Subcomplexes on filtered Riemannian manifolds
Abstract.
In this paper, we present a general construction to extract subcomplexes from two distinct complexes on filtered Riemannian manifolds. The first subcomplex computes the de Rham cohomology of the underlying manifold. On regular subRiemannian manifold equipped with a compatible Riemannian metric, it aligns locally with the so-called Rumin complex. The second complex instead generalises the Chevalley-Eilenberg complex computing Lie algebra cohomology of a nilpotent Lie group. Our approach offers key insights on the role of the Riemannian metric when extracting subcomplexes, opening up potential new applications in more general geometric settings, such as singular subRiemannian manifolds.
Key words and phrases:
Differential forms on filtered manifolds, Subcomplexes in sub-Riemannian geometry, Osculating nilpotent Lie groups2010 Mathematics Subject Classification:
58A10, 58J10, 58H99, 53C17, 43A80, 22E251. Introduction
1.1. Motivation
The de Rham complex and its cohomology have inspired many mathematical ideas in the twentieth century, including Hodge theory and the Atiyah-Singer index theorem on Riemannian manifolds. More recently, an increasing number of techniques have been developed in order to extract subcomplexes with the aim of extending such classical results to more general geometric settings [CCY16, Tri54, DH22, Cas11, ČH23, LT23], especially in subRiemannian geometry.
In [FT23], we proposed two constructions on homogeneous nilpotent Lie groups, one by adapting techniques previously developed in the context of parabolic geometry [ČSS01, CD01, DH22] and the other influenced by the ideas in subRiemannian geometry and spectral sequences [Jul95, Rum05]. In the present paper, we push these ideas further to obtain subcomplexes from two different initial complexes and with two equivalent but different constructions. One one hand, we start from the usual de Rham complex to extract a subcomplex computing the de Rham cohomology of the underlying manifold , while on the other, we extract from the “algebraic” complex . Here denotes the algebraic part of the de Rham differential , and so the complex generalises the Chevalley-Eilenberg complex computing the Lie algebra cohomology of a given group to the manifold setting.
1.2. Sketch of the constructions of the complexes
Here, we explain our setting and briefly sketch our two equivalent constructions of each of the two complexes and . The various claims will be proved in Section 4.
We consider a filtered manifold and denote by its osculating nilpotent Lie algebra (also called nilpotentisation) above each point (see Section 3.1). We then define the osculating Chevalley-Eilenberg differential on the space of osculating forms over , defined above each point as the Chevalley-Eilenberg differential of the osculating group (see Section 3.1.2).
We now further assume that is also equipped with a metric on . This allows us to construct the natural isomorphism , providing an identification between forms and osculating forms (see Section 2.5). We are then able to define the map acting on , which we call the base differential.
1.2.1. Constructing using
The transpose of makes sense unambiguously and globally thanks to the metric on . We consider the projection onto along and the projection onto along . The proper definitions are in fact given by being the orthogonal projection onto the kernel of the algebraic operator , and being the generalised kernel projection of the differential operator . Once we define the invertible differential operator , we have the decomposition
yielding two complexes and whose cohomologies are described as follows:
has trivial cohomology, since is conjugated to on , and is acyclic.
The cohomology of is linearly isomorphic to the de Rham cohomology of via the homotopically invertible chain map .
1.2.2. Constructing using
Relying again on the metric on , we can define the orthogonal projection onto . We then consider the partial inverse of and thus the corresponding partial inverse of . Then the projection above may be constructed explicitly in terms of as:
The complex is then obtained as .
1.2.3. Constructions of
We can modify the two constructions above by applying the same steps, but substituting the initial differential operator with the algebraic part of the de Rham differential instead (see Section 2.2.1). This yields two complexes and , with acyclic and linearly isomorphic to .
1.3. Novelty and future works
As already mentioned above, constructions related to the ones presented in this paper for the complex appeared before, in fact more than two decades ago in parabolic geometry [ČSS01, CD01, DH22], and in relation to subRiemannian geometry and spectral sequences [Rum90, Jul95]. Already at that time, it was conjectured that these constructions should align in the resulting subcomplex in some sense (see Section 5.3 in [Rum05], Section 5 in [JK95], Section 2.1 in [Jul19], Section 5.3.3 in [GAJV19], and Section 1.3 in [DH22]), but no proof was offered. An important difficulty incomparing the constructions is that the considerations in the subRiemannian settings were either local or at the level of osculating objects (see Section 5.1 and Remark 5.2 in [Rum05]).
The constructions presented in this paper are set on a filtered Riemannian manifold ; they yield complexes globally defined and acting on - not its osculating counterpart . We obtain two equivalent constructions for the complex , which in the subRiemannian world and together with its osculating couterparts are often referred to as the Rumin complex. We are then able to provide a clearer context and a proof to the conjecture explained above. We also show that these two equivalent constructions may be adapted to yield a different complex, that we have denoted by . In Appendix A, we give an explicit construction for the operators and on a 3D contact manifold.
We have two main examples in mind for our geometric setting of a filtered Riemannian manifold. We discuss them in turn.
Example 1.1.
A regular subRiemannian manifold equipped with a compatible Riemannian metric (in the sense of Definition 4.13) is naturally a filtered Riemannian manifold.
Within the setting of Example 1.1, we are able to show that the subcomplex coincides with what is now customarily referred to as the Rumin complex on regular -structures, first introduced in [Rum90, Rum99, Rum05]. Part of the motivation behind the present work is to shed a new light on Rumin’s construction on regular subRiemannian manifolds and on the nature of constructed objects (e.g. complexes acting on forms or on osculating forms). We also want to address the potential impact that the choice of a Riemannian metric can have on the resulting subcomplex. In particular, in this paper, we emphasise the nontrivial role that the extra hypothesis of compatibility between the Riemannian and subRiemannian structures plays in the construction, see Section 4.6.2. Our hope is to highlight the behaviour of subcomplexes such as the Rumin complex, under changes of variables on regular subRiemannian manifolds [FT12] or pullback by Pansu-differentiable maps on Carnot groups [KMX28].
Example 1.2.
A nilpotent Lie group equipped with a left-invariant filtration is naturally a filtered Riemannian manifold.
Within the setting of Example 1.2, the complex computes the Lie algebra cohomology of the given Lie group. A more thorough study of this setting, especially concerning the impact of the particular choice of a filtration on the resulting subcomplex , will be presented in a forthcoming paper [FT24].
