Differential operators on -algebras and applications to smooth functional calculus and Schwartz functions on the tangent groupoid
Abstract
We introduce the notion of a differential operator on -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 be a compact smooth manifold. In [Con94a, Chapter 2.5], Connes introduced the tangent groupoid
The space 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 is equipped with a convolution law making it a -algebra. The convolution law is a deformation of the usual convolution law on which is used in Schwartz kernels of operators acting on and the convolution law on for which comes from the commutative group structure. As Connes shows, the space 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 on introduced by Carrillo-Rouse [Car08a]. The -algebra can be naturally completed into a -algebra . In [Car08a], Carrillo-Rouse proves that is a -subalgebra of . In this article, we improve his result by showing that is closed under smooth functional calculus.
Theorem A.
-
1.
If is a Schwartz function and is normal () and is a smooth function on some open neighbourhood of with , then .
-
2.
If is a Schwartz function and is a holomorphic function on some open neighbourhood of with , then .
Of course in Theorem A means the functional calculus in applied to . The main difficulty in Theorem A is the derivability of in the direction of because as , the structure of changes from compact operators on to . We remark that we suppose because isn’t unital.
To prove Theorem A we introduce differential operators on an arbitrary -algebra . Recall that a derivation on is a map defined on a -subalgebra such that
Derivations usually come from the derivative of an action on . If and where is a smooth manifold , then a derivation is a vector field on . Therefore, one can think of derivations as differential operators of order on with domain . We introduce the notion of a differential operator of higher order on an arbitrary -algebra . Our differential operators coincide with differential operators on if and . We show that the derivative in the direction of is a differential operator on of order . The main theorem of this article is the following, see Definition 1.8 for the precise definition of . We remark that is a -subalgebra of which contains .
Theorem B.
Let be a dense -subalgebra and be the intersection of the domains of the closures of all differential operators defined on . Then
-
1.
If , is normal () and is a smooth function defined on some open neighbourhood of with , then .
-
2.
If , and is a holomorphic function defined on some open neighbourhood of with , then .
Here is the spectrum of as an element of , the -algebra of matrices with coefficients in .
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 .
The main interest in subalgebras closed under holomorphic calculus in the theory of operator algebras is that they give isomorphisms in -theory, thus allowing tools from cyclic cohomology to be used. Cyclic cocycles are rarely defined on the whole -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 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 such that is bounded whenever is a differential operator with polynomial coefficients. Analogously, on the space , 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.
1 Differential operators on -algebras
Throughout the article stands for , for and for .
Example 1.1.
Let
be the Heisenberg group and
(1) |
The maps and satisfy the Leibniz rule
but satisfies
(2) |
where 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 be a -algebra equipped with an involution, a -subalgebra of -linear endomorphisms of which contains the identity.
-
1.
We say that is a -differential operator of order if for all .
-
2.
We define -differential operators of higher order by recurrence. A map is called a -differential operator of order , if there exists a finite family of linear maps
which are -differential operators of order such that
(3) Furthermore, we suppose that
(4)
We denote by the space of -differential operators of any order.
Remarks 1.3.
Let us explain Definition 1.2 in more details.
-
1.
We use a subspace of linear maps instead of the space of all linear endomorphisms of , because in applications, it seems unhelpful to consider all linear endomorphisms. Usually, one has a given subspace of linear endomorphisms of interest.
-
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 -differential operators we consider have the property with and thus trivially satisfy (4). We include the general case of arbitrary satisfying (4) because the arguments in this section don’t change with its inclusion.
-
3.
It is worth nothing that if is a -differential operator of order , then it is also a -differential operator of order for any . It might be tempting to define the order of a -differential operator to be the minimal such that is of order . We won’t do so because the order behaves in a rather unintuitive way. For example, the sum of a -differential operator of order and another of order isn’t necessarily of order . It is always of order . See the proof of Proposition 1.5.
Remark 1.4.
The following observation will be used in the proof of Proposition 1.5. If such that there exists -differential operators such that (3) and (4) hold, then is a -differential operator of order one plus the maximum of the orders of . Whenever we apply this remark, to make it clearer how it is applied, we colored the term by red, the term by green, and the rest by blue.
Proposition 1.5.
The space is a unital subalgebra of and thus of as well.
