneiξxn)=m=0(iξ)mlimn+δ(xnm+1)m!\displaystyle\delta(ae^{i\xi a}):=\lim_{n\to+\infty}\delta(x_{n}e^{i\xi x_{n}})=\sum_{m=0}^{\infty}(i\xi)^{m}\frac{\lim_{n\to+\infty}\delta(x_{n}^{m+1})}{m!} (27)

exists.

Lemma 1.15.

For every ll\in\mathbb{N}, δ𝒟\delta\in\mathcal{D} of order ll, there exists constant C>0C>0 such that

δ(xneiξxn)C(|ξ|l+1),n,ξ.\left\lVert\delta(x_{n}e^{i\xi x_{n}})\right\rVert\leq C(|\xi|^{l}+1),\quad\forall n\in\mathbb{N},\xi\in\mathbb{R}.
Proof.

We prove the lemma by induction on ll. If l=0l=0, then by (19)

δ(xneiξxn)=m=0(iξ)mδ(xnm+1)m!=δ(xn)m=0(iξ)mxnmm!=δ(xn)eiξxn\displaystyle\delta(x_{n}e^{i\xi x_{n}})=\sum_{m=0}^{\infty}(i\xi)^{m}\frac{\delta(x_{n}^{m+1})}{m!}=\delta(x_{n})\sum_{m=0}^{\infty}(i\xi)^{m}\frac{x_{n}^{m}}{m!}=\delta(x_{n})e^{i\xi x_{n}} (28)

Since xnx_{n} are self-adjoint, it follows that eiξxn=1\left\lVert e^{i\xi x_{n}}\right\rVert=1. The case l=0l=0 follows.

Let δ𝒟\delta\in\mathcal{D} of order ll. We suppose the lemma holds for 𝒟\mathcal{D}-differential operators of order <l<l. By (24), and the identity

1(j1)!(mj)!01sj1(1s)mj𝑑s=1m!,1jm,\displaystyle\frac{1}{(j-1)!(m-j)!}\int_{0}^{1}s^{j-1}(1-s)^{m-j}ds=\frac{1}{m!},\quad 1\leq j\leq m, (29)

it follows that

δ(xneiξxn)=δ(xn)eiξδ20(xn)\displaystyle\delta(x_{n}e^{i\xi x_{n}})=\delta(x_{n})e^{i\xi\delta_{20}(x_{n})} +iξ01δ10(xneisξxn)δ(xn)ei(1s)ξδ20(xn)𝑑s\displaystyle+i\xi\int_{0}^{1}\delta_{10}(x_{n}e^{is\xi x_{n}})\delta(x_{n})e^{i(1-s)\xi\delta_{20}(x_{n})}ds
+iξk=1r01δ1k(xneisξxn)δ2k(xn)ei(1s)ξδ20(xn)𝑑s\displaystyle+i\xi\sum_{k=1}^{r}\int_{0}^{1}\delta_{1k}(x_{n}e^{is\xi x_{n}})\delta_{2k}(x_{n})e^{i(1-s)\xi\delta_{20}(x_{n})}ds

Now we use crucially (4), to deduce that δ20(xn)\delta_{20}(x_{n}) is self-adjoint. Therefore,

eiξδ20(xn)=ei(1s)ξδ20(xn)=1.\left\lVert e^{i\xi\delta_{20}(x_{n})}\right\rVert=\left\lVert e^{i(1-s)\xi\delta_{20}(x_{n})}\right\rVert=1.

The lemma follows from the induction hypothesis on δ1k(xneisξxn)\delta_{1k}(x_{n}e^{is\xi x_{n}}) for k{0,,r}k\in\{0,\cdots,r\}. ∎

Proof of Lemma 1.13.

Let ϕ:+\phi:\mathbb{R}_{+}\to\mathbb{N} be any function such which satisfies

ϕ(ξ)<mξmm!1m!,ξ+,m.\displaystyle\phi(\xi)<m\implies\frac{\xi^{m}}{m!}\leq\frac{1}{\sqrt{m!}},\quad\forall\xi\in\mathbb{R}_{+},m\in\mathbb{N}. (30)

We take

xn,ξ:=m=0ϕ(|ξ|)+n(iξ)mxnm+1m!.\displaystyle x_{n,\xi}:=\sum_{m=0}^{\phi(|\xi|)+n}(i\xi)^{m}\frac{x_{n}^{m+1}}{m!}. (31)

The element xn,ξx_{n,\xi} is a finite sum of elements of 𝒜\mathcal{A}. Hence, it belongs to 𝒜\mathcal{A}. We now check that xn,ξx_{n,\xi} satisfies the required properties.

  1. 1.

    It is rather straightforward to show that xn,ξx_{n,\xi} converges to aeiξaae^{i\xi a} uniformly on each compact subset of \mathbb{R}. Let us show (12). Let C=sup{xn:n}C=\sup\{\left\lVert x_{n}\right\rVert:n\in\mathbb{N}\}. It is finite because xnax_{n}\to a. Since

    m=0(iξ)mxnm+1m!=xneiξxn,\sum_{m=0}^{\infty}(i\xi)^{m}\frac{x_{n}^{m+1}}{m!}=x_{n}e^{i\xi x_{n}},

    and xnx_{n} is self-adjoint, we deduce that

    m=0(iξ)mxnm+1m!xnC.\displaystyle\left\lVert\sum_{m=0}^{\infty}(i\xi)^{m}\frac{x_{n}^{m+1}}{m!}\right\rVert\leq\left\lVert x_{n}\right\rVert\leq C. (32)

    By

    xn,ξm=0(iξ)mxnm+1m!\displaystyle\left\lVert x_{n,\xi}-\sum_{m=0}^{\infty}(i\xi)^{m}\frac{x_{n}^{m+1}}{m!}\right\rVert =m=ϕ(|ξ|)+n+1(iξ)mxnm+1m!\displaystyle=\left\lVert\sum_{m=\phi(|\xi|)+n+1}^{\infty}(i\xi)^{m}\frac{x_{n}^{m+1}}{m!}\right\rVert (33)
    m=ϕ(|ξ|)+n+1|ξ|mCm+1m!\displaystyle\leq\sum_{m=\phi(|\xi|)+n+1}^{\infty}|\xi|^{m}\frac{C^{m+1}}{m!}
    m=ϕ(|ξ|)+n+1Cm+1m!\displaystyle\leq\sum_{m=\phi(|\xi|)+n+1}^{\infty}\frac{C^{m+1}}{\sqrt{m!}}
    m=0Cm+1m!<+\displaystyle\leq\sum_{m=0}^{\infty}\frac{C^{m+1}}{\sqrt{m!}}<+\infty

    we deduce (12).

  2. 2.

    By Lemma 1.14, it is again straightforward to show that δ(xn,ξ)\delta(x_{n,\xi}) converges to δ(aeiξa)\delta(ae^{i\xi a}) uniformly on each compact subset of \mathbb{R}.

  3. 3.

    The function ξδ(aeiξa)\xi\mapsto\delta(ae^{i\xi a}) is a continuous function because it is the uniform limit (over compact subsets of \mathbb{R}) of continuous function by (27). Let us show (14). Let δ𝒟\delta\in\mathcal{D} of order ll. By Lemma 1.14, there exists C1,C2C_{1},C_{2} such that δ(xnm)C1mlC2m\left\lVert\delta(x_{n}^{m})\right\rVert\leq C_{1}m^{l}C_{2}^{m}. Therefore, we have

    δ(xn,ξ)δ(xneiξxn)\displaystyle\left\lVert\delta(x_{n,\xi})-\delta(x_{n}e^{i\xi x_{n}})\right\rVert =m=ϕ(|ξ|)+n+1(iξ)mδ(xnm+1)m!\displaystyle=\left\lVert\sum_{m=\phi(|\xi|)+n+1}^{\infty}(i\xi)^{m}\frac{\delta(x_{n}^{m+1})}{m!}\right\rVert (34)
    m=ϕ(|ξ|)+n+1|ξ|mC1(m+1)lC2m+1m!\displaystyle\leq\sum_{m=\phi(|\xi|)+n+1}^{\infty}|\xi|^{m}\frac{C_{1}(m+1)^{l}C_{2}^{m+1}}{m!}
    m=ϕ(|ξ|)+n+1C1(m+1)lC2m+1m!\displaystyle\leq\sum_{m=\phi(|\xi|)+n+1}^{\infty}\frac{C_{1}(m+1)^{l}C_{2}^{m+1}}{\sqrt{m!}}
    m=0C1(m+1)lC2m+1m!<+\displaystyle\leq\sum_{m=0}^{\infty}\frac{C_{1}(m+1)^{l}C_{2}^{m+1}}{\sqrt{m!}}<+\infty

    By Lemma 1.15, we deduce (12).

  4. 4.

    The identity xn,ξ=xn,ξx_{n,\xi}^{*}=x_{n,-\xi} is trivial to verify.

This finishes the proof of Lemma 1.13. ∎

We now proceed to prove Theorem 1.10.2. The proof of Theorem 1.10.2 relies on the fact that Theorem 1.10.1 is true even in a parametrized setting. By this, we mean the following

Lemma 1.16.

Let KK be a compact topological space, (aw)wK𝒟(𝒜)(a_{w})_{w\in K}\subseteq\mathcal{D}(\mathcal{A}), (xn,w)n,wK𝒜(x_{n,w})_{n\in\mathbb{N},w\in K}\subseteq\mathcal{A} a family of elements which satisfies the following:

  1. 1.

    For every n,wKn\in\mathbb{N},w\in K, xn,w=xn,wx_{n,w}^{*}=x_{n,w}.