We selected the two specific settings in Examples 1.1 and 1.2 due to the exciting recent advances in their applications:
- •
- •
- •
We believe that the techniques behind the subcomplexes and presented in Sections 1.2 and 4 are general enough that their construction will be adaptable to other settings in the future. Interesting generalisations would be to singular subRiemannian manifolds (e.g. Martinet distributions) and to the quasi-subRiemannian setting, for instance the Grushin plane.
1.4. Organisation of the paper
In order to present the hypotheses of our setting and to remove any ambiguity on the objects we consider, the paper starts with some foundational preliminaries on filtrations over vector spaces and on tangent bundles of manifolds (Section 2). When the manifold is filtered (i.e. when the tangent bundle is filtered and the commutator brackets of vector fields respect this filtration), then we can define the osculating Chevalley-Eilenberg differential and other osculating objects (see Section 3). In Section 4, we present the general scheme for our construction, discussing the particular case of regular subRiemannian manifold at the end in Section 4.6, and the particular case of 3D contact manifolds in Appendix A.
1.5. Acknowledgements
Both authors acknowledge the support of The Leverhulme Trust through the Research Project Grant 2020-037, Quantum limits for sub-elliptic operators. We are also glad to thank the Centro di Ricerca Matematica Ennio De Giorgi and the Scuola Normale Superiore for the hospitality and support in the early stages of the article, and Professor Giuseppe Tinaglia for hosting us at King’s College London in October 2023.
2. Preliminaries
The main objective of this section is to set some notation for well-known notions regarding manifolds (Section 2.2) and filtrations. For the latter, the filtrations are on vector spaces and vector bundles (Sections 2.1 and 2.4 respectively), which will come in handy when considering a manifold equipped with a metric (Section 2.5).
2.1. Filtrations of vector spaces
Here, we briefly recall the construction of graded objects associated with a filtration of vector spaces. The notions presented below extend naturally to smooth vector bundles and their smooth sections, as well as to modules.
In this paper, all the vector spaces are taken over the field of real numbers.
2.1.1. Filtered vector spaces
A filtered vector space is a vector space with a filtration by vector subspaces:
(2.1) |
In general, a filtration need not be finite; however, in this paper, all the filtrations considered will be finite as above.
A graded vector space is a vector space equipped with a gradation by vector subspaces, i.e. a direct sum decomposition by vector subspaces :
(2.2) |
If is graded, there is a natural filtration associated with its gradation (2.2), given by , where , .
2.1.2. The vector space
Given a filtered vector space with the filtration (2.1), its associated graded space is the vector space
(2.3) |
2.1.3. The linear map
Consider a linear map between two vector spaces, each equipped with a filtration and . We may assume .
The map is said to respect the filtrations when for every . In this case, one can define the linear map as
2.1.4. The basis
If is finite dimensional, then has the same dimension as . Given and , we say that a basis of is adapted to the filtration (2.4), when is a basis of , is a basis of , and so on. For such a basis, we have that, for each ,
gives a basis of .
In particular, we get that for each , the -tuple is a basis of . In this sense, the basis is graded and is said to be the graded basis associated with .
2.1.5. Matrix representation of linear maps
In this paper, we will often use matrix representations of certain maps. To fix some notation, given a morphism between two finite dimensional vector spaces, once we fix and two bases of and respectively, we denote by
the matrix representing in the bases and .
Let us take and two bases adapted to the filtrations (2.4) of two finite dimensional filtered vector spaces and respectively. Assuming , and , if a linear map respects the filtrations, the matrix representation of in the bases is block-upper-triangular, i.e.
with , with . Moreover, the matrix representing in is block-diagonal, i.e.
2.1.6. Subfiltration
We say that a vector subspace admits a subfiltration when , yields a filtration of . In this case, the injection map respects the filtrations and is a subspace of .
2.1.7. Decreasing labelling and duality
One can also consider a filtered vector space with a decreasing filtration, that is a filtration with decreasing labelling:
(2.5) |
It is not difficult to adapt all the constructions considered previously to a decreasing filtration. Indeed, it suffices to change the labels by setting , so that, for instance, the associated graded space becomes . However, in this setting, the matrix representations will differ from those in Subsection 2.1.5, as the matrix representing a linear map that respects such decreasing filtrations will be block-lower-triangular.
We observe that if has a decreasing filtration, then we obtain an increasing filtration of the dual space of by considering the filtration by the subspaces , the annihilators of the subspaces , that is
also known as the dual filtration of (2.5).
Given two finite dimensional filtered vector spaces and , let us take and two bases adapted to the decreasing filtrations and . Then for any linear map that respects these decreasing filtrations, the dual map
respects the dual filtrations, and
where and denote the dual bases of and respectively (one can easily show that if a basis is adapted to a decreasing filtration (2.5), then is adapted to its increasing dual filtration). In particular, this readily implies that the matrix representation of is block-upper-triangular, as pointed out in Subection 2.1.5.
Moreover, the vector spaces and are canonically isomorphic.
2.1.8. Graded vector spaces
As explained earlier, given a graded vector space , its gradation (2.2) naturally produces a filtration through , while a filtration (2.1) of a vector space yields a graded space (2.3).
Definition 2.1.
Let and be two graded vector spaces. We may assume . A linear map respects the gradations when for any .
For instance, if is a linear map between two filtered vector spaces that respects the filtrations, then respects the gradations.
Consider a graded vector space as above. Then its dual is also graded with the dual gradation . The exterior algebra is also graded via the (possibly infinite) gradation
This leads to a (possibly infinite) gradation of the exterior algebra which respects the wedge product:
with , and for . If is finite dimensional, the gradations of and are finite.
In the graded setting, there is a natural notion of weight: an element in has weight , and more generally, an element in has weight .
2.1.9. Filtration on the exterior algebra.
A filtration (2.1) on a vector space also induces a decreasing filtration on its exterior algebra:
(2.6) |
where
The following lemma is easily checked.
Lemma 2.2.
Let be a finite dimensional vector space equipped with a filtration (2.1), and let be a basis of adapted to the filtration. Define the linear map
via
Then is a linear isomorphism that respects the filtration (2.6) of and the one induced by the gradation of . Moreover, provides an isomorphism between
and respecting their gradations.
Note that the isomorphism depends on the choice of the basis .
2.1.10. Euclidean filtered vector spaces
Here, we consider a Euclidean space, that is, a finite dimensional vector space equipped with a scalar product .
Any subspace is naturally Euclidean, its scalar product being obtained by restricting . Moreover, its orthogonal complement in is naturally identified with the quotient
This quotient is also Euclidean if we equip it with the scalar product corresponding to .
