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

Subcomplexes on filtered Riemannian manifolds

Véronique Fischer and Francesca Tripaldi University of Bath, Department of Mathematical Sciences, Bath, BA2 7AY, UK v.c.m.fischer@bath.ac.uk Centro De Giorgi, SNS, Piazza dei Cavalieri 3, 56 126 Pisa, Italy. francesca.tripaldi@sns.it
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 groups
2010 Mathematics Subject Classification:
58A10, 58J10, 58H99, 53C17, 43A80, 22E25

1. 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 (Ω(M),d)(\Omega^{\bullet}(M),d) to extract a subcomplex (E0,D)(E_{0}^{\bullet},D) computing the de Rham cohomology of the underlying manifold MM, while on the other, we extract (E~0,D~)(\tilde{E}_{0}^{\bullet},\tilde{D}) from the “algebraic” complex (Ω(M),d~)(\Omega^{\bullet}(M),\tilde{d}). Here d~\tilde{d} denotes the algebraic part of the de Rham differential dd, and so the complex (Ω(M),d~)(\Omega^{\bullet}(M),\tilde{d}) 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 (E0,D)(E_{0}^{\bullet},D) and (E~0,D~)(\tilde{E}_{0}^{\bullet},\tilde{D}). The various claims will be proved in Section 4.

We consider a filtered manifold MM and denote by 𝔤xM\mathfrak{g}_{x}M its osculating nilpotent Lie algebra (also called nilpotentisation) above each point xMx\in M (see Section 3.1). We then define the osculating Chevalley-Eilenberg differential d𝔤Md_{\mathfrak{g}M} on the space GΩ(M)G\Omega^{\bullet}(M) of osculating forms over MM, defined above each point xMx\in M as the Chevalley-Eilenberg differential of the osculating group GxMG_{x}M (see Section 3.1.2).

We now further assume that MM is also equipped with a metric on TMTM. This allows us to construct the natural isomorphism Φ:Ω(M)GΩ(M)\Phi:\Omega^{\bullet}(M)\to G\Omega^{\bullet}(M), providing an identification between forms and osculating forms (see Section 2.5). We are then able to define the map d0:=Φ1d𝔤MΦd_{0}:=\Phi^{-1}\circ d_{\mathfrak{g}M}\circ\Phi acting on Ω(M)\Omega^{\bullet}(M), which we call the base differential.

1.2.1. Constructing DD using d0td_{0}^{t}

The transpose d0td_{0}^{t} of d0d_{0} makes sense unambiguously and globally thanks to the metric on TMTM. We consider the projection Π0\Pi_{0} onto E0:=kerd0kerd0tE_{0}:=\ker d_{0}\cap\ker d_{0}^{t} along F0:=Imd0+Imd0t,F_{0}:={\rm Im}\,d_{0}+{\rm Im}\,d_{0}^{t}, and the projection PP onto kerd0tkerd0td\ker d_{0}^{t}\cap\ker d_{0}^{t}d along Imd0t+Imdd0t{\rm Im}\,d_{0}^{t}+{\rm Im}\,dd_{0}^{t}. The proper definitions are in fact given by Π0\Pi_{0} being the orthogonal projection onto the kernel of the algebraic operator 0:=d0d0t+d0td0\Box_{0}:=d_{0}d_{0}^{t}+d_{0}^{t}d_{0}, and PP being the generalised kernel projection of the differential operator :=dd0t+d0td\Box:=dd_{0}^{t}+d_{0}^{t}d. Once we define the invertible differential operator L:=PΠ0+(IP)(IΠ0)L:=P\Pi_{0}+(\text{\rm I}-P)(\text{\rm I}-\Pi_{0}), we have the decomposition

L1dL=:D+C,L^{-1}dL=:D+C,

yielding two complexes (E0,D)(E_{0}^{\bullet},D) and (F0,C)(F_{0}^{\bullet},C) whose cohomologies are described as follows:

\bullet (F0,C)(F_{0}^{\bullet},C) has trivial cohomology, since CC is conjugated to d0d_{0} on F0F_{0}, and (F0,d0)(F_{0}^{\bullet},d_{0}) is acyclic.

\bullet The cohomology of (E0,D)(E_{0}^{\bullet},D) is linearly isomorphic to the de Rham cohomology of MM via the homotopically invertible chain map Π0L1\Pi_{0}L^{-1}.

1.2.2. Constructing DD using d01d_{0}^{-1}

Relying again on the metric on TMTM, we can define the orthogonal projection prImd𝔤M{\rm pr}_{{\rm Im}\,d_{\mathfrak{g}M}} onto Imd𝔤M{\rm Im}\,d_{\mathfrak{g}M}. We then consider the partial inverse d𝔤M1:=(d𝔤xM)1prImd𝔤xMd_{\mathfrak{g}M}^{-1}:=(d_{\mathfrak{g}_{x}M})^{-1}{\rm pr}_{{\rm Im}\,d_{\mathfrak{g}_{x}M}} of d𝔤xMd_{\mathfrak{g}_{x}M} and thus the corresponding partial inverse d01:=Φ1d𝔤M1Φd_{0}^{-1}:=\Phi^{-1}\circ d_{\mathfrak{g}M}^{-1}\circ\Phi of d0d_{0}. Then the projection PP above may be constructed explicitly in terms of d01d_{0}^{-1} as:

P=(Ib)1d01d+d(Ib)1d01,whereb:=d01d0d01d.P=(\text{\rm I}-b)^{-1}d_{0}^{-1}d+d(\text{\rm I}-b)^{-1}d_{0}^{-1},\quad\mbox{where}\quad b:=d_{0}^{-1}d_{0}-d_{0}^{-1}d.

The complex DD is then obtained as D=Π0d(IP)Π0D=\Pi_{0}d(\text{\rm I}-P)\Pi_{0}.

1.2.3. Constructions of D~\tilde{D}

We can modify the two constructions above by applying the same steps, but substituting the initial differential operator dd with the algebraic part d~\tilde{d} of the de Rham differential instead (see Section 2.2.1). This yields two complexes (E~0,D~)(\tilde{E}_{0}^{\bullet},\tilde{D}) and (F~0,C~)(\tilde{F}_{0}^{\bullet},\tilde{C}), with (F~0,C~)(\tilde{F}_{0}^{\bullet},\tilde{C}) acyclic and (E~0,D~)(\tilde{E}_{0}^{\bullet},\tilde{D}) linearly isomorphic to (Ω(M),d~)(\Omega^{\bullet}(M),\tilde{d}).

1.3. Novelty and future works

As already mentioned above, constructions related to the ones presented in this paper for the complex DD 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 MM; they yield complexes globally defined and acting on Ω(M)\Omega^{\bullet}(M) - not its osculating counterpart GΩ(M)G\Omega^{\bullet}(M). We obtain two equivalent constructions for the complex (E0,D)(E_{0}^{\bullet},D), 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 (E~0,D~)(\tilde{E}_{0}^{\bullet},\tilde{D}). In Appendix A, we give an explicit construction for the operators DD and D~\tilde{D} 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 (E0,D)(E_{0}^{\bullet},D) coincides with what is now customarily referred to as the Rumin complex on regular CCCC-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 (E~0,D~)(\tilde{E}_{0}^{\bullet},\tilde{D}) 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 (E0,D)(E_{0}^{\bullet},D), 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:

  • large scale geometry (especially the open conjecture classifying nilpotent Lie groups by quasi-isometries [PR18, PT19, BFP22]),

  • analytic torsion [RS12, Kit20, Hal22] and currents [Can21, Vit22, JP12] for subRiemannian geometry, and

  • KK and index theories for subelliptic operators and CC^{*}-algebras associated with subRiemannian structures [vE10, BvE14].

We believe that the techniques behind the subcomplexes (E0,D)(E_{0}^{\bullet},D) and (E~0,D~)(\tilde{E}_{0}^{\bullet},\tilde{D}) 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 d𝔤Md_{\mathfrak{g}M} 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 WW is a vector space WW with a filtration by vector subspaces:

(2.1) {0}=W0W1Ws=W.\{0\}=W_{0}\subseteq W_{1}\subseteq\ldots\subseteq W_{s}=W.

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 GG is a vector space GG equipped with a gradation by vector subspaces, i.e. a direct sum decomposition by vector subspaces GjGG_{j}\subset G:

(2.2) G=j=1sGj.\displaystyle G=\oplus_{j=1}^{s}G_{j}\,.

If GG is graded, there is a natural filtration associated with its gradation (2.2), given by {0}=W0W1Ws=G\{0\}=W_{0}\subseteq W_{1}\subseteq\cdots\subseteq W_{s}=G, where Wj:=kjGkW_{j}:=\oplus_{k\leq j}G_{k}, j=1,,sj=1,\ldots,s.

2.1.2. The vector space gr(W){\rm gr}(W)

Given a filtered vector space WW with the filtration (2.1), its associated graded space is the vector space

(2.3) gr(W):=j=1sWj/Wj1.{\rm gr}(W):=\oplus_{j=1}^{s}W_{j}/W_{j-1}\,.

Many of the summands above may be trivial. To avoid dealing with zero quotient spaces, it is enough to consider an appropriate choice of indices w0,w1,,ws0w_{0},w_{1},\ldots,w_{s_{0}}\in\mathbb{N} for which the summands in (2.3) are not trivial, or equivalently for which the inclusions in (2.1) are strict:

(2.4) {0}=Ww0Ww1Wws0=W,andgr(W)=j=1s0Wwj/Wwj1.\displaystyle\{0\}=W_{w_{0}}\subsetneq W_{w_{1}}\subsetneq\ldots\subsetneq W_{w_{s_{0}}}=W\ ,\ \mbox{and}\ {\rm gr}\,(W)=\oplus_{j=1}^{s_{0}}W_{w_{j}}/W_{w_{j-1}}.

2.1.3. The linear map gr(T){\rm gr}(T)

Consider a linear map T:WWT:W\to W^{\prime} between two vector spaces, each equipped with a filtration {0}=W0W1Ws=W\{0\}=W_{0}\subseteq W_{1}\subseteq\ldots\subseteq W_{s}=W and {0}=W0W1Ws=W\{0\}=W^{\prime}_{0}\subseteq W^{\prime}_{1}\subseteq\ldots\subseteq W^{\prime}_{s^{\prime}}=W^{\prime}. We may assume s=ss=s^{\prime}.

The map TT is said to respect the filtrations when T(Wj)WjT(W_{j})\subseteq W^{\prime}_{j} for every j=0,,sj=0,\ldots,s. In this case, one can define the linear map gr(T):gr(W)gr(W){\rm gr}(T):{\rm gr}\,(W)\to{\rm gr}\,(W^{\prime}) as

gr(T)(vmodWj):=(Tv)modWj,vWj+1,j=0,,s1.{\rm gr}(T)\,(v\ {\rm mod}\ W_{j}):=(Tv)\ {\rm mod}\ W^{\prime}_{j}\ \ ,\ v\in W_{j+1}\ ,\ j=0,\ldots,s-1\,.

2.1.4. The basis e\langle e\rangle

If WW is finite dimensional, then gr(W){\rm gr}(W) has the same dimension as WW. Given n:=dimW=dimgr(W)n:=\dim W=\dim{\rm gr}(W) and nj=dimWwj/Wwj1n_{j}=\dim W_{w_{j}}/W_{w_{j-1}}, we say that a basis e=(e1,,en)e=(e_{1},\ldots,e_{n}) of WW is adapted to the filtration (2.4), when (e1,,en1)(e_{1},\ldots,e_{n_{1}}) is a basis of Ww1W_{w_{1}}, (e1,,en2+n1)(e_{1},\ldots,e_{n_{2}+n_{1}}) is a basis of Ww2W_{w_{2}}, and so on. For such a basis, we have that, for each j=0,,s01j=0,\ldots,s_{0}-1,

ei=eimodWwj,i=dj+1,,dj+1, with dj=n0+n1++nj,\langle e_{i}\rangle=e_{i}\ {\rm mod}\ W_{w_{j}},\quad i=d_{j}+1,\ldots,d_{j+1}\ ,\text{ with }d_{j}=n_{0}+n_{1}+\cdots+n_{j}\,,

gives a basis e:=(e1,,en)\langle e\rangle:=(\langle e_{1}\rangle,\ldots,\langle e_{n}\rangle) of gr(W){\rm gr}(W).

In particular, we get that for each j=0,,s01j=0,\ldots,s_{0}-1, the njn_{j}-tuple (edj+1,,edj+1)(\langle e_{d_{j}+1}\rangle,\ldots,\langle e_{d_{j+1}}\rangle) is a basis of Wwj+1/WwjW_{w_{j+1}}/W_{w_{j}}. In this sense, the basis e\langle e\rangle is graded and is said to be the graded basis associated with ee.

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 φ:𝒱1𝒱2\varphi:\mathcal{V}_{1}\to\mathcal{V}_{2} between two finite dimensional vector spaces, once we fix β1\beta_{1} and β2\beta_{2} two bases of 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2} respectively, we denote by

Matβ1β2φ{\rm Mat}_{\beta_{1}}^{\beta_{2}}\,\varphi

the matrix representing φ\varphi in the bases β1\beta_{1} and β2\beta_{2}.

Let us take ee and ee^{\prime} two bases adapted to the filtrations (2.4) of two finite dimensional filtered vector spaces WW and WW^{\prime} respectively. Assuming s0=s0s_{0}=s^{\prime}_{0}, nj=dimWwj/Wwj1n_{j}=\dim W_{w_{j}}/W_{w_{j-1}} and nj=dimWwj/Wwj1n^{\prime}_{j}=\dim W^{\prime}_{w_{j}}/W^{\prime}_{w_{j-1}}, if a linear map T:WWT\colon W\to W^{\prime} respects the filtrations, the matrix representation of TT in the bases e,ee,e^{\prime} is block-upper-triangular, i.e.

M:=MateeT=(M100Ms),M:={\rm Mat}_{e}^{e^{\prime}}\,T=\left(\begin{array}[]{ccccc}M_{1}&*&*\\ 0&\ddots&*\\ 0&&M_{s}\\ \end{array}\right),

with Mjnj×njM_{j}\in\mathbb{R}^{n^{\prime}_{j}}\times\mathbb{R}^{n_{j}}, with j=1,,s0j=1,\ldots,s_{0}. Moreover, the matrix representing gr(T){\rm gr}(T) in e,e\langle e\rangle,\langle e^{\prime}\rangle is block-diagonal, i.e.

gr(M):=Mateegr(T)=(M100000Ms).{\rm gr}(M):={\rm Mat}_{\langle e\rangle}^{\langle e^{\prime}\rangle}\,{\rm gr}(T)=\left(\begin{array}[]{ccccc}M_{1}&0&0\\ 0&\ddots&0\\ 0&&M_{s}\\ \end{array}\right)\,.

2.1.6. Subfiltration

We say that a vector subspace W~W\tilde{W}\subseteq W admits a subfiltration when W~wj:=W~Wwj\tilde{W}_{w_{j}}:=\tilde{W}\cap W_{w_{j}}, j=1,,s0j=1,\ldots,s_{0} yields a filtration of W~\tilde{W}. In this case, the injection map W~W\tilde{W}\hookrightarrow W respects the filtrations and gr(W~){\rm gr}\,(\tilde{W}) is a subspace of gr(W){\rm gr}\,(W).

2.1.7. Decreasing labelling and duality

One can also consider a filtered vector space WW with a decreasing filtration, that is a filtration with decreasing labelling:

(2.5) W=V0V1Vt={0}.\displaystyle W=V_{0}\supseteq V_{1}\supseteq\ldots\supseteq V_{t}=\{0\}\,.

It is not difficult to adapt all the constructions considered previously to a decreasing filtration. Indeed, it suffices to change the labels by setting Wj:=VtjW_{j}:=V_{t-j}, so that, for instance, the associated graded space becomes gr(V):=j=0t1Vj/Vj+1{\rm gr}\,(V):=\oplus_{j=0}^{t-1}V_{j}/V_{j+1}. 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 WW has a decreasing filtration, then we obtain an increasing filtration of the dual space WW^{\ast} of WW by considering the filtration by the subspaces VjWV^{\perp}_{j}\subset W^{\ast}, the annihilators of the subspaces VjWV_{j}\subset W, that is

{0}=V0V1Vt=W,Vj={W|Vj0},\{0\}=V_{0}^{\perp}\subseteq V_{1}^{\perp}\subseteq\ldots\subseteq V_{t}^{\perp}=W^{\ast}\ ,\ V_{j}^{\perp}=\{\ell\in W^{\ast}\mid\ell|_{V_{j}}\equiv 0\}\,,

also known as the dual filtration of (2.5).

Given two finite dimensional filtered vector spaces WW and WW^{\prime}, let us take ee and ee^{\prime} two bases adapted to the decreasing filtrations W=V0V1Vt={0}W=V_{0}\supseteq V_{1}\supseteq\cdots\supseteq V_{t}=\{0\} and W=V0V1Vt={0}W^{\prime}=V^{\prime}_{0}\supseteq V^{\prime}_{1}\supseteq\cdots\supseteq V^{\prime}_{t}=\{0\}. Then for any linear map T:WWT\colon W\to W^{\prime} that respects these decreasing filtrations, the dual map

T:WW where T()=T,WT^{\ast}\colon W^{\prime\ast}\to W^{\ast}\ \text{ where }\ T^{\ast}(\ell)=\ell\circ T\ ,\ \forall\ \ell\in W^{\prime\ast}

respects the dual filtrations, and

MateeT=(MateeT)t,\displaystyle{\rm Mat}_{e^{\prime\ast}}^{e^{\ast}}T^{\ast}=\big{(}{\rm Mat}_{e}^{e^{\prime}}T\big{)}^{t}\,,

where ee^{\ast} and ee^{\prime\ast} denote the dual bases of ee and ee^{\prime} respectively (one can easily show that if a basis ee is adapted to a decreasing filtration (2.5), then ee^{\ast} is adapted to its increasing dual filtration). In particular, this readily implies that the matrix representation of T:WWT^{\ast}\colon W^{\prime\ast}\to W^{\ast} is block-upper-triangular, as pointed out in Subection 2.1.5.

Moreover, the vector spaces gr(V){\rm gr}(V^{*}) and (gr(V))({\rm gr}\,(V))^{*} are canonically isomorphic.

2.1.8. Graded vector spaces

As explained earlier, given a graded vector space GG, its gradation (2.2) naturally produces a filtration through kjGk\oplus_{k\leq j}G_{k}, while a filtration (2.1) of a vector space WW yields a graded space gr(W){\rm gr}\,(W) (2.3).

Definition 2.1.

Let G=j=1sGjG=\oplus_{j=1}^{s}G_{j} and G=j=1sGjG^{\prime}=\oplus_{j^{\prime}=1}^{s^{\prime}}G^{\prime}_{j} be two graded vector spaces. We may assume s=ss=s^{\prime}. A linear map T:GGT:G\to G^{\prime} respects the gradations when T(Gj)GjT(G_{j})\subset G^{\prime}_{j} for any j=1,,sj=1,\ldots,s.

For instance, if TT is a linear map between two filtered vector spaces that respects the filtrations, then gr(T){\rm gr}\,(T) respects the gradations.

Consider a graded vector space GG as above. Then its dual GG^{*} is also graded with the dual gradation G=j=1sGjG^{*}=\oplus_{j=1}^{s}G_{j}^{\perp}. The kthk^{th} exterior algebra is also graded via the (possibly infinite) gradation

kG=w=1Gk,w, where Gk,w:=j1++jk=wGj1Gjk.\wedge^{k}G=\oplus_{w=1}^{\infty}G^{k,w}\,\text{, where }\ G^{k,w}:=\sum_{j_{1}+\cdots+j_{k}=w}G_{j_{1}}\wedge\cdots\wedge G_{j_{k}}\,.

This leads to a (possibly infinite) gradation of the exterior algebra G\wedge^{\bullet}G which respects the wedge product:

G=w=0G,w,G,w=k=0Gk,w,\wedge^{\bullet}G=\oplus_{w=0}^{\infty}G^{\bullet,w},\qquad G^{\bullet,w}=\oplus_{k=0}^{\infty}G^{k,w},

with 0G=G0,0=\wedge^{0}G=G^{0,0}=\mathbb{R}, and G0,w={0}G^{0,w}=\{0\} for w>0w>0. If GG is finite dimensional, the gradations of kG\wedge^{k}G and G\wedge^{\bullet}G are finite.

In the graded setting, there is a natural notion of weight: an element in GjG_{j} has weight jj, and more generally, an element in G,wG^{\bullet,w} has weight ww.

2.1.9. Filtration on the exterior algebra.

A filtration (2.1) on a vector space WW also induces a decreasing filtration on its kthk^{th} exterior algebra:

(2.6) kW=Wk,0Wk,1Wk,ks+1={0},\wedge^{k}W=W^{k,\geq 0}\supseteq W^{k,\geq 1}\supseteq\ldots\supseteq W^{k,\geq k\cdot s+1}=\{0\},

where

Wk,w:=j1++jkwWj1Wjk.W^{k,\geq w}:=\sum_{j_{1}+\cdots+j_{k}\geq w}W_{j_{1}}\wedge\cdots\wedge W_{j_{k}}\,.

The following lemma is easily checked.

Lemma 2.2.