  2. 2.

    For every nn\in\mathbb{N}, the function nxn,wn\mapsto x_{n,w} is continuous in wKw\in K.

  3. 3.

    The limit of xn,wx_{n,w} as n+n\to+\infty is equal to an,wa_{n,w} uniformly in wKw\in K.

  4. 4.

    For every δ𝒟\delta\in\mathcal{D}, the limit δ(xn,w)\delta(x_{n,w}) as n+n\to+\infty exists in AA uniformly in wKw\in K

  5. 5.

    For every δ𝒟\delta\in\mathcal{D}, the limit δ(xn,w)\delta(x_{n,w}^{*}) as n+n\to+\infty exists in AA uniformly in wKw\in K

Let ff be a continuous complex-valued function defined in an open neighbourhood of

{(w,λ)K×:λSpec(aw)}\{(w,\lambda)\in K\times\mathbb{C}:\lambda\in\operatorname{Spec}(a_{w})\}

which is smooth in the λ\lambda variable, and which satisfies f(w,0)=0f(w,0)=0. Then, there exists a family (yn,w)n,wK𝒜(y_{n,w})_{n\in\mathbb{N},w\in K}\subseteq\mathcal{A} which satisfies the following:

  1. 1.

    For every nn\in\mathbb{N}, the function nyn,wn\mapsto y_{n,w} is continuous in wKw\in K.

  2. 2.

    The limit of yn,wy_{n,w} as n+n\to+\infty is equal to f(w,aw)f(w,a_{w}) uniformly in wKw\in K.

  3. 3.

    For every δ𝒟\delta\in\mathcal{D}, the limit δ(yn,w)\delta(y_{n,w}) as n+n\to+\infty exists in AA uniformly in wKw\in K

  4. 4.

    For every δ𝒟\delta\in\mathcal{D}, the limit δ(yn,w)\delta(y_{n,w}^{*}) as n+n\to+\infty exists in AA uniformly in wKw\in K

To prove Lemma 1.16, one follows the proof of Theorem 1.10.1 carefully making sure that adding a parameter doesn’t cause any problems. We leave this verification to the reader.

The following is a counterpart to Lemma 1.13 for non self-adjoint elements.

Lemma 1.17.

Let a𝒟(𝒜)a\in\mathcal{D}(\mathcal{A}). For every wspec(a){0}w\notin\mathrm{spec(a)}\cup\{0\}, awa𝒟(𝒜)\frac{a}{w-a}\in\mathcal{D}(\mathcal{A}). Furthermore, if K\(spec(a){0})K\subseteq\mathbb{C}\backslash(\mathrm{spec(a)}\cup\{0\}) is a compact subset, then we can find a family (xn,w)n,wK𝒜(x_{n,w})_{n\in\mathbb{N},w\in K}\in\mathcal{A} such that the following is satisfied:

  • The sequence (xn,w)n(x_{n,w})_{n\in\mathbb{N}} converges uniformly in wKw\in K to awa\frac{a}{w-a}.

  • For every δ𝒟\delta\in\mathcal{D}, δ(xn,w)\delta(x_{n,w}) converges uniformly in wKw\in K to an element denoted by δ(awa)\delta(\frac{a}{w-a}).

  • For every δ𝒟\delta\in\mathcal{D}, δ(xn,w)\delta(x_{n,w}^{*}) converges uniformly in wKw\in K to an element denoted by δ(aw¯a)\delta(\frac{a^{*}}{\bar{w}-a^{*}}).

  • For every δ𝒟\delta\in\mathcal{D}, the maps wδ(awa)w\mapsto\delta(\frac{a}{w-a}) and wδ(aw¯a)w\mapsto\delta(\frac{a^{*}}{\bar{w}-a^{*}}) are continuous.

Proof.

Let wspec(a){0}w\notin\mathrm{spec(a)}\cup\{0\}. The element

aw:=(wa)(wa)|w|2=waw¯a+aa𝒟(𝒜)a_{w}:=(w-a)^{*}(w-a)-|w|^{2}=-wa^{*}-\bar{w}a+a^{*}a\in\mathcal{D}(\mathcal{A})

is self-adjoint whose spectrum is contained in ]|w|2,+[]-|w|^{2},+\infty[. By Theorem 1.10.1, applied to f(x)=1x+|w|21|w|2f(x)=\frac{1}{x+|w|^{2}}-\frac{1}{|w|^{2}}, we deduce that f(aw)𝒟(𝒜)f(a_{w})\in\mathcal{D}(\mathcal{A}). It follows that

awa=(f(aw)+1|w|2)(wa)a𝒟(𝒜).\frac{a}{w-a}=\left(f(a_{w})+\frac{1}{|w|^{2}}\right)(w-a)^{*}a\in\mathcal{D}(\mathcal{A}).

Lemma 1.17 now follows from Lemma 1.16. ∎

Proof of Theorem 1.10.2.

Let a𝒟(𝒜)a\in\mathcal{D}(\mathcal{A}), and ff a holomorphic function on a neighbourhood of Spec(a)\operatorname{Spec}(a) with f(0)=0f(0)=0. Since f(0)=0f(0)=0, f(z)=zg(z)f(z)=zg(z) for some holomorphic function gg. Let γ\gamma be a contour around spec(a)\mathrm{spec}(a) such that γ\gamma is in the domain of ff and the Cauchy integral formula applies

f(z)=z2πiγg(w)wz𝑑w.f(z)=\frac{z}{2\pi i}\int_{\gamma}\frac{g(w)}{w-z}dw.

We thus have

f(a)=12πiγg(w)awa𝑑w.f(a)=\frac{1}{2\pi i}\int_{\gamma}g(w)\frac{a}{w-a}dw.

Without loss of generality, we can suppose that 0 doesn’t belong to the image of γ\gamma.

Let xn,wx_{n,w} be as in Lemma 1.17 applied to KK equal to the image of γ\gamma. We approximate f(a)f(a) by a Riemann sum, then approximate each term awa\frac{a}{w-a} by xn,wx_{n,w}. This way we obtain a sequence sn𝒜s_{n}\in\mathcal{A} which converges to f(a)f(a) and such that δ(sn)\delta(s_{n}) converges to 12πiγg(w)δ(awa)𝑑w\frac{1}{2\pi i}\int_{\gamma}g(w)\delta(\frac{a}{w-a})dw and δ(sn)\delta(s_{n}^{*}) converges to 12πiγg¯(w)δ(aw¯a)𝑑w\frac{1}{2\pi i}\int_{\gamma}\bar{g}(w)\delta(\frac{a^{*}}{\bar{w}-a^{*}})dw. The result follows. ∎

Corollary 1.18.

Let 𝔤\mathfrak{g}, AA be as in Proposition 1.7, 𝒮(𝔤)A\mathcal{S}(\mathfrak{g})\subseteq A be the *-subalgebra of Schwartz functions. Then

  1. 1.

    If a𝒮(𝔤)a\in\mathcal{S}(\mathfrak{g}) is normal and ff is a smooth function defined on some neighbourhood of spec(a)\mathrm{spec}(a) with f(0)=0f(0)=0, then f(a)𝒮(𝔤)f(a)\in\mathcal{S}(\mathfrak{g}).

  2. 2.

    If a𝒮(𝔤)a\in\mathcal{S}(\mathfrak{g}) and ff is a holomorphic function defined on some neighbourhood of spec(a)\mathrm{spec}(a) with f(0)=0f(0)=0, then f(a)𝒮(𝔤)f(a)\in\mathcal{S}(\mathfrak{g}).

Corollary 1.18 is well known, see for example [HJ84a]. A well known argument using holomorphic functional calculus (see for example [Bla98a]) implies the following corollary.

Corollary 1.19.

The inclusion i:𝒟(𝒜)Ai:\mathcal{D}(\mathcal{A})\to A induces an isomorphism in KK-theory K0(𝒟(𝒜))K0(A)K_{0}(\mathcal{D}(\mathcal{A}))\simeq K_{0}(A).

2 Schwartz functions on Connes’s tangent groupoid

Let MM be a compact smooth manifold. The set

𝕋M=M×M×+×TM×{0}\mathbb{T}M=M\times M\times\mathbb{R}_{+}^{\times}\sqcup TM\times\{0\}

can be equipped with the structure of a smooth manifold with boundary TM×{0}TM\times\{0\} as follows. The subset M×M×+×M\times M\times\mathbb{R}_{+}^{\times} is declared an open subset of 𝕋M\mathbb{T}M with its usual smooth structure. If UMU\subseteq M is an open set, ϕ:Udim(M)\phi:U\to\mathbb{R}^{\dim(M)} is a diffeomorphism, then U×U×+×TU×{0}U\times U\times\mathbb{R}_{+}^{\times}\sqcup TU\times\{0\} is declared an open subset of 𝕋M\mathbb{T}M and the following bijection is declared a diffeomorphism

U×U×+×TU×{0}\displaystyle U\times U\times\mathbb{R}_{+}^{\times}\sqcup TU\times\{0\} dim(M)×dim(M)×+\displaystyle\to\mathbb{R}^{\dim(M)}\times\mathbb{R}^{\dim(M)}\times\mathbb{R}_{+} (35)
(y,x,t)\displaystyle(y,x,t) (ϕ(y)ϕ(x)t,ϕ(x),t),x,yU,t+×\displaystyle\mapsto\left(\frac{\phi(y)-\phi(x)}{t},\phi(x),t\right),\quad x,y\in U,t\in\mathbb{R}_{+}^{\times}
(v,x,0)\displaystyle(v,x,0) (dϕx(v),ϕ(x),0),xU,vTxM\displaystyle\mapsto(d\phi_{x}(v),\phi(x),0),\quad x\in U,v\in T_{x}M

One can check that this defines a smooth structure on 𝕋M\mathbb{T}M. A more conceptual way to define the smooth structure is to notice that the smooth structure (and the topology) is uniquely determined by requiring that the maps

𝕋MM×M×+,(y,x,t)(y,x,t),(v,x,0)(x,x,0)\displaystyle\mathbb{T}M\to M\times M\times\mathbb{R}_{+},\quad(y,x,t)\mapsto(y,x,t),\quad(v,x,0)\mapsto(x,x,0) (36)

and

DNC(f):𝕋M,(y,x,t)f(y)f(x)t,(v,x,0)dfx(v),\displaystyle\operatorname{DNC}(f):\mathbb{T}M\to\mathbb{C},\quad(y,x,t)\mapsto\frac{f(y)-f(x)}{t},\quad(v,x,0)\mapsto df_{x}(v), (37)

be smooth, where fC(M)f\in C^{\infty}(M).