Beside being Euclidean, we further assume the vector space also be filtered with filtration (2.4). The resulting graded space inherits a natural scalar product by imposing that the decomposition (2.3) is orthogonal, and that restricted to each is given by . Indeed, each is naturally isomorphic to the orthogonal complement of in . Therefore, by construction, the gradation by quotients on is isomorphic to the following gradation of :
which we will refer to as the orthogonal gradation or orthogonal graded decomposition of (associated with the given scalar product and filtration).
We observe that when considering a basis of adapted to its filtration, after a Graham-Schmidt process, we may assume that is an orthonormal basis of , is an orthonormal basis of , etc. In other words, after applying a Graham-Schmidt process, we can always obtain a basis adapted to the given orthogonal gradation of .
2.2. Preliminaries on manifolds
2.2.1. The maps and on a general manifold
In this section, we recall briefly some well-known facts about the de Rham differential which are valid on any smooth manifold (without any further assumptions). As is customary, denotes the exterior differential on the space of smooth forms on . It is well known that the map is a smooth differential map that satisfies the Leibniz property and . The associated cochain complex is traditionally referred to as the de Rham complex.
The explicit formula for is given for any and by
the hats denoting omissions.
In this paper, the algebraic part of is denoted by . This is the map given by on , and for and any , , by
(2.7) |
Lemma 2.3.
The map is smooth and algebraic. It satisfies the Leibniz property and is a complex, i.e. . The action of over is given by the explicit formula
Sketch of the proof.
Since and for any
(2.8) |
the Leibniz property of follows from the fact that both vector fields and satisfy the Leibniz property.
The formula on 1-forms is easily computed from (2.7) by taking . This formula, together with the Jacobi identity for the commutator bracket, implies on . Recursively and by the Leibniz rule, we get that on forms of any degree, since for any and any
∎
2.3. Some natural concepts
In this subsection, we recall some basic concepts in differential geometry.
2.3.1. Frame and coframe
Let be a (real, finite dimensional, and smooth) vector bundle over a (smooth) manifold with . A frame for is a smooth section of such that is a basis of for every . A coframe for is frame for the dual vector bundle .
Throughout this paper, if the bundle is not specified, we will take to be the tangent bundle over the smooth manifold and to be the cotangent bundle. Frames for may not exist globally on the whole manifold , but they can always be constructed locally, i.e. over an open neighbourhood of every point in .
2.3.2. Algebraic maps on bundles
Let and be two smooth vector bundles over a manifold . We say that a smooth map is linear and algebraic when, for each , there exists a linear map satisfying for any , . We will call the pointwise restriction of above .
Remark 2.4.
A (smooth) differential operator of order 0 is an algebraic linear map.
2.3.3. Metric on bundles
Let be a vector bundle over a manifold . By definition, a metric on is a family of scalar products on depending smoothly on . The smooth dependence means that
When is the tangent bundle of a smooth manifold , the metric is said to be Riemannian and the couple is referred to as a Riemannian manifold.
2.4. Filtered tangent bundles
In this section, we consider an -dimensional smooth manifold whose tangent bundle is filtered by vector subbundles
(2.9) |
Our aim is to extend the concepts of Section 2.1 to this setting. After an appropriate choice of indices like in (2.4), one can define the smooth vector bundle
with , and .
2.4.1. Adapted frames
The notion of adapted bases in the setting of filtered vector spaces naturally leads to the notion of adapted frames and coframes.
Definition 2.5.
A frame for is said to be adapted to the filtration (2.9) when, for every , the basis is adapted to the filtration .
If is a frame for adapted to the filtration (2.9), then the associated graded basis on
yields a frame for adapted to the gradation in the following sense.
Definition 2.6.
A frame for is adapted to the gradation, or graded, when is a frame for , is a frame for , and so on.
By duality, we obtain the following decreasing filtration of the cotangent bundle by vector bundles
(2.10) |
by considering the annihilators of for , as follows.
Definition 2.7.
A coframe for is adapted to the filtration when, for every , the basis is adapted to the filtration of .
This definition is equivalent to saying that, at each point , the final covectors annihilate . Furthermore, these final covectors together with the preceding covectors annihilate , and so on. The associated frame is a frame for that is adapted to the gradation in the following sense.
Definition 2.8.
A frame for is said to be adapted to the gradation (or simply graded) when is a frame for , is a frame for , and so on.
One can readily check that if is a frame adapted to a gradation for , then its dual is an adapted coframe for . Conversely, if is a coframe for adapted to a gradation of , then its dual is a graded frame for .
As mentioned in Subsection 2.3.1, frames for can always be constructed locally. Furthermore, through a pivoting process, one may easily construct a local frame for adapted to a given gradation. The inverse function theorem implies that, given a local frame of adapted to the gradation, we can then construct a local frame of adapted to the filtration such that .
2.4.2. Weights of forms
Building on top of Section 2.4.1, we define below the natural notion of forms of weights at least in this context.
Given an increasing filtration (2.9) of , we denote by the set of (based) vectors of weight at most (shortened with ). By duality, we say that the (based) covectors in have weight at least (shortened as ). We extend this vocabulary to the space of smooth sections and , that is vector fields and 1-forms respectively.
With our definition, a 1-form is of weight when for any . We denote the space of 1-forms of weight at least by
The filtration in (2.10) of the cotangent bundle then determines the following strict filtration
Example 2.9.
If is a frame for adapted to the filtration (2.10), then all the 1-forms are of weight . The 1-forms are of weight , and more generally are of weight . By taking , we can state this as
In general, we will not have a global coframe as in Example 2.9, but just local ones.
For , we set
For , we define the space of -forms of weight at least (shortened as ) by
In other words, a -form is of weight at least when it can be written locally as a linear combination of -wedges of 1-form of weights at least with . From the definition, it follows that a -form is of weight when
Example 2.10.
Consider an adapted coframe for . We will often use the shorthand
for the multi-index , always assuming . Then is of weight , where each is of weight .
Definition 2.11.
The family where runs above all multi-indices of length (taken with strictly increasing indices as in example 2.10 above) is a free basis of the -module called the basis associated with the coframe .
By construction, we have the inclusion for any . Moreover, given the increasing filtration (2.4) of , for any degree
In particular, the module over the ring admits the filtration
(2.11) |
The basis introduced in Definition 2.11 is adapted to this filtration.
2.4.3. The bundle and its weights
Given the filtration (2.9), by duality the smooth vector bundle
is graded. We say that an element of is of weight when it is in . We extend the same notion of weight to smooth sections of .
Example 2.12.