Let WW be a finite dimensional vector space equipped with a filtration (2.1), and let e=(e1,,en)e=(e_{1},\ldots,e_{n}) be a basis of WW adapted to the filtration. Define the linear map

Ψ:kWkgr(W),\Psi:\wedge^{k}W\to\wedge^{k}{\rm gr}\,(W),

via

Ψ(ei1eik)=ei1eik,i1<<in.\Psi(e_{i_{1}}\wedge\cdots\wedge e_{i_{k}})=\langle e_{i_{1}}\rangle\wedge\cdots\wedge\langle e_{i_{k}}\rangle,\quad i_{1}<\ldots<i_{n}.

Then Ψ\Psi is a linear isomorphism that respects the filtration (2.6) of kW\wedge^{k}W and the one induced by the gradation of kgr(W)\wedge^{k}{\rm gr}\,(W). Moreover, gr(Ψ){\rm gr}\,(\Psi) provides an isomorphism between

gr(kW)=w=0ksWk,w/Wk,w+1{\rm gr}(\wedge^{k}W)=\oplus_{w=0}^{k\cdot s}W^{k,\geq w}/W^{k,\geq w+1}

and kgr(W)\wedge^{k}{\rm gr}\,(W) respecting their gradations.

Note that the isomorphism Ψ\Psi depends on the choice of the basis ee.

2.1.10. Euclidean filtered vector spaces

Here, we consider a Euclidean space, that is, a finite dimensional vector space WW equipped with a scalar product (,)W(\cdot,\cdot)_{W}.

Any subspace WWW^{\prime}\subset W is naturally Euclidean, its scalar product (,)W(\cdot,\cdot)_{W^{\prime}} being obtained by restricting (,)W(\cdot,\cdot)_{W}. Moreover, its orthogonal complement W(,W)W^{\prime(\perp,W)} in WW is naturally identified with the quotient

W/WW(,W).W/W^{\prime}\cong W^{\prime(\perp,W)}.

This quotient W/WW/W^{\prime} is also Euclidean if we equip it with the scalar product (,)W/W(\cdot,\cdot)_{W/W^{\prime}} corresponding to (,)W(,W)(\cdot,\cdot)_{W^{\prime(\perp,W)}}.

Beside being Euclidean, we further assume the vector space WW also be filtered with filtration (2.4). The resulting graded space grW{\rm gr}\,W inherits a natural scalar product (,)grW(\cdot,\cdot)_{{\rm gr}W} by imposing that the decomposition (2.3) is orthogonal, and that (,)grW(\cdot,\cdot)_{{\rm gr}W} restricted to each Wwj/Wwj1W_{w_{j}}/W_{w_{j-1}} is given by (,)Wwj/Wwj1(\cdot,\cdot)_{W_{w_{j}}/W_{w_{j-1}}}. Indeed, each Wwj/Wwj1W_{w_{j}}/W_{w_{j-1}} is naturally isomorphic to the orthogonal complement Vj:=Wnj1(,Wnj)V_{j}:=W_{n_{j-1}}^{(\perp,W_{n_{j}})} of Wwj1W_{w_{j-1}} in WwjW_{w_{j}}. Therefore, by construction, the gradation by quotients on gr(W){\rm gr}\,(W) is isomorphic to the following gradation of WW:

W=V1V2Vs0,W=V_{1}\oplus^{\perp}V_{2}\oplus^{\perp}\ldots\oplus^{\perp}V_{s_{0}},

which we will refer to as the orthogonal gradation or orthogonal graded decomposition of WW (associated with the given scalar product and filtration).

We observe that when considering a basis e=(e1,,en)e=(e_{1},\ldots,e_{n}) of WW adapted to its filtration, after a Graham-Schmidt process, we may assume that (e1,,en1)(e_{1},\ldots,e_{n_{1}}) is an orthonormal basis of V1=W1V_{1}=W_{1}, (en1+1,,en2+n1)(e_{n_{1}+1},\ldots,e_{n_{2}+n_{1}}) is an orthonormal basis of V2V_{2}, etc. In other words, after applying a Graham-Schmidt process, we can always obtain a basis ee adapted to the given orthogonal gradation of WW.

2.2. Preliminaries on manifolds

2.2.1. The maps dd and d~\tilde{d} 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 MM (without any further assumptions). As is customary, dd denotes the exterior differential on the space of smooth forms Ω(M)\Omega^{\bullet}(M) on MM. It is well known that the map d:Ωk(M)Ωk+1(M)d\colon\Omega^{k}(M)\to\Omega^{k+1}(M) is a smooth differential map that satisfies the Leibniz property and d2=0d^{2}=0. The associated cochain complex (Ω(M),d)(\Omega^{\bullet}(M),d) is traditionally referred to as the de Rham complex.

The explicit formula for dd is given for any ωΩk(M)\omega\in\Omega^{k}(M) and V0,,VkΓ(TM)V_{0},\ldots,V_{k}\in\Gamma(TM) by

dω(V0,,Vk)\displaystyle d\omega(V_{0},\ldots,V_{k}) =i=0k(1)iVi(ω(V0,,V^i,,Vk))\displaystyle=\sum_{i=0}^{k}(-1)^{i}V_{i}\left(\omega(V_{0},\ldots,\hat{V}_{i},\ldots,V_{k})\right)
+0i<jk(1)i+jω([Vi,Vj],V0,,V^i,,V^j,,Vk),\displaystyle\qquad+\sum_{0\leq i<j\leq k}(-1)^{i+j}\omega([V_{i},V_{j}],V_{0},\ldots,\hat{V}_{i},\ldots,\hat{V}_{j},\ldots,V_{k}),

the hats denoting omissions.

In this paper, the algebraic part of dd is denoted by d~\tilde{d}. This is the map given by d~=0\tilde{d}=0 on Ω0(M)\Omega^{0}(M), and for k>0k>0 and any ωΩk(M)\omega\in\Omega^{k}(M), V0,,VkΓ(TM)V_{0},\ldots,V_{k}\in\Gamma(TM), by

(2.7) d~ω(V0,,Vk)=0i<jk(1)i+jω([Vi,Vj],V0,,V^i,,V^j,,Vk).\tilde{d}\omega(V_{0},\ldots,V_{k})=\sum_{0\leq i<j\leq k}(-1)^{i+j}\omega([V_{i},V_{j}],V_{0},\ldots,\hat{V}_{i},\ldots,\hat{V}_{j},\ldots,V_{k}).
Lemma 2.3.

The map d~:Ωk(M)Ωk+1(M)\tilde{d}:\Omega^{k}(M)\to\Omega^{k+1}(M) is smooth and algebraic. It satisfies the Leibniz property and (Ω(M),d~)(\Omega^{\bullet}(M),\tilde{d}) is a complex, i.e. d~2=0\tilde{d}^{2}=0. The action of d~\tilde{d} over Ω1(M)\Omega^{1}(M) is given by the explicit formula

d~ω(V0,V1)=ω([V0,V1]),ωΩ1(M),V0,V1Γ(TM).\tilde{d}\omega(V_{0},V_{1})=-\omega([V_{0},V_{1}])\ ,\ \forall\,\omega\in\Omega^{1}(M),\ V_{0},V_{1}\in\Gamma(TM)\,.
Sketch of the proof.

Since d~=d(dd~)\tilde{d}=d-(d-\tilde{d}) and for any ωΩk(M),V0,,VkΓ(TM)\omega\in\Omega^{k}(M),\,V_{0},\ldots,V_{k}\in\Gamma(TM)

(2.8) (dd~)ω(V0,,Vk)=i=0k(1)iVi(ω(V0,,V^i,,Vk)),(d-\tilde{d})\,\omega(V_{0},\ldots,V_{k})=\sum_{i=0}^{k}(-1)^{i}V_{i}\left(\omega(V_{0},\ldots,\hat{V}_{i},\ldots,V_{k})\right)\ ,

the Leibniz property of d~\tilde{d} follows from the fact that both vector fields and dd satisfy the Leibniz property.

The formula on 1-forms is easily computed from (2.7) by taking k=1k=1. This formula, together with the Jacobi identity for the commutator bracket, implies d~2=0\tilde{d}^{2}=0 on Ω1(M)\Omega^{1}(M). Recursively and by the Leibniz rule, we get that d~2=0\tilde{d}^{2}=0 on forms of any degree, since for any αΩk(M)\alpha\in\Omega^{k}(M) and any βΩ(M)\beta\in\Omega^{\bullet}(M)

d~2(αβ)=\displaystyle\tilde{d}^{2}(\alpha\wedge\beta)= d~(d~αβ+(1)kαd~β)\displaystyle\tilde{d}\big{(}\tilde{d}\alpha\wedge\beta+(-1)^{k}\alpha\wedge\tilde{d}\beta\big{)}
=\displaystyle= (d~2α)β+(1)k+1d~αd~β+(1)kd~αd~β+αd~2β=0.\displaystyle(\tilde{d}^{2}\alpha)\wedge\beta+(-1)^{k+1}\tilde{d}\alpha\wedge\tilde{d}\beta+(-1)^{k}\tilde{d}\alpha\wedge\tilde{d}\beta+\alpha\wedge\tilde{d}^{2}\beta=0\,.

2.3. Some natural concepts

In this subsection, we recall some basic concepts in differential geometry.

2.3.1. Frame and coframe

Let EE be a (real, finite dimensional, and smooth) vector bundle over a (smooth) manifold MM with dimM=n\dim M=n. A frame for EE is a smooth section (S1,,Sn)(S_{1},\ldots,S_{n}) of E××E=EnE\times\ldots\times E=E^{n} such that (S1(x),,Sn(x))(S_{1}(x),\ldots,S_{n}(x)) is a basis of ExE_{x} for every xMx\in M. A coframe for EE is frame for the dual vector bundle EE^{*}.

Throughout this paper, if the bundle EE is not specified, we will take EE to be the tangent bundle TMTM over the smooth manifold MM and EE^{*} to be the cotangent bundle. Frames for TMTM may not exist globally on the whole manifold MM, but they can always be constructed locally, i.e. over an open neighbourhood of every point in MM.

2.3.2. Algebraic maps on bundles

Let EE and FF be two smooth vector bundles over a manifold MM. We say that a smooth map T:Γ(E)Γ(F)T:\Gamma(E)\to\Gamma(F) is linear and algebraic when, for each xMx\in M, there exists a linear map (T)x:ExFx(T)_{x}:E_{x}\to F_{x} satisfying T(α)(x)=(T)x(α(x))T(\alpha)(x)=(T)_{x}(\alpha(x)) for any αE\alpha\in E, xMx\in M. We will call (T)x(T)_{x} the pointwise restriction of TT above xMx\in M.

Remark 2.4.

A (smooth) differential operator of order 0 is an algebraic linear map.

2.3.3. Metric on bundles

Let EE be a vector bundle over a manifold MM. By definition, a metric gg on EE is a family of scalar products gx:=(,)Exg_{x}:=(\cdot,\cdot)_{E_{x}} on ExE_{x} depending smoothly on xMx\in M. The smooth dependence means that

gΓ(Sym(EE)).g\in\Gamma({\rm Sym}(E\otimes E))\,.

When EE is the tangent bundle TMTM of a smooth manifold MM, the metric is said to be Riemannian and the couple (M,g)(M,g) is referred to as a Riemannian manifold.

2.4. Filtered tangent bundles

In this section, we consider an nn-dimensional smooth manifold MM whose tangent bundle TMTM is filtered by vector subbundles

(2.9) M×{0}=H0H1Hs=TM.M\times\{0\}=H^{0}\subseteq H^{1}\subseteq\ldots\subseteq H^{s}=TM\,.

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

gr(TM)=i=1sHi/Hi1=j=1s0Hwj/Hwj1,{\rm gr}(TM)=\oplus_{i=1}^{s}H^{i}/H^{i-1}=\oplus_{j=1}^{s_{0}}H^{w_{j}}/H^{w_{j-1}}\ ,

with nj:=dimHwj/Hwj1>0,j=1,,s0n_{j}:=\dim H^{w_{j}}/H^{w_{j-1}}>0,\ j=1,\ldots,s_{0}, and dim(M)=n=n1+n2++ns0dim(M)=n=n_{1}+n_{2}+\cdots+n_{s_{0}}.

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 𝕏=(X1,,Xn)\mathbb{X}=(X_{1},\ldots,X_{n}) for TMTM is said to be adapted to the filtration (2.9) when, for every xMx\in M, the basis (X1(x),,Xn(x))(X_{1}(x),\ldots,X_{n}(x)) is adapted to the filtration TxM=Hxs0Hxw1Hxw0={0}T_{x}M=H^{s_{0}}_{x}\supseteq\ldots\supseteq H^{w_{1}}_{x}\supseteq H^{w_{0}}_{x}=\{0\}.

If 𝕏=(X1,,Xn)\mathbb{X}=(X_{1},\ldots,X_{n}) is a frame for TMTM adapted to the filtration (2.9), then the associated graded basis on gr(TxM){\rm gr}\,(T_{x}M)

𝕏x=(X1x,,Xnx),xM,\langle\mathbb{X}\rangle_{x}=(\langle X_{1}\rangle_{x},\ldots,\langle X_{n}\rangle_{x})\,,\ x\in M\,,

yields a frame 𝕏\langle\mathbb{X}\rangle for gr(TM){\rm gr}(TM) adapted to the gradation in the following sense.

Definition 2.6.

A frame 𝕐=(Y1,,Yn)\mathbb{Y}=(Y_{1},\ldots,Y_{n}) for gr(TM){\rm gr}(TM) is adapted to the gradation, or graded, when Y1,,Yn1Y_{1},\ldots,Y_{n_{1}} is a frame for Hw1/Hw0H^{w_{1}}/H^{w_{0}}, Yn1+1,,Yn2+n1Y_{n_{1}+1},\ldots,Y_{n_{2}+n_{1}} is a frame for Hw2/Hw1H^{w_{2}}/H^{w_{1}}, and so on.

By duality, we obtain the following decreasing filtration of the cotangent bundle TMT^{*}M by vector bundles

(2.10) TM=(Hw0)(Hw1)(Hws0)=M×{0},T^{*}M=(H^{w_{0}})^{\perp}\supseteq(H^{w_{1}})^{\perp}\supseteq\ldots\supseteq(H^{w_{s_{0}}})^{\perp}=M\times\{0\}\,,

by considering the annihilators (Hwj)(H^{w_{j}})^{\perp} of HwjH^{w_{j}} for j=1,,s0j=1,\ldots,s_{0}, as follows.

Definition 2.7.

A coframe Θ\Theta for TMTM is adapted to the filtration when, for every xMx\in M, the basis Θ(x):=(θ1(x),,θn(x))\Theta(x):=(\theta^{1}(x),\ldots,\theta^{n}(x)) is adapted to the filtration of TxM=(Hxw0)(Hxw1)(Hxws0)={0}T_{x}^{*}M=(H^{w_{0}}_{x})^{\perp}\supseteq(H^{w_{1}}_{x})^{\perp}\supseteq\ldots\supseteq(H^{w_{s_{0}}}_{x})^{\perp}=\{0\}.

This definition is equivalent to saying that, at each point xMx\in M, the final ns0n_{s_{0}} covectors annihilate Hxws01H^{w_{s_{0}-1}}_{x}. Furthermore, these final ns0n_{s_{0}} covectors together with the preceding ns01n_{s_{0}-1} covectors annihilate Hxs02H^{s_{0}-2}_{x}, and so on. The associated frame Θ=(θ1,,θn)\langle\Theta\rangle=(\langle\theta^{1}\rangle,\ldots,\langle\theta^{n}\rangle) is a frame for gr(TM){\rm gr}(T^{*}M) that is adapted to the gradation in the following sense.

Definition 2.8.

A frame Ξ=(ξ1,,ξn)\Xi=(\xi^{1},\ldots,\xi^{n}) for gr(TM){\rm gr}(T^{*}M) is said to be adapted to the gradation (or simply graded) when ξn,,ξnns0+1\xi^{n},\ldots,\xi^{n-n_{s_{0}}+1} is a frame for (Hws01)/(Hws0)(H^{w_{s_{0}-1}})^{\perp}/(H^{w_{s_{0}}})^{\perp}, ξnns0,,\xi^{n-n_{s_{0}}},\ldots, ξnns01+1\xi^{n-n_{s_{0}-1}+1} is a frame for (Hws02)/(Hws01)(H^{w_{s_{0}-2}})^{\perp}/(H^{w_{s_{0}-1}})^{\perp}, and so on.

One can readily check that if 𝕏\mathbb{X} is a frame adapted to a gradation for TMTM, then its dual Θ\Theta is an adapted coframe for TMTM. Conversely, if Θ\Theta is a coframe for TMT^{\ast}M adapted to a gradation of TMTM, then its dual Θ=𝕏\Theta^{\ast}=\mathbb{X} is a graded frame for TMTM.

As mentioned in Subsection 2.3.1, frames for TMTM can always be constructed locally. Furthermore, through a pivoting process, one may easily construct a local frame for TMTM adapted to a given gradation. The inverse function theorem implies that, given a local frame 𝕐\mathbb{Y} of gr(TM){\rm gr}(TM) adapted to the gradation, we can then construct a local frame 𝕏\mathbb{X} of TMTM adapted to the filtration such that 𝕏=𝕐\langle\mathbb{X}\rangle=\mathbb{Y}.

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 ww in this context.

Given an increasing filtration (2.9) of TMTM, we denote by HjH^{j} the set of (based) vectors of weight at most jj (shortened with j\leq j). By duality, we say that the (based) covectors in (Hj1)(H^{j-1})^{\perp} have weight at least jj (shortened as j\geq j). We extend this vocabulary to the space of smooth sections Γ(TM)\Gamma(TM) and Γ(TM)=Ω1(M)\Gamma(T^{*}M)=\Omega^{1}(M), that is vector fields and 1-forms respectively.

With our definition, a 1-form αΩ1(M)\alpha\in\Omega^{1}(M) is of weight w\geq w when α(V)=0\alpha(V)=0 for any VΓ(Hw1)V\in\Gamma(H^{w-1}). We denote the space of 1-forms of weight at least jj by

Ω1,j(M):=Γ((Hj1)),j=1,2,,s0+1.\Omega^{1,\geq j}(M):=\Gamma((H^{j-1})^{\perp}),\qquad j=1,2,\ldots,s_{0}+1\,.

The filtration in (2.10) of the cotangent bundle then determines the following strict filtration

Ω1(M)=Ω1,w1(M)Ω1,w2(M)Ω1,ws0+1(M)={0}.\Omega^{1}(M)=\Omega^{1,\geq w_{1}}(M)\supsetneq\Omega^{1,\geq w_{2}}(M)\supsetneq\ldots\supsetneq\Omega^{1,\geq w_{s_{0}+1}}(M)=\{0\}.
Example 2.9.

If Θ:=(θ1,,θn)\Theta:=(\theta^{1},\ldots,\theta^{n}) is a frame for TMT^{\ast}M adapted to the filtration (2.10), then all the 1-forms θ1,,θn\theta^{1},\ldots,\theta^{n} are of weight w1\geq w_{1}. The 1-forms θn1+1,,θn\theta^{n_{1}+1},\ldots,\theta^{n} are of weight w2\geq w_{2}, and more generally θn1++ni1+1,,θn\theta^{n_{1}+\cdots+n_{i-1}+1},\ldots,\theta^{n} are of weight wi\geq w_{i}. By taking n0=dimHw0=0n_{0}=\dim H^{w_{0}}=0, we can state this as

θjΩ1,wi(M),j=n0++ni1+1,,n,i=1,,s0+1.\theta^{j}\in\Omega^{1,\geq w_{i}}(M),\quad j=n_{0}+\cdots+n_{i-1}+1,\ldots,n\ ,\quad i=1,\ldots,s_{0}+1.

In general, we will not have a global coframe as in Example 2.9, but just local ones.

For k=0k=0, we set

C(M)=Ω0(M):=Ω0,0(M),andΩ0,1(M):={0}.C^{\infty}(M)=\Omega^{0}(M):=\Omega^{0,\geq 0}(M),\qquad\mbox{and}\qquad\Omega^{0,\geq 1}(M):=\{0\}.

For k=1,2,k=1,2,\ldots, we define the space of kk-forms of weight at least ww (shortened as w\geq w) by

Ωk,w(M):=i1++ikwΩ1,i1(M)Ω1,ik(M).\Omega^{k,\geq w}(M):=\oplus_{i_{1}+\ldots+i_{k}\geq w}\Omega^{1,\geq i_{1}}(M)\wedge\ldots\wedge\Omega^{1,\geq i_{k}}(M).

In other words, a kk-form is of weight at least ww when it can be written locally as a linear combination of kk-wedges of 1-form of weights at least i1,,iki_{1},\ldots,i_{k} with wi1++ikw\leq i_{1}+\ldots+i_{k}. From the definition, it follows that a kk-form αΩk(M)\alpha\in\Omega^{k}(M) is of weight w\geq w when

V1Γ(Hi1),,VkΓ(Hik)i1++ik<wα(V1,,Vk)=0.\forall\ V_{1}\in\Gamma(H^{i_{1}}),\ldots,V_{k}\in\Gamma(H^{i_{k}})\qquad i_{1}+\ldots+i_{k}<w\Longrightarrow\alpha(V_{1},\ldots,V_{k})=0.
Example 2.10.

Consider an adapted coframe Θ\Theta for TMTM. We will often use the shorthand

ΘI=θi1θik\Theta^{\wedge I}=\theta^{i_{1}}\wedge\ldots\wedge\theta^{i_{k}}

for the multi-index I=(i1,,ik)I=(i_{1},\ldots,i_{k}), always assuming i1<<iki_{1}<\ldots<i_{k}. Then ΘI\Theta^{\wedge I} is of weight υ1++υk\geq\upsilon_{1}+\ldots+\upsilon_{k}, where each θij\theta^{i_{j}} is of weight υj\geq\upsilon_{j}.