Differential operators.

Let X𝒳(M)X\in\mathcal{X}(M) be a vector field. We can define a vector field 𝕏\mathbb{X} on 𝕋M\mathbb{T}M as follows. If fC(𝕋M)f\in C^{\infty}(\mathbb{T}M) is a smooth function, then we define 𝕏fCc(𝕋M)\mathbb{X}\cdot f\in C^{\infty}_{c}(\mathbb{T}M) by

𝕏f(y,x,t)=tXyf(y,x,t),𝕏f(v,x,0)=X(x)f(v,x,0),\mathbb{X}\cdot f(y,x,t)=tX_{y}f(y,x,t),\quad\mathbb{X}\cdot f(v,x,0)=X(x)\cdot f(v,x,0),

where XyX_{y} means that XX acts on the yy variable, and X(x)f(v,x,0)X(x)\cdot f(v,x,0) means that we view X(x)X(x) as a constant vector field on TxMT_{x}M which acts on the smooth function f(,x,0)f(\cdot,x,0). By the description of charts above, one checks that 𝕏\mathbb{X} is indeed a vector field.

More generally, let DD be a differential operator of order dd acting on MM. Then we can define a differential operator denoted 𝔻\mathbb{D} of order dd acting on 𝕋M\mathbb{T}M by the formula

𝔻f(y,x,t)=tdDyf(y,x,t),𝔻f(v,x,0)=D(x)f(v,x,0),\mathbb{D}\cdot f(y,x,t)=t^{d}D_{y}f(y,x,t),\quad\mathbb{D}\cdot f(v,x,0)=D(x)\cdot f(v,x,0), (38)

where DyD_{y} means that DD acts on the yy variable, and D(x)f(v,x,0)D(x)\cdot f(v,x,0) means that we view the principal part of DD as a constant coefficient differential operator on TxMT_{x}M acting on the smooth function f(,x,0)f(\cdot,x,0). Locally by the principal part, we mean that if D=|I|dgIxID=\sum_{|I|\leq d}g_{I}\frac{\partial}{\partial x_{I}}, then D(x)=|I|=dgI(x)xID(x)=\sum_{|I|=d}g_{I}(x)\frac{\partial}{\partial x_{I}} seen as a constant coefficient differential operator on TxMT_{x}M. Using the local charts of 𝕋M\mathbb{T}M, one can check that 𝔻\mathbb{D} is indeed a differential operator.

Convolution and adjoint.

From now on, we suppose that MM is equipped with a Riemannian metric. If f,gc(𝕋M)f,g\in\mathbb{C}^{\infty}_{c}(\mathbb{T}M), then we define their convolution fgCc(𝕋M)f\star g\in C^{\infty}_{c}(\mathbb{T}M) by the formula

fg(y,x,t)\displaystyle f\star g(y,x,t) =tdim(M)Mf(y,z,t)g(z,x,t)𝑑z,t0\displaystyle=t^{-\dim(M)}\int_{M}f(y,z,t)g(z,x,t)dz,\quad t\neq 0 (39)
fg(v,x,0)\displaystyle f\star g(v,x,0) =TxMf(vw,x,0)g(w,x,0)𝑑w,\displaystyle=\int_{T_{x}M}f(v-w,x,0)g(w,x,0)dw,

where the integral over TxMT_{x}M is with respect to the constant Riemannian metric on TxMT_{x}M induced from the Riemannian metric on MM. We also define the adjoint by

f(y,x,t)=f¯(x,y,t),f(v,x,0)=f¯(v,x,0)f^{*}(y,x,t)=\bar{f}(x,y,t),\quad f^{*}(v,x,0)=\bar{f}(-v,x,0)

We leave it to the reader to check that Cc(𝕋M)C^{\infty}_{c}(\mathbb{T}M) with the operations defined above is a *-algebra. We refer the reader to [Con94a] and [DL10a] for more details on the convolution algebra of the tangent groupoid.

Adjoint of Differential operators on 𝕋M\mathbb{T}M.

Let DD be a differential operator on MM of order dd, 𝔻\mathbb{D} the associated differential operator on 𝕋M\mathbb{T}M. We sometimes denote 𝔻(f)\mathbb{D}(f) by 𝔻f\mathbb{D}\star f. We also define f𝔻f\star\mathbb{D} by the formula

f𝔻:=(𝔻f),f\star\mathbb{D}:=(\mathbb{D}^{*}\star f^{*})^{*},

where 𝔻\mathbb{D}^{*} is the differential operator on 𝕋M\mathbb{T}M associated to the formal adjoint DD^{*} of DD. Equivalently

f𝔻(y,x,t)=tdDxf(y,x,t),f𝔻(v,x,0)=D(x)f(v,x,0),f\star\mathbb{D}(y,x,t)=t^{d}D^{\prime}_{x}f(y,x,t),\quad f\star\mathbb{D}(v,x,0)=D(x)\cdot f(v,x,0),

where DD^{\prime} is the formal transpose given by

MDf(x)g(x)𝑑x=Mf(x)Dg(x)𝑑x,f,gC(M).\int_{M}Df(x)g(x)dx=\int_{M}f(x)D^{\prime}g(x)dx,\quad f,g\in C^{\infty}(M).

One can check that

(𝔻f)g=𝔻(fg),(f𝔻)g=f(𝔻g)f,gCc(𝕋M).(\mathbb{D}\star f)\star g=\mathbb{D}\star(f\star g),\quad(f\star\mathbb{D})\star g=f\star(\mathbb{D}\star g)\quad f,g\in C^{\infty}_{c}(\mathbb{T}M). (40)

This justifies our notation f𝔻f\star\mathbb{D} and 𝔻f\mathbb{D}\star f.

Derivations in direction of MM.

The functions 𝔻f\mathbb{D}\star f and f𝔻f\star\mathbb{D} agree at t=0t=0. It follows that the map

δD:Cc(𝕋M)Cc(𝕋M),f1t(𝔻ff𝔻)\displaystyle\delta_{D}:C^{\infty}_{c}(\mathbb{T}M)\to C^{\infty}_{c}(\mathbb{T}M),\quad f\mapsto\frac{1}{t}(\mathbb{D}\star f-f\star\mathbb{D})

is a well-defined differential operator on 𝕋M\mathbb{T}M of order dd. If DD is an order 0 differential operator given by multiplication by gC(M)g\in C^{\infty}(M), then δg\delta_{g} is an order 0 differential operator given by multiplication by DNC(g)\operatorname{DNC}(g), see (37). For X𝒳(M)X\in\mathcal{X}(M), one can describe δX\delta_{X} as follows. The flow of XX gives an +×\mathbb{R}_{+}^{\times} action on 𝕋M\mathbb{T}M by

βλ(y,x,t)=(exp(λX)y,exp(λX)x,t),βλ(v,x,0)=(dexp(λX)(x)(v),exp(λX)x,0).\beta_{\lambda}(y,x,t)=(\exp(\lambda X)\cdot y,\exp(\lambda X)\cdot x,t),\quad\beta_{\lambda}(v,x,0)=(d\exp(\lambda X)(x)(v),\exp(\lambda X)\cdot x,0).

The differential operator δX\delta_{X} is the derivative of the action β\beta (plus multiplication by the divergence of XX). Since δD\delta_{D} is given by a commutator, it satisfies the Leibniz rule

δD(fg)=δD(f)g+fδD(g),f,gCc(𝕋M)\delta_{D}(f\star g)=\delta_{D}(f)\star g+f\star\delta_{D}(g),\quad f,g\in C^{\infty}_{c}(\mathbb{T}M) (41)

and

δD1D2(f)=δD1(f)𝔻2+𝔻1δD2(f),fCc(𝕋M)\delta_{D_{1}D_{2}}(f)=\delta_{D_{1}}(f)\star\mathbb{D}_{2}+\mathbb{D}_{1}\star\delta_{D_{2}}(f),\quad f\in C^{\infty}_{c}(\mathbb{T}M) (42)

Derivation in direction of tt.

The group +×\mathbb{R}_{+}^{\times} acts smoothly on the left on 𝕋M\mathbb{T}M by the formula

αλ:𝕋M𝕋M,αλ(y,x,t)=(y,x,λ1t),αλ(v,x,0)=(λv,x,0),λ+×.\alpha_{\lambda}:\mathbb{T}M\to\mathbb{T}M,\quad\alpha_{\lambda}(y,x,t)=(y,x,\lambda^{-1}t),\quad\alpha_{\lambda}(v,x,0)=(\lambda v,x,0),\quad\lambda\in\mathbb{R}_{+}^{\times}. (43)

We also let +×\mathbb{R}_{+}^{\times} acts on CC^{\infty} functions by the formula

αλ(f)=λdim(M)fαλ1,fCc(𝕋M),λ+×.\displaystyle\alpha_{\lambda}(f)=\lambda^{-\dim(M)}f\circ\alpha_{\lambda^{-1}},\quad f\in C^{\infty}_{c}(\mathbb{T}M),\lambda\in\mathbb{R}_{+}^{\times}.