Proof.
It is clear that is a -differential operator of order . Let and be -differential operators of order and respectively with and as in Definition 1.2. Then,
Hence, is a -differential operator of order by Remark 1.4. Notice here that there are only blue terms. This is possible because, one already supposes that and are -differential operators.
We prove that is a -differential operator by induction on the order. The case is clear. We prove the case by induction on . We have
(5) |
All the terms are -differential operators by recurrence. Hence, by Remark 1.4, is also a -differential operator.
The case is also proved by induction on . We have
(6) |
All the terms are -differential operators by recurrence. Hence, by Remark 1.4, is again a -differential operator.
We can now prove the general case of the composition of -differential operators by induction on . We suppose that it is true for any such that and and . One has
(7) | ||||
Again by Remark 1.4, is a -differential operator.
Finally, since the map is a differential operator of order for any , it follows that , which is a composition of two -differential operators, is also a -differential operator. This finishes the proof that is a unital subalgebra of .: ∎
Definition 1.6.
We say that a unital subalgebra of is stable if .
It is clear from Definition 1.2 that for any unital subalgebra of , is stable.
We now give natural examples of stable subalgebras. Let be a nilpotent Lie algebra. The Baker-Campbell-Hausdorff formula is a finite sum because is nilpotent. Hence, we can view as a nilpotent Lie group. Let . We equip with the convolution product after fixing some Haar measure on . We remark that a Haar measure is equivalently a Lebesgue measure because is nilpotent.
Proposition 1.7.
For a differential operator on with polynomial coefficients, let
The space of all maps as varies is a stable unital subalgebra of , i.e., is a -differential operator for every .
Proof.
It is clear that is a unital subalgebra. We now show that it is stable. For , let be the associated right invariant vector field. One easily sees that is a -differential operator of order . We can write as a sum of terms of the form , where is a polynomial and are right-invariant vector fields. Hence, by Proposition 1.5, it suffices to prove that if , then is a -differential operator. Again, by Proposition 1.5, we can suppose that is linear. Let and be the central series of . Since is nilpotent, there exists such that . We choose a linear decomposition such that . It is enough to consider for some . We proceed by induction on . For , is a -differential operator of order . In fact
(8) |
The induction case follows by using the Baker-Campbell-Hausdorff formula like in (2). ∎
From now on, we suppose that is a dense -subalgebra of a -algebra .
Definition 1.8.
Let be a stable unital subalgebra of . We denote by the set of such that there exists a sequence such that and for every , and converge in .
By taking constant sequences, we see that .
Remark 1.9.
Let . Since is generally an unbounded operator, it is desirable to take the closure of the graph of . 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, , and denotes the space of matrices with coefficients in , and respectively.
Theorem 1.10.
Let be a stable unital subalgebra of . The set is a dense -subalgebra of . It is also closed under smooth functional calculus, i.e,
-
1.
If , is normal () and is a smooth function defined on some neighborhood of with , then .
-
2.
If , and is a holomorphic function defined on some neighborhood of with , then .
Here is the spectrum of as an element of .
For clarity of the exposition, the proof of Theorem 1.10 will be divided into several lemmas
Lemma 1.11.
The space is a dense -subalgebra of .
Proof.
The space is dense because . Proving that is closed under addition, involution, multiplication by scalars is straightforward. Let , such that and and , , , converge for every . For any , we can find , such that
(9) |
Since and converge for every , we obtain that converges. By a similar argument, we deduce that converges. Hence, . Therefore, is a dense -algebra. ∎
Let ,
(10) |
Clearly is an algebra homomorphism.
Lemma 1.12.
For any , is a stable subalgebra of . Furthermore, .
Proof.
Lemma 1.13.
Let be self-adjoint and . Then, . Furthermore, we can find a family such that
-
1.
For every compact subset of , the sequence converges uniformly in to as . Moreover,
(12) -
2.
For every compact subset of and for every , the sequence converges uniformly in to an element in as . We denote the limit by .
-
3.
For every , the function
(13) is continuous. Furthermore, if is of order , then
(14) -
4.
For any , , .
We remark that the notation is an abuse of notation. In general doesn’t belong to , therefore the value of usually depends on the choice of a sequence converging to 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.