Definition 2.11.

The family (ΘI)I(\Theta^{I})_{I} where II runs above all multi-indices of length kk (taken with strictly increasing indices as in example 2.10 above) is a free basis of the C(M)C^{\infty}(M)-module Ω(M)\Omega^{\bullet}(M) called the basis associated with the coframe Θ\Theta.

By construction, we have the inclusion Ωk,w(M)Ωj,w(M)Ωk+j,w+w(M)\Omega^{k,\geq w}(M)\wedge\Omega^{j,\geq w^{\prime}}(M)\subseteq\Omega^{k+j,\geq w+w^{\prime}}(M) for any k,j0k,j\geq 0. Moreover, given the increasing filtration (2.4) of TMTM, for any degree k0k\geq 0

Ωk,kws0+1(M)={0} and wwΩk,w(M)Ωk,w(M).\Omega^{k,\geq k\cdot w_{s_{0}}+1}(M)=\{0\}\ \text{ and }\ w\leq w^{\prime}\ \Longrightarrow\ \Omega^{k,\geq w}(M)\supseteq\Omega^{k,\geq w^{\prime}}(M)\,.

In particular, the module Ωk(M)\Omega^{k}(M) over the ring C(M)C^{\infty}(M) admits the filtration

(2.11) Ωk(M)=Ωk,0(M)Ωk,1(M)Ωk,w(M)Ωk,kws0+1(M)={0}.\displaystyle\Omega^{k}(M)=\Omega^{k,\geq 0}(M)\supset\Omega^{k,\geq 1}(M)\supset\ldots\supset\Omega^{k,\geq w}(M)\supset\ldots\supset\Omega^{k,\geq k\cdot w_{s_{0}}+1}(M)=\{0\}.

The basis (ΘI)I(\Theta^{I})_{I} introduced in Definition 2.11 is adapted to this filtration.

2.4.3. The bundle kgr(TM)\wedge^{k}{\rm gr}(T^{*}M) and its weights

Given the filtration (2.9), by duality the smooth vector bundle

gr(TM)=i=1s((Hi1)/(Hi))=j=1s0((Hwj1)/(Hwj)){\rm gr}(T^{*}M)=\oplus_{i=1}^{s}((H^{i-1})^{\perp}/(H^{i})^{\perp})=\oplus_{j=1}^{s_{0}}((H^{w_{j}-1})^{\perp}/(H^{w_{j}})^{\perp})

is graded. We say that an element of gr(TM){\rm gr}(T^{*}M) is of weight jj when it is in (Hj1)/(Hj)(H^{j-1})^{\perp}/(H^{j})^{\perp}. We extend the same notion of weight to smooth sections of gr(TM){\rm gr}(T^{*}M).

Example 2.12.

Let us consider the same adapted coframe Θ\Theta for TMTM as in Example 2.9. Then θ1,,θn1\langle\theta^{1}\rangle,\ldots,\langle\theta^{n_{1}}\rangle have weight w1w_{1}, and for i=2,,s0i=2,\ldots,s_{0}, the sections θn1++ni1+1,,\langle\theta^{n_{1}+\cdots+n_{i-1}+1}\rangle,\ldots, θn1++ni\langle\theta^{n_{1}+\cdots+n_{i}}\rangle of gr(TM){\rm gr}(T^{*}M) have weight wiw_{i}. If we denote by υj\upsilon_{j} the weight of θj\langle\theta^{j}\rangle, we have:

(2.12) υn1++ni1+1==υn1++ni=wi,i=1,,s0.\upsilon_{n_{1}+\cdots+n_{i-1}+1}=\ldots=\upsilon_{n_{1}+\cdots+n_{i}}=w_{i},\qquad i=1,\ldots,s_{0}.

There is a natural identification between the following smooth vector bundles

kgr(TM):=xMkgr(TxM)xM(kgr(TxM))=:kgr(TM),\wedge^{k}{\rm gr}(T^{*}M):=\cup_{x\in M}\wedge^{k}{\rm gr}(T_{x}^{*}M)\cong\cup_{x\in M}(\wedge^{k}{\rm gr}(T_{x}M))^{*}=:\wedge^{k}{\rm gr}(TM)^{*},

allowing us to consider directly kgr(TM)\wedge^{k}{\rm gr}(T^{*}M) in our discussion.

It is straightforward to check that kgr(TM)\wedge^{k}{\rm gr}(T^{*}M) is a module over C(M)C^{\infty}(M) which, by Section 2.1.8, is naturally equipped with a gradation and a notion of weight. In other words,

kgr(TM)=wk,wgr(TM),\wedge^{k}{\rm gr}(T^{*}M)=\oplus_{w\in\mathbb{N}}\wedge^{k,w}{\rm gr}(T^{*}M),

where the elements of weight ww are in the linear subbundle

k,wgr(TM):=i1++ik=w((Hi11)/(Hi1))((Hik1)/(Hik)).\wedge^{k,w}{\rm gr}(T^{*}M):=\sum_{i_{1}+\ldots+i_{k}=w}((H^{i_{1}-1})^{\perp}/(H^{i_{1}})^{\perp})\wedge\ldots\wedge((H^{i_{k}-1})^{\perp}/(H^{i_{k}})^{\perp}).

Once we extend this construction to the space of smooth sections of gr(TM)\wedge^{\bullet}{\rm gr}\,(T^{\ast}M), we get the gradation

(2.13) Γ(gr(TM))=k,w0Γ(k,wgr(TM)),\displaystyle\Gamma(\wedge^{\bullet}{\rm gr}\,(T^{\ast}M))=\oplus_{k,w\in\mathbb{N}_{0}}\ \Gamma(\wedge^{k,w}{\rm gr}\,(T^{\ast}M))\,,

where the kk-forms of weight ww are given by

Γ(k,wgr(TM))=\displaystyle\Gamma(\wedge^{k,w}{\rm gr}\,(T^{\ast}M))= i1++ik=wΓ((Hi11))/Γ((Hi1))Γ((Hik1))/Γ((Hik)).\displaystyle\sum_{i_{1}+\cdots+i_{k}=w}\Gamma((H^{i_{1}-1})^{\perp})/\Gamma((H^{i_{1}})^{\perp})\wedge\cdots\wedge\Gamma((H^{i_{k}-1})^{\perp})/\Gamma((H^{i_{k}})^{\perp})\,.
Example 2.13.

We continue with the setting of Example 2.12 and examine the case of kk-forms, k>1k>1. The sections θi1θik\langle\theta^{i_{1}}\rangle\wedge\ldots\wedge\langle\theta^{i_{k}}\rangle of kgr(TM)\wedge^{k}{\rm gr}(T^{*}M) are of weight

|(υ1,,υk)|:=υi1++υik.|(\upsilon_{1},\ldots,\upsilon_{k})|:=\upsilon_{i_{1}}+\ldots+\upsilon_{i_{k}}.

We may use the shorthand notation ΘI=θi1θik\langle\Theta\rangle^{\wedge I}=\langle\theta^{i_{1}}\rangle\wedge\ldots\wedge\langle\theta^{i_{k}}\rangle where II is the multi-index I=(i1,,ik)I=(i_{1},\ldots,i_{k}) with i1<<iki_{1}<\ldots<i_{k}.

We readily check the following properties:

Lemma 2.14.

The family (ΘI)I(\langle\Theta\rangle^{I})_{I}, where II runs above all multi-indices of length kk and strictly increasing indices, is a basis of Γ(kgr(TM))\Gamma(\wedge^{k}{\rm gr}(T^{*}M)) adapted to the gradation. Moreover, it is the basis associated to the basis (ΘI)I(\Theta^{I})_{I} adapted to the filtration (2.11) of Ω(M)\Omega^{\bullet}(M) in the sense described in Section 2.1.4.

2.4.4. The gr{\rm gr} operation

Definition 2.15.

A morphism D:Ω(M)Ω(M)D:\Omega^{\bullet}(M)\to\Omega^{\bullet}(M) of C(M)C^{\infty}(M)-modules respects the filtration (2.11), when

w0DΩ,w(M)Ω,w(M).\forall w\in\mathbb{N}_{0}\qquad D\,\Omega^{\bullet,\geq w}(M)\subseteq\Omega^{\bullet,\geq w}(M).

The following property follows readily from Section 2.1:

Lemma 2.16.

Let D:Ω(M)Ω(M)D:\Omega^{\bullet}(M)\to\Omega^{\bullet}(M) be a morphism of C(M)C^{\infty}(M)-modules that respects the filtration (2.11). Then the map

gr(D):Γ(gr(TM))Γ(gr(TM)),{\rm gr}(D):\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M))\to\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M)),

is a morphism of C(M)C^{\infty}(M)-modules that respects the gradation (2.13). If in addition, DD is a differential operator, then gr(D){\rm gr}(D) is also a differential operator of the same order or lower.

As an operation, gr{\rm gr} is a module morphism from {D:Ω(M)Ω(M):\{D:\Omega^{\bullet}(M)\to\Omega^{\bullet}(M): module morphism}\} to {T:Γ(gr(TM))Γ(gr(TM)):\{T:\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M))\to\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M)): module morphism}\}.

By construction, gr(D)=0{\rm gr}(D)=0 if and only if DD increases the weight in the following sense.

Definition 2.17.

We say that a linear map D:Ω(M)Ω(M)D:\Omega^{\bullet}(M)\to\Omega^{\bullet}(M) increases the weights of the filtration when

w0DΩ,w(M)Ω,w+1(M).\forall w\in\mathbb{N}_{0}\qquad D\,\Omega^{\bullet,\geq w}(M)\subseteq\Omega^{\bullet,\geq w+1}(M).

As there are only a finite number of weights, a linear map that increases the weights is necessarily nilpotent, that is

(2.14) grD=0N0 such that DN0=0.{\rm gr}\,D=0\Longrightarrow\quad\exists\,N_{0}\in\mathbb{N}\ \text{ such that }D^{N_{0}}=0.

This N0N_{0} depends only on MM and its filtration (2.9) (via the filtrations (2.11)), not on DD. Throughout this paper, we will assume it to be a fixed integer.

2.5. Tangent bundles equipped with a filtration and a metric

Let MM be a manifold whose tangent bundle admits a filtration (2.9) by subbundles and is equipped with a metric gTMg_{TM}.

2.5.1. Identification of Ω(M)\Omega^{\bullet}(M) and Γ(gr(TM))\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M))

The considerations in Section 2.1.10 show that gr(TM){\rm gr}(TM) inherits a metric ggr(TM)g_{{\rm gr}(TM)}. The same section also implies that, at each point xMx\in M, there exists an open neighbourhood UU of xx and a frame 𝕏\mathbb{X} of TMTM on UU that is gTMg_{TM}-orthonormal and adapted to the natural orthogonal gradation

TM=jGj,TM=\oplus_{j}^{\perp}G^{j},

where GjG^{j} is the orthogonal complement of Hwj1H^{w_{j-1}} in HwjH^{w_{j}} for each j=1,,s0j=1,\ldots,s_{0}.

Let xMx\in M. For any j=1,,s0j=1,\ldots,s_{0}, if ξ\xi is a covector in GxjG^{j}_{x} (i.e. a linear form on TxMT_{x}M vanishing on every GxkG^{k}_{x}, kjk\neq j), we denote by Φx(ξ)\Phi_{x}(\xi) the corresponding covector in Hxwj/Hxwj1H^{w_{j}}_{x}/H^{w_{j-1}}_{x} (i.e. the corresponding linear form on gr(TxM){\rm gr}(T^{*}_{x}M) vanishing on Hxwk/Hxwk1H^{w_{k}}_{x}/H^{w_{k-1}}_{x}, kjk\neq j). This extends uniquely to a linear bijective map Φx:TxMgr(TxM)\Phi_{x}:T_{x}^{*}M\to{\rm gr}(T_{x}^{*}M), and then to a C(M)C^{\infty}(M)-module isomorphism

Φ:Ω(M)Γ(gr(TM))\Phi:\Omega^{\bullet}(M)\to\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M))

respecting the Leibniz property.

Lemma 2.18.

Let MM be a manifold whose tangent bundle admits a filtration by subbundles (2.9) and is equipped with a metric gTMg_{TM}. The map Φ\Phi constructed above is an isomorphism of C(M)C^{\infty}(M)-modules that respects the filtrations, even when restricted to Ωk(M)Γ(kgr(TM))\Omega^{k}(M)\to\Gamma(\wedge^{k}{\rm gr}(T^{*}M)) for each k=0,1,,nk=0,1,\ldots,n. Moreover, gr(Φ){\rm gr}(\Phi) is an isomorphism between the C(M)C^{\infty}(M)-modules gr(Ωk(M)){\rm gr}(\Omega^{k}(M)) and Γ(kgr(TM))\Gamma(\wedge^{k}{\rm gr}(T^{*}M)) that respects the gradation.

If 𝕏\mathbb{X} is a gTMg_{TM}-orthonormal local frame of TMTM on an open subset UMU\subset M that is adapted to the natural orthogonal gradation TM=GjTM=\oplus^{\perp}G^{j}, then we have

Φ(ΘI)=ΘI\Phi(\Theta^{I})=\langle\Theta\rangle^{I}

for any increasing multi-index II of length kk, where Θ\Theta is the coframe associated to 𝕏\mathbb{X} over UU.

Note that the map Φ\Phi is in fact algebraic, with

Φx:Γ(TxM)Γ(gr(TM))defined viaΦx(ΘI(x))=ΘI(x),\Phi_{x}:\Gamma(\wedge^{\bullet}T_{x}^{*}M)\to\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M))\quad\mbox{defined via}\quad\Phi_{x}(\Theta^{I}(x))=\langle\Theta\rangle^{I}(x),

using the notation of Lemma 2.18. When restricted to Ω1(M)\Omega^{1}(M), the matrix representing the map Φ\Phi is just the identity matrix of n\mathbb{R}^{n}:

MatΘΘ(Φ:Ω1(M)Γ(gr(TM)))=I.{\rm Mat}_{\Theta}^{\langle\Theta\rangle}(\Phi:\Omega^{1}(M)\to\Gamma({\rm gr}(T^{\ast}M)))=\text{\rm I}.

Moreover, for each k=2,,nk=2,\ldots,n, using the notation Θk=(ΘI)I\Theta^{k}=(\Theta^{I})_{I} and Θk=(ΘI)I\langle\Theta\rangle^{k}=(\langle\Theta\rangle^{I})_{I} for the corresponding bases of Ωk(M)\Omega^{k}(M) and of Γ(kgr(TM))\Gamma(\wedge^{k}{\rm gr}(T^{*}M)), the matrix representing Φ:Ωk(M)Γ(kgr(TM))\Phi\colon\Omega^{k}(M)\to\Gamma(\wedge^{k}{\rm gr}(T^{\ast}M)) is the identity matrix of dimΩk(M)\mathbb{R}^{\dim\Omega^{k}(M)}:

MatΘkΘk(Φ)=I.{\rm Mat}_{\Theta^{k}}^{\langle\Theta\rangle^{k}}(\Phi)=\text{\rm I}\,.

Intertwining with Φ\Phi provides forms with a reverse operation to the operation gr{\rm gr} described in Lemma 2.16:

Lemma 2.19.

We continue with the setting of Lemma 2.18.

  1. (1)

    If T:Γ(gr(TM))Γ(gr(TM))T:\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M))\to\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M)) is a morphism of C(M)C^{\infty}(M)-modules, then

    TΦ:=Φ1TΦ:Ω(M)Ω(M)T^{\Phi}:=\Phi^{-1}\circ T\circ\Phi:\Omega^{\bullet}(M)\to\Omega^{\bullet}(M)

    is a morphism of C(M)C^{\infty}(M)-modules. If in addition TT is a differential operator, then TΦT^{\Phi} is also a differential operator of the same order.

  2. (2)

    If T:Γ(gr(TM))Γ(gr(TM))T:\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M))\to\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M)) is a morphism of C(M)C^{\infty}(M)-modules that respects the gradation, then TΦ:Ω(M)Ω(M)T^{\Phi}:\Omega^{\bullet}(M)\to\Omega^{\bullet}(M) respects the filtration and we have

    gr(TΦ)=T.{\rm gr}(T^{\Phi})=T.
  3. (3)

    If D:Ω(M)Ω(M)D:\Omega^{\bullet}(M)\to\Omega^{\bullet}(M) respects the filtration, then D(gr(D))ΦD-({\rm gr}(D))^{\Phi} strictly increases weights, and so

    gr(D)=gr((gr(D))Φ).{\rm gr}(D)={\rm gr}\left(({\rm gr}(D))^{\Phi}\right).
Proof.

These properties are easily checked on ΘI\Theta^{I} and ΘI\langle\Theta\rangle^{I} having fixed a local orthonormal adapted coframe Θ\Theta. ∎

2.5.2. Scalar products on TxM\wedge^{\bullet}T_{x}^{*}M and of gr(TxM)\wedge^{\bullet}{\rm gr}(T_{x}^{*}M)

For each xMx\in M, the scalar products gTxMg_{T_{x}M} and ggr(TxM)g_{{\rm gr}(T_{x}M)} on TxMT_{x}M and gr(TxM){\rm gr}(T_{x}M) respectively induce scalar products on their duals TxMT_{x}^{*}M and gr(TxM)gr(TxM){\rm gr}(T_{x}M)^{*}\cong{\rm gr}(T_{x}^{\ast}M), and then scalar products gTxMg_{\wedge^{\bullet}T_{x}^{*}M} on TxM\wedge^{\bullet}T_{x}^{*}M and ggr(TxM)g_{\wedge^{\bullet}{\rm gr}(T_{x}^{*}M)} on gr(TxM)\wedge^{\bullet}{\rm gr}(T_{x}^{*}M) via the formula in (2.15) below and the orthogonal decompositions:

TxM=kTxMandgr(TxM)=kgr(TxM).\wedge^{\bullet}T_{x}^{*}M=\oplus^{\perp}\wedge^{k}T_{x}^{*}M\quad\mbox{and}\quad\wedge^{\bullet}{\rm gr}(T_{x}^{*}M)=\oplus^{\perp}\wedge^{k}{\rm gr}(T_{x}^{*}M)\,.

We will denote all these scalar product as ,\langle\cdot,\cdot\rangle, their meaning being clear from the context or from an indexation. We have

(2.15) α1αk,β1βkkTxM:=det(αi,βjTxM)1i,jk,\langle\alpha_{1}\wedge\ldots\wedge\alpha_{k},\beta_{1}\wedge\ldots\wedge\beta_{k}\rangle_{\wedge^{k}T_{x}^{*}M}:=\det\left(\langle\alpha_{i},\beta_{j}\rangle_{T_{x}^{*}M}\right)_{1\leq i,j\leq k},

where α1,,αk,β1,βkTxM\alpha_{1},\ldots,\alpha_{k},\beta_{1}\ldots,\beta_{k}\in T_{x}^{*}M, and similarly for gr(kTxM)kgr(TxM){\rm gr}(\wedge^{k}T_{x}^{*}M)\cong\wedge^{k}{\rm gr}(T_{x}^{\ast}M).

We check readily that the map Φx\Phi_{x} defined above sends the scalar product of TxM\wedge^{\bullet}T_{x}^{*}M onto the one of gr(TxM)\wedge^{\bullet}{\rm gr}(T_{x}^{*}M):

(2.16) Φx:TxMgr(TxM),Φx(α),Φx(β)=α,βα,βTxM.\Phi_{x}\colon\wedge^{\bullet}T_{x}^{\ast}M\to\wedge^{\bullet}{\rm gr}(T_{x}^{\ast}M)\ ,\quad\langle\Phi_{x}(\alpha),\Phi_{x}(\beta)\rangle=\langle\alpha,\beta\rangle\ \forall\,\alpha,\beta\in\wedge^{\bullet}T_{x}^{*}M\,.

Given an orthonormal local frame 𝕏\mathbb{X} or equivalently an orthonormal local coframe Θ\Theta, we check readily that the bases Θk(x)=(ΘI)I(x)\Theta^{k}(x)=(\Theta^{I})_{I}(x) and Θk(x)=(ΘI)I(x)\langle\Theta^{k}\rangle(x)=(\langle\Theta^{I}\rangle)_{I}(x) of kTxM\wedge^{k}T_{x}^{*}M and of kgr(TxM)\wedge^{k}{\rm gr}(T_{x}^{*}M) respectively are orthonormal. We will use the following notation to denote the elements of the bases in degree 0 and nn:

1=Θ for 0TxM and volΘ(x):=θ1(x)θn(x) for nTxM1=\Theta^{\emptyset}\text{ for }\wedge^{0}T^{\ast}_{x}M\ \text{ and }{\rm vol}_{\Theta}(x):=\theta^{1}(x)\wedge\ldots\wedge\theta^{n}(x)\text{ for }\wedge^{n}T_{x}^{\ast}M

and

1=Θ for 0gr(TxM) and volΘ(x):=θ1(x)θn(x) for ngr(TxM).1=\langle\Theta\rangle^{\emptyset}\text{ for }\wedge^{0}{\rm gr}(T_{x}^{\ast}M)\text{ and }\langle{\rm vol}_{\Theta}\rangle(x):=\langle\theta^{1}\rangle(x)\wedge\ldots\wedge\langle\theta^{n}\rangle(x)\text{ for }\wedge^{n}{\rm gr}(T_{x}^{\ast}M)\,.