One has

αλ(fg)=αλ(f)αλ(g),αλ(f)=αλ(f).\alpha_{\lambda}(f\star g)=\alpha_{\lambda}(f)\star\alpha_{\lambda}(g),\quad\alpha_{\lambda}(f)^{*}=\alpha_{\lambda}(f^{*}). (44)

We define δα\delta_{\alpha} to be the derivative of the +×\mathbb{R}_{+}^{\times}-action at λ=1\lambda=1. One can check that

δα(f)(y,x,t)\displaystyle\delta_{\alpha}(f)(y,x,t) =dim(M)f(y,x,t)+ttf(y,x,t)\displaystyle=-\dim(M)f(y,x,t)+t\frac{\partial}{\partial t}f(y,x,t)
δα(f)(v,x,0)\displaystyle\delta_{\alpha}(f)(v,x,0) =dim(M)f(v,x,0)ivivif(v,x,0),\displaystyle=-\dim(M)f(v,x,0)-\sum_{i}v_{i}\frac{\partial}{\partial v_{i}}f(v,x,0),

where viv_{i} are any local coordinates for TxMT_{x}M. It follows from (44) that δα\delta_{\alpha} satisfies the Leibniz rule

δα(fg)=δα(f)g+fδα(g),f,gCc(𝕋M).\delta_{\alpha}(f\star g)=\delta_{\alpha}(f)\star g+f\star\delta_{\alpha}(g),\quad f,g\in C^{\infty}_{c}(\mathbb{T}M). (45)
Lemma 2.1.

One can find a finite family of smooth functions f1,,fkC(M)f_{1},\cdots,f_{k}\in C^{\infty}(M) and vector fields X1,,Xk𝒳(M)X_{1},\cdots,X_{k}\in\mathcal{X}(M) such that for any X𝒳(M)X\in\mathcal{X}(M), X=i=1kX(fi)XiX=\sum_{i=1}^{k}X(f_{i})X_{i}.

Proof.

Let f1,,fk:Mf_{1},\cdots,f_{k}:M\to\mathbb{R} be smooth functions such that for all xMx\in M, df1,x,,dfk,xdf_{1,x},\cdots,df_{k,x} span TxMT_{x}^{*}M. Let θ=M×k\theta=M\times\mathbb{R}^{k} be the trivial vector bundle on MM of rank kk. The forms dfidf_{i} define a surjective bundle map θTM\theta\to T^{*}M. Its dual is an injective bundle map i:TMθi:TM\to\theta. Let p:θTMp:\theta\to TM be any bundle map such that pi=Idp\circ i=\mathrm{Id}. The map pp determines the vector fields X1,,XkX_{1},\cdots,X_{k}. ∎

We fix a choice of fi,Xif_{i},X_{i} for the rest of this section. By Lemma 2.1, if xMx\in M, then i=1kdfi(x)Xi(x)\sum_{i=1}^{k}df_{i}(x)X_{i}(x) is equal to the identity TxMTxMT_{x}M\to T_{x}M. By taking the trace of i=1kdfi(x)Xi(x)\sum_{i=1}^{k}df_{i}(x)X_{i}(x), we deduce that

i=1kXi(fi)(x)=dim(M),xM\sum_{i=1}^{k}X_{i}(f_{i})(x)=\dim(M),\quad\forall x\in M (46)

For fCc(𝕋M)f\in C^{\infty}_{c}(\mathbb{T}M), the functions δα(f)\delta_{\alpha}(f) and dim(M)fi=1kδfi(𝕏if)-\dim(M)f-\sum_{i=1}^{k}\delta_{f_{i}}(\mathbb{X}_{i}\star f) agree at t=0t=0. Hence, we can define

δ^(f):\displaystyle\hat{\delta}(f): =1t(δα(f)+dim(M)f+i=1kδfi(𝕏if))\displaystyle=\frac{1}{t}\left(\delta_{\alpha}(f)+\dim(M)f+\sum_{i=1}^{k}\delta_{f_{i}}(\mathbb{X}_{i}\star f)\right) (47)
=1t(δα(f)+i=1k𝕏iδfi(f)),\displaystyle=\frac{1}{t}\left(\delta_{\alpha}(f)+\sum_{i=1}^{k}\mathbb{X}_{i}\star\delta_{f_{i}}(f)\right),

where in the second equality we used (46).

Lemma 2.2.

The map δ^\hat{\delta} satisfies

δ^(fg)=δ^(f)g+fδ^(g)+i=1kδXi(f)δfi(g),f,gCc(𝕋M)\hat{\delta}(f\star g)=\hat{\delta}(f)\star g+f\star\hat{\delta}(g)+\sum_{i=1}^{k}\delta_{X_{i}}(f)\star\delta_{f_{i}}(g),\quad f,g\in C^{\infty}_{c}(\mathbb{T}M)

and hence a derivation of order 22 on Cc(𝕋M)C^{\infty}_{c}(\mathbb{T}M).

Proof.

By (41) and (45)

δ^(fg)δ^(f)gfδ^(g)=1t(i=1k𝕏ifδfi(g)f𝕏iδfi(g))=i=1kδXi(f)δfi(g)\hat{\delta}(f\star g)-\hat{\delta}(f)\star g-f\star\hat{\delta}(g)=\frac{1}{t}\left(\sum_{i=1}^{k}\mathbb{X}_{i}\star f\star\delta_{f_{i}}(g)-f\star\mathbb{X}_{i}\star\delta_{f_{i}}(g)\right)=\sum_{i=1}^{k}\delta_{X_{i}}(f)\star\delta_{f_{i}}(g)\qed

The reader is invited to see the similarity between (47) and [HH18a, The formula for TT on Page 4].

Schwartz functions.

Definition 2.3.

We define 𝒮(𝕋M)\mathcal{S}(\mathbb{T}M) to be the space of smooth function fC(𝕋M)f\in C^{\infty}(\mathbb{T}M) such that ff is bounded and all iterated applications of the following differential operators give bounded functions on 𝕋M\mathbb{T}M

  • the operator ftff\mapsto tf, where t:𝕋M+t:\mathbb{T}M\to\mathbb{R}_{+} is the natural projection.

  • the operator f𝔻ff\mapsto\mathbb{D}\star f, where DD is any differential operator on MM

  • the operator fδD(f)f\mapsto\delta_{D}(f), where DD is any differential operator on MM

  • the operator fδ^(f)f\mapsto\hat{\delta}(f).

We can add the application fδα(f)f\mapsto\delta_{\alpha}(f) to the above list, but this is redundant as it follows from the others using (47).

Proposition 2.4.

Let f𝒮(𝕋M)f\in\mathcal{S}(\mathbb{T}M). Then for every xMx\in M, the function vf(v,x,0)v\mapsto f(v,x,0) is a Schwartz function on the vector space TxMT_{x}M in the classical sense.

Proof.

If X𝒳(M)X\in\mathcal{X}(M) is a vector field, then v(𝕏f)(v,x,0)v\mapsto(\mathbb{X}\star f)(v,x,0) is the application of the constant vector field X(x)X(x) to vf(v,x,0)v\mapsto f(v,x,0). While if gC(M)g\in C^{\infty}(M), then the map vδg(f)(v,x,0)v\mapsto\delta_{g}(f)(v,x,0) is the pointwise product of the map vf(v,x,0)v\mapsto f(v,x,0) with the linear map dgx:TxMdg_{x}:T_{x}M\to\mathbb{C}. By iterated application of the previous two operations, it follows that vf(v,x,0)v\mapsto f(v,x,0) is Schwartz. ∎

Example 2.5.

Let g1,,gl:Mg_{1},\cdots,g_{l}:M\to\mathbb{R} be smooth functions such that the map

Ml,x(g1(x),,gl(x))M\mapsto\mathbb{R}^{l},\quad x\mapsto(g_{1}(x),\cdots,g_{l}(x))

is an embedding. Then, the function

ei=1lDNC(gi)2t2C(𝕋M)e^{-\sum_{i=1}^{l}\operatorname{DNC}(g_{i})^{2}-t^{2}}\in C^{\infty}(\mathbb{T}M)

is Schwartz, where DNC(gi)\operatorname{DNC}(g_{i}) is defined in (37).

The following proposition summarizes the main properties of Schwartz functions. In its proof, we remark that by (42), in Definition 2.3 the second and third conditions can be replaced by 𝔻f\mathbb{D}\star f and δD(f)\delta_{D}(f) are bounded where DD is either a vector field or a smooth function on MM.

Proposition 2.6.

The following holds

  1. 1.

    The definition of Schwartz functions doesn’t depend on the choice of fi,Xif_{i},X_{i} in Lemma 2.1.

  2. 2.

    If f𝒮(𝕋M)f\in\mathcal{S}(\mathbb{T}M), then fC0(𝕋M)f\in C_{0}(\mathbb{T}M), i.e. ff vanishes at infinity.

  3. 3.

    If f,g𝒮(𝕋M)f,g\in\mathcal{S}(\mathbb{T}M), then the integral in fgf\star g is absolutely convergent and fg𝒮(𝕋M)f\star g\in\mathcal{S}(\mathbb{T}M)

  4. 4.