Let us consider the same adapted coframe for as in Example 2.9. Then have weight , and for , the sections of have weight . If we denote by the weight of , we have:
(2.12) |
There is a natural identification between the following smooth vector bundles
allowing us to consider directly in our discussion.
It is straightforward to check that is a module over which, by Section 2.1.8, is naturally equipped with a gradation and a notion of weight. In other words,
where the elements of weight are in the linear subbundle
Once we extend this construction to the space of smooth sections of , we get the gradation
(2.13) |
where the -forms of weight are given by
Example 2.13.
We continue with the setting of Example 2.12 and examine the case of -forms, . The sections of are of weight
We may use the shorthand notation where is the multi-index with .
We readily check the following properties:
2.4.4. The operation
Definition 2.15.
A morphism of -modules respects the filtration (2.11), when
The following property follows readily from Section 2.1:
Lemma 2.16.
Let be a morphism of -modules that respects the filtration (2.11). Then the map
is a morphism of -modules that respects the gradation (2.13). If in addition, is a differential operator, then is also a differential operator of the same order or lower.
As an operation, is a module morphism from module morphism to module morphism.
By construction, if and only if increases the weight in the following sense.
Definition 2.17.
We say that a linear map increases the weights of the filtration when
2.5. Tangent bundles equipped with a filtration and a metric
Let be a manifold whose tangent bundle admits a filtration (2.9) by subbundles and is equipped with a metric .
2.5.1. Identification of and
The considerations in Section 2.1.10 show that inherits a metric . The same section also implies that, at each point , there exists an open neighbourhood of and a frame of on that is -orthonormal and adapted to the natural orthogonal gradation
where is the orthogonal complement of in for each .
Let . For any , if is a covector in (i.e. a linear form on vanishing on every , ), we denote by the corresponding covector in (i.e. the corresponding linear form on vanishing on , ). This extends uniquely to a linear bijective map , and then to a -module isomorphism
respecting the Leibniz property.
Lemma 2.18.
Let be a manifold whose tangent bundle admits a filtration by subbundles (2.9) and is equipped with a metric . The map constructed above is an isomorphism of -modules that respects the filtrations, even when restricted to for each . Moreover, is an isomorphism between the -modules and that respects the gradation.
If is a -orthonormal local frame of on an open subset that is adapted to the natural orthogonal gradation , then we have
for any increasing multi-index of length , where is the coframe associated to over .
Note that the map is in fact algebraic, with
using the notation of Lemma 2.18. When restricted to , the matrix representing the map is just the identity matrix of :
Moreover, for each , using the notation and for the corresponding bases of and of , the matrix representing is the identity matrix of :
Intertwining with provides forms with a reverse operation to the operation described in Lemma 2.16:
Lemma 2.19.
We continue with the setting of Lemma 2.18.
-
(1)
If is a morphism of -modules, then
is a morphism of -modules. If in addition is a differential operator, then is also a differential operator of the same order.
-
(2)
If is a morphism of -modules that respects the gradation, then respects the filtration and we have
-
(3)
If respects the filtration, then strictly increases weights, and so
Proof.
These properties are easily checked on and having fixed a local orthonormal adapted coframe . ∎
2.5.2. Scalar products on and of
For each , the scalar products and on and respectively induce scalar products on their duals and , and then scalar products on and on via the formula in (2.15) below and the orthogonal decompositions:
We will denote all these scalar product as , their meaning being clear from the context or from an indexation. We have
(2.15) |
where , and similarly for .
We check readily that the map defined above sends the scalar product of onto the one of :
(2.16) |
Given an orthonormal local frame or equivalently an orthonormal local coframe , we check readily that the bases and of and of respectively are orthonormal. We will use the following notation to denote the elements of the bases in degree 0 and :
and
With these scalar products, we can now define transposes and obtain some of their properties: if and are algebraic maps acting linearly on and respectively, then we define their transpose and as the algebraic maps acting linearly on and by their pointwise transpose, that is, and are the transpose of and for the scalar products and at each point . We check readily that if is an algebraic map acting linearly on respecting the gradation, then so is . Moreover, the map and its transpose are both algebraic map acting linearly on respecting the gradation, and satisfy:
2.5.3. Hodge-star, scalar product on and , and transpose
The Hodge star operator on or is the map defined via
and similarly on . As maps the scalar product of to the one of , (see (2.16)), it also maps the Hodge star operator of to the one of :
(2.17) |
The operator naturally extends to an algebraic operator on and on respectively. Moreover, from the definition of , we have that
and
Moreover, if a local orthonormal adapted coframe is fixed, the Hodge star operators on and with multi-index are given by
where with is the multi-index complementing , and is the sign of the permutation . This completely characterises the Hodge star operators on both and on .
The previous properties imply readily that for any in either or , we have
We can now define the so-called -inner products on the subspaces and of forms with compact support in and respectively via
We have and
(2.18) |
2.5.4. Properties of the transpose for the -inner product
The above definitions of the Hodge operator and of the scalar product show that if is a differential operator on either or , then so are and , as well as the formal transpose defined as for any or and any or .
In the particular case of the de Rham differential on forms, we observe that the transpose of is a differential operator that satisfies
We can also consider the transpose of the algebraic linear map of Subsection 2.2.1. The linear map is then algebraic and satisfies
It is important to realise that the transpose may not respect the filtration in the following sense: for a general differential operator , it is true that the transpose is a differential operator, but respecting the filtration (2.11) will not imply always that does so. This may not be true even for and , and two explicit ‘counter-examples’ of this situation are given in Appendix B. However, the following lemma guarantees that will also respect the filtration when is of the form with a homomorphism of the -module :
Lemma 2.20.
We continue with the setting of Lemma 2.19, where is a morphism of -modules. We consider the map acting on . Let us further assume that respects the gradation.
-
(1)
The -module morphism also respects the gradation, and Consequently, the -module morphism respects the filtration and we have .
-
(2)
If in addition, is a differential operator, then is also a differential operator of the same order as , it respects the gradation, and satisfies . Consequently, the differential operators on smooth forms and its transpose respect the filtration, and
Proof.
The equality follows from (2.17).
Let be a local adapted frame. We observe that is of weight while is of weight where is the weight of the volume form.
Since respects the gradation and the weight of is , has the same weight . Consequently, is a -linear combination of basis elements where the multi-indices have weight . Finally, has weight . We have therefore shown that the morphism maps elements of weight to elements of weight , and so respects the gradation. Lemma 2.19 implies the rest of Part (1).
Part (2) follows from the definition of and its link with covered in Subsection 2.5.3. ∎
2.5.5. Properties of for algebraic maps
Here, we describe some properties of in relation with algebraic maps (for the latter concept see Section 2.3.2).