With these scalar products, we can now define transposes and obtain some of their properties: if TT and DD are algebraic maps acting linearly on gr(TM)\wedge^{\bullet}{\rm gr}(T^{*}M) and Ω(M)\Omega^{\bullet}(M) respectively, then we define their transpose TtT^{t} and DtD^{t} as the algebraic maps acting linearly on gr(TM)\wedge^{\bullet}{\rm gr}(T^{*}M) and Ω(M)\Omega^{\bullet}(M) by their pointwise transpose, that is, (Tt)x(T^{t})_{x} and (Dt)x(D^{t})_{x} are the transpose of TxT_{x} and DxD_{x} for the scalar products ggr(TxM)g_{\wedge^{\bullet}{\rm gr}(T_{x}^{*}M)} and gTxMg_{\wedge^{\bullet}T_{x}^{*}M} at each point xMx\in M. We check readily that if TT is an algebraic map acting linearly on gr(TM)\wedge^{\bullet}{\rm gr}(T^{*}M) respecting the gradation, then so is TtT^{t}. Moreover, the map D:=Φ1TΦD:=\Phi^{-1}\circ T\circ\Phi and its transpose DtD^{t} are both algebraic map acting linearly on Ω(M)\Omega^{\bullet}(M) respecting the gradation, and satisfy:

Dt=Φ1TtΦandgr(Dt)=(gr(D))t.D^{t}=\Phi^{-1}\circ T^{t}\circ\Phi\qquad\mbox{and}\qquad{\rm gr}(D^{t})=({\rm gr}(D))^{t}.

2.5.3. Hodge-star, scalar product on Ωc(M)\Omega_{c}^{\bullet}(M) and Γc(gr(TM))\Gamma_{c}(\wedge^{\bullet}{\rm gr}(T^{*}M)), and transpose

The Hodge star operator on TxM\wedge^{\bullet}T_{x}^{*}M or gr(TxM)\wedge^{\bullet}{\rm gr}(T_{x}^{*}M) is the map defined via

:kTxMnkTxM,αβ:=α,βvolΘ(x)α,βkTxM,k=1,,n,\star:\wedge^{k}T_{x}^{*}M\overset{\cong}{\longrightarrow}\wedge^{n-k}T_{x}^{*}M\,,\ \alpha\wedge\star\beta:=\langle\alpha,\beta\rangle{\rm vol}_{\Theta}(x)\ \forall\,\alpha,\beta\in\wedge^{k}T_{x}^{*}M,\ k=1,\ldots,n\,,

and similarly on gr(TxM)\wedge^{\bullet}{\rm gr}(T_{x}^{*}M). As Φx\Phi_{x} maps the scalar product of TxM\wedge^{\bullet}T_{x}^{*}M to the one of gr(TxM)\wedge^{\bullet}{\rm gr}(T_{x}^{*}M), (see (2.16)), it also maps the Hodge star operator of TxM\wedge^{\bullet}T_{x}^{*}M to the one of gr(TxM)\wedge^{\bullet}{\rm gr}(T_{x}^{*}M):

(2.17) α,βTxMΦx(α)=Φx(α).\forall\alpha,\beta\in\wedge^{\bullet}T_{x}^{*}M\qquad\star\Phi_{x}(\alpha)=\Phi_{x}(\star\alpha).

The \star operator naturally extends to an algebraic operator on Ω(M)\Omega^{\bullet}(M) and on Γ(gr(TM))\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M)) respectively. Moreover, from the definition of \star, we have that

αβ=βαonΩ(M)andΓ(gr(TM)),\alpha\wedge\star\beta=\beta\wedge\star\alpha\quad\mbox{on}\ \Omega^{\bullet}(M)\ \mbox{and}\ \Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M)),

and

1=volΘ,volΘ=1,and1=volΘ,volΘ=1.\star 1={\rm vol}_{\Theta},\ \star{\rm vol}_{\Theta}=1,\quad\mbox{and}\quad\star 1=\langle{\rm vol}_{\Theta}\rangle,\ \star\langle{\rm vol}_{\Theta}\rangle=1.

Moreover, if a local orthonormal adapted coframe Θ\Theta is fixed, the Hodge star operators on ΘI\Theta^{I} and ΘI\langle\Theta\rangle^{I} with multi-index I=(i1,,ik)I=(i_{1},\ldots,i_{k}) are given by

(ΘI)=(1)σ(I)ΘI¯,(ΘI)=(1)σ(I)ΘI¯\star(\Theta^{I})=(-1)^{\sigma(I)}\Theta^{\bar{I}},\qquad\star(\langle\Theta\rangle^{I})=(-1)^{\sigma(I)}\langle\Theta\rangle^{\bar{I}}

where I¯=(i¯1,,i¯nk)\bar{I}=(\bar{i}_{1},\ldots,\bar{i}_{n-k}) with i¯1<i¯2<<i¯nk\bar{i}_{1}<\bar{i}_{2}<\ldots<\bar{i}_{n-k} is the multi-index complementing II, and (1)σ(I)(-1)^{\sigma(I)} is the sign of the permutation σ(I)=i1iki¯1i¯nk\sigma(I)=i_{1}\cdots i_{k}\,\bar{i}_{1}\cdots\bar{i}_{n-k}. This completely characterises the Hodge star operators on both Ω(M)\Omega^{\bullet}(M) and on Γ(gr(TM))\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M)).

The previous properties imply readily that for any α,β\alpha,\beta in either Ωk(M)\Omega^{k}(M) or Γ(kgr(TM))\Gamma(\wedge^{k}{\rm gr}(T^{*}M)), we have

β=(1)k(nk)βandα,β=α,β.\star\star\beta=(-1)^{k(n-k)}\beta\qquad\mbox{and}\qquad\langle\star\alpha,\star\beta\rangle=\langle\alpha,\beta\rangle\,.

We can now define the so-called L2L^{2}-inner products on the subspaces Ωc(M)\Omega^{\bullet}_{c}(M) and Γc(gr(TM))\Gamma_{c}(\wedge^{\bullet}{\rm gr}(T^{*}M)) of forms with compact support in Ω(M)\Omega^{\bullet}(M) and Γ(gr(TM))\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M)) respectively via

α,βL2:=Mαβ=Mα,βvol,α,βΩc(M) or Γc(gr(TM)).\langle\alpha,\beta\rangle_{L^{2}}:=\int_{M}\alpha\wedge\star\beta=\int_{M}\langle\alpha,\beta\rangle\,{\rm vol}\ ,\ \forall\,\alpha,\beta\in\Omega^{\bullet}_{c}(M)\text{ or }\Gamma_{c}(\wedge^{\bullet}{\rm gr}(T^{*}M))\,.

We have Φ(Ωc(M))=Γc(gr(TM))\Phi(\Omega^{\bullet}_{c}(M))=\Gamma_{c}(\wedge^{\bullet}{\rm gr}(T^{*}M)) and

(2.18) α,βΩc(M)α,βL2=Φ(α),Φ(β)L2.\forall\alpha,\beta\in\Omega^{\bullet}_{c}(M)\qquad\langle\alpha,\beta\rangle_{L^{2}}=\langle\Phi(\alpha),\Phi(\beta)\rangle_{L^{2}}.

2.5.4. Properties of the transpose for the L2L^{2}-inner product

The above definitions of the Hodge operator and of the scalar product show that if DD is a differential operator on either Ω(M)\Omega^{\bullet}(M) or Γ(gr(TM))\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M)), then so are D\star D and DD\star, as well as the formal transpose DtD^{t} defined as Dα,βL2=α,DtβL2\langle D\alpha,\beta\rangle_{L^{2}}=\langle\alpha,D^{t}\beta\rangle_{L^{2}} for any αΩck(M)\alpha\in\Omega_{c}^{k}(M) or Γc(kgr(TM))\Gamma_{c}(\wedge^{k}{\rm gr}(T^{\ast}M)) and any βΩck+1(M)\beta\in\Omega^{k+1}_{c}(M) or Γc(k+1gr(TM))\Gamma_{c}(\wedge^{k+1}{\rm gr}(T^{\ast}M)).

In the particular case of the de Rham differential dd on forms, we observe that the transpose of d(k)=d:Ωk(M)Ωk+1(M)d^{(k)}=d:\Omega^{k}(M)\to\Omega^{k+1}(M) is a differential operator that satisfies

d(k,t)=(1)kn+1d(nk1):Ωk+1(M)Ωk(M).d^{(k,t)}=(-1)^{kn+1}\star d^{(n-k-1)}\star\colon\Omega^{k+1}(M)\to\Omega^{k}(M)\,.

We can also consider the transpose of the algebraic linear map d~(k)=d~:Ωk(M)Ωk+1(M)\tilde{d}^{(k)}=\tilde{d}:\Omega^{k}(M)\to\Omega^{k+1}(M) of Subsection 2.2.1. The linear map d~t\tilde{d}^{t} is then algebraic and satisfies

d~(k,t)=(1)kn+1d~(nk1):Ωk+1(M)Ωk(M).\tilde{d}^{(k,t)}=(-1)^{kn+1}\star\tilde{d}^{(n-k-1)}\star\colon\Omega^{k+1}(M)\to\Omega^{k}(M)\,.

It is important to realise that the transpose may not respect the filtration in the following sense: for a general differential operator D:Ω(M)Ω(M)D\colon\Omega^{\bullet}(M)\to\Omega^{\bullet}(M), it is true that the transpose DtD^{t} is a differential operator, but DD respecting the filtration (2.11) will not imply always that DtD^{t} does so. This may not be true even for dd and d~\tilde{d}, and two explicit ‘counter-examples’ of this situation are given in Appendix B. However, the following lemma guarantees that DtD^{t} will also respect the filtration when DD is of the form D=TΦ=Φ1TΦD=T^{\Phi}=\Phi^{-1}\circ T\circ\Phi with TT a homomorphism of the C(M)C^{\infty}(M)-module Γ(gr(TM))\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M)):

Lemma 2.20.

We continue with the setting of Lemma 2.19, where T:Γ(gr(TM))Γ(gr(TM))T:\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M))\to\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M)) is a morphism of C(M)C^{\infty}(M)-modules. We consider the map D:=TΦ=Φ1TΦD:=T^{\Phi}=\Phi^{-1}\circ T\circ\Phi acting on Ω(M)\Omega^{\bullet}(M). Let us further assume that TT respects the gradation.

  1. (1)

    The C(M)C^{\infty}(M)-module morphism T:Γ(gr(TM))Γ(gr(TM))\star T\star\colon\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M))\to\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M)) also respects the gradation, and (T)Φ=TΦ.(\star T\star)^{\Phi}=\star T^{\Phi}\star. Consequently, the C(M)C^{\infty}(M)-module morphism D:=TΦ:Ω(M)Ω(M)D:=T^{\Phi}:\Omega^{\bullet}(M)\to\Omega^{\bullet}(M) respects the filtration and we have gr(D)=gr(D){\rm gr}(\star D\star)=\star{\rm gr}(D)\star.

  2. (2)

    If in addition, TT is a differential operator, then TtT^{t} is also a differential operator of the same order as TT, it respects the gradation, and satisfies (Tt)Φ=(TΦ)t(T^{t})^{\Phi}=(T^{\Phi})^{t}. Consequently, the differential operators on smooth forms D:=TΦD:=T^{\Phi} and its transpose DtD^{t} respect the filtration, and gr(Dt)=(gr(D))t.{\rm gr}(D^{t})=({\rm gr}(D))^{t}.

Proof.

The equality (T)Φ=TΦ(\star T\star)^{\Phi}=\star T^{\Phi}\star follows from (2.17).

Let Θ\Theta be a local adapted frame. We observe that ΘI\langle\Theta\rangle^{I} is of weight |I||I| while ΘI=(1)σ(I)ΘI¯\star\langle\Theta\rangle^{I}=(-1)^{\sigma(I)}\langle\Theta\rangle^{\bar{I}} is of weight |I¯|=Q|I||\bar{I}|=Q-|I| where Q:=υ1++υnQ:=\upsilon_{1}+\ldots+\upsilon_{n} is the weight of the volume form.

Since TT respects the gradation and the weight of ΘI¯\langle\Theta\rangle^{\bar{I}} is |I¯||\bar{I}|, T(ΘI¯)T(\langle\Theta\rangle^{\bar{I}}) has the same weight |I¯||\bar{I}|. Consequently, T(ΘI¯)T(\langle\Theta\rangle^{\bar{I}}) is a C(M)C^{\infty}(M)-linear combination of basis elements ΘJ\langle\Theta\rangle^{J} where the multi-indices JJ have weight |J|=|I¯|=Q|I||J|=|\bar{I}|=Q-|I|. Finally, T(ΘI¯)\star T(\Theta^{\bar{I}}) has weight Q|J|=|I|Q-|J|=|I|. We have therefore shown that the morphism T\star T\star maps elements of weight |I||I| to elements of weight |I||I|, and so T\star T\star respects the gradation. Lemma 2.19 implies the rest of Part (1).

Part (2) follows from the definition of DtD^{t} and its link with \star covered in Subsection 2.5.3. ∎

2.5.5. Properties of gr{\rm gr} for algebraic maps

Here, we describe some properties of gr{\rm gr} in relation with algebraic maps (for the latter concept see Section 2.3.2).

The following properties are easily checked.

Lemma 2.21.

Let MM be a smooth manifold whose tangent bundle is filtered by vector subbundles as in (2.9).

  1. (1)

    Any algebraic morphism of C(M)C^{\infty}(M)-module DD acting either on Ω(M)\Omega^{\bullet}(M) or Γ(gr(TM))\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M)) is determined by its pointwise restriction DxD_{x} at each xMx\in M. Conversely, any smooth map on TM\wedge^{\bullet}T^{*}M or gr(TM)\wedge^{\bullet}{\rm gr}(T^{*}M) that is linear at each fibre yields an algebraic morphism of C(M)C^{\infty}(M)-modules.

  2. (2)

    If a morphism of C(M)C^{\infty}(M)-modules DD acting on Ω(M)\Omega^{\bullet}(M) is algebraic and respects the filtration, then gr(D){\rm gr}(D) is a morphism of C(M)C^{\infty}(M)-module acting on Γ(gr(TM))\Gamma(\wedge^{\bullet}{\rm gr}(T^{*}M)) which is also an algebraic map that respects the gradation. Moreover, if DD is invertible, then gr(D){\rm gr}(D) is also invertible, and D1D^{-1} and gr(D)1{\rm gr}(D)^{-1} are algebraic.

The isomorphism Φ\Phi 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 MM be a smooth manifold whose tangent bundle is filtered by vector subbundles as in (2.9). Assume that MM is equipped with a metric gTMg_{TM}. We consider the corresponding map Φ\Phi defined in Section 2.5.1.

Let D:Ω(M)Ω(M)D:\Omega^{\bullet}(M)\to\Omega^{\bullet}(M) be a morphism of C(M)C^{\infty}(M)-modules that respects the filtration and such that gr(D){\rm gr}(D) is algebraic. Assume also that gr(D){\rm gr}(D) is fiberwise invertible in the sense that its restriction gr(D)x{\rm gr}(D)_{x} at each xMx\in M is a linear invertible map on TxM\wedge^{\bullet}T_{x}^{*}M. Then the map DD is invertible and the map D1:Ω(M)Ω(M)D^{-1}:\Omega^{\bullet}(M)\to\Omega^{\bullet}(M) respects the filtration. Moreover, gr(D1)=gr(D)1{\rm gr}(D^{-1})={\rm gr}(D)^{-1}, and if DD is in addition a differential operator acting on Ω(M)\Omega^{\bullet}(M), then so is its inverse.

Proof.

By (2) in Lemma 2.19, the map T:=(gr(D))Φ:Ω(M)Ω(M)T:=({\rm gr}(D))^{\Phi}\colon\Omega^{\bullet}(M)\to\Omega^{\bullet}(M) respects the filtration and gr(T)1=gr(D)1{\rm gr}(T)^{-1}={\rm gr}(D)^{-1}. Moreover, by (2) of Lemma 2.21, TT is invertible and its inverse is T1=(gr(D)1)ΦT^{-1}=({\rm gr}(D)^{-1})^{\Phi}.

By Lemma 2.16, we have

gr(T1D)=gr(T1)gr(D)=gr(D)1gr(D)=I, so gr(T1DI)=gr(T1D)I=0.{\rm gr}(T^{-1}D)={\rm gr}(T^{-1}){\rm gr}(D)={\rm gr}(D)^{-1}{\rm gr}(D)=\text{\rm I}\ ,\mbox{ so }{\rm gr}(T^{-1}D-\text{\rm I})={\rm gr}(T^{-1}D)-\text{\rm I}=0\ .

Hence T1DI=:BT^{-1}D-\text{\rm I}=:B is nilpotent with BN0=0B^{N_{0}}=0 by (2.14). Since T1D=I+BT^{-1}D=\text{\rm I}+B and I+B\text{\rm I}+B is invertible with (I+B)1=j=0N0(1)jBj(\text{\rm I}+B)^{-1}=\sum_{j=0}^{N_{0}}(-1)^{j}B^{j}, we have (I+B)1T1D=I(\text{\rm I}+B)^{-1}T^{-1}D=\text{\rm I}. This provides a left inverse for DD and the formula below with B=BLB=B_{L}. We can do the same on the right. ∎

Corollary 2.23.

We continue with the setting of Proposition 2.22 and its proof. The following formulae holds for D1:Ω(M)Ω(M)D^{-1}:\Omega^{\bullet}(M)\to\Omega^{\bullet}(M):

D1=j=0N0(1)jBLjT1=T1j=0N0(1)jBRjD^{-1}=\sum_{j=0}^{N_{0}}(-1)^{j}B_{L}^{j}T^{-1}=T^{-1}\sum_{j=0}^{N_{0}}(-1)^{j}B_{R}^{j}

where T=gr(D)Φ=Φgr(D)1Φ1T={\rm gr}(D)^{\Phi}=\Phi{\rm gr}(D)^{-1}\Phi^{-1}, and BL=T1DIB_{L}=T^{-1}D-\text{\rm I} while BR=DT1IB_{R}=DT^{-1}-\text{\rm I}.

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 d𝔤Md_{\mathfrak{g}M} and study its relation to the de Rham complex (Ω(M),d)(\Omega^{\bullet}(M),d). We will also use the algebraic part d~\tilde{d} of dd. The main result of this section is in Proposition 3.3 stating that gr(d)=gr(d~)=d𝔤M{\rm gr}(d)={\rm gr}(\tilde{d})=d_{\mathfrak{g}M}.

3.1. Setting and definitions

A filtered manifold is a smooth manifold MM equipped with a filtration of the tangent bundle TMTM by vector subbundles

M×{0}=H0H1Hs=TMM\times\{0\}=H^{0}\subseteq H^{1}\subseteq\ldots\subseteq H^{s}=TM

satisfying

(3.1) [Γ(Hi),Γ(Hj)]Γ(Hi+j),[\Gamma(H^{i}),\Gamma(H^{j})]\subseteq\Gamma(H^{i+j})\,,

with the convention that Hi=TMH^{i}=TM when i>si>s.

Example 3.1.

Any contact manifold, or more generally any regular subRiemannian manifold, is a filtered manifold, see Section 4.6.1. A related class of examples is given by nilpotent Lie groups equipped with a left-invariant filtration [LDT22].

3.1.1. Bundle of osculating Lie groups and algebras

For each xMx\in M, the quotient gr(TxM)=i=1s0(Hxi/Hxi1){\rm gr}(T_{x}M)=\oplus_{i=1}^{s_{0}}\left(H^{i}_{x}/H^{i-1}_{x}\right) is naturally equipped with a Lie bracket [,]𝔤xM[\cdot,\cdot]_{\mathfrak{g}_{x}M}, since

f,gC(M),XΓ(Hi),YΓ(Hj)[fX,gY]=fg[X,Y]+Γ(Hi+j1).\forall\,f,g\in C^{\infty}(M),\ X\in\Gamma(H^{i}),\ Y\in\Gamma(H^{j})\qquad[fX,gY]=fg[X,Y]+\Gamma(H^{i+j-1}).

When gr(TxM){\rm gr}(T_{x}M) is equipped with this Lie bracket, we denote the resulting Lie algebra as

𝔤xM:=(gr(TxM),[,]𝔤xM).\mathfrak{g}_{x}M:=({\rm gr}(T_{x}M),[\cdot,\cdot]_{\mathfrak{g}_{x}M})\,.

It is naturally graded by

𝔤xM=i=1s0(Hxi/Hxi1),\mathfrak{g}_{x}M=\oplus_{i=1}^{s_{0}}\left(H^{i}_{x}/H^{i-1}_{x}\right),

and is therefore nilpotent. We denote by GxMG_{x}M the corresponding connected simply connected nilpotent Lie group (sometimes called the nilpotentisation or the tangent cone [Mit85] of MM at xx). The unions

GM:=xMGxMand𝔤M:=xM𝔤xM,GM:=\cup_{x\in M}G_{x}M\qquad\mbox{and}\qquad\mathfrak{g}M:=\cup_{x\in M}\mathfrak{g}_{x}M,

are naturally equipped with a smooth bundle structure that are called [RS76, vEY17] the bundles of osculating Lie groups and Lie algebras over MM.

We observe that the notions of weights defined in Section 2.4.3 and for the graded Lie algebra 𝔤xM\mathfrak{g}_{x}M coincide. We therefore obtain an analogous decomposition of 𝔤xM\wedge^{\bullet}\mathfrak{g}_{x}^{*}M in terms of weights and degrees:

(3.2) 𝔤xM=k,w0k,w𝔤xM,xM.\wedge^{\bullet}\mathfrak{g}_{x}^{*}M=\oplus_{k,w\in\mathbb{N}_{0}}\wedge^{k,w}\mathfrak{g}_{x}^{*}M\ ,\quad\forall\,x\in M\ .