Let be self-adjoint, a function on a neighbourhood of with . Since we are only interested in , without loss of generality, we can suppose . Since , we can write where . The Fourier transform of is a Schwartz function. By the Fourier inversion formula
Hence,
(15) and the integral is absolutely convergent. We approximate the integral by the finite Riemann sums
It is clear that as . Let be as in Lemma 1.13. We define
It is clear that . Since converges to uniformly in on compact sets, and is Schwartz, and by (12), we deduce that . Hence, . Let . By Lemma 1.13.3, the integral
(16) is absolutely convergent. By using Riemann sums like we did with Integral (15), we deduce that
Since , we can reuse our argument to deduce that
Hence, .
-
2.
Let be a normal element, a function on a neighbourhood of with . Like before, we can suppose that . We write for the coordinates in . Consider the function . Then, . It follows from the self-adjoint case that . Hence, by replacing with , we can suppose that on . By a similar argument, we can suppose that on both and . Hence, there exists such that . Since is smooth compactly supported, it follows that is a Schwartz function. Let and . By the Fourier inversion formula
(17)
The proof of Lemma 1.13 requires a few lemmas. Let be a self-adjoint element. We fix a sequence such that and and converge for every . By replacing with , we can suppose that .
Lemma 1.14.
For every and of order , there exists such that
(18) |
Proof.
We prove the lemma by induction on . For and , we have
(19) |
Since and converge, it follows that they are bounded. The case is finished.
Suppose the lemma holds for of order . Let be of order . For each , the Lemma holds for . Therefore, there exists constants such that
(20) |
Let be a constant such that
(21) |
Let be a constant such that
(22) |
Some constants exist because , converge in .
Let , . By Lemma 1.14, the sum
(26) |
converges in norm uniformly in . By an abuse of notation, we denote the sum by . Since exists for all and the convergence in (26) is uniform, it follows that
(27) |
exists.
Lemma 1.15.
For every , of order , there exists constant such that
Proof.
We prove the lemma by induction on . If , then by (19)
(28) |
Since are self-adjoint, it follows that . The case follows.
Proof of Lemma 1.13.
Let be any function such which satisfies
(30) |
We take
(31) |
The element is a finite sum of elements of . Hence, it belongs to . We now check that satisfies the required properties.
- 1.
-
2.
By Lemma 1.14, it is again straightforward to show that converges to uniformly on each compact subset of .
- 3.
-
4.
The identity 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 be a compact topological space, , a family of elements which satisfies the following:
-
1.
For every , .
-
2.
For every , the function is continuous in .
-
3.
The limit of as is equal to uniformly in .
-
4.
For every , the limit as exists in uniformly in
-
5.
For every , the limit as exists in uniformly in
Let be a continuous complex-valued function defined in an open neighbourhood of
which is smooth in the variable, and which satisfies . Then, there exists a family which satisfies the following:
-
1.
For every , the function is continuous in .
-
2.
The limit of as is equal to uniformly in .
-
3.
For every , the limit as exists in uniformly in
-
4.
For every , the limit as exists in uniformly in
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 . For every , . Furthermore, if is a compact subset, then we can find a family such that the following is satisfied:
-
•
The sequence converges uniformly in to .
-
•
For every , converges uniformly in to an element denoted by .
-
•
For every , converges uniformly in to an element denoted by .
-
•
For every , the maps and are continuous.
Proof.
Proof of Theorem 1.10.2.
Let , and a holomorphic function on a neighbourhood of with . Since , for some holomorphic function . Let be a contour around such that is in the domain of and the Cauchy integral formula applies
We thus have
Without loss of generality, we can suppose that doesn’t belong to the image of .
Let be as in Lemma 1.17 applied to equal to the image of . We approximate by a Riemann sum, then approximate each term by . This way we obtain a sequence which converges to and such that converges to and converges to . The result follows. ∎
Corollary 1.18.
Let , be as in Proposition 1.7, be the -subalgebra of Schwartz functions. Then
-
1.
If is normal and is a smooth function defined on some neighbourhood of with , then .
-
2.
If and is a holomorphic function defined on some neighbourhood of with , then .
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 induces an isomorphism in -theory .