The following properties are easily checked.
Lemma 2.21.
Let be a smooth manifold whose tangent bundle is filtered by vector subbundles as in (2.9).
-
(1)
Any algebraic morphism of -module acting either on or is determined by its pointwise restriction at each . Conversely, any smooth map on or that is linear at each fibre yields an algebraic morphism of -modules.
-
(2)
If a morphism of -modules acting on is algebraic and respects the filtration, then is a morphism of -module acting on which is also an algebraic map that respects the gradation. Moreover, if is invertible, then is also invertible, and and are algebraic.
The isomorphism together with the concept of algebraic maps allow us to construct inverses in some cases, and this provides a partial converse to (2) in Lemma 2.21 as follows.
Proposition 2.22.
Let be a smooth manifold whose tangent bundle is filtered by vector subbundles as in (2.9). Assume that is equipped with a metric . We consider the corresponding map defined in Section 2.5.1.
Let be a morphism of -modules that respects the filtration and such that is algebraic. Assume also that is fiberwise invertible in the sense that its restriction at each is a linear invertible map on . Then the map is invertible and the map respects the filtration. Moreover, , and if is in addition a differential operator acting on , then so is its inverse.
Proof.
Corollary 2.23.
We continue with the setting of Proposition 2.22 and its proof. The following formulae holds for :
where , and while .
3. Filtered manifolds and osculating objects
In this section, we discuss the setting of filtered manifolds and some complexes naturally associated with them. In particular, we will define the osculating differential and study its relation to the de Rham complex . We will also use the algebraic part of . The main result of this section is in Proposition 3.3 stating that .
3.1. Setting and definitions
A filtered manifold is a smooth manifold equipped with a filtration of the tangent bundle by vector subbundles
satisfying
(3.1) |
with the convention that when .
Example 3.1.
3.1.1. Bundle of osculating Lie groups and algebras
For each , the quotient is naturally equipped with a Lie bracket , since
When is equipped with this Lie bracket, we denote the resulting Lie algebra as
It is naturally graded by
and is therefore nilpotent. We denote by the corresponding connected simply connected nilpotent Lie group (sometimes called the nilpotentisation or the tangent cone [Mit85] of at ). The unions
are naturally equipped with a smooth bundle structure that are called [RS76, vEY17] the bundles of osculating Lie groups and Lie algebras over .
We observe that the notions of weights defined in Section 2.4.3 and for the graded Lie algebra coincide. We therefore obtain an analogous decomposition of in terms of weights and degrees:
(3.2) |
We introduce the following vocabulary.
Definition 3.2.
The elements of are called osculating forms. Moreover, for any ,
are called osculating -forms (or osculating forms of degree ) and osculating -forms of weights respectively.
We say that a map respects the weights of osculating forms when , and it increases the weights when .
3.1.2. The osculating Chevalley-Eilenberg differential
For each , denotes the Chevalley-Eilenberg differential on the Lie group viewed as the map
defined via for , for
(3.3) |
and more generally for , for any and ,
(3.4) |
It is well known [FT23] that is a linear map such that
and that it satisfies the Leibniz property on , i.e.
Moreover, also preserves the weights of :
By construction, the algebraic map defined by
is smooth. We call the osculating complex or osculating Chevalley-Eilenberg differential.
3.1.3. The osculating box operator
Here we assume that the vector bundle is equipped with a metric : each is equipped with a scalar product with a smooth dependence in . This scalar product allows us to consider, at each point , the transpose of the osculating differential , and then to define
We call this operator the osculating box. As is algebraic, so are and .
Let us now assume that the decomposition (3.2) is orthogonal for the scalar product for each . This arises naturally when the manifold is filtered and Riemannian (see Section 2.5). With this assumption, we readily check that preserves the weights, i.e. , and respects the degree and weights of osculating forms:
Moreover, given the scalar product , the linear map
is symmetric. We denote by the spectral orthogonal projection onto its kernel
Note that the image of is
By [FT23, Lemma 2.1], the complex is acyclic, i.e. , and the resulting cohomology is trivial:
For small enough, is invertible, and by the Cauchy residue formula we have
where the contour integration is over a circle about 0 of radius small enough. This defines a smooth algebraic map acting on that respects degrees and weights.
3.1.4. The partial inverse
Proceeding as in [FT23], at each , a partial inverse of can be defined as
where we write to denote the orthogonal projection onto any closed subspace of .
Note that in our construction in Section 4, we will not use . However, it is used in Rumin’s construction, see Section 4.5.
where we show that the two constructions coincide.
3.2. The maps and on a filtered manifold
Let us show that and are well-defined, crucial operators, and that both and coincide with the osculating complex.
Proposition 3.3.
The maps and respect the filtration with
for any degree and weight . Moreover, the maps and acting on coincide with :
Proof.
From the definitions of and , both differentials as maps respect the corresponding filtrations. As they obey the Leibniz rule, both and respect the filtration on . Moreover, and satisfy the Leibniz property on .
Let and for each take with . Since , we have that , and so by (2.8).
This shows , and implies . We are left to prove that .
4. Construction of subcomplexes on Riemannian filtered manifolds
Here, we present the general scheme to construct subcomplexes on a filtered manifold whose tangent bundle is equipped with a metric . It relies on the notions of a base differential and codifferential that we introduce in Section 4.1 (this terminology is borrowed from [GT53]). In Section 4.2, we construct certain operators which allow us to define the subcomplexes and , with the latter computing the same cohomology as de Rham’s (see Section 4.3). In Section 4.4, we also present an analogous construction, obtaining the operators and then the subcomplexes and , with the latter computing the same cohomology as the complex . We present an alternative construction for and in Section 4.5 that follows Rumin’s ideas.
4.1. Base differential and codifferential and
Definition 4.1.
Let and be two differential algebraic complexes on a filtered manifold equipped with a metric . We say that is a pair of base differential and codifferential when the following properties are satisfied:
-
•
increases the degree by one while decreases the degree by one, and both respect the filtration of :
-
•
and are disjoint [Kos61], that is, for any
-
•
and .
Above, the transpose is defined using a metric on the osculating bundle induced by .
Given a pair of base differential and codifferential on a filtered manifold , we define the associated base box via
For any , we define the operator
for small enough.
Proposition 4.2.
-
(1)
The operators and are smooth and algebraic on . They respect its filtration and keep the degrees of the forms constant:
They also satisfy
The algebraic subbundles
admit subfiltrations and we have:
-
(2)
We have
Moreover, is the projection onto along .
Proof.