We introduce the following vocabulary.

Definition 3.2.

The elements of GΩ(M):=Γ(𝔤M)G\Omega^{\bullet}(M):=\Gamma(\wedge^{\bullet}\mathfrak{g}^{\ast}M) are called osculating forms. Moreover, for any k,w0k,w\in\mathbb{N}_{0},

GΩk(M):=Γ(k𝔤M)andGΩk,w(M):=Γ(k,w𝔤M)G\Omega^{k}(M):=\Gamma(\wedge^{k}\mathfrak{g}^{*}M)\qquad\mbox{and}\qquad G\Omega^{k,w}(M):=\Gamma(\wedge^{k,w}\mathfrak{g}^{*}M)

are called osculating kk-forms (or osculating forms of degree kk) and osculating kk-forms of weights ww respectively.

We say that a map T:GΩ(M)GΩ(M)T\colon G\Omega^{\bullet}(M)\to G\Omega^{\bullet}(M) respects the weights of osculating forms when T(GΩ,w(M))GΩ,w(M)T(G\Omega^{\bullet,w}(M))\subseteq G\Omega^{\bullet,w}(M), and it increases the weights when T(GΩ,w(M))wwGΩ,w(M)T(G\Omega^{\bullet,w}(M))\subseteq\oplus_{w^{\prime}\geq w}G\Omega^{\bullet,w^{\prime}}(M).

3.1.2. The osculating Chevalley-Eilenberg differential

For each xMx\in M, d𝔤xMd_{\mathfrak{g}_{x}M} denotes the Chevalley-Eilenberg differential on the Lie group GxMG_{x}M viewed as the map

d𝔤xM:𝔤xM𝔤xMd_{\mathfrak{g}_{x}M}:\wedge^{\bullet}\mathfrak{g}_{x}^{*}M\to\wedge^{\bullet}\mathfrak{g}_{x}^{*}M

defined via d𝔤xM(0𝔤xM)={0}d_{\mathfrak{g}_{x}M}(\wedge^{0}\mathfrak{g}_{x}^{*}M)=\{0\} for k=0k=0, for k=1k=1

(3.3) α1𝔤xM,V0,V1𝔤xM,d𝔤xMα(V0,V1)=α([V0,V1]𝔤xM),\forall\,\alpha\in\wedge^{1}\mathfrak{g}_{x}^{*}M\,,\ V_{0},V_{1}\in\mathfrak{g}_{x}M\ ,\quad d_{\mathfrak{g}_{x}M}\alpha(V_{0},V_{1})=-\alpha([V_{0},V_{1}]_{\mathfrak{g}_{x}M}),

and more generally for k>0k>0, for any αk𝔤xM\alpha\in\wedge^{k}\mathfrak{g}_{x}^{*}M and V0,Vk𝔤xMV_{0},\ldots V_{k}\in\mathfrak{g}_{x}M,

(3.4) d𝔤xMα(V0,,Vk)=0i<jk(1)i+jα([Vi,Vj]𝔤xM,V0,,V^i,,V^j,,Vk).d_{\mathfrak{g}_{x}M}\alpha(V_{0},\ldots,V_{k})=\sum_{0\leq i<j\leq k}(-1)^{i+j}\alpha([V_{i},V_{j}]_{\mathfrak{g}_{x}M},V_{0},\ldots,\hat{V}_{i},\ldots,\hat{V}_{j},\ldots,V_{k}).

It is well known [FT23] that d𝔤xMd_{\mathfrak{g}_{x}M} is a linear map such that

d𝔤xM2=0,d𝔤xM(k𝔤xM)k+1𝔤xM,d_{\mathfrak{g}_{x}M}^{2}=0\ ,\quad d_{\mathfrak{g}_{x}M}(\wedge^{k}\mathfrak{g}_{x}^{*}M)\subseteq\wedge^{k+1}\mathfrak{g}_{x}^{*}M,

and that it satisfies the Leibniz property on 𝔤xM\wedge^{\bullet}\mathfrak{g}_{x}^{*}M, i.e.

αk𝔤xM,β𝔤xM,d𝔤xM(αβ)=(d𝔤xMα)β+(1)kα(d𝔤xMβ).\forall\,\alpha\in\wedge^{k}\mathfrak{g}_{x}^{*}M,\ \beta\in\wedge^{\bullet}\mathfrak{g}_{x}^{*}M\ ,\quad d_{\mathfrak{g}_{x}M}(\alpha\wedge\beta)=(d_{\mathfrak{g}_{x}M}\alpha)\wedge\beta+(-1)^{k}\alpha\wedge(d_{\mathfrak{g}_{x}M}\beta).

Moreover, d𝔤xMd_{\mathfrak{g}_{x}M} also preserves the weights of 𝔤xM\wedge^{\bullet}\mathfrak{g}_{x}^{*}M:

d𝔤xM(k,w𝔤xM)k+1,w𝔤xM.d_{\mathfrak{g}_{x}M}(\wedge^{k,w}\mathfrak{g}_{x}^{*}M)\subseteq\wedge^{k+1,w}\mathfrak{g}_{x}^{*}M.

By construction, the algebraic map d𝔤Md_{\mathfrak{g}M} defined by

d𝔤M:GΩ(M)GΩ(M)(d𝔤M)x=d𝔤xM,xM,d_{\mathfrak{g}M}:G\Omega^{\bullet}(M)\to G\Omega^{\bullet}(M)\qquad(d_{\mathfrak{g}M})_{x}=d_{\mathfrak{g}_{x}M},\ \forall\,x\in M,

is smooth. We call (GΩ(M),d𝔤M)(G\Omega^{\bullet}(M),d_{\mathfrak{g}M}) the osculating complex or osculating Chevalley-Eilenberg differential.

3.1.3. The osculating box operator

Here we assume that the vector bundle 𝔤M\wedge^{\bullet}\mathfrak{g}^{*}M is equipped with a metric g𝔤Mg_{\wedge^{\bullet}\mathfrak{g}^{*}M}: each 𝔤xM\wedge^{\bullet}\mathfrak{g}_{x}^{*}M is equipped with a scalar product g𝔤xMg_{\wedge^{\bullet}\mathfrak{g}_{x}^{*}M} with a smooth dependence in xMx\in M. This scalar product allows us to consider, at each point xMx\in M, the transpose of the osculating differential d𝔤Md_{\mathfrak{g}M}, and then to define

𝔤M:=d𝔤Md𝔤Mt+d𝔤Mtd𝔤M.\Box_{\mathfrak{g}M}:=d_{\mathfrak{g}M}d_{\mathfrak{g}M}^{t}+d_{\mathfrak{g}M}^{t}d_{\mathfrak{g}M}\,.

We call this operator the osculating box. As d𝔤Md_{\mathfrak{g}_{M}} is algebraic, so are d𝔤Mtd_{\mathfrak{g}M}^{t} and 𝔤M\Box_{\mathfrak{g}M}.

Let us now assume that the decomposition (3.2) is orthogonal for the scalar product g𝔤xMg_{\wedge^{\bullet}\mathfrak{g}_{x}^{*}M} for each xMx\in M. This arises naturally when the manifold is filtered and Riemannian (see Section 2.5). With this assumption, we readily check that d𝔤Mtd_{\mathfrak{g}M}^{t} preserves the weights, i.e. d𝔤xMt(k,w𝔤xM)k1,w𝔤xMd_{\mathfrak{g}_{x}M}^{t}(\wedge^{k,w}\mathfrak{g}_{x}^{*}M)\subseteq\wedge^{k-1,w}\mathfrak{g}_{x}^{*}M, and 𝔤M\Box_{\mathfrak{g}M} respects the degree and weights of osculating forms:

𝔤xM(k,w𝔤xM)k,w𝔤xM.\Box_{\mathfrak{g}_{x}M}(\wedge^{k,w}\mathfrak{g}_{x}^{*}M)\subseteq\wedge^{k,w}\mathfrak{g}_{x}^{*}M.

Moreover, given the scalar product g𝔤xMg_{\wedge^{\bullet}\mathfrak{g}_{x}^{*}M}, the linear map

𝔤xM:=(𝔤M)x=d𝔤xMd𝔤xMt+d𝔤xMtd𝔤xM:𝔤xM𝔤xM\Box_{\mathfrak{g}_{x}M}:=(\Box_{\mathfrak{g}M})_{x}=d_{\mathfrak{g}_{x}M}d_{\mathfrak{g}_{x}M}^{t}+d_{\mathfrak{g}_{x}M}^{t}d_{\mathfrak{g}_{x}M}\colon\wedge^{\bullet}\mathfrak{g}_{x}^{*}M\to\wedge^{\bullet}\mathfrak{g}_{x}^{*}M

is symmetric. We denote by Π𝔤xM\Pi_{\mathfrak{g}_{x}M} the spectral orthogonal projection onto its kernel

E𝔤xM:=ker(𝔤xM)=kerd𝔤xMkerd𝔤xMt=ImΠ𝔤xM.E_{\mathfrak{g}_{x}M}:=\ker(\Box_{\mathfrak{g}_{x}M})=\ker d_{\mathfrak{g}_{x}M}\cap\ker d_{\mathfrak{g}_{x}M}^{t}={\rm Im}\,\Pi_{\mathfrak{g}_{x}M}.

Note that the image of 𝔤xM\Box_{\mathfrak{g}_{x}M} is

F𝔤xM:=Im(𝔤xM)=Imd𝔤xM+Imd𝔤xMt=kerΠ𝔤xM=E𝔤xM.F_{\mathfrak{g}_{x}M}:={\rm Im}\,(\Box_{\mathfrak{g}_{x}M})={\rm Im}\,d_{\mathfrak{g}_{x}M}+{\rm Im}\,d_{\mathfrak{g}_{x}M}^{t}=\ker\Pi_{\mathfrak{g}_{x}M}=E_{\mathfrak{g}_{x}M}^{\perp}.

By [FT23, Lemma 2.1], the complex (F𝔤M,d𝔤M)(F_{\mathfrak{g}M}^{\bullet},d_{\mathfrak{g}M}) is acyclic, i.e. d𝔤M(F𝔤M)F𝔤Md_{\mathfrak{g}M}(F_{\mathfrak{g}M})\subseteq F_{\mathfrak{g}M}, and the resulting cohomology is trivial:

d𝔤xM(F𝔤xM)=ker(d𝔤xM:F𝔤xMF𝔤xM)𝔤xM.d_{\mathfrak{g}_{x}M}(F_{\mathfrak{g}_{x}M})=\ker(d_{\mathfrak{g}_{x}M}:F_{\mathfrak{g}_{x}M}\to F_{\mathfrak{g}_{x}M})\subseteq\wedge^{\bullet}\mathfrak{g}_{x}^{*}M.

For |z|=ε|z|=\varepsilon small enough, z𝔤xMz-\Box_{\mathfrak{g}_{x}M} is invertible, and by the Cauchy residue formula we have

Π𝔤xM=12πi|z|=ε(z𝔤xM)1𝑑z,\Pi_{\mathfrak{g}_{x}M}=\frac{1}{2\pi i}\oint_{|z|=\varepsilon}(z-\Box_{\mathfrak{g}_{x}M})^{-1}dz,

where the contour integration is over a circle about 0 of radius ε>0\varepsilon>0 small enough. This defines a smooth algebraic map Π𝔤M\Pi_{\mathfrak{g}M} acting on GΩ(M)G\Omega^{\bullet}(M) that respects degrees and weights.

3.1.4. The partial inverse d𝔤M1d^{-1}_{\mathfrak{g}M}

Proceeding as in [FT23], at each xMx\in M, a partial inverse d𝔤xM1d_{\mathfrak{g}_{x}M}^{-1} of d𝔤xMd_{\mathfrak{g}_{x}M} can be defined as

d𝔤xM1:=(d𝔤xM)1prImd𝔤xM,d_{\mathfrak{g}_{x}M}^{-1}:=(d_{\mathfrak{g}_{x}M})^{-1}{\rm pr}_{{\rm Im}\,d_{\mathfrak{g}_{x}M}}\,,

where we write prS{\rm pr}_{S} to denote the orthogonal projection onto any SS closed subspace of 𝔤xM\wedge^{\bullet}\mathfrak{g}_{x}^{*}M.

Note that in our construction in Section 4, we will not use d𝔤xM1d_{\mathfrak{g}_{x}M}^{-1}. However, it is used in Rumin’s construction, see Section 4.5.

where we show that the two constructions coincide.

The resulting map d𝔤M1d_{\mathfrak{g}M}^{-1} on GΩ(M)G\Omega(M) is a smooth map that respects the filtration, and such that

(d𝔤M1)2=0 and d𝔤xM1(k+1,w𝔤xM)k,w𝔤xM,k,w0.(d_{\mathfrak{g}M}^{-1})^{2}=0\ \text{ and }\,d_{\mathfrak{g}_{x}M}^{-1}(\wedge^{k+1,w}\mathfrak{g}_{x}^{*}M)\subseteq\wedge^{k,w}\mathfrak{g}_{x}^{*}M\ ,\ \forall\,k,w\in\mathbb{N}_{0}\,.

By [FT23, Section 2], the kernel and image of d𝔤M1d_{\mathfrak{g}M}^{-1} coincide with the one of d𝔤Mtd_{\mathfrak{g}M}^{t}:

kerd𝔤xMt=kerd𝔤xM1andImd𝔤xMt=Imd𝔤xM1,\ker d_{\mathfrak{g}_{x}M}^{t}=\ker d_{\mathfrak{g}_{x}M}^{-1}\quad\mbox{and}\quad{\rm Im}\,d_{\mathfrak{g}_{x}M}^{t}={\rm Im}\,d_{\mathfrak{g}_{x}M}^{-1},

so

E𝔤xM=kerd𝔤xMkerd𝔤xM1andF𝔤xM=Imd𝔤xM+Imd𝔤xM1.E_{\mathfrak{g}_{x}M}=\ker d_{\mathfrak{g}_{x}M}\cap\ker d_{\mathfrak{g}_{x}M}^{-1}\quad\mbox{and}\quad F_{\mathfrak{g}_{x}M}={\rm Im}\,d_{\mathfrak{g}_{x}M}+{\rm Im}\,d_{\mathfrak{g}_{x}M}^{-1}.

By [FT23, Proposition 2.2],

Π𝔤M=Id𝔤M1d𝔤Md𝔤Md𝔤M1\Pi_{\mathfrak{g}M}=\text{\rm I}-d_{\mathfrak{g}M}^{-1}d_{\mathfrak{g}M}-d_{\mathfrak{g}M}d_{\mathfrak{g}M}^{-1}

and we have

d𝔤M1Π𝔤M=Π𝔤Md𝔤M1=0,d𝔤MΠ𝔤M=Π𝔤Md𝔤M=0.d_{\mathfrak{g}M}^{-1}\Pi_{\mathfrak{g}M}=\Pi_{\mathfrak{g}M}d_{\mathfrak{g}M}^{-1}=0\ ,\quad d_{\mathfrak{g}M}\Pi_{\mathfrak{g}M}=\Pi_{\mathfrak{g}M}d_{\mathfrak{g}M}=0\,.

3.2. The maps dd and d~\tilde{d} on a filtered manifold

Let us show that gr(d){\rm gr}(d) and gr(d~){\rm gr}(\tilde{d}) are well-defined, crucial operators, and that both (Γ(gr(TM)),gr(d))(\Gamma(\wedge^{\bullet}{\rm gr}(T^{\ast}M)),{\rm gr}(d)) and (Γ(gr(TM)),gr(d~))(\Gamma(\wedge^{\bullet}{\rm gr}(T^{\ast}M)),{\rm gr}(\tilde{d})) coincide with the osculating complex.

Proposition 3.3.

The maps dd and d~\tilde{d} respect the filtration with

(dd~)Ωk,w(M)Ωk,w+1(M),(d-\tilde{d})\Omega^{k,\geq w}(M)\subseteq\Omega^{k,\geq w+1}(M),

for any degree k0k\in\mathbb{N}_{0} and weight w0w\in\mathbb{N}_{0}. Moreover, the maps gr(d){\rm gr}(d) and gr(d~){\rm gr}(\tilde{d}) acting on GΩ(M)Γ(gr(TM))G\Omega^{\bullet}(M)\cong\Gamma(\wedge^{\bullet}{\rm gr}(T^{\ast}M)) coincide with d𝔤Md_{\mathfrak{g}M}:

gr(d)=gr(d~)=d𝔤M.{\rm gr}(d)={\rm gr}(\tilde{d})=d_{\mathfrak{g}M}\,.
Proof.

From the definitions of dd and d~\tilde{d}, both differentials as maps Ω1(M)Ω2(M)\Omega^{1}(M)\to\Omega^{2}(M) respect the corresponding filtrations. As they obey the Leibniz rule, both dd and d~\tilde{d} respect the filtration on Ω(M)\Omega^{\bullet}(M). Moreover, gr(d){\rm gr}(d) and gr(d~){\rm gr}(\tilde{d}) satisfy the Leibniz property on GΩ(M)Γ(gr(TM))G\Omega^{\bullet}(M)\cong\Gamma(\wedge^{\bullet}{\rm gr}(T^{\ast}M)).

Let ωΩk,w(M)\omega\in\Omega^{k,\geq w}(M) and for each j=0,,kj=0,\ldots,k take VjHwi(j)V_{j}\in H^{w_{i(j)}} with wi(0)++wi(k)ww_{i(0)}+\ldots+w_{i(k)}\leq w. Since wi(0)++w^i(l)++wi(k)<ww_{i(0)}+\ldots+\hat{w}_{i(l)}+\ldots+w_{i(k)}<w, we have that ω(V0,,V^l,,Vk)=0\omega(V_{0},\ldots,\hat{V}_{l},\ldots,V_{k})=0, and so (dd~)ω(V0,,Vk)=0(d-\tilde{d})\omega(V_{0},\ldots,V_{k})=0 by (2.8).

This shows (dd~)Ωk,w(M)Ωk,w+1(M)(d-\tilde{d})\Omega^{k,\geq w}(M)\subseteq\Omega^{k,\geq w+1}(M), and implies gr(d)=gr(d~){\rm gr}(d)={\rm gr}(\tilde{d}). We are left to prove that gr(d~)=d𝔤M{\rm gr}(\tilde{d})=d_{\mathfrak{g}M}.

If ωΩ1(M)\omega\in\Omega^{1}(M) and Vi1Γ(Hwi(1)),Vi2Γ(Hwi(2))V_{i_{1}}\in\Gamma(H^{w_{i(1)}}),V_{i_{2}}\in\Gamma(H^{w_{i(2)}}), then by Lemma 2.3 and (3.1).

d~ω(V1modHwi(1)1,V2modHwi(2)1)\displaystyle\tilde{d}\omega(V_{1}\,{\rm mod}\,H^{w_{i(1)}-1},V_{2}\,{\rm mod}\,H^{w_{i(2)}-1}) =ω([V1modHwi(1)1,V2modHwi(2)1])\displaystyle=-\omega([V_{1}\,{\rm mod}\,H^{w_{i(1)}-1},V_{2}\,{\rm mod}\,H^{w_{i(2)}-1}])
=ω([V1,V2]modHwi(1)+wi(2)1).\displaystyle=-\omega([V_{1},V_{2}]\,{\rm mod}\,H^{w_{i(1)}+w_{i(2)}-1})\,.

Hence, we have that for any ω(Hwj)\omega\in(H^{w_{j}})^{\perp}, j=0,1,,s01j=0,1,\ldots,s_{0}-1

gr(d~)(ωmod(Hwj+1))=(ωmod(Hwj+1))([,]𝔤M).{\rm gr}(\tilde{d})(\omega\,{\rm mod}\,(H^{w_{j+1}})^{\perp})=-(\omega\,{\rm mod}\,(H^{w_{j+1}})^{\perp})([\cdot,\cdot]_{\mathfrak{g}M}).

We recognise d𝔤M(ωmod(Hwj+1))d_{\mathfrak{g}M}(\omega\,{\rm mod}\,(H^{w_{j+1}})^{\perp}) from (3.3). Hence the maps d𝔤Md_{\mathfrak{g}_{M}} and gr(d~){\rm gr}(\tilde{d}) coincide on GΩ1(M)G\Omega^{1}(M). Since they both vanish on GΩ0(M)G\Omega^{0}(M) and satisfy the Leibniz property, they coincide in fact on the whole GΩ(M)G\Omega^{\bullet}(M). ∎

4. Construction of subcomplexes on Riemannian filtered manifolds

Here, we present the general scheme to construct subcomplexes on a filtered manifold MM whose tangent bundle is equipped with a metric gTMg_{TM}. 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 ,P,L\Box,P,L which allow us to define the subcomplexes (F0,C)(F_{0}^{\bullet},C) and (E0,D)(E_{0}^{\bullet},D), 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 ~,P~,L~\tilde{\Box},\tilde{P},\tilde{L} and then the subcomplexes (F~0,C~)(\tilde{F}_{0}^{\bullet},\tilde{C}) and (E~0,D~)(\tilde{E}_{0}^{\bullet},\tilde{D}), with the latter computing the same cohomology as the complex (Ω(M),d~)(\Omega^{\bullet}(M),\tilde{d}). We present an alternative construction for (E0,D)(E_{0}^{\bullet},D) and (E~0,D~)(\tilde{E}_{0}^{\bullet},\tilde{D}) in Section 4.5 that follows Rumin’s ideas.