2 Schwartz functions on Connes’s tangent groupoid
Let be a compact smooth manifold. The set
can be equipped with the structure of a smooth manifold with boundary as follows. The subset is declared an open subset of with its usual smooth structure. If is an open set, is a diffeomorphism, then is declared an open subset of and the following bijection is declared a diffeomorphism
(35) | ||||
One can check that this defines a smooth structure on . 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
(36) |
and
(37) |
be smooth, where .
Differential operators.
Let be a vector field. We can define a vector field on as follows. If is a smooth function, then we define by
where means that acts on the variable, and means that we view as a constant vector field on which acts on the smooth function . By the description of charts above, one checks that is indeed a vector field.
More generally, let be a differential operator of order acting on . Then we can define a differential operator denoted of order acting on by the formula
(38) |
where means that acts on the variable, and means that we view the principal part of as a constant coefficient differential operator on acting on the smooth function . Locally by the principal part, we mean that if , then seen as a constant coefficient differential operator on . Using the local charts of , one can check that is indeed a differential operator.
Convolution and adjoint.
From now on, we suppose that is equipped with a Riemannian metric. If , then we define their convolution by the formula
(39) | ||||
where the integral over is with respect to the constant Riemannian metric on induced from the Riemannian metric on . We also define the adjoint by
We leave it to the reader to check that 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 .
Let be a differential operator on of order , the associated differential operator on . We sometimes denote by . We also define by the formula
where is the differential operator on associated to the formal adjoint of . Equivalently
where is the formal transpose given by
One can check that
(40) |
This justifies our notation and .
Derivations in direction of .
The functions and agree at . It follows that the map
is a well-defined differential operator on of order . If is an order differential operator given by multiplication by , then is an order differential operator given by multiplication by , see (37). For , one can describe as follows. The flow of gives an action on by
The differential operator is the derivative of the action (plus multiplication by the divergence of ). Since is given by a commutator, it satisfies the Leibniz rule
(41) |
and
(42) |
Derivation in direction of .
The group acts smoothly on the left on by the formula
(43) |
We also let acts on functions by the formula
One has
(44) |
We define to be the derivative of the -action at . One can check that
where are any local coordinates for . It follows from (44) that satisfies the Leibniz rule
(45) |
Lemma 2.1.
One can find a finite family of smooth functions and vector fields such that for any , .
Proof.
Let be smooth functions such that for all , span . Let be the trivial vector bundle on of rank . The forms define a surjective bundle map . Its dual is an injective bundle map . Let be any bundle map such that . The map determines the vector fields . ∎
We fix a choice of for the rest of this section. By Lemma 2.1, if , then is equal to the identity . By taking the trace of , we deduce that
(46) |
For , the functions and agree at . Hence, we can define
(47) | ||||
where in the second equality we used (46).
Lemma 2.2.
The map satisfies
and hence a derivation of order on .
Schwartz functions.
Definition 2.3.
We define to be the space of smooth function such that is bounded and all iterated applications of the following differential operators give bounded functions on
-
•
the operator , where is the natural projection.
-
•
the operator , where is any differential operator on
-
•
the operator , where is any differential operator on
-
•
the operator .
We can add the application to the above list, but this is redundant as it follows from the others using (47).
Proposition 2.4.
Let . Then for every , the function is a Schwartz function on the vector space in the classical sense.
Proof.
If is a vector field, then is the application of the constant vector field to . While if , then the map is the pointwise product of the map with the linear map . By iterated application of the previous two operations, it follows that is Schwartz. ∎
Example 2.5.
Let be smooth functions such that the map
is an embedding. Then, the function
is Schwartz, where 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 and are bounded where is either a vector field or a smooth function on .
Proposition 2.6.
The following holds
-
1.
The definition of Schwartz functions doesn’t depend on the choice of in Lemma 2.1.
-
2.
If , then , i.e. vanishes at infinity.
-
3.
If , then the integral in is absolutely convergent and
-
4.
If , then
Proof.
For the first part, we first need some lemmas.
Lemma 2.7.
If and , then the pointwise product , where is the map in (36)
Proof.
is bounded because and are bounded. We argue that each of the applications in Definition 2.3 gives functions of the form for some and . The result then follows by induction. We thus have
-
•
the function
-
•
If is a function on , then . If is a vector field, then we have
(48) where is the action of on the variable in . The equality can be checked directly on and by density equality follows on .