    If f𝒮(𝕋M)f\in\mathcal{S}(\mathbb{T}M), then f𝒮(𝕋M)f^{*}\in\mathcal{S}(\mathbb{T}M)

Proof.

For the first part, we first need some lemmas.

Lemma 2.7.

If fCc(M×M×+)f\in C^{\infty}_{c}(M\times M\times\mathbb{R}_{+}) and g𝒮(𝕋M)g\in\mathcal{S}(\mathbb{T}M), then the pointwise product (fπ)g𝒮(𝕋M)(f\circ\pi)g\in\mathcal{S}(\mathbb{T}M), where π:𝕋MM×M×+\pi:\mathbb{T}M\to M\times M\times\mathbb{R}_{+} is the map in (36)

Proof.

(fπ)g(f\circ\pi)g is bounded because gg and ff are bounded. We argue that each of the applications in Definition 2.3 gives functions of the form (fπ)g(f^{\prime}\circ\pi)g^{\prime} for some fCc(M×M×+)f^{\prime}\in C^{\infty}_{c}(M\times M\times\mathbb{R}_{+}) and g𝒮(𝕋M)g^{\prime}\in\mathcal{S}(\mathbb{T}M). The result then follows by induction. We thus have

  • the function t(fπ)g=(fπ)(tg)t(f\circ\pi)g=(f\circ\pi)(tg)

  • If DD is a function on MM, then 𝔻((fπ)g)=(fπ)(𝔻g)\mathbb{D}\star((f\circ\pi)g)=(f\circ\pi)(\mathbb{D}\star g). If DD is a vector field, then we have

    𝔻((fπ)g)=(Dy(f)π)(tg)+(fπ)(𝔻g),\mathbb{D}\star((f\circ\pi)g)=(D_{y}(f)\circ\pi)(tg)+(f\circ\pi)(\mathbb{D}\star g), (48)

    where Dy(f)D_{y}(f) is the action of DD on the yy variable in f(y,x,t)f(y,x,t). The equality can be checked directly on M×M×+×M\times M\times\mathbb{R}_{+}^{\times} and by density equality follows on 𝕋M\mathbb{T}M.

  • If DD is a function, then δD((fπ)g)=(fπ)δD(g)\delta_{D}((f\circ\pi)g)=(f\circ\pi)\delta_{D}(g). If DD is a vector field, then one can as in (48) show that

    δD((fπ)g)=(Dy(f)πDx(f)π)(g)+(fπ)δD(g)\delta_{D}((f\circ\pi)g)=(D_{y}(f)\circ\pi-D_{x}^{\prime}(f)\circ\pi)(g)+(f\circ\pi)\delta_{D}(g)
  • one has

    δα((fπ)g)=(fπ)δα(g)+(t(f)π)(tg)\delta_{\alpha}((f\circ\pi)g)=(f\circ\pi)\delta_{\alpha}(g)+(\frac{\partial}{\partial t}(f)\circ\pi)(tg)

    and so

    δ^((fπ)g)=(fπ)δ^(g)+(t(f)π)(g)+i=1k(Xiy(f)π)(δfi(g))\hat{\delta}((f\circ\pi)g)=(f\circ\pi)\hat{\delta}(g)+(\frac{\partial}{\partial t}(f)\circ\pi)(g)+\sum_{i=1}^{k}(X_{iy}(f)\circ\pi)(\delta_{f_{i}}(g))

This finishes the proof. ∎

Lemma 2.8.

Let fC(M×M)f\in C^{\infty}(M\times M) be a smooth function that vanishes to the order 11 on the diagonal, and let f~C(𝕋M)\tilde{f}\in C^{\infty}(\mathbb{T}M) be the smooth function given by

(y,x,t)f(y,x)t2,(v,x,0)12d2fx(v),(y,x,t)\mapsto\frac{f(y,x)}{t^{2}},\quad(v,x,0)\mapsto\frac{1}{2}d^{2}f_{x}(v),

where d2fd^{2}f is the Hessian of ff. Then if g𝒮(𝕋M)g\in\mathcal{S}(\mathbb{T}M) then the pointwise product f~g𝒮(𝕋M)\tilde{f}g\in\mathcal{S}(\mathbb{T}M).

Proof.

Any function ff which vanishes to the order 11 can be written as a finite sum of functions of the form h(y,x)(h1(y)h1(x))(h2(y)h2(x))h(y,x)(h_{1}(y)-h_{1}(x))(h_{2}(y)-h_{2}(x)) where hC(M×M),h1,h2C(M)h\in C^{\infty}(M\times M),h_{1},h_{2}\in C^{\infty}(M). It follows that the pointwise product f~g\tilde{f}g is sum of pointwise product of hπh\circ\pi with δh1(δh2(g))\delta_{h_{1}}(\delta_{h_{2}}(g)) where π:𝕋MM×M\pi:\mathbb{T}M\to M\times M is the natural map. The result then follows from Lemma 2.7. ∎

Lemma 2.9.

Let g1,,gkg_{1},\cdots,g_{k^{\prime}} and Y1,,YkY_{1},\cdots,Y_{k^{\prime}} be another family satisfying Lemma 2.1. Then there exists a family of smooth functions hij:Mh_{ij}:M\to\mathbb{R} for 1ik1\leq i\leq k and 1jk1\leq j\leq k^{\prime} such that

dfi=j=1khijdgj,Yj=i=1khijXi.df_{i}=\sum_{j=1}^{k^{\prime}}h_{ij}dg_{j},\quad Y_{j}=\sum_{i=1}^{k}h_{ij}X_{i}. (49)
Proof.

Let x1,,xdim(M)x_{1},\cdots,x_{\dim(M)} be local coordinates in MM. Then dfi=l=1dim(M)ϕildxldf_{i}=\sum_{l=1}^{\dim(M)}\phi_{il}dx_{l} and Xi=l=1dim(M)ϕilxlX_{i}=\sum_{l=1}^{\dim(M)}\phi^{\prime}_{il}\frac{\partial}{\partial x_{l}} for some functions ϕil\phi_{il} and ϕil\phi_{il}^{\prime}. Similarly, dgj=l=1dim(M)ψjldxldg_{j}=\sum_{l=1}^{{\dim(M)}}\psi_{jl}dx_{l} and Yj=l=1dim(M)ψjlxlY_{j}=\sum_{l=1}^{{\dim(M)}}\psi^{\prime}_{jl}\frac{\partial}{\partial x_{l}}. Then one can take hij=l=1dim(M)ϕilψjlh_{ij}=\sum_{l=1}^{{\dim(M)}}\phi_{il}\psi_{jl}^{\prime}. To define hh globally one takes a partition of unity. ∎

We can now prove Proposition 2.6.1. Let δ^\hat{\delta}^{\prime} be the operator associated to g1,,gkg_{1},\cdots,g_{k^{\prime}} and Y1,,YkY_{1},\cdots,Y_{k^{\prime}}. We have

δ^(f)δ^(f)\displaystyle\hat{\delta}(f)-\hat{\delta}^{\prime}(f) =1t(i=1kδfi(𝕏if)j=1kδgj(𝕐jf))\displaystyle=\frac{1}{t}\left(\sum_{i=1}^{k}\delta_{f_{i}}(\mathbb{X}_{i}\star f)-\sum_{j=1}^{k^{\prime}}\delta_{g_{j}}(\mathbb{Y}_{j}\star f)\right)
=i=1k1t2((fi(y)fi(x)j=1k(gj(y)gj(x))hij(y))(𝕏if)).\displaystyle=\sum_{i=1}^{k}\frac{1}{t^{2}}\left(\left(f_{i}(y)-f_{i}(x)-\sum_{j=1}^{k^{\prime}}(g_{j}(y)-g_{j}(x))h_{ij}(y)\right)(\mathbb{X}_{i}\star f)\right).

By (49), the function fi(y)fi(x)j=1k(gj(y)gj(x))hij(y)f_{i}(y)-f_{i}(x)-\sum_{j=1}^{k^{\prime}}(g_{j}(y)-g_{j}(x))h_{ij}(y) vanishes to order 11 on the diagonal. By Lemma 2.8, δ^(f)𝒮(𝕋M)\hat{\delta}^{\prime}(f)\in\mathcal{S}(\mathbb{T}M). Proposition 2.6.1 follows.

For Proposition 2.6.2, suppose that fC0(𝕋M)f\notin C_{0}(\mathbb{T}M). Then there exists ϵ>0\epsilon>0 a sequence in 𝕋M\mathbb{T}M and which goes to infinity yet |f|ϵ|f|\geq\epsilon for every element in the sequence. By passing to a subsequence we can suppose that the sequence is either of the form (yn,xn,tn)(y_{n},x_{n},t_{n}) or of the form (vn,xn,0)(v_{n},x_{n},0). Suppose we have the first case. Then by taking a subsequence, we can suppose that xnxx_{n}\to x and ynyy_{n}\to y and tnt[0,+]t_{n}\to t\in[0,+\infty]. If t=+t=+\infty, we get a contradiction to the fact that tftf is bounded. If t]0,+[t\in]0,+\infty[, then the sequence (yn,xn,tn)(y_{n},x_{n},t_{n}) converges in 𝕋M\mathbb{T}M to (y,x,t)(y,x,t), again a contradiction. If xyx\neq y then we get a contradiction to the fact that δg(f)\delta_{g}(f) is bounded where gC(M)g\in C^{\infty}(M) is any smooth function with g(x)=g(y)g(x)=g(y). So we have t=0t=0 and x=yx=y. The sequence (yn,xn,tn)(y_{n},x_{n},t_{n}) converging to infinity implies that there exists gC(M)g\in C^{\infty}(M) such that |g(yn)g(xn)tn|+\left|\frac{g(y_{n})-g(x_{n})}{t_{n}}\right|\to+\infty. We get then a contradiction to the fact that δg(f)\delta_{g}(f) is bounded. The case of a sequence (vn,xn,0)(v_{n},x_{n},0) is similar.