4.1. Base differential and codifferential d0d_{0} and δ0\delta_{0}

Definition 4.1.

Let d0d_{0} and δ0\delta_{0} be two differential algebraic complexes on a filtered manifold MM equipped with a metric gTMg_{TM}. We say that (d0,δ0)(d_{0},\delta_{0}) is a pair of base differential and codifferential when the following properties are satisfied:

  • d0d_{0} increases the degree by one while δ0\delta_{0} decreases the degree by one, and both respect the filtration of Ω(M)\Omega^{\bullet}(M):

    d0(Ωk,w(M))Ωk+1,w(M)andδ0(Ωk+1,w(M))Ωk,w(M),d_{0}(\Omega^{k,\geq w}(M))\subset\Omega^{k+1,\geq w}(M)\quad\mbox{and}\quad\delta_{0}(\Omega^{k+1,\geq w}(M))\subset\Omega^{k,\geq w}(M),
  • d0d_{0} and δ0\delta_{0} are disjoint [Kos61], that is, for any αΩ(M)\alpha\in\Omega^{\bullet}(M)

    δ0d0α=0d0α=0 and d0δ0α=0δ0α=0,\displaystyle\delta_{0}d_{0}\alpha=0\Longrightarrow d_{0}\alpha=0\ \text{ and }\ d_{0}\delta_{0}\alpha=0\Longrightarrow\delta_{0}\alpha=0\,,
  • gr(d0)=d𝔤M{\rm gr}(d_{0})=d_{\mathfrak{g}M} and gr(δ0)=d𝔤Mt{\rm gr}(\delta_{0})=d_{\mathfrak{g}M}^{t}.

Above, the transpose is defined using a metric g𝔤Mg_{\wedge^{\bullet}\mathfrak{g}^{*}M} on the osculating bundle 𝔤M\wedge^{\bullet}\mathfrak{g}^{*}M induced by gTMg_{TM}.

Given a pair (d0,δ0)(d_{0},\delta_{0}) of base differential and codifferential on a filtered manifold MM, we define the associated base box via

0:=d0δ0+δ0d0.\Box_{0}:=d_{0}\,\delta_{0}+\delta_{0}\,d_{0}.

For any xMx\in M, we define the operator

Π0,x:=12πi|z|=ϵ(z0,x)1𝑑z,\Pi_{0,x}:=\frac{1}{2\pi i}\oint_{|z|=\epsilon}(z-\Box_{0,x})^{-1}dz,

for ε>0\varepsilon>0 small enough.

Proposition 4.2.
  1. (1)

    The operators 0\Box_{0} and Π0\Pi_{0} are smooth and algebraic on Ω(M)\Omega^{\bullet}(M). They respect its filtration and keep the degrees of the forms constant:

    0(Ωk,w(M))Ωk,w(M)andΠ0(Ωk,w(M))Ωk,w(M).\Box_{0}(\Omega^{k,\geq w}(M))\subset\Omega^{k,\geq w}(M)\quad\mbox{and}\quad\Pi_{0}(\Omega^{k,\geq w}(M))\subset\Omega^{k,\geq w}(M).

    They also satisfy

    gr(0)=𝔤Mandgr(Π0)=Π𝔤M.{\rm gr}(\Box_{0})=\Box_{\mathfrak{g}M}\qquad\mbox{and}\qquad{\rm gr}(\Pi_{0})=\Pi_{\mathfrak{g}M}.

    The algebraic subbundles

    E0:=kerd0kerδ0andF0:=Imd0+Imδ0,E_{0}:=\ker d_{0}\cap\ker\delta_{0}\quad\mbox{and}\quad F_{0}:={\rm Im}\,d_{0}+{\rm Im}\,\delta_{0},

    admit subfiltrations and we have:

    gr(E0)=E𝔤Mandgr(F0)=F𝔤M.{\rm gr}(E_{0})=E_{\mathfrak{g}M}\qquad\mbox{and}\qquad{\rm gr}(F_{0})=F_{\mathfrak{g}M}.
  2. (2)

    We have

    Ω(M)=E0F0withE0=ker0andF0=Im0.\Omega^{\bullet}(M)=E_{0}^{\bullet}\oplus F_{0}^{\bullet}\qquad\mbox{with}\qquad E_{0}=\ker\Box_{0}\qquad\mbox{and}\qquad F_{0}={\rm Im}\,\Box_{0}.

    Moreover, Π0\Pi_{0} is the projection onto E0E_{0} along F0F_{0}.

Proof.

Part (1) is satisfied by construction. Disjointedness implies readily that E0F0={0}E_{0}\cap F_{0}=\{0\}. Since E𝔤MF𝔤M=𝔤ME_{\mathfrak{g}M}\oplus F_{\mathfrak{g}M}=\wedge^{\bullet}\mathfrak{g}^{*}M, Part (1) implies that dimE0+dimF0=dimE𝔤M+dimF𝔤M=dimΩ(M)\dim E_{0}+\dim F_{0}=\dim E_{\mathfrak{g}M}+\dim F_{\mathfrak{g}M}=\dim\Omega^{\bullet}(M). Hence E0F0=Ω(M)E_{0}\oplus F_{0}=\Omega^{\bullet}(M).

Disjointedness also implies E0=ker0E_{0}=\ker\Box_{0}. Clearly, we also have Im0F0{\rm Im}\,\Box_{0}\subset F_{0}. From the Cauchy residue formula, proceeding as in the proof of [FT23, Proposition 4.1], we check that Π02=Π0\Pi_{0}^{2}=\Pi_{0}. In other words, Π0\Pi_{0} is a projection of Ω(M)\Omega^{\bullet}(M). We also check using the Cauchy residue that Π0=I\Pi_{0}=\text{\rm I} on ker0\ker\Box_{0} while Π00=0\Pi_{0}\Box_{0}=0 so Π0=0\Pi_{0}=0 on Im0{\rm Im}\,\Box_{0}. We obtain Part (2) from Ω(M)=E0F0\Omega^{\bullet}(M)=E_{0}\oplus F_{0}. ∎

We call Π0\Pi_{0} the base kernel projection.

4.2. The operators \Box, PP and LL

We define the differential operator acting on Ω(M)\Omega^{\bullet}(M):

:=dδ0+δ0d.\Box:=d\,\delta_{0}+\delta_{0}\,d.
Remark 4.3.

The de Rham differential dd will be replaced with its algebraic part d~\tilde{d} in Section 4.4. The new construction will closely follow the steps below, as the only property that it relies on is that gr(d)=d𝔤M{\rm gr}(d)=d_{\mathfrak{g}M}.

Proposition 4.4.

The operator \Box commutes with dd and δ0\delta_{0}. It is a differential operator that preserves the degrees and the weights of forms:

(Ωk,w(M))Ωk,w(M).\Box(\Omega^{k,\geq w}(M))\subseteq\Omega^{k,\geq w}(M).

We have

gr()=𝔤M.{\rm gr}(\Box)=\Box_{\mathfrak{g}M}.
Proof.

Since dd and δ0\delta_{0} are complexes, they commute with \Box. Moreover, dd and d0d_{0} increase the degrees of forms by one, while δ0\delta_{0} decreases the degrees by one. Hence, \Box and 0\Box_{0} preserve the degrees of forms. As dd, d0d_{0} and δ0\delta_{0} respect the filtration, so do \Box and 0\Box_{0}. We conclude with

gr()=gr(d)gr(δ0)+gr(δ0)gr(d)=d𝔤Md𝔤Mt+d𝔤Mtd𝔤M=𝔤M.{\rm gr}(\Box)={\rm gr}(d){\rm gr}(\delta_{0})+{\rm gr}(\delta_{0}){\rm gr}(d)=d_{\mathfrak{g}M}d_{\mathfrak{g}M}^{t}+d_{\mathfrak{g}M}^{t}d_{\mathfrak{g}M}=\Box_{\mathfrak{g}M}.

Applying Proposition 4.4 together with Proposition 2.22 to zz-\Box, the map

P:=12πi|z|=ε(z)1𝑑z,P:=\frac{1}{2\pi i}\oint_{|z|=\varepsilon}(z-\Box)^{-1}dz,

is well-defined for |z|=ε|z|=\varepsilon small enough.

Proposition 4.5.
  1. (1)

    The map PP is a differential operator acting on Ω(M)\Omega^{\bullet}(M) that respects the filtration and the degree of the forms with gr(P)=Π𝔤M{\rm gr}(P)=\Pi_{\mathfrak{g}M}. It commutes with \Box, dd and δ0\delta_{0}.

  2. (2)

    We have N0P=0\Box^{N_{0}}P=0 while N0\Box^{N_{0}} acts on ImP{\rm Im}\,P where it is invertible. Moreover, PP is the projection onto

    E:=kerδ0ker(δ0d)=ImP=kerN0,E:=\ker\delta_{0}\cap\ker(\delta_{0}d)={\rm Im}\,P=\ker\Box^{N_{0}},

    along

    F:=Imδ0+Imdδ0=kerP=ImN0.F:={\rm Im}\,\delta_{0}+{\rm Im}\,d\delta_{0}=\ker P={\rm Im}\,\Box^{N_{0}}.

    These modules satisfy gr(E)=E0{\rm gr}(E)=E_{0} and gr(F)=F0{\rm gr}(F)=F_{0}.

Proof.

Part (1) follows from Proposition 4.4 and applying gr{\rm gr}. From the Cauchy residue formula, proceeding as in the proof of [FT23, Proposition 4.1], we check that P2=PP^{2}=P. In other words, P:Ω(M)Ω(M)P\colon\Omega^{\bullet}(M)\to\Omega^{\bullet}(M) is a projection.

For Part (2), gr(P)=0{\rm gr}(\Box P)=0 by the Cauchy formula and the properties of gr{\rm gr}. Therefore (P)N0=0(\Box P)^{N_{0}}=0, but (P)N0=N0P(\Box P)^{N_{0}}=\Box^{N_{0}}P since PP is a projection commuting with \Box. Therefore ImPkerN0{\rm Im}\,P\subseteq\ker\Box^{N_{0}}.

Naturally, \Box acts on ImN0{\rm Im}\,\Box^{N_{0}} which is a submodule of Ω(M)\Omega^{\bullet}(M) that inherits a filtration. Moreover, we check

gr(ImN0)=Imgr(N0)=Im𝔤M=kerΠ𝔤M,{\rm gr}({\rm Im}\,\Box^{N_{0}})={\rm Im}\,{\rm gr}(\Box^{N_{0}})={\rm Im}\,\Box_{\mathfrak{g}M}=\ker\Pi_{\mathfrak{g}M},

since 𝔤M\Box_{\mathfrak{g}M} is symmetric and the orthogonal projection onto its kernel is Π𝔤M\Pi_{\mathfrak{g}M}, and

gr|ImN0(:ImN0ImN0)=𝔤M:ImΠ𝔤MImΠ𝔤M.{\rm gr}|_{{\rm Im}\,\Box^{N_{0}}}(\Box:{\rm Im}\,\Box^{N_{0}}\to{\rm Im}\,\Box^{N_{0}})=\Box_{\mathfrak{g}M}:{\rm Im}\,\Pi_{\mathfrak{g}M}\to{\rm Im}\,\Pi_{\mathfrak{g}M}.

Adapting the proof of Proposition 2.22 to a submodule, we obtain that |ImN0\Box|_{{\rm Im}\,\Box^{N_{0}}} is invertible on ImN0{\rm Im}\,\Box^{N_{0}}. This implies that kerN0ImN0={0}\ker\Box^{N_{0}}\cap{\rm Im}\,\Box^{N_{0}}=\{0\} and kerN0ImN0=Ω(M)\ker\Box^{N_{0}}\oplus{\rm Im}\,\Box^{N_{0}}=\Omega^{\bullet}(M). We then readily check that PP is the projection onto kerN0\ker\Box^{N_{0}} along ImN0{\rm Im}\,\Box^{N_{0}}.

As Π𝔤Md𝔤Mt=0\Pi_{\mathfrak{g}M}d_{\mathfrak{g}M}^{t}=0, we have Π0δ0=0\Pi_{0}\delta_{0}=0. Hence, since PP commutes with δ0\delta_{0}, we have

(PΠ0)δ0=Pδ0=δ0P,(P-\Pi_{0})\delta_{0}=P\delta_{0}=\delta_{0}P\,,

and so recursively, for any k1k\geq 1,

(PΠ0)kδ0=Pkδ0=δ0Pk.(P-\Pi_{0})^{k}\delta_{0}=P^{k}\delta_{0}=\delta_{0}P^{k}\,.

For kN0k\geq N_{0}, (PΠ0)k=0(P-\Pi_{0})^{k}=0 since gr(P)=Π𝔤M=gr(Π0){\rm gr}(P)=\Pi_{\mathfrak{g}M}={\rm gr}(\Pi_{0}). As PP is a projection, we have obtained 0=Pδ0=δ0P0=P\delta_{0}=\delta_{0}P. Moreover, since PP commutes with dd, we also have Pdδ0=dPδ0=0Pd\delta_{0}=dP\delta_{0}=0, which implies the inclusion kerPImδ0+Imdδ0=:F\ker P\supset{\rm Im}\,\delta_{0}+{\rm Im}\,d\delta_{0}=:F.

Since d2=0d^{2}=0 and (δ0)2=0(\delta_{0})^{2}=0, we compute easily

N0=(dδ0)N0+(δ0d)N0,\Box^{N_{0}}=(d\delta_{0})^{N_{0}}+(\delta_{0}d)^{N_{0}},

As PP is the projection onto kerN0\ker\Box^{N_{0}} along ImN0{\rm Im}\,\Box^{N_{0}}, we have

ImP\displaystyle{\rm Im}\,P =kerN0kerδ0ker(δ0d)=:E,\displaystyle=\ker\Box^{N_{0}}\supset\ker\delta_{0}\cap\ker(\delta_{0}d)=:E,
kerP\displaystyle\ker P =ImN0Imδ0+Im(dδ0)=:F.\displaystyle={\rm Im}\,\Box^{N_{0}}\subset{\rm Im}\,\delta_{0}+{\rm Im}\,(d\delta_{0})=:F\,.

These last inclusions then imply that kerP=F\ker P=F, but also ImP=E{\rm Im}\,P=E. ∎

Note that in our construction below, we do not need the characterisations of ImP{\rm Im}\,P and kerP\ker P as E=kerδ0ker(δ0d)E=\ker\delta_{0}\cap\ker(\delta_{0}d) and F=Imδ0+Imdδ0F={\rm Im}\,\delta_{0}+{\rm Im}\,d\delta_{0}. 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 LL acting on Ω(M)\Omega^{\bullet}(M) and defined as

L:=PΠ0+(IP)(IΠ0),L:=P\Pi_{0}+(\text{\rm I}-P)(\text{\rm I}-\Pi_{0}),

preserves the degrees of the forms and respects the filtration with gr(L)=I{\rm gr}(L)=\text{\rm I}. It is invertible and its inverse is a differential operator acting on Ω(M)\Omega^{\bullet}(M). We have

PL=PΠ0=LΠ0PL=P\Pi_{0}=L\Pi_{0}

and this implies

P=LΠ0L1and(L1L)Π0=Π0(L1L).P=L\Pi_{0}L^{-1}\qquad\mbox{and}\qquad(L^{-1}\Box L)\Pi_{0}=\Pi_{0}(L^{-1}\Box L).

Consequently,

F=kerP=L(kerΠ0)L1=LF0L1,E=ImP=L(ImΠ0)L1=LE0L1.F=\ker P=L(\ker\Pi_{0})L^{-1}=LF_{0}L^{-1},\qquad E={\rm Im}\,P=L({\rm Im}\,\Pi_{0})L^{-1}=LE_{0}L^{-1}.
Proof.

By Lemma 2.16, we have

gr(L)=gr(P)gr(Π0)+gr(IP)gr(IΠ0)=Π𝔤M2+(IΠ𝔤M)2=I.{\rm gr}(L)={\rm gr}(P)\ {\rm gr}(\Pi_{0})+{\rm gr}(\text{\rm I}-P)\ {\rm gr}(\text{\rm I}-\Pi_{0})=\Pi_{\mathfrak{g}M}^{2}+(\text{\rm I}-\Pi_{\mathfrak{g}M})^{2}=\text{\rm I}.

We conclude with Proposition 2.22 and straightforward computations. ∎

4.3. The complexes (E0,D)(E_{0}^{\bullet},D) and (F0,C)(F_{0}^{\bullet},C)

By Propositions 4.5 and 4.6, the de Rham differential dd commutes with PP, so L1dLL^{-1}dL commutes with L1PL=Π0L^{-1}PL=\Pi_{0}. Hence, we can decompose the differential operator L1dLL^{-1}dL as

L1dL=L1dLΠ0+L1dL(IΠ0)=D+C,L^{-1}dL\ =\ L^{-1}dL\Pi_{0}\ +\ L^{-1}dL(\text{\rm I}-\Pi_{0})\ =\ D\ +\ C\,,

where DD and CC are the differential operators

D\displaystyle D :=L1dLΠ0=Π0L1dLΠ0,\displaystyle:=L^{-1}dL\Pi_{0}\ =\ \Pi_{0}L^{-1}dL\Pi_{0},
C\displaystyle C :=L1dL(IΠ0)=(IΠ0)L1dL(IΠ0).\displaystyle:=L^{-1}dL(\text{\rm I}-\Pi_{0})\ =\ (\text{\rm I}-\Pi_{0})L^{-1}dL(\text{\rm I}-\Pi_{0}).

Since (L1dL)2=L1d2L=0(L^{-1}dL)^{2}=L^{-1}d^{2}L=0 and DC=0=CDDC=0=CD, it follows that D2=C2=0D^{2}=C^{2}=0. Hence, the chain complex L1dLL^{-1}dL decomposes into the direct sum of the two chain complexes: DD acting on E0=ImΠ0E_{0}={\rm Im}\,\Pi_{0}, and CC acting on F0=kerΠ0F_{0}=\ker\Pi_{0}.

Proposition 4.7.

The maps CC and d0d_{0} are conjugated on F0F_{0}:

Cg=gd0onF0,Cg=gd_{0}\quad\mbox{on}\ F_{0}\,,

where g:F0F0g:F_{0}\to F_{0} is the invertible operator defined as

g:=Cδ001+δ001d0:F0F0.g:=C\delta_{0}\Box_{0}^{-1}+\delta_{0}\Box_{0}^{-1}d_{0}\colon F_{0}\to F_{0}\,.

Hence, the complexes (F0,C)(F_{0}^{\bullet},C) and (F0,d0)(F_{0}^{\bullet},d_{0}) have the same cohomology.

Proof.

Note that Imd0Im0=F0{\rm Im}\,d_{0}\subseteq{\rm Im}\,\Box_{0}=F_{0} and that 0\Box_{0} is invertible on Im0=F0{\rm Im}\,\Box_{0}=F_{0}, so gg is well-defined on F0F_{0}.

Since gr(L)=I{\rm gr}(L)=\text{\rm I}, gr(d)=d𝔤M{\rm gr}(d)=d_{\mathfrak{g}_{M}} and gr(Π0)=Π𝔤M{\rm gr}(\Pi_{0})=\Pi_{\mathfrak{g}M}, we have

gr(C)=gr(L1dL(IΠ0))=d𝔤M(IΠ𝔤M),{\rm gr}(C)={\rm gr}(L^{-1}dL(\text{\rm I}-\Pi_{0}))=d_{\mathfrak{g}_{M}}(\text{\rm I}-\Pi_{\mathfrak{g}M}),

and on F𝔤MF_{\mathfrak{g}M}

gr|F0(g)\displaystyle{\rm gr}|_{F_{0}}(g) =gr(C)d𝔤Mt𝔤M1+d𝔤Mt𝔤M1d𝔤M\displaystyle={\rm gr}(C)\,d_{\mathfrak{g}_{M}}^{t}\Box_{\mathfrak{g}M}^{-1}+d_{\mathfrak{g}_{M}}^{t}\Box_{\mathfrak{g}M}^{-1}d_{\mathfrak{g}_{M}}
=d𝔤Md𝔤Mt𝔤M1+d𝔤Mt𝔤M1d𝔤M\displaystyle=d_{\mathfrak{g}_{M}}d_{\mathfrak{g}_{M}}^{t}\Box_{\mathfrak{g}M}^{-1}+d_{\mathfrak{g}_{M}}^{t}\Box_{\mathfrak{g}M}^{-1}d_{\mathfrak{g}_{M}}
=(d𝔤Md𝔤Mt+d𝔤Mtd𝔤M)𝔤M1=IF𝔤M.\displaystyle=(d_{\mathfrak{g}_{M}}d_{\mathfrak{g}_{M}}^{t}+d_{\mathfrak{g}_{M}}^{t}d_{\mathfrak{g}_{M}})\Box_{\mathfrak{g}M}^{-1}=\text{\rm I}_{F_{\mathfrak{g}M}}\,.

Applying the proof of Proposition 2.22 to the submodule F0F_{0}, gg is invertible on ImP0=F0{\rm Im}\,P_{0}=F_{0}.

Since C2=0C^{2}=0 and d02=0d_{0}^{2}=0, it is a straightfoward to check that Cg=gd0Cg=gd_{0} holds on F0F_{0}. ∎

Proposition 4.8.