Part (1) is satisfied by construction. Disjointedness implies readily that . Since , Part (1) implies that . Hence .
Disjointedness also implies . Clearly, we also have . From the Cauchy residue formula, proceeding as in the proof of [FT23, Proposition 4.1], we check that . In other words, is a projection of . We also check using the Cauchy residue that on while so on . We obtain Part (2) from . ∎
We call the base kernel projection.
4.2. The operators , and
We define the differential operator acting on :
Remark 4.3.
The de Rham differential will be replaced with its algebraic part in Section 4.4. The new construction will closely follow the steps below, as the only property that it relies on is that .
Proposition 4.4.
The operator commutes with and . It is a differential operator that preserves the degrees and the weights of forms:
We have
Proof.
Since and are complexes, they commute with . Moreover, and increase the degrees of forms by one, while decreases the degrees by one. Hence, and preserve the degrees of forms. As , and respect the filtration, so do and . We conclude with
∎
Applying Proposition 4.4 together with Proposition 2.22 to , the map
is well-defined for small enough.
Proposition 4.5.
-
(1)
The map is a differential operator acting on that respects the filtration and the degree of the forms with . It commutes with , and .
-
(2)
We have while acts on where it is invertible. Moreover, is the projection onto
along
These modules satisfy and .
Proof.
Part (1) follows from Proposition 4.4 and applying . From the Cauchy residue formula, proceeding as in the proof of [FT23, Proposition 4.1], we check that . In other words, is a projection.
For Part (2), by the Cauchy formula and the properties of . Therefore , but since is a projection commuting with . Therefore .
Naturally, acts on which is a submodule of that inherits a filtration. Moreover, we check
since is symmetric and the orthogonal projection onto its kernel is , and
Adapting the proof of Proposition 2.22 to a submodule, we obtain that is invertible on . This implies that and . We then readily check that is the projection onto along .
As , we have . Hence, since commutes with , we have
and so recursively, for any ,
For , since . As is a projection, we have obtained . Moreover, since commutes with , we also have , which implies the inclusion .
Since and , we compute easily
As is the projection onto along , we have
These last inclusions then imply that , but also . ∎
Note that in our construction below, we do not need the characterisations of and as and . We will only need it when proving that this construction coincides with Rumin’s.
We follow the same strategy as in the case of homogeneous groups [FT23, Section 4.1]:
Proposition 4.6.
The differential operator acting on and defined as
preserves the degrees of the forms and respects the filtration with . It is invertible and its inverse is a differential operator acting on . We have
and this implies
Consequently,
4.3. The complexes and
By Propositions 4.5 and 4.6, the de Rham differential commutes with , so commutes with . Hence, we can decompose the differential operator as
where and are the differential operators
Since and , it follows that . Hence, the chain complex decomposes into the direct sum of the two chain complexes: acting on , and acting on .
Proposition 4.7.
The maps and are conjugated on :
where is the invertible operator defined as
Hence, the complexes and have the same cohomology.
Proof.
Note that and that is invertible on , so is well-defined on .
Since , and , we have
and on
Applying the proof of Proposition 2.22 to the submodule , is invertible on .
Since and , it is a straightfoward to check that holds on . ∎
Proposition 4.8.
The differential operator is a chain map between and , that is . This chain map is homotopically invertible, with homotopic inverse given by since we have
where is the differential operator acting on and defined as
with as in Proposition 4.7.
Proof.
By construction, we have , with and , so .
It remains to prove the properties regarding . We first point out that makes sense because and acts on in an invertible way. The definitions of and together with Proposition 4.7 then yield:
The conclusion follows. ∎
Corollary 4.9.
The cohomology of is linearly isomorphic to the de Rham cohomology of the manifold .
4.4. The complexes and
In this section, we consider the complexes obtained by considering instead of in the construction above (see Remark 4.3). The proofs are omitted, as the arguments are essentially the same as above. We start by defining the algebraic operator acting on :
It commutes with and and preserves the degrees and the weights of forms:
We have
Applying Proposition 2.22 to locally at , together with Proposition 4.4, the map
is well-defined for small enough. is then an algebraic operator acting on that respects the filtration and the degree of forms, and . It commutes with , and . We have , while acts on where it is invertible. Moreover, is the projection onto
along
These modules satisfy and .
The algebraic operator acting on and defined as
preserves the degree of forms, respects the filtration, and . It is invertible and its inverse is an algebraic operator acting on . We have
Consequently,
We define the two differentials on , and on via
The maps and are conjugated on :
where is the invertible operator defined as
The complexes and have the same cohomology.
The differential operator is a chain map between and , that is . This chain map is homotopically invertible, with homotopic inverse given by since we have
where is the algebraic operator acting on and defined as
Consequently, the cohomology of is linearly isomorphic to the cohomology of of the manifold .
4.5. Equivalent constructions for the subcomplexes and
Here, we present an interpretation of Rumin’s construction [Rum99, Rum05] of the complex that bears his name. In our presentation, we highlight the difference between objects living on the manifold and their osculating counterparts.
We first define the map
(4.1) |
which is defined via the partial inverse of the osculating differential (see Section 3.1.4). From the properties of , we see that , it respects the filtration and decreases the degree by one:
We have
so
Moreover,
(4.2) |
and we have
Lemma 4.10.
Let us consider the differential operators acting on defined by:
-
(1)
The maps and are nilpotent. Consequently, and are invertible, and and are well-defined differential operators acting on . We have:
-
(2)
The differential operator defined as
is the projection onto
along
Proof.
Corollary 4.11.
Following Rumin’s notation, the two projections onto and along and are denoted respectively as
We then obtain Rumin’s construction and its equivalence with the one presented in this paper.
Theorem 4.12.
-
(1)
(M. Rumin) The de Rham complex splits into two subcomplexes and . Moreover, the differential operator defined as
satisfies
Moreover, we have
and so , and the complex is conjugated to via .
-
(2)
The maps and coincide, i.e. .
Proof.
This equivalent construction also works if we replace with , yielding a complex which coincides with .
4.6. Case of regular subRiemannian manifolds
Regular subRiemannian manifolds are filtered manifolds (see below). If we equip such a manifold with a Riemannian metric, then our construction applies and we obtain two complexes and . Here, we show that if the Riemannian metric is compatible with the subRiemannian structure as explained in Section 4.6.2 below, then coincides at least locally with what is now customarily referred to as the Rumin complex. We would like to stress that, when introducing what Rumin calls the “Carnot complex of an -regular -structure” in [Rum05], he only briefly mentions the potential impact that the choice of a Riemannian metric can have on the resulting subcomplex of because he considers only local or osculating objects. In particular, he does not address any compatibility conditions between the Riemannian and subRiemannian structures, although it seems to be an important ingredient in the construction. We hope that our approach will lead to a better understanding of the constructed subcomplexes.