For Proposition 2.6.3, first the integral in fgf\star g is absolutely convergent by Proposition 2.4. Hence, fgf\star g is a well-defined function on 𝕋M\mathbb{T}M. We now show its continuity. Let gig_{i} be as in Example 2.5. Then consider the function

ϕ=(1+DNC(g1)2++DNC(gl)2)dim(M)2+1C(𝕋M).\phi=(1+\operatorname{DNC}(g_{1})^{2}+\cdots+\operatorname{DNC}(g_{l})^{2})^{\frac{\dim(M)}{2}+1}\in C^{\infty}(\mathbb{T}M). (50)

The function ff being Schwartz implies that ϕf\phi f is bounded. One can easily show by looking at local coordinates of 𝕋M\mathbb{T}M that there exists C>0C>0 (only depends on ϕ\phi) such that

fgCϕfg.\left\lVert f\star g\right\rVert_{\infty}\leq C\left\lVert\phi f\right\rVert_{\infty}\left\lVert g\right\rVert_{\infty}.

Since ϕf\phi f and gg are Schwartz functions, ϕf,gC0(𝕋M)\phi f,g\in C_{0}(\mathbb{T}M). We can thus approximate them uniformly with compactly supported functions. It follows that fgC0(𝕋M)f\star g\in C_{0}(\mathbb{T}M). Smoothness of fgf\star g as well as the fact that fg𝒮(𝕋M)f\star g\in\mathcal{S}(\mathbb{T}M) follow easily from (40), (41), (44) and Lemma 2.2.

For Proposition 2.6.4, since ff^{*} is bounded if ff is bounded, the result follows from the following identities

tf=(tf),𝔻f=(tδD(f)+𝔻f),δD(f)=δD(f),δα(f)=δα(f),\displaystyle tf^{*}=(tf)^{*},\quad\mathbb{D}\star f^{*}=(-t\delta_{D^{*}}(f)+\mathbb{D}^{*}\star f)^{*},\quad\delta_{D}(f^{*})=-\delta_{D^{*}}(f)^{*},\quad\delta_{\alpha}(f^{*})=\delta_{\alpha}(f)^{*},

and the identity

δ^(f)δ^(f)\displaystyle\hat{\delta}(f^{*})^{*}-\hat{\delta}(f) =1t(i=1kδfi(f𝕏i)δfi(𝕏if))\displaystyle=\frac{1}{t}\left(\sum_{i=1}^{k}-\delta_{f_{i}}(f\star\mathbb{X}_{i}^{*})-\delta_{f_{i}}(\mathbb{X}_{i}\star f)\right)
=1t(i=1kδfi(f𝕏i)δfi(𝕏if))+i=1kδfi(fdiv(Xi))\displaystyle=\frac{1}{t}\left(\sum_{i=1}^{k}\delta_{f_{i}}(f\star\mathbb{X}_{i})-\delta_{f_{i}}(\mathbb{X}_{i}\star f)\right)+\sum_{i=1}^{k}\delta_{f_{i}}(f\star\mathrm{div}(X_{i}))
=i=1kδfi(δXi(f))+δfi(fdiv(Xi))\displaystyle=\sum_{i=1}^{k}-\delta_{f_{i}}(\delta_{X_{i}}(f))+\delta_{f_{i}}(f\star\mathrm{div}(X_{i}))\qed
Remark 2.10.

There are different definitions in the literature of Schwartz functions. Our definition doesn’t precisely agree with [Car08a]. In [Car08a], the author adds a conical support condition which we don’t need. Our definition agrees with the one proposed by Debord and Skandalis [DS14a]. We refer the reader to [DS14a, Section 1.6] for a treatment of the Schwartz functions defined here using classical semi-norm estimates.

We give here a definition of Schwartz functions which is equivalent to ours by looking in local coordinates. A function f𝒮(𝕋M)f\in\mathcal{S}(\mathbb{T}M) is Schwartz if and only if it satisfies the following:

  1. 1.

    For every k,lk,l\in\mathbb{N}, DD a differential operator on M×MM\times M,

    sup{|tkddtlDf(y,x,t)|:(y,x,t)M×M×[1,+[}<+\sup\left\{\left|t^{k}\frac{d}{dt^{l}}Df(y,x,t)\right|:(y,x,t)\in M\times M\times[1,+\infty[\right\}<+\infty
  2. 2.

    For every k,lk,l\in\mathbb{N}, DD a differential operator on M×MM\times M, KM×M\MK\subseteq M\times M\backslash M a compact subset outside the diagonal,

    sup{|tkddtlDf(y,x,t)|:(y,x,t)K×]0,1]}<+\sup\left\{\left|t^{-k}\frac{d}{dt^{l}}Df(y,x,t)\right|:(y,x,t)\in K\times]0,1]\right\}<+\infty
  3. 3.

    For every UMU\subseteq M open subset diffeomorphic to dim(M)\mathbb{R}^{\dim(M)} by a map ϕ:Udim(M)\phi:U\to\mathbb{R}^{\dim(M)}, Φ\Phi the local chart of 𝕋M\mathbb{T}M associated to ϕ\phi defined in (35), Kdim(M)K\subseteq\mathbb{R}^{\dim(M)} a compact subset, k,lk,l\in\mathbb{N}, α,βdim(M)\alpha,\beta\in\mathbb{N}^{\dim(M)}, one has

    sup{|vkddtlddvαddxβfΦ1(v,x,t)|:(v,x,t)K^}<+,\sup\left\{\left|\left\lVert v\right\rVert^{k}\frac{d}{dt^{l}}\frac{d}{dv^{\alpha}}\frac{d}{dx^{\beta}}f\circ\Phi^{-1}(v,x,t)\right|:(v,x,t)\in\hat{K}\right\}<+\infty,

    where K^={(v,x,t)2dim(M)+1:x,x+tvK}\hat{K}=\{(v,x,t)\in\mathbb{R}^{2\dim(M)+1}:x,x+tv\in K\}.

CC^{*}-algebra of the tangent groupoid.

Let fCc(𝕋M)f\in C^{\infty}_{c}(\mathbb{T}M), t0t\neq 0. We define the operator

πt(f):L2M\displaystyle\pi_{t}(f):L^{2}M L2M\displaystyle\to L^{2}M
g\displaystyle g (ytdim(M)Mf(y,x,t)g(x)𝑑x),gL2(M)\displaystyle\mapsto(y\mapsto t^{-\dim(M)}\int_{M}f(y,x,t)g(x)dx),\quad g\in L^{2}(M)

For each xMx\in M, we also define the operator

πx(f):L2TxM\displaystyle\pi_{x}(f):L^{2}T_{x}M L2TxM\displaystyle\to L^{2}T_{x}M
g\displaystyle g (vTxMf(vw,x,0)g(w)𝑑w),gL2(TxM)\displaystyle\mapsto(v\mapsto\int_{T_{x}M}f(v-w,x,0)g(w)dw),\quad g\in L^{2}(T_{x}M)

We then define

f:=max(supt+×πt(f),supxMπx(f)).\left\lVert f\right\rVert:=\max{\left(\sup_{t\in\mathbb{R}_{+}^{\times}}\left\lVert\pi_{t}(f)\right\rVert,\sup_{x\in M}\left\lVert\pi_{x}(f)\right\rVert\right)}.

We define the CC^{*}-algebra C𝕋MC^{*}\mathbb{T}M to be the completion of Cc(𝕋M)C^{\infty}_{c}(\mathbb{T}M) with respect to \left\lVert\cdot\right\rVert. The CC^{*}-algebra C𝕋MC^{*}\mathbb{T}M lies in a short exact sequence

0𝒦(L2M)C0(+×)C𝕋MC0(TM)0,0\to\mathcal{K}(L^{2}M)\otimes C_{0}(\mathbb{R}_{+}^{\times})\to C^{*}\mathbb{T}M\to C_{0}(T^{*}M)\to 0,

where 𝒦(L2M)\mathcal{K}(L^{2}M) is the CC^{*}-algebra of compact operators on L2ML^{2}M, see [Con94a, Proposition 5 Page 108]

Proposition 2.11.

The space 𝒮(𝕋M)\mathcal{S}(\mathbb{T}M) is a *-subalgebra of C𝕋MC^{*}\mathbb{T}M.

Proof.

It is well known that

πt(f)supxMmax(tdim(M)M|f(x,y,t)|𝑑y,tdim(M)M|f(y,x,t)|𝑑y),t+×,fCc(𝕋M)\left\lVert\pi_{t}(f)\right\rVert\leq\sup_{x\in M}\max\left(t^{-\dim(M)}\int_{M}|f(x,y,t)|dy,t^{-\dim(M)}\int_{M}|f(y,x,t)|dy\right),\ t\in\mathbb{R}_{+}^{\times},f\in C^{\infty}_{c}(\mathbb{T}M)

and

πx(f)supxM(TxM|f(v,x,0)|𝑑v),xM,fCc(𝕋M).\left\lVert\pi_{x}(f)\right\rVert\leq\sup_{x\in M}\left(\int_{T_{x}M}|f(v,x,0)|dv\right),\quad x\in M,f\in C^{\infty}_{c}(\mathbb{T}M).