The differential operator Π0L1\Pi_{0}L^{-1} is a chain map between (Ω(M),d)(\Omega^{\bullet}(M),d) and (E0,D)(E_{0},D), that is (Π0L1)d=D(Π0L1)(\Pi_{0}L^{-1})d=D(\Pi_{0}L^{-1}). This chain map is homotopically invertible, with homotopic inverse given by LL since we have

(Π0L1)L=Π0,andIL(Π0L1)=dh+hd,(\Pi_{0}L^{-1})L=\Pi_{0},\quad\mbox{and}\quad\text{\rm I}-L(\Pi_{0}L^{-1})=dh+hd\,,

where hh is the differential operator acting on Ω(M)\Omega^{\bullet}(M) and defined as

h:=Lgδ001g1(IΠ0)L1,h:=L\,g\delta_{0}\Box_{0}^{-1}\,g^{-1}(\text{\rm I}-\Pi_{0})\,L^{-1},

with gg as in Proposition 4.7.

Proof.

By construction, we have d=L(D+C)L1d=L(D+C)L^{-1}, with D=Π0DΠ0D=\Pi_{0}D\Pi_{0} and C=(IΠ0)C(IΠ0)C=(\text{\rm I}-\Pi_{0})C(\text{\rm I}-\Pi_{0}), so Π0L1d=Π0DΠ0L1=DΠ0L1\Pi_{0}L^{-1}d=\Pi_{0}D\Pi_{0}L^{-1}=D\Pi_{0}L^{-1}.

It remains to prove the properties regarding hh. We first point out that hh makes sense because Imδ0F0{\rm Im}\,\delta_{0}\subset F_{0} and gg acts on F0=Im(IΠ0)F_{0}={\rm Im}\,(\text{\rm I}-\Pi_{0}) in an invertible way. The definitions of hh and CC together with Proposition 4.7 then yield:

L1(dh+hd)L\displaystyle L^{-1}(dh+hd)L =L1dLgδ001g1(IΠ0)+gδ001g1(IΠ0)L1dL\displaystyle=L^{-1}dL\,g\delta_{0}\Box_{0}^{-1}\,g^{-1}(\text{\rm I}-\Pi_{0})+g\delta_{0}\Box_{0}^{-1}\,g^{-1}(\text{\rm I}-\Pi_{0})\,L^{-1}dL
=Cgδ001g1(IΠ0)+gδ001g1C\displaystyle=C\,g\delta_{0}\Box_{0}^{-1}\,g^{-1}(\text{\rm I}-\Pi_{0})+g\delta_{0}\Box_{0}^{-1}\,g^{-1}C
=gd0δ001g1(IΠ0)+gδ001d0g1(IΠ0)\displaystyle=gd_{0}\delta_{0}\Box_{0}^{-1}\,g^{-1}(\text{\rm I}-\Pi_{0})+g\delta_{0}\Box_{0}^{-1}\,d_{0}g^{-1}(\text{\rm I}-\Pi_{0})
=g(d0δ001+δ001d0)g1(IΠ0)=IΠ0.\displaystyle=g\left(d_{0}\delta_{0}\Box_{0}^{-1}+\delta_{0}\Box_{0}^{-1}d_{0}\right)g^{-1}(\text{\rm I}-\Pi_{0})=\text{\rm I}-\Pi_{0}.

The conclusion follows. ∎

Corollary 4.9.

The cohomology of (E0,D)(E_{0}^{\bullet},D) is linearly isomorphic to the de Rham cohomology of the manifold MM.

4.4. The complexes (E0,D~)(E_{0}^{\bullet},\tilde{D}) and (F0,C~)(F_{0}^{\bullet},\tilde{C})

In this section, we consider the complexes obtained by considering d~\tilde{d} instead of dd 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 ~\tilde{\Box} operator acting on Ω(M)\Omega^{\bullet}(M):

~:=d~δ0+δ0d~.\tilde{\Box}:=\tilde{d}\,\delta_{0}+\delta_{0}\,\tilde{d}.

It commutes with d~\tilde{d} and δ0\delta_{0} and preserves the degrees and the weights of forms:

~(Ωk,w(M))Ωk,w(M).\tilde{\Box}(\Omega^{k,\geq w}(M))\subseteq\Omega^{k,\geq w}(M).

We have

gr(~)=𝔤M.{\rm gr}(\tilde{\Box})=\Box_{\mathfrak{g}M}.

Applying Proposition 2.22 to z~z-\tilde{\Box} locally at xMx\in M, together with Proposition 4.4, the map

P~:=12πi|z|=ε(z~)1𝑑z,\tilde{P}:=\frac{1}{2\pi i}\oint_{|z|=\varepsilon}(z-\tilde{\Box})^{-1}dz,

is well-defined for |z|=ε|z|=\varepsilon small enough. P~\tilde{P} is then an algebraic operator acting on Ω(M)\Omega^{\bullet}(M) that respects the filtration and the degree of forms, and gr(P~)=Π𝔤M{\rm gr}(\tilde{P})=\Pi_{\mathfrak{g}M}. It commutes with ~\tilde{\Box}, d~\tilde{d} and δ0\delta_{0}. We have ~N0P~=0\tilde{\Box}^{N_{0}}\tilde{P}=0, while ~N0\tilde{\Box}^{N_{0}} acts on ImP~{\rm Im}\,\tilde{P} where it is invertible. Moreover, P~\tilde{P} is the projection onto

E~:=kerδ0ker(δ0d~)=ImP~=ker~N0,\tilde{E}:=\ker\delta_{0}\cap\ker(\delta_{0}\tilde{d})={\rm Im}\,\tilde{P}=\ker\tilde{\Box}^{N_{0}},

along

F~:=Imδ0+Imd~δ0=kerP~=Im~N0.\tilde{F}:={\rm Im}\,\delta_{0}+{\rm Im}\,\tilde{d}\delta_{0}=\ker\tilde{P}={\rm Im}\,\tilde{\Box}^{N_{0}}.

These modules satisfy gr(E~)=E0{\rm gr}(\tilde{E})=E_{0} and gr(F~)=F0{\rm gr}(\tilde{F})=F_{0}.

The algebraic operator L~\tilde{L} acting on Ω(M)\Omega^{\bullet}(M) and defined as

L~:=P~Π0+(IP~)(IΠ0),\tilde{L}:=\tilde{P}\Pi_{0}+(\text{\rm I}-\tilde{P})(\text{\rm I}-\Pi_{0}),

preserves the degree of forms, respects the filtration, and gr(L~)=I{\rm gr}(\tilde{L})=\text{\rm I}. It is invertible and its inverse is an algebraic operator acting on Ω(M)\Omega^{\bullet}(M). We have

P~L~=P~Π0=L~Π0,P~=L~Π0L1and(L~1~L~)Π0=Π0(L~1~L~).\tilde{P}\tilde{L}=\tilde{P}\Pi_{0}=\tilde{L}\Pi_{0},\qquad\tilde{P}=\tilde{L}\Pi_{0}L^{-1}\qquad\mbox{and}\qquad(\tilde{L}^{-1}\tilde{\Box}\tilde{L})\Pi_{0}=\Pi_{0}(\tilde{L}^{-1}\tilde{\Box}\tilde{L}).

Consequently,

F~=kerP~=L~(kerΠ0)L~1=L~F0L~1,E~=ImP~=L~(ImΠ0)L~1=L~E0L~1.\tilde{F}=\ker\tilde{P}=\tilde{L}(\ker\Pi_{0})\tilde{L}^{-1}=\tilde{L}F_{0}\tilde{L}^{-1},\qquad\tilde{E}={\rm Im}\,\tilde{P}=\tilde{L}({\rm Im}\,\Pi_{0})\tilde{L}^{-1}=\tilde{L}E_{0}\tilde{L}^{-1}.

We define the two differentials D~\tilde{D} on E0=ImΠ0E_{0}={\rm Im}\,\Pi_{0}, and C~\tilde{C} on F0=kerΠ0F_{0}=\ker\Pi_{0} via

L~1d~L~=L~1d~L~Π0+L~1d~L~(IΠ0)=:D~+C~\tilde{L}^{-1}\tilde{d}\tilde{L}=\tilde{L}^{-1}\tilde{d}\tilde{L}\Pi_{0}+\tilde{L}^{-1}\tilde{d}\tilde{L}(\text{\rm I}-\Pi_{0})=:\tilde{D}+\tilde{C}

The maps C~\tilde{C} and d0d_{0} are conjugated on F0F_{0}:

C~g~=g~d0onF0,\tilde{C}\tilde{g}=\tilde{g}d_{0}\quad\mbox{on}\ F_{0}\,,

where g~:F0F0\tilde{g}:F_{0}\to F_{0} is the invertible operator defined as

g~:=C~δ001+δ001d0:F0F0.\tilde{g}:=\tilde{C}\delta_{0}\Box_{0}^{-1}+\delta_{0}\Box_{0}^{-1}d_{0}\colon F_{0}\to F_{0}\,.

The complexes (F0,C~)(F_{0}^{\bullet},\tilde{C}) and (F0,d0)(F_{0}^{\bullet},d_{0}) have the same cohomology.

The differential operator Π0L~1\Pi_{0}\tilde{L}^{-1} is a chain map between (Ω(M),d~)(\Omega^{\bullet}(M),\tilde{d}) and (E0,D~)(E_{0},\tilde{D}), that is (Π0L~1)d~=D~(Π0L~1)(\Pi_{0}\tilde{L}^{-1})\tilde{d}=\tilde{D}(\Pi_{0}\tilde{L}^{-1}). This chain map is homotopically invertible, with homotopic inverse given by L~\tilde{L} since we have

(Π0L~1)L~=Π0,andIL~(Π0L~1)=d~h~+h~d~,(\Pi_{0}\tilde{L}^{-1})\tilde{L}=\Pi_{0},\quad\mbox{and}\quad\text{\rm I}-\tilde{L}(\Pi_{0}\tilde{L}^{-1})=\tilde{d}\tilde{h}+\tilde{h}\tilde{d}\,,

where h~\tilde{h} is the algebraic operator acting on Ω(M)\Omega^{\bullet}(M) and defined as

h~:=L~g~δ001g~1(IΠ0)L~1.\tilde{h}:=\tilde{L}\,\tilde{g}\delta_{0}\Box_{0}^{-1}\,\tilde{g}^{-1}(\text{\rm I}-\Pi_{0})\,\tilde{L}^{-1}.

Consequently, the cohomology of (E0,D~)(E_{0}^{\bullet},\tilde{D}) is linearly isomorphic to the cohomology of d~\tilde{d} of the manifold MM.

4.5. Equivalent constructions for the subcomplexes (E0,D)(E_{0}^{\bullet},D) and (E0,D~)(E_{0}^{\bullet},\tilde{D})

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) d01:=Φ1d𝔤M1Φ:Ω(M)Ω(M),d_{0}^{-1}:=\Phi^{-1}\circ d_{\mathfrak{g}M}^{-1}\circ\Phi\colon\Omega^{\bullet}(M)\to\Omega^{\bullet}(M)\,,

which is defined via the partial inverse d𝔤M1d_{\mathfrak{g}M}^{-1} of the osculating differential d𝔤Md_{\mathfrak{g}M} (see Section 3.1.4). From the properties of d𝔤M1d_{\mathfrak{g}M}^{-1}, we see that (d01)2=0(d_{0}^{-1})^{2}=0, it respects the filtration and decreases the degree by one:

k,w0d01(Ωk+1,w(M))Ωk,w(M).\forall\,k,w\in\mathbb{N}_{0}\qquad d_{0}^{-1}(\Omega^{k+1,\geq w}(M))\subseteq\Omega^{k,\geq w}(M)\,.

We have

kerd0t=kerd01andImd0t=Imd01,\ker d_{0}^{t}=\ker d_{0}^{-1}\quad\mbox{and}\quad{\rm Im}\,d_{0}^{t}={\rm Im}\,d_{0}^{-1},

so

E0=kerd0kerd01andF0=Imd0+Imd01.E_{0}=\ker d_{0}\cap\ker d_{0}^{-1}\quad\mbox{and}\quad F_{0}={\rm Im}\,d_{0}+{\rm Im}\,d_{0}^{-1}.

Moreover,

(4.2) Π0=Id01d0d0d01,\Pi_{0}=\text{\rm I}-d_{0}^{-1}d_{0}-d_{0}d_{0}^{-1},

and we have

d01Π0=Π0d01=0,d0Π0=Π0d0=0.d_{0}^{-1}\Pi_{0}=\Pi_{0}d_{0}^{-1}=0\ ,\quad d_{0}\Pi_{0}=\Pi_{0}d_{0}=0\,.
Lemma 4.10.

Let us consider the differential operators acting on Ω(M)\Omega^{\bullet}(M) defined by:

b:=d01d0d01d=d01(dd0)andb1:=d0d01dd01=(dd0)d01.b:=d_{0}^{-1}d_{0}-d_{0}^{-1}d=-d_{0}^{-1}(d-d_{0})\quad\mbox{and}\quad b_{1}:=d_{0}d_{0}^{-1}-dd_{0}^{-1}=-(d-d_{0})d_{0}^{-1}.
  1. (1)

    The maps bb and b1b_{1} are nilpotent. Consequently, Ib\text{\rm I}-b and Ib1\text{\rm I}-b_{1} are invertible, and (Ib)1(\text{\rm I}-b)^{-1} and (Ib1)1(\text{\rm I}-b_{1})^{-1} are well-defined differential operators acting on Ω(M)\Omega^{\bullet}(M). We have:

    bd01=d01b1,and(Ib)1d01=d01(Ib1)1.bd_{0}^{-1}=d_{0}^{-1}b_{1},\quad\mbox{and}\quad(\text{\rm I}-b)^{-1}d_{0}^{-1}=d_{0}^{-1}(\text{\rm I}-b_{1})^{-1}.
  2. (2)

    The differential operator Π:Ω(M)Ω(M)\Pi\colon\Omega^{\bullet}(M)\to\Omega^{\bullet}(M) defined as

    Π\displaystyle\Pi :=(Ib)1d01d+d(Ib)1d01=d01(Ib1)1d+dd01(Ib1)1,\displaystyle:=(\text{\rm I}-b)^{-1}d_{0}^{-1}d+d(\text{\rm I}-b)^{-1}d_{0}^{-1}=d_{0}^{-1}(\text{\rm I}-b_{1})^{-1}d+dd_{0}^{-1}(\text{\rm I}-b_{1})^{-1}\,,

    is the projection onto

    F:=Imd01+Imdd01=Imd0t+Imdd0t,F:={\rm Im}\,d_{0}^{-1}+{\rm Im}\,dd_{0}^{-1}={\rm Im}\,d_{0}^{t}+{\rm Im}\,dd_{0}^{t}\,,

    along

    E:=kerd01ker(d01d)=kerd0tker(d0td).E:=\ker d_{0}^{-1}\cap\ker(d_{0}^{-1}d)=\ker d_{0}^{t}\cap\ker(d_{0}^{t}d)\,.
Proof.

By Proposition 3.3 and Lemma 2.19, dd and d0d_{0} increase the degree of forms by one, respect the filtration, and gr(d)=d𝔤M=gr(d0){\rm gr}(d)=d_{\mathfrak{g}M}={\rm gr}(d_{0}). As d01d_{0}^{-1} decreases the degree of forms by one, bb respects the filtration, and gr(b)=gr(d01)gr(dd0)=0{\rm gr}(b)=-{\rm gr}(d_{0}^{-1})\ {\rm gr}(d-d_{0})=0. Consequently, it is nilpotent with bN0=0b^{N_{0}}=0. The proof then follows as in [FT23, Lemma 3.9]. ∎

Corollary 4.11.

The projections Π\Pi and PP constructed in Lemma 4.10 and Proposition 4.5 respectively coincide.

Following Rumin’s notation, the two projections onto FF and EE along EE and FF are denoted respectively as

ΠF:=ΠandΠE:=IΠ.\Pi_{F}:=\Pi\qquad\mbox{and}\qquad\Pi_{E}:=\text{\rm I}-\Pi\,.

We then obtain Rumin’s construction and its equivalence with the one presented in this paper.

Theorem 4.12.
  1. (1)

    (M. Rumin) The de Rham complex (Ω(M),d)(\Omega^{\bullet}(M),d) splits into two subcomplexes (E,d)(E^{\bullet},d) and (F,d)(F^{\bullet},d). Moreover, the differential operator defined as

    dc:=Π0dΠEΠ0:Ω(M)Ω(M),d_{c}:=\Pi_{0}d\,\Pi_{E}\,\Pi_{0}\colon\Omega^{\bullet}(M)\to\Omega^{\bullet}(M)\ ,

    satisfies

    dc2=0,dc(Ωk(M))Ωk+1(M).d_{c}^{2}=0\quad,\quad d_{c}(\Omega^{k}(M))\subset\Omega^{k+1}(M)\,.

    Moreover, we have

    ΠE=ΠEΠ0ΠEandΠ0ΠEΠ0=Π0.\Pi_{E}=\Pi_{E}\Pi_{0}\Pi_{E}\quad\mbox{and}\quad\Pi_{0}\Pi_{E}\Pi_{0}=\Pi_{0}.

    and so E0=ImΠ0=Π0EE_{0}={\rm Im}\,\Pi_{0}=\Pi_{0}E, and the complex (E,d)(E^{\bullet},d) is conjugated to (E0,dc)(E_{0}^{\bullet},d_{c}) via Π0\Pi_{0}.

  2. (2)

    The maps dcd_{c} and DD coincide, i.e. dc=Dd_{c}=D.

Proof.

Adapting the proof of [FT23, Lemma 3.10], it is easy to check that the operators d01d_{0}^{-1}, ΠE\Pi_{E} and ΠF\Pi_{F} satisfy the following properties:

  • 1.

    d01ΠE=ΠEd01=0d_{0}^{-1}\Pi_{E}=\Pi_{E}d_{0}^{-1}=0;

  • 2.

    dΠF=ΠFdd\Pi_{F}=\Pi_{F}d, and dΠE=ΠEdd\Pi_{E}=\Pi_{E}d;

  • 3.

    ΠE(IΠ0)ΠE=0\Pi_{E}(\text{\rm I}-\Pi_{0})\Pi_{E}=0;

  • 4.

    on Ωk(M)\Omega^{k}(M), we have ΠFt=(1)k(nk)ΠF\Pi_{F}^{t}=(-1)^{k(n-k)}\star\Pi_{F}\star and ΠEt=(1)k(nk)ΠE\Pi_{E}^{t}=(-1)^{k(n-k)}\star\Pi_{E}\star.

These readily imply Part (1) and Part (2), after adapting the proofs of Theorems 3.11 and 4.9 in [FT23] respectively. ∎

This equivalent construction also works if we replace dd with d~\tilde{d}, yielding a complex (E0,d~c)(E_{0}^{\bullet},\tilde{d}_{c}) which coincides with D~\tilde{D}.

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 (E0,D)(E_{0}^{\bullet},D) and (E0,D~)(E_{0}^{\bullet},\tilde{D}). Here, we show that if the Riemannian metric is compatible with the subRiemannian structure as explained in Section 4.6.2 below, then (E0,D)(E_{0}^{\bullet},D) 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 E0E_{0}-regular CCCC-structure” in [Rum05], he only briefly mentions the potential impact that the choice of a Riemannian metric can have on the resulting subcomplex of MM 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 MM equipped with a bracket generating distribution 𝒟TM\mathcal{D}\subset TM and with a metric g𝒟g_{\mathcal{D}} on 𝒟\mathcal{D}. Let Γ1=Γ(𝒟)\Gamma^{1}=\Gamma(\mathcal{D}) be the set of smooth sections of 𝒟\mathcal{D}, and Γi:=[Γ1,Γi1]+Γi1\Gamma^{i}:=[\Gamma^{1},\Gamma^{i-1}]+\Gamma^{i-1}, for i>1.i>1. As 𝒟\mathcal{D} is bracket generating, there exists ii such that Γi=TM\Gamma^{i}=TM, and we denote by ss the smallest such integer ii. If, for every i=1,,si=1,\ldots,s, there exists a subbundle HiTMH^{i}\subset TM for which Γi=Γ(Hi)\Gamma^{i}=\Gamma(H^{i}) is the set of smooth sections on HiH^{i}, then MM is said to have an regular subRiemannian structure. Clearly, the HiH^{i}s provide the structure of filtered manifold.

Consider a regular subRiemannian manifold MM as above. In this case, the osculating Lie algebras are stratified

𝔤xM=i=1s𝔤i,x,𝔤i,x=Hxi/Hxi+1,xM,\displaystyle\mathfrak{g}_{x}M=\oplus_{i=1}^{s}\mathfrak{g}_{i,x},\qquad\mathfrak{g}_{i,x}=H^{i}_{x}/H^{i+1}_{x}\ ,\ \forall\,x\in M\,,

and the subspaces 𝔤i,x𝔤xM\mathfrak{g}_{i,x}\subset\mathfrak{g}_{x}M are given by imposing 𝔤1:=𝒟\mathfrak{g}_{1}:=\mathcal{D}, and

𝔤2,x=[𝔤1,x,𝔤1,x]𝔤xM,𝔤3,x=[𝔤1,x,𝔤2,x]𝔤xM,,𝔤i,x=[𝔤1,x,[𝔤i1,x]𝔤xM,\mathfrak{g}_{2,x}=[\mathfrak{g}_{1,x},\mathfrak{g}_{1,x}]_{\mathfrak{g}_{x}M},\quad\mathfrak{g}_{3,x}=[\mathfrak{g}_{1,x},\mathfrak{g}_{2,x}]_{\mathfrak{g}_{x}M},\ldots,\ \mathfrak{g}_{i,x}=[\mathfrak{g}_{1,x},[\mathfrak{g}_{i-1,x}]_{\mathfrak{g}_{x}M},\ldots

The metric g𝒟g_{\mathcal{D}} on 𝒟\mathcal{D} naturally induces a scalar product on each osculating Lie algebra fibre 𝔤xM\mathfrak{g}_{x}M [Mon02, p. 188]. Let us briefly recall its construction.