4.6.1. The osculating metric of a regular subRiemannian manifold
Recall that a subRiemannian manifold is a smooth manifold equipped with a bracket generating distribution and with a metric on . Let be the set of smooth sections of , and , for As is bracket generating, there exists such that , and we denote by the smallest such integer . If, for every , there exists a subbundle for which is the set of smooth sections on , then is said to have an regular subRiemannian structure. Clearly, the s provide the structure of filtered manifold.
Consider a regular subRiemannian manifold as above. In this case, the osculating Lie algebras are stratified
and the subspaces are given by imposing , and
The metric on naturally induces a scalar product on each osculating Lie algebra fibre [Mon02, p. 188]. Let us briefly recall its construction.
At every point , the map
is -linear and surjective. Hence, the metric on induces a scalar product on and then on (which is the image of the above map, and therefore inherits the scalar product from the orthogonal complement of the kernel). Constructing this for , we obtain a scalar product on . By construction, the dependence in is smooth, and , , defines an element . Therefore, is a metric on the osculating bundle of Lie algebras. We call the osculating metric induced by .
4.6.2. Compatible Riemannian metrics
We now assume that, in addition to being a regular subRiemannian manifold, is also equipped with a Riemannian metric, that is, with a metric on its tangent bundle . As already seen in Sections 2.1.10 and 2.5.1, will also induce a metric on . The two metrics and will be different in general.
Definition 4.13.
A Riemannian metric is compatible with the subRiemannian structure of a regular subRiemannian manifold (or compatible for short) when the two metrics and on the osculating Lie algebra bundle coincide.
By construction, coincides on with . Hence, a necessary condition for to be compatible is for to coincide with on . However, it is not sufficient generally. For example, let us consider the 6-dimensional free nilpotent Lie group of rank 3 and step 2, denoted by in [LDT22]. Keeping the notation of [LDT22], if we take and as global left-invariant orthogonal coframes of , then the corresponding subcomplexes and do not coincide [FT24]. For clarity, we stress that and are isomorphic subspaces of , and this is necessarily true in general once we assume is a regular subRiemannian manifold. However, they need not be the same subspace, and in this explicit example they are not (it is sufficient to compare and ). In other words, just extending the metric from to an arbitrary Riemannian metric on is not sufficient to ensure compatibility. However, compatible metrics can always be constructed.
Lemma 4.14.
Let be a (second countable) regular subRiemannian manifold. We can always construct a compatible Riemannian metric.
Proof.
Consider a sequence of non-negative functions , , such that the sum on is locally finite and constantly equal to 1. We may assume that the support of each is small enough so that it is included in an open subset where a frame of exists. After a Graham-Schmidt procedure, we may assume that each is such that the corresponding frame of the vector bundle is -orthonormal. Denoting by the corresponding metric on such that is orthonormal, one can easily check that is a compatible Riemannian metric. ∎
4.6.3. The base differential and codifferential associated with a compatible Riemannian metric
Let be a regular subRiemannian manifold equipped with a Riemannian metric . The compatibility implies that the scalar product on (defined as in Section 2.5.1 using ) coincides with the scalar product on induced by the osculating metric . Hence, the map identifying forms and osculating forms and the objects built upon it (for instance the base differential and co-differential and ) are defined globally as objects leaving on the manifolds. These removes ambiguities and improves on earlier constructions [Rum90, Rum99, Rum05].
Appendix A Case of a three-dimensional contact manifold (locally)
On a three-dimensional contact manifold, the existence of Darboux coordinates implies that the manifold is filtered and can be locally described as follows. The manifold is equipped with a frame satisfying
with , , with nowhere vanishing.
The associated free basis of is given by
The numbers in parenthesis refer to the weight of the form.
Description of
For , . For , let us describe in matrix form with respect to the canonical basis described above.
Description of
For , we have , while for , we have , . Hence,
Description of
The osculating group is isomorphic to the Heisenberg group at any point :
The associated free basis of is
The number in parenthesis refers to the weight of the osculating form.
For , . For , let us describe in matrix form with respect to the canonical basis described above.
Description of and
For , while for , the description of in matrix form with respect to the canonical basis of given above coincides with . For , . For , the matrix description of coincides with the transpose of . Consequently,
Description of ,
For , , while and are represented by the following diagonal matrices:
Consequently, is the identity for , while for ,
Consequently, and while
Description of , ,
For , we have , while for
Consequently, is the identity on for . Since for , the formulae for Cauchy residues and the von Neumann series simplify into
where denotes the projection onto the -eigenspace of . Hence, for ,
We now describe
For , is the identity on while for ,
where denotes the 3-by-3 identity matrix.
Computation of
We now describe
for .
The zeros in the matrices above illustrate acting on while being trivial on .
Description of , ,
For , we have , while for
Consequently, is the identity on for . Since for , the formulae for Cauchy residues and the von Neumann series simplify into
where denotes the projection onto the -eigenspace of . Hence, for ,
We now describe
For , is the identity on while for ,
where denotes the 3-by-3 identity matrix.
Computation of
We now describe
for .
Appendix B Examples
In this appendix, we give two examples of a differential operator that respects the filtration (2.11) whose transpose does not necessarily respect the filtration.
Example B.1.
We consider the 3-dimensional Heisenberg group , and the canonical basis of its Lie algebra . The only non-trivial bracket is . Identifying with the space of left-invariant vector fields on , we obtain a filtration of the tangent bundle of in subbundles:
The dual of the left-invariant frame of yields the global coframe (as well as the natural linear isomorphism between and ).
Let us now consider the de Rham differential acting on the 1-form with :
If we now consider its formal transpose acting on the 2-form with :
and so , but not .
Applying similar computations, one can easily check that the transpose of the operator instead respects the filtration.
A situation similar to Example B.1 also holds for the algebraic part of the de Rham differential , when considering a nilpotent Lie group with a filtration on its Lie algebra that is not coming from a homogeneous structure as in the following example:
Example B.2.
Let us consider the 4-dimensional Engel group , that is, the connected simply connected nilpotent Lie group with Lie algebra and for . We identify with the space of left-invariant vector fields, and we consider the following left-invariant filtration of
The coframe of the canonical frame is denoted by Then, for an arbitrary form in , with , we have
However, given a form with , we obtain .