Hence,

fsupxM,t+×max(tdim(M)M|f(x,y,t)|𝑑y,tdim(M)|f(y,x,t)|𝑑y,TxM|f(v,x,0)|𝑑v)\left\lVert f\right\rVert\leq\sup_{x\in M,t\in\mathbb{R}_{+}^{\times}}\max\left(t^{-\dim(M)}\int_{M}|f(x,y,t)|dy,t^{-\dim(M)}\int|f(y,x,t)|dy,\int_{T_{x}M}|f(v,x,0)|dv\right) (51)

It follows that C𝕋MC^{*}\mathbb{T}M contains measurable functions on 𝕋M\mathbb{T}M for which the right-hand side of (51) is finite (usually denoted L1(𝕋M)L^{1}(\mathbb{T}M)). Let ϕ\phi be as in (50). If f𝒮(𝕋M)f\in\mathcal{S}(\mathbb{T}M), then the right-hand side of (51) is bounded by ϕf\left\lVert\phi f\right\rVert_{\infty}. Hence, fC𝕋Mf\in C^{*}\mathbb{T}M. Finally, 𝒮(𝕋M)\mathcal{S}(\mathbb{T}M) is a *-subalgebra by Proposition 2.6. ∎

Relation between uniform norm and CC^{*}-norm.

The following theorem which is essentially just the Sobolev embedding theorem will be very useful in allowing us to replace the uniform norm with the CC^{*}-norm, which is more convenient to use.

Theorem 2.12.

Let Δ\Delta be the positive Laplace-Beltrami operator on MM, ΔΔ\Delta\!\!\!\!\Delta the corresponding differential operator on 𝕋M\mathbb{T}M as in (38), and kk\in\mathbb{N} with 2k>dim(M)22k>\frac{\dim(M)}{2}. There exists C>0C>0, such that

fC(1+tdim(M))(1+ΔΔ)kf(1+ΔΔ)k,fCc(𝕋M)\left\lVert f\right\rVert_{\infty}\leq C\left\lVert(1+t^{\dim(M)})(1+\Delta\!\!\!\!\Delta)^{k}\star f\star(1+\Delta\!\!\!\!\Delta)^{k}\right\rVert,\quad\forall f\in C^{\infty}_{c}(\mathbb{T}M)
Lemma 2.13.

Let kk\in\mathbb{N} with 2k>dim(M)22k>\frac{\dim(M)}{2}. There exists a constant C>0C>0 such that for all t>0t>0 and xMx\in M, if uxu_{x} denotes the Dirac delta distribution on MM at xx, then

(1+t2Δ)k(ux)L2MCmax(tdim(M)2,1).\left\lVert(1+t^{2}\Delta)^{-k}(u_{x})\right\rVert_{L^{2}M}\leq C\max(t^{-\frac{\dim(M)}{2}},1).
Proof.

The Sobolev embedding lemma implies that uxH2k(M)u_{x}\in H^{-2k}(M) where Hs(M)H^{s}(M) denotes the ss Sobolev space. Hence, (1+t2Δ)k(ux)L2M(1+t^{2}\Delta)^{-k}(u_{x})\in L^{2}M. For t>1t>1 we have and use the inequality

(1+t2Δ)k(ux)L2M\displaystyle\left\lVert(1+t^{2}\Delta)^{-k}(u_{x})\right\rVert_{L^{2}M} =(1+t2Δ)k(1+Δ)k(1+Δ)k(ux)L2M\displaystyle=\left\lVert(1+t^{2}\Delta)^{-k}(1+\Delta)^{k}(1+\Delta)^{-k}(u_{x})\right\rVert_{L^{2}M}
(1+t2Δ)k(1+Δ)kB(L2M)(1+Δ)k(ux)L2M\displaystyle\leq\left\lVert(1+t^{2}\Delta)^{-k}(1+\Delta)^{k}\right\rVert_{B(L^{2}M)}\left\lVert(1+\Delta)^{-k}(u_{x})\right\rVert_{L^{2}M}
(1+Δ)k(ux)L2M,\displaystyle\leq\left\lVert(1+\Delta)^{-k}(u_{x})\right\rVert_{L^{2}M},

where in the last inequality we used the fact that (1+x)k(1+t2x)k1\frac{(1+x)^{k}}{(1+t^{2}x)^{k}}\leq 1 for all x[0,+[x\in[0,+\infty[. For t<1t<1, we proceed differently. If gL2Mg\in L^{2}M, then

(1+t2Δ)k(ux),gL2M=|(1+t2Δ)k(g)(x)|.\langle(1+t^{2}\Delta)^{-k}(u_{x}),g\rangle_{L^{2}M}=|(1+t^{2}\Delta)^{-k}(g)(x)|.

We need to maximize |(1+t2Δ)k(g)(x)|gL2M\frac{|(1+t^{2}\Delta)^{-k}(g)(x)|}{\left\lVert g\right\rVert_{L^{2}M}} as gg varies in L2ML^{2}M. The function gg needs to vanish outside any given neighbourhood of xx to reach the supremum. We can thus assume that M=dim(M)M=\mathbb{R}^{\dim(M)} and x=0x=0 and the metric on MM is Euclidean outside the unit ball. The inequality follows from a change of variable xxtx\to\frac{x}{t}. ∎

Proof of Theorem 2.12.

Let t>0t>0. We apply

πt((1+ΔΔ)kf(1+ΔΔ)k)=tdim(M)(1+t2Δ)kf(1+t2Δ)k\pi_{t}((1+\Delta\!\!\!\!\Delta)^{k}\star f\star(1+\Delta\!\!\!\!\Delta)^{k})=t^{-\dim(M)}(1+t^{2}\Delta)^{k}\star f\star(1+t^{2}\Delta)^{k}

to (1+t2Δ)k(ux)(1+t^{2}\Delta)^{-k}(u_{x}) to deduce that

tdim(M)(1+t2Δ)k(f(,x,t))L2M\displaystyle t^{-\dim(M)}\left\lVert(1+t^{2}\Delta)^{k}(f(\cdot,x,t))\right\rVert_{L^{2}M} Cmax(tdim(M)2,1)(1+ΔΔ)kf(1+ΔΔ)k\displaystyle\leq C\max(t^{-\frac{\dim(M)}{2}},1)\left\lVert(1+\Delta\!\!\!\!\Delta)^{k}\star f\star(1+\Delta\!\!\!\!\Delta)^{k}\right\rVert

One has

|f(y,x,t)|\displaystyle|f(y,x,t)| =|uy,f(,x,t)H2k(M)×H2k(M)|\displaystyle=|\left\langle u_{y},f(\cdot,x,t)\right\rangle_{H^{-2k}(M)\times H^{2k}(M)}|
=|(1+t2Δ)k(uy),(1+t2Δ)k(f(,x,t))L2(M)×L2(M)|\displaystyle=|\langle(1+t^{2}\Delta)^{-k}(u_{y}),(1+t^{2}\Delta)^{k}(f(\cdot,x,t))\rangle_{L^{2}(M)\times L^{2}(M)}|
(1+t2Δ)k(uy)L2M(1+t2Δ)k(f(,x,t))L2M\displaystyle\leq\left\lVert(1+t^{2}\Delta)^{-k}(u_{y})\right\rVert_{L^{2}M}\left\lVert(1+t^{2}\Delta)^{k}(f(\cdot,x,t))\right\rVert_{L^{2}M}
Cmax(tdim(M)2,1)(1+t2Δ)k(f(,x,t))L2M\displaystyle\leq C\max(t^{-\frac{\dim(M)}{2}},1)\left\lVert(1+t^{2}\Delta)^{k}(f(\cdot,x,t))\right\rVert_{L^{2}M}

It follows that

|f(y,x,t)|\displaystyle|f(y,x,t)| C2tdim(M)max(tdim(M),1)(1+ΔΔ)kf(1+ΔΔ)k\displaystyle\leq C^{2}t^{\dim(M)}\max(t^{-\dim(M)},1)\left\lVert(1+\Delta\!\!\!\!\Delta)^{k}\star f\star(1+\Delta\!\!\!\!\Delta)^{k}\right\rVert
=C2max(1,tdim(M))(1+ΔΔ)kf(1+ΔΔ)k.\displaystyle=C^{2}\max(1,t^{\dim(M)})\left\lVert(1+\Delta\!\!\!\!\Delta)^{k}\star f\star(1+\Delta\!\!\!\!\Delta)^{k}\right\rVert.

By taking the limit as t0+t\to 0^{+}, one obtains a bound of ff on TM×{0}TM\times\{0\}. The theorem follows. ∎

Schwartz functions are closed under smooth calculus.

Proof of Theorem A.