-
•
If is a function, then . If is a vector field, then one can as in (48) show that
-
•
one has
and so
This finishes the proof. ∎
Lemma 2.8.
Let be a smooth function that vanishes to the order on the diagonal, and let be the smooth function given by
where is the Hessian of . Then if then the pointwise product .
Proof.
Any function which vanishes to the order can be written as a finite sum of functions of the form where . It follows that the pointwise product is sum of pointwise product of with where is the natural map. The result then follows from Lemma 2.7. ∎
Lemma 2.9.
Let and be another family satisfying Lemma 2.1. Then there exists a family of smooth functions for and such that
(49) |
Proof.
Let be local coordinates in . Then and for some functions and . Similarly, and . Then one can take . To define globally one takes a partition of unity. ∎
We can now prove Proposition 2.6.1. Let be the operator associated to and . We have
By (49), the function vanishes to order on the diagonal. By Lemma 2.8, . Proposition 2.6.1 follows.
For Proposition 2.6.2, suppose that . Then there exists a sequence in and which goes to infinity yet for every element in the sequence. By passing to a subsequence we can suppose that the sequence is either of the form or of the form . Suppose we have the first case. Then by taking a subsequence, we can suppose that and and . If , we get a contradiction to the fact that is bounded. If , then the sequence converges in to , again a contradiction. If then we get a contradiction to the fact that is bounded where is any smooth function with . So we have and . The sequence converging to infinity implies that there exists such that . We get then a contradiction to the fact that is bounded. The case of a sequence is similar.
For Proposition 2.6.3, first the integral in is absolutely convergent by Proposition 2.4. Hence, is a well-defined function on . We now show its continuity. Let be as in Example 2.5. Then consider the function
(50) |
The function being Schwartz implies that is bounded. One can easily show by looking at local coordinates of that there exists (only depends on ) such that
Since and are Schwartz functions, . We can thus approximate them uniformly with compactly supported functions. It follows that . Smoothness of as well as the fact that follow easily from (40), (41), (44) and Lemma 2.2.
For Proposition 2.6.4, since is bounded if is bounded, the result follows from the following identities
and the identity
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 is Schwartz if and only if it satisfies the following:
-
1.
For every , a differential operator on ,
-
2.
For every , a differential operator on , a compact subset outside the diagonal,
-
3.
For every open subset diffeomorphic to by a map , the local chart of associated to defined in (35), a compact subset, , , one has
where .
-algebra of the tangent groupoid.
Let , . We define the operator
For each , we also define the operator
We then define
We define the -algebra to be the completion of with respect to . The -algebra lies in a short exact sequence
where is the -algebra of compact operators on , see [Con94a, Proposition 5 Page 108]
Proposition 2.11.
The space is a -subalgebra of .
Relation between uniform norm and -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 -norm, which is more convenient to use.
Theorem 2.12.
Let be the positive Laplace-Beltrami operator on , the corresponding differential operator on as in (38), and with . There exists , such that
Lemma 2.13.
Let with . There exists a constant such that for all and , if denotes the Dirac delta distribution on at , then
Proof.
The Sobolev embedding lemma implies that where denotes the Sobolev space. Hence, . For we have and use the inequality
where in the last inequality we used the fact that for all . For , we proceed differently. If , then
We need to maximize as varies in . The function needs to vanish outside any given neighbourhood of to reach the supremum. We can thus assume that and and the metric on is Euclidean outside the unit ball. The inequality follows from a change of variable . ∎
Proof of Theorem 2.12.
Let . We apply
to to deduce that
One has
It follows that
By taking the limit as , one obtains a bound of on . The theorem follows. ∎
Schwartz functions are closed under smooth calculus.
Proof of Theorem A.
Let be the subalgebra of generated by the maps in Definition 2.3. We will prove that . Theorem A would then follow from Theorem 1.10. Let , as in Definition 1.8. Since
is a differential operator on which belongs to , it follows that converges in . By Theorem 2.12, it follows that is a bounded continuous function and in uniform norm. Let . Since the map belongs to , again by Theorem 2.12, converges in the uniform norm to a bounded continuous function on . It follows that is smooth and bounded for every . Hence, . ∎
Remark 2.14.
[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 “-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 -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 -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 -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 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 -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 -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 “-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 -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 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]