References
- [BFP22] Annalisa Baldi, Bruno Franchi, and Pierre Pansu. Poincaré and Sobolev inequalities for differential forms in Heisenberg groups and contact manifolds. J. Inst. Math. Jussieu, 21(3):869–920, 2022.
- [BvE14] Paul F. Baum and Erik van Erp. -homology and index theory on contact manifolds. Acta Math., 213(1):1–48, 2014.
- [Can21] Giovanni Canarecci. Sub-Riemannian currents and slicing of currents in the Heisenberg group . J. Geom. Anal., 31(5):5166–5200, 2021.
- [Cas11] Jeffrey S. Case. The bigraded rumin complex via differential forms, arXiv:2108.13911.
- [CCY16] Jeffrey S. Case, Sagun Chanillo, and Paul Yang. The CR Paneitz operator and the stability of CR pluriharmonic functions. Adv. Math., 287:109–122, 2016.
- [CD01] David M. J. Calderbank and Tammo Diemer. Differential invariants and curved Bernstein-Gelfand-Gelfand sequences. J. Reine Angew. Math., 537:67–103, 2001.
- [ČH23] Andreas Čap and Kaibo Hu. Bounded Poincaré operators for twisted and BGG complexes. J. Math. Pures Appl. (9), 179:253–276, 2023.
- [ČSS01] Andreas Čap, Jan Slovák, and Vladimír Souček. Bernstein-Gelfand-Gelfand sequences. Ann. of Math. (2), 154(1):97–113, 2001.
- [DH22] Shantanu Dave and Stefan Haller. Graded hypoellipticity of BGG sequences. Ann. Global Anal. Geom., 62(4):721–789, 2022.
- [FT12] Bruno Franchi and Maria Carla Tesi. Wave and Maxwell’s equations in Carnot groups. Commun. Contemp. Math., 14(5):1250032, 62, 2012.
- [FT23] Véronique Fischer and Francesca Tripaldi. An alternative construction of the Rumin complex on homogeneous nilpotent Lie groups. Adv. Math., 429:Paper No. 109192, 39, 2023.
- [FT24] Véronique Fischer and Francesca Tripaldi. Extracting subcomplexes on nilpotent groups without relying on homogeneous structures, in preparation, 2024.
- [GAJV19] Maria Paula Gomez Aparicio, Pierre Julg, and Alain Valette. The Baum-Connes conjecture: an extended survey. In Advances in noncommutative geometry—on the occasion of Alain Connes’ 70th birthday, pages 127–244. Springer, Cham, [2019] ©2019.
- [GT53] Erlend Grong and Francesca Tripaldi. Filtered complexes and cohomologically equivalent subcomplexes, arXiv:2308.11353.
- [Hal22] Stefan Haller. Analytic torsion of generic rank two distributions in dimension five. J. Geom. Anal., 32(10):Paper No. 248, 66, 2022.
- [JK95] Pierre Julg and Gennadi Kasparov. Operator -theory for the group . J. Reine Angew. Math., 463:99–152, 1995.
- [JP12] Antoine Julia and Pierre Pansu. Flat compactness of normal currents, and charges in carnot groups, arXiv:2303.02012.
- [Jul95] Pierre Julg. Complexe de Rumin, suite spectrale de Forman et cohomologie des espaces symétriques de rang . C. R. Acad. Sci. Paris Sér. I Math., 320(4), 1995.
- [Jul19] Pierre Julg. How to prove the Baum-Connes conjecture for the groups ? J. Geom. Phys., 141:105–119, 2019.
- [Kit20] Akira Kitaoka. Analytic torsions associated with the Rumin complex on contact spheres. Internat. J. Math., 31(13):2050112, 16, 2020.
- [KMX28] Bruce Kleiner, Stefan Muller, and Xiangdong Xie. Sobolev mappings and the rumin complex, arXiv:2101.04528.
- [Kos61] Bertram Kostant. Lie algebra cohomology and the generalized Borel-Weil theorem. Ann. of Math. (2), 74:329–387, 1961.
- [LDT22] Enrico Le Donne and Francesca Tripaldi. A cornucopia of Carnot groups in low dimensions. Anal. Geom. Metr. Spaces, 10(1):155–289, 2022.
- [LT23] Antonio Lerario and Francesca Tripaldi. Multicomplexes on Carnot groups and their associated spectral sequence. J. Geom. Anal., 33(7):Paper No. 199, 22, 2023.
- [Mit85] John Mitchell. On Carnot-Carathéodory metrics. J. Differential Geom., 21(1):35–45, 1985.
- [Mon02] Richard Montgomery. A tour of subriemannian geometries, their geodesics and applications, volume 91 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2002.
- [PR18] Pierre Pansu and Michel Rumin. On the cohomology of Carnot groups. Ann. H. Lebesgue, 1:267–295, 2018.
- [PT19] Pierre Pansu and Francesca Tripaldi. Averages and the cohomology of Heisenberg groups. Ann. Math. Blaise Pascal, 26(1):81–100, 2019.
- [RS76] Linda Preiss Rothschild and E. M. Stein. Hypoelliptic differential operators and nilpotent groups. Acta Math., 137(3-4):247–320, 1976.
- [RS12] Michel Rumin and Neil Seshadri. Analytic torsions on contact manifolds. Ann. Inst. Fourier (Grenoble), 62(2):727–782, 2012.
- [Rum90] Michel Rumin. Un complexe de formes différentielles sur les variétés de contact. C. R. Acad. Sci. Paris Sér. I Math., 310(6):401–404, 1990.
- [Rum99] Michel Rumin. Differential geometry on C-C spaces and application to the Novikov-Shubin numbers of nilpotent Lie groups. C. R. Acad. Sci. Paris Sér. I Math., 329(11):985–990, 1999.
- [Rum05] Michel Rumin. An introduction to spectral and differential geometry in Carnot-Carathéodory spaces. Rend. Circ. Mat. Palermo (2) Suppl., (75):139–196, 2005.
- [Tri54] Francesca Tripaldi. The Rumin complex on nilpotent lie groups, arXiv:2009.10154.
- [vE10] Erik van Erp. The Atiyah-Singer index formula for subelliptic operators on contact manifolds. Parts I and II. Ann. of Math. (2), 171(3):1683–1706 and 1647–1681, 2010.
- [vEY17] Erik van Erp and Robert Yuncken. On the tangent groupoid of a filtered manifold. Bull. Lond. Math. Soc., 49(6):1000–1012, 2017.
- [Vit22] Davide Vittone. Lipschitz graphs and currents in Heisenberg groups. Forum Math. Sigma, 10:Paper No. e6, 104, 2022.