Let 𝒟End(𝒮(𝕋M))\mathcal{D}\subseteq\operatorname{End}(\mathcal{S}(\mathbb{T}M)) be the subalgebra of End(𝒮(𝕋M))\operatorname{End}(\mathcal{S}(\mathbb{T}M)) generated by the maps in Definition 2.3. We will prove that 𝒟(𝒮(𝕋M))=𝒮(𝕋M)\mathcal{D}(\mathcal{S}(\mathbb{T}M))=\mathcal{S}(\mathbb{T}M). Theorem A would then follow from Theorem 1.10. Let a𝒟(𝒮(𝕋M))a\in\mathcal{D}(\mathcal{S}(\mathbb{T}M)), xn𝒮(𝕋M)x_{n}\in\mathcal{S}(\mathbb{T}M) as in Definition 1.8. Since

x(1+tdim(M))(1+ΔΔ)kx(1+ΔΔ)kx\mapsto(1+t^{\dim(M)})(1+\Delta\!\!\!\!\Delta)^{k}\star x\star(1+\Delta\!\!\!\!\Delta)^{k}

is a differential operator on 𝒮(𝕋M)\mathcal{S}(\mathbb{T}M) which belongs to 𝒟\mathcal{D}, it follows that (1+tdim(M))(1+ΔΔ)kxn(1+ΔΔ)k(1+t^{\dim(M)})(1+\Delta\!\!\!\!\Delta)^{k}\star x_{n}\star(1+\Delta\!\!\!\!\Delta)^{k} converges in C𝕋MC^{*}\mathbb{T}M. By Theorem 2.12, it follows that aa is a bounded continuous function and xnax_{n}\to a in uniform norm. Let δ𝒟\delta\in\mathcal{D}. Since the map x(1+tdim(M))(1+ΔΔ)kδ(x)(1+ΔΔ)kx\mapsto(1+t^{\dim(M)})(1+\Delta\!\!\!\!\Delta)^{k}\star\delta(x)\star(1+\Delta\!\!\!\!\Delta)^{k} belongs to 𝒟\mathcal{D}, again by Theorem 2.12, δ(xn)\delta(x_{n}) converges in the uniform norm to a bounded continuous function on 𝕋M\mathbb{T}M. It follows that aa is smooth and δ(a)\delta(a) bounded for every δ\delta. Hence, a𝒮(𝕋M)a\in\mathcal{S}(\mathbb{T}M). ∎

Remark 2.14.

In [Ewe21a], Ewert defined an algebra of Schwartz functions for the inhomogeneous tangent groupoid defined in [EY17a, Moh21a, HH18a, CP19a]. Does Theorem A hold for this Schwartz algebra, or a similarly defined one?

{refcontext}

[sorting=nyt]

References

  • [AS68] M.. Atiyah and I.. Singer “The index of elliptic operators. I” In Ann. of Math. (2) 87, 1968, pp. 484–530 DOI: 10.2307/1970715
  • [Bla98] B. Blackadar KK-theory for operator algebras” 5, Mathematical Sciences Research Institute Publications Cambridge University Press, Cambridge, 1998, pp. xx+300
  • [BR75] O. Bratteli and D.. Robinson “Unbounded derivations of CC^{\ast}-algebras” In Comm. Math. Phys. 42, 1975, pp. 253–268 URL: http://projecteuclid.org.ezproxy.universite-paris-saclay.fr/euclid.cmp/1103899048
  • [BR76] O. Bratteli and D.. Robinson “Unbounded derivations of CC^{\ast}-algebras. II” In Comm. Math. Phys. 46.1, 1976, pp. 11–30 URL: http://projecteuclid.org.ezproxy.universite-paris-saclay.fr/euclid.cmp/1103899543
  • [Car08] P. Carrillo Rouse “A Schwartz type algebra for the tangent groupoid” In KK-theory and noncommutative geometry, EMS Ser. Congr. Rep. Eur. Math. Soc., Zürich, 2008, pp. 181–199 DOI: 10.4171/060-1/7
  • [CP19] W. Choi and R. Ponge “Tangent maps and tangent groupoid for Carnot manifolds” In Differential Geom. Appl. 62, 2019, pp. 136–183 DOI: 10.1016/j.difgeo.2018.11.002
  • [Con86] A. Connes “Cyclic cohomology and the transverse fundamental class of a foliation” In Geometric methods in operator algebras (Kyoto, 1983) 123, Pitman Res. Notes Math. Ser. Longman Sci. Tech., Harlow, 1986, pp. 52–144
  • [Con94] A. Connes “Noncommutative geometry” Academic Press, Inc., San Diego, CA, 1994, pp. xiv+661
  • [DL10] C. Debord and J.-M. Lescure “Index theory and groupoids” In Geometric and topological methods for quantum field theory Cambridge Univ. Press, Cambridge, 2010, pp. 86–158 DOI: 10.1017/CBO9780511712135.004
  • [DS14] C. Debord and G. Skandalis “Adiabatic groupoid, crossed product by +\mathbb{R}_{+}^{\ast} and pseudodifferential calculus” In Adv. Math. 257, 2014, pp. 66–91 DOI: 10.1016/j.aim.2014.02.012
  • [EY17] E. Erp and R. Yuncken “On the tangent groupoid of a filtered manifold” In Bull. Lond. Math. Soc. 49.6, 2017, pp. 1000–1012 DOI: 10.1112/blms.12096
  • [Ewe21] E. Ewert “Pseudodifferential operators on filtered manifolds as generalized fixed points” In To appear in Journal of noncommutative Geometry arXiv, 2021 DOI: 10.48550/ARXIV.2110.03548
  • [Ful84] W. Fulton “Intersection theory” 2, Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)] Springer-Verlag, Berlin, 1984, pp. xi+470 DOI: 10.1007/978-3-662-02421-8
  • [HH18] A. Haj Saeedi Sadegh and N. Higson “Euler-like vector fields, deformation spaces and manifolds with filtered structure” In Doc. Math. 23, 2018, pp. 293–325
  • [HJ84] A. Hulanicki and J.. Jenkins “Nilpotent Lie groups and summability of eigenfunction expansions of Schrödinger operators” In Studia Math. 80.3, 1984, pp. 235–244 DOI: 10.4064/sm-80-3-235-244
  • [Moh21] O. Mohsen “On the deformation groupoid of the inhomogeneous pseudo-differential calculus” In Bull. Lond. Math. Soc. 53.2, 2021, pp. 575–592 DOI: 10.1112/blms.12443
  • [Ped00] G.. Pedersen “Operator differentiable functions” In Publ. Res. Inst. Math. Sci. 36.1, 2000, pp. 139–157 DOI: 10.2977/prims/1195143229

References

  • [AS68a] M.. Atiyah and I.. Singer “The index of elliptic operators. I” In Ann. of Math. (2) 87, 1968, pp. 484–530 DOI: 10.2307/1970715
  • [BR75a] O. Bratteli and D.. Robinson “Unbounded derivations of CC^{\ast}-algebras” In Comm. Math. Phys. 42, 1975, pp. 253–268 URL: http://projecteuclid.org.ezproxy.universite-paris-saclay.fr/euclid.cmp/1103899048
  • [BR76a] O. Bratteli and D.. Robinson “Unbounded derivations of CC^{\ast}-algebras. II” In Comm. Math. Phys. 46.1, 1976, pp. 11–30 URL: http://projecteuclid.org.ezproxy.universite-paris-saclay.fr/euclid.cmp/1103899543
  • [Ful84a] W. Fulton “Intersection theory” 2, Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)] Springer-Verlag, Berlin, 1984, pp. xi+470 DOI: 10.1007/978-3-662-02421-8
  • [HJ84a] A. Hulanicki and J.. Jenkins “Nilpotent Lie groups and summability of eigenfunction expansions of Schrödinger operators” In Studia Math. 80.3, 1984, pp. 235–244 DOI: 10.4064/sm-80-3-235-244
  • [Con86a] A. Connes “Cyclic cohomology and the transverse fundamental class of a foliation” In Geometric methods in operator algebras (Kyoto, 1983) 123, Pitman Res. Notes Math. Ser. Longman Sci. Tech., Harlow, 1986, pp. 52–144
  • [Con94a] A. Connes “Noncommutative geometry” Academic Press, Inc., San Diego, CA, 1994, pp. xiv+661
  • [Bla98a] B. Blackadar KK-theory for operator algebras” 5, Mathematical Sciences Research Institute Publications Cambridge University Press, Cambridge, 1998, pp. xx+300
  • [Ped00a] G.. Pedersen “Operator differentiable functions” In Publ. Res. Inst. Math. Sci. 36.1, 2000, pp. 139–157 DOI: 10.2977/prims/1195143229
  • [Car08a] P. Carrillo Rouse “A Schwartz type algebra for the tangent groupoid” In KK-theory and noncommutative geometry, EMS Ser. Congr. Rep. Eur. Math. Soc., Zürich, 2008, pp. 181–199 DOI: 10.4171/060-1/7
  • [DL10a] C. Debord and J.-M. Lescure “Index theory and groupoids” In Geometric and topological methods for quantum field theory Cambridge Univ. Press, Cambridge, 2010, pp. 86–158 DOI: 10.1017/CBO9780511712135.004
  • [DS14a] C. Debord and G. Skandalis “Adiabatic groupoid, crossed product by +\mathbb{R}_{+}^{\ast} and pseudodifferential calculus” In Adv. Math. 257, 2014, pp. 66–91 DOI: 10.1016/j.aim.2014.02.012
  • [EY17a] E. Erp and R. Yuncken “On the tangent groupoid of a filtered manifold” In Bull. Lond. Math. Soc. 49.6, 2017, pp. 1000–1012 DOI: 10.1112/blms.12096
  • [HH18a] A. Haj Saeedi Sadegh and N. Higson “Euler-like vector fields, deformation spaces and manifolds with filtered structure” In Doc. Math. 23, 2018, pp. 293–325
  • [CP19a] W. Choi and R. Ponge “Tangent maps and tangent groupoid for Carnot manifolds” In Differential Geom. Appl. 62, 2019, pp. 136–183 DOI: 10.1016/j.difgeo.2018.11.002
  • [Ewe21a] E. Ewert “Pseudodifferential operators on filtered manifolds as generalized fixed points” In To appear in Journal of noncommutative Geometry arXiv, 2021 DOI: 10.48550/ARXIV.2110.03548
  • [Moh21a] O. Mohsen “On the deformation groupoid of the inhomogeneous pseudo-differential calculus” In Bull. Lond. Math. Soc. 53.2, 2021, pp. 575–592 DOI: 10.1112/blms.12443

(Omar Mohsen) Paris-Saclay University, Paris, France e-mail: [email protected]