At every point xMx\in M, the map

(𝔤1,x)j=𝔤1,x××𝔤1,x𝔤j,x,(V1,,Vj)[V1,[V2,[,Vj]𝔤xM]𝔤xM]𝔤xM,(\mathfrak{g}_{1,x})^{j}=\mathfrak{g}_{1,x}\times\ldots\times\mathfrak{g}_{1,x}\to\mathfrak{g}_{j,x}\ ,\ \,(V_{1},\ldots,V_{j})\longmapsto[V_{1},[V_{2},[\cdots,V_{j}]_{\mathfrak{g}_{x}M}]_{\mathfrak{g}_{x}M}\cdots]_{\mathfrak{g}_{x}M}\ ,

is jj-linear and surjective. Hence, the metric g𝒟g_{\mathcal{D}} on 𝔤1=H1=𝒟\mathfrak{g}_{1}=H^{1}=\mathcal{D} induces a scalar product on (𝔤1,x)j(\mathfrak{g}_{1,x})^{j} and then on 𝔤j,x\mathfrak{g}_{j,x} (which is the image of the above map, and therefore inherits the scalar product from the orthogonal complement of the kernel). Constructing this for j=2,,sj=2,\ldots,s, we obtain a scalar product g𝔤xMg_{\mathfrak{g}_{x}M} on 𝔤xM\mathfrak{g}_{x}M. By construction, the dependence in xx is smooth, and (g𝔤M)(x):=g𝔤xM(g_{\mathfrak{g}M})(x):=g_{\mathfrak{g}_{x}M}, xMx\in M, defines an element g𝔤MΓ(Sym(𝔤M𝔤M))g_{\mathfrak{g}M}\in\Gamma({\rm Sym}(\mathfrak{g}M\otimes\mathfrak{g}M)). Therefore, g𝔤Mg_{\mathfrak{g}M} is a metric on the osculating bundle 𝔤M\mathfrak{g}M of Lie algebras. We call g𝔤Mg_{\mathfrak{g}M} the osculating metric induced by g𝒟g_{\mathcal{D}}.

4.6.2. Compatible Riemannian metrics

We now assume that, in addition to being a regular subRiemannian manifold, MM is also equipped with a Riemannian metric, that is, with a metric gTMg_{TM} on its tangent bundle TMTM. As already seen in Sections 2.1.10 and 2.5.1, gTMg_{TM} will also induce a metric ggr(TM)g_{{\rm gr}(TM)} on gr(TM)𝔤M{\rm gr}(TM)\cong\mathfrak{g}M. The two metrics g𝔤Mg_{\mathfrak{g}M} and ggr(TM)g_{{\rm gr}(TM)} will be different in general.

Definition 4.13.

A Riemannian metric gTMg_{TM} is compatible with the subRiemannian structure of a regular subRiemannian manifold MM (or compatible for short) when the two metrics ggr(TM)g_{{\rm gr}(TM)} and g𝔤Mg_{\mathfrak{g}M} on the osculating Lie algebra bundle 𝔤Mgr(TM)\mathfrak{g}M\cong{\rm gr}(TM) coincide.

By construction, g𝔤Mg_{\mathfrak{g}M} coincides on 𝒟=H1=𝔤1\mathcal{D}=H^{1}=\mathfrak{g}_{1} with g𝒟g_{\mathcal{D}}. Hence, a necessary condition for gTMg_{TM} to be compatible is for gTMg_{TM} to coincide with g𝒟g_{\mathcal{D}} on 𝒟\mathcal{D}. 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 N6,3,6N_{6,3,6} in [LDT22]. Keeping the notation of [LDT22], if we take Θ={θ1,θ2,θ3,θ4,θ5,θ6}\Theta=\{\theta^{1},\theta^{2},\theta^{3},\theta^{4},\theta^{5},\theta^{6}\} and Θ^={θ1,θ2,θ3,θ4,θ4+θ5,θ5+θ6}\hat{\Theta}=\{\theta^{1},\theta^{2},\theta^{3},\theta^{4},\theta^{4}+\theta^{5},\theta^{5}+\theta^{6}\} as global left-invariant orthogonal coframes of TN6,3,6TN_{6,3,6}, then the corresponding subcomplexes (E0,D)(E_{0}^{\bullet},D) and (E^0,D^)(\hat{E}_{0}^{\bullet},\hat{D}) do not coincide [FT24]. For clarity, we stress that E0E_{0}^{\bullet} and E^0\hat{E}_{0}^{\bullet} are isomorphic subspaces of Ω(N6,3,6)\Omega^{\bullet}(N_{6,3,6}), and this is necessarily true in general once we assume MM 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 E02E_{0}^{2} and E^02\hat{E}_{0}^{2}). In other words, just extending the metric g𝒟g_{\mathcal{D}} from 𝒟\mathcal{D} to an arbitrary Riemannian metric on TMTM is not sufficient to ensure compatibility. However, compatible metrics can always be constructed.

Lemma 4.14.

Let MM be a (second countable) regular subRiemannian manifold. We can always construct a compatible Riemannian metric.

Proof.

Consider a sequence of non-negative functions φiCc(M)\varphi_{i}\in C_{c}^{\infty}(M), ii\in\mathbb{N}, such that the sum iφi\sum_{i\in\mathbb{N}}\varphi_{i} on MM is locally finite and constantly equal to 1. We may assume that the support of each φi\varphi_{i} is small enough so that it is included in an open subset UiMU_{i}\subset M where a frame 𝕏i\mathbb{X}_{i} of TUiTU_{i} exists. After a Graham-Schmidt procedure, we may assume that each 𝕏i\mathbb{X}_{i} is such that the corresponding frame 𝕏i\langle\mathbb{X}_{i}\rangle of the vector bundle 𝔤M|Ui\mathfrak{g}M|_{U_{i}} is g𝔤Mg_{\mathfrak{g}M}-orthonormal. Denoting by gig_{i} the corresponding metric on UiU_{i} such that 𝕏i\mathbb{X}_{i} is orthonormal, one can easily check that g=iφigig=\sum_{i\in\mathbb{N}}\varphi_{i}g_{i} is a compatible Riemannian metric. ∎

4.6.3. The base differential and codifferential associated with a compatible Riemannian metric

Let MM be a regular subRiemannian manifold equipped with a Riemannian metric gTMg_{TM}. The compatibility implies that the scalar product on gr(TxM)\wedge^{\bullet}{\rm gr}(T_{x}^{*}M) (defined as in Section 2.5.1 using gTMg_{TM}) coincides with the scalar product on 𝔤xM\wedge^{\bullet}\mathfrak{g}_{x}^{*}M induced by the osculating metric g𝔤Mg_{\mathfrak{g}M}. Hence, the map Φ\Phi identifying forms and osculating forms and the objects built upon it (for instance the base differential and co-differential d0d_{0} and δ0\delta_{0}) 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 MM is equipped with a frame X,Y,TX,Y,T satisfying

[X,Y]\displaystyle[X,Y] =c0T+c1X+c2Y,\displaystyle=c_{0}T+c_{1}X+c_{2}Y,
[X,T]\displaystyle[X,T] =c3X+c4Y,\displaystyle=c_{3}X+c_{4}Y,
[Y,T]\displaystyle[Y,T] =c5X+c6Y,\displaystyle=c_{5}X+c_{6}Y,

with cjC(M)c_{j}\in C^{\infty}(M), j=0,,6j=0,\ldots,6, with c0c_{0} nowhere vanishing.

The associated free basis Θ\Theta^{\bullet} of Ω(M)\Omega^{\bullet}(M) is given by

k=0\displaystyle k=0 1(0),\displaystyle\qquad 1(\geq 0),
k=1\displaystyle k=1 X(1),Y(1),T(2),\displaystyle\qquad X^{*}(\geq 1),\ Y^{*}(\geq 1),\ T^{*}(\geq 2),
k=2\displaystyle k=2 XY(2),XT(3),YT(3),\displaystyle\qquad X^{*}\wedge Y^{*}(\geq 2),\ X^{*}\wedge T^{*}(\geq 3),\ Y^{*}\wedge T^{*}(\geq 3),
k=3\displaystyle k=3 vol=XYT(4).\displaystyle\qquad{\rm vol}=X^{*}\wedge Y^{*}\wedge T^{*}(\geq 4).

The numbers in parenthesis refer to the \geq weight of the form.

Description of d~\tilde{d}

For k=0,3k=0,3, d~(k)=0\tilde{d}^{(k)}=0. For k=1,2k=1,2, let us describe d~\tilde{d} in matrix form with respect to the canonical basis described above.

Mat(d~(1))=(c1c2c0c3c40c5c60),Mat(d~(2))=(c3+c6c1c2).{\rm Mat}(\tilde{d}^{(1)})=\left(\begin{array}[]{ccc}-c_{1}&-c_{2}&-c_{0}\\ -c_{3}&-c_{4}&0\\ -c_{5}&-c_{6}&0\end{array}\right),\qquad{\rm Mat}(\tilde{d}^{(2)})=\left(\begin{array}[]{ccc}c_{3}+c_{6}&-c_{1}&-c_{2}\end{array}\right).

Description of dd

For k=3k=3, we have d(3)=0d^{(3)}=0, while for k=0k=0, we have df(V)=Vfdf(V)=Vf, fC(M)f\in C^{\infty}(M). Hence,

Mat(d(0))\displaystyle{\rm Mat}(d^{(0)}) =(XYT)\displaystyle=\left(\begin{array}[]{c}X\\ Y\\ T\end{array}\right)
Mat(d(1))\displaystyle{\rm Mat}(d^{(1)}) =(YX0T0X0TY)+Mat(d~(1))\displaystyle=\left(\begin{array}[]{ccc}-Y&X&0\\ -T&0&X\\ 0&-T&Y\end{array}\right)+{\rm Mat}(\tilde{d}^{(1)})
Mat(d(2))\displaystyle{\rm Mat}(d^{(2)}) =(TYX)+Mat(d~(2))\displaystyle=\left(\begin{array}[]{ccc}T&-Y&X\end{array}\right)+{\rm Mat}(\tilde{d}^{(2)})

Description of d𝔤Md_{\mathfrak{g}M}

The osculating group is isomorphic to the Heisenberg group at any point pMp\in M:

[Xp,Yp]𝔤xM=c0(p)Tp,[Xp,Tp]𝔤xM=0=[Yp,Tp]𝔤xM.[\langle X\rangle_{p},\langle Y\rangle_{p}]_{\mathfrak{g}_{x}M}=c_{0}(p)\langle T\rangle_{p},\qquad[\langle X\rangle_{p},\langle T\rangle_{p}]_{\mathfrak{g}_{x}M}=0=[\langle Y\rangle_{p},\langle T\rangle_{p}]_{\mathfrak{g}_{x}M}.

The associated free basis of GΩ(M)G\Omega^{\bullet}(M) is

k=0\displaystyle k=0 1(=0),\displaystyle\qquad 1(=0),
k=1\displaystyle k=1 X(=1),Y(=1),T(=2),\displaystyle\qquad\langle X^{*}\rangle(=1),\ \langle Y^{*}\rangle(=1),\ \langle T^{*}\rangle(=2),
k=2\displaystyle k=2 XY(=2),XT(=3),YT(=3),\displaystyle\qquad\langle X^{*}\rangle\wedge\langle Y^{*}\rangle(=2),\ \langle X^{*}\rangle\wedge\langle T^{*}\rangle(=3),\ \langle Y^{*}\rangle\wedge\langle T^{*}\rangle(=3),
k=3\displaystyle k=3 XYT(=4).\displaystyle\qquad\langle X^{*}\rangle\wedge\langle Y^{*}\rangle\wedge\langle T^{*}\rangle(=4).

The number in parenthesis refers to the weight of the osculating form.

For k=0,2,3k=0,2,3, d𝔤M(k)=0d_{\mathfrak{g}M}^{(k)}=0. For k=1k=1, let us describe d𝔤Md_{\mathfrak{g}M} in matrix form with respect to the canonical basis described above.

Mat(d𝔤M(1))=(00c0000000).{\rm Mat}(d_{\mathfrak{g}M}^{(1)})=\left(\begin{array}[]{ccc}0&0&-c_{0}\\ 0&0&0\\ 0&0&0\end{array}\right).

Description of d0d_{0} and d0td_{0}^{t}

For k=0,2,3k=0,2,3, d0(k)=0d_{0}^{(k)}=0 while for k=1k=1, the description of d0(1)d_{0}^{(1)} in matrix form with respect to the canonical basis of Ω(M)\Omega^{\bullet}(M) given above coincides with Mat(d𝔤M(1)){\rm Mat}(d_{\mathfrak{g}M}^{(1)}). For k=0,1,3k=0,1,3, d0(k,t)=0d_{0}^{(k,t)}=0. For k=2k=2, the matrix description of d0td_{0}^{t} coincides with the transpose of Mat(d𝔤M(1)){\rm Mat}(d_{\mathfrak{g}M}^{(1)}). Consequently,

Mat(d0(1))=(00c0000000),Mat(d0(1,t))=(000000c000).{\rm Mat}(d_{0}^{(1)})=\left(\begin{array}[]{ccc}0&0&-c_{0}\\ 0&0&0\\ 0&0&0\end{array}\right)\ ,\quad{\rm Mat}(d_{0}^{(1,t)})=\left(\begin{array}[]{ccc}0&0&0\\ 0&0&0\\ -c_{0}&0&0\end{array}\right).

Description of 0\Box_{0}, P0P_{0}

For k=0,3k=0,3, 0(k)=0\Box^{(k)}_{0}=0, while 0(1)=d0(1,t)d0(1)\Box^{(1)}_{0}=d_{0}^{(1,t)}d_{0}^{(1)} and 0(2)=d0(1)d0(1,t)\Box^{(2)}_{0}=d_{0}^{(1)}d_{0}^{(1,t)} are represented by the following diagonal matrices:

Mat(0(1))\displaystyle{\rm Mat}\big{(}\Box_{0}^{(1)}\big{)} =Mat(d0(1,t))Mat(d0(1))=diag(0,0,c02),\displaystyle={\rm Mat}\big{(}d_{0}^{(1,t)}\big{)}{\rm Mat}\big{(}d_{0}^{(1)}\big{)}={\rm diag}\big{(}0,0,c_{0}^{2}\big{)}\ ,
Mat(0(2))\displaystyle{\rm Mat}\big{(}\Box_{0}^{(2)}\big{)} =Mat(d0(1))Mat(d0(1,t))=diag(c02,0,0).\displaystyle={\rm Mat}\big{(}d_{0}^{(1)}\big{)}{\rm Mat}\big{(}d_{0}^{(1,t)}\big{)}={\rm diag}\big{(}c_{0}^{2},0,0\big{)}\ .

Consequently, Π0(k)\Pi_{0}^{(k)} is the identity for k=0,3k=0,3, while for k=1,2k=1,2,

Mat(Π0(1))=diag(1,1,0)andMat(Π0(2))=diag(0,1,1).{\rm Mat}\big{(}\Pi_{0}^{(1)}\big{)}={\rm diag}(1,1,0)\qquad\mbox{and}\qquad{\rm Mat}\big{(}\Pi_{0}^{(2)}\big{)}={\rm diag}(0,1,1)\ .

Consequently, E0(0)=Ω0(M)=Φ1(1)E_{0}^{(0)}=\Omega^{0}(M)=\Phi^{-1}\big{(}\langle 1\rangle\big{)} and E0(3)=Ω3(M)=Φ1(XYT)E_{0}^{(3)}=\Omega^{3}(M)=\Phi^{-1}\big{(}\langle X^{*}\wedge Y^{*}\wedge T^{*}\rangle\big{)} while

E0(1)=Φ1(XY)andE0(2)=Φ1(XTYT).E_{0}^{(1)}=\Phi^{-1}\big{(}\langle X^{*}\rangle\oplus\langle Y^{*}\rangle\big{)}\quad\mbox{and}\quad E_{0}^{(2)}=\Phi^{-1}\big{(}\langle X^{*}\wedge T^{*}\rangle\oplus\langle Y^{*}\wedge T^{*}\rangle\big{)}.

Description of \Box, PP, LL

For k=0,3k=0,3, we have (k)=0\Box^{(k)}=0, while for k=1,2k=1,2

Mat((1))\displaystyle{\rm Mat}\big{(}\Box^{(1)}\big{)} =Mat(d0(1,t))Mat(d(1))=c0(000000Y+c1X+c2c0),\displaystyle={\rm Mat}\big{(}d_{0}^{(1,t)}\big{)}{\rm Mat}\big{(}d^{(1)}\big{)}=c_{0}\left(\begin{array}[]{ccc}0&0&0\\ 0&0&0\\ Y+c_{1}&-X+c_{2}&c_{0}\end{array}\right),
Mat((2))\displaystyle{\rm Mat}\big{(}\Box^{(2)}\big{)} =Mat(d(1))Mat(d0(1,t))=c0(c000X00Y00).\displaystyle={\rm Mat}\big{(}d^{(1)}\big{)}{\rm Mat}\big{(}d_{0}^{(1,t)}\big{)}=-c_{0}\left(\begin{array}[]{ccc}-c_{0}&0&0\\ X&0&0\\ Y&0&0\end{array}\right).

Consequently, P(k)P^{(k)} is the identity on Ωk(M)\Omega^{k}(M) for k=0,3k=0,3. Since ((z0)(0))j=0\big{(}(z-\Box_{0})(\Box_{0}-\Box)\big{)}^{j}=0 for j>1j>1, the formulae for Cauchy residues and the von Neumann series simplify into

P=Π0+Π0(0)c02prc02+c02prc02(0)Π0,P=\Pi_{0}+\Pi_{0}(\Box_{0}-\Box)c_{0}^{-2}{\rm pr}_{c_{0}^{2}}+c_{0}^{-2}{\rm pr}_{c_{0}^{2}}(\Box_{0}-\Box)\Pi_{0}\ ,

where prc02{\rm pr}_{c_{0}^{2}} denotes the projection onto the c02c_{0}^{2}-eigenspace of 0\Box_{0}. Hence, for k=1,2k=1,2,

Mat(P(1))\displaystyle{\rm Mat}\big{(}P^{(1)}\big{)} =diag(1,1,0)c01(000000Y+c1X+c20),\displaystyle={\rm diag}(1,1,0)-c_{0}^{-1}\left(\begin{array}[]{ccc}0&0&0\\ 0&0&0\\ Y+c_{1}&-X+c_{2}&0\end{array}\right)\ ,
Mat(P(2))\displaystyle{\rm Mat}\big{(}P^{(2)}\big{)} =diag(0,1,1)+c01(000X00Y00).\displaystyle={\rm diag}(0,1,1)+c_{0}^{-1}\left(\begin{array}[]{ccc}0&0&0\\ X&0&0\\ Y&0&0\end{array}\right)\ .

We now describe

L=PΠ0+(IP)(IΠ0)=I+(PΠ0)(I+2Π0).L=P\Pi_{0}+(\text{\rm I}-P)(\text{\rm I}-\Pi_{0})=\text{\rm I}+(P-\Pi_{0})(-\text{\rm I}+2\Pi_{0}).

For k=0,3k=0,3, L(k)L^{(k)} is the identity on Ωk(M)\Omega^{k}(M) while for k=1,2k=1,2,

Mat(L(1))\displaystyle{\rm Mat}\big{(}L^{(1)}\big{)} =I3c01(000000Y+c1X+c20)\displaystyle=\text{\rm I}_{3}-c_{0}^{-1}\left(\begin{array}[]{ccc}0&0&0\\ 0&0&0\\ Y+c_{1}&-X+c_{2}&0\end{array}\right)
Mat(L(2))\displaystyle{\rm Mat}\big{(}L^{(2)}\big{)} =I3c01(000X00Y00),\displaystyle=\text{\rm I}_{3}-c_{0}^{-1}\left(\begin{array}[]{ccc}0&0&0\\ X&0&0\\ Y&0&0\end{array}\right),

where I3=diag(1,1,1)\text{\rm I}_{3}={\rm diag}(1,1,1) denotes the 3-by-3 identity matrix.

Computation of DD

We now describe

D(k)=(L(k+1))1d(k)L(k)Π0(k),D^{(k)}=\big{(}L^{(k+1)}\big{)}^{-1}d^{(k)}L^{(k)}\Pi_{0}^{(k)},

for k=0,1,2k=0,1,2.

Mat(D(0))\displaystyle{\rm Mat}\big{(}D^{(0)}\big{)} =(XYT+c01((Y+c1)X+(X+c2)Y))=(XY0),\displaystyle=\left(\begin{array}[]{c}X\\ Y\\ T+c_{0}^{-1}((Y+c_{1})X+(-X+c_{2})Y)\end{array}\right)=\left(\begin{array}[]{c}X\\ Y\\ 0\end{array}\right),
Mat(D(1))\displaystyle{\rm Mat}(D^{(1)}) =(000Tc3c01X(Y+c1)c4+c01X(Xc2)0c5c01Y(Y+c1)Tc6+Yc01(Xc2)0),\displaystyle=\left(\begin{array}[]{ccc}0&0&0\\ -T-c_{3}-c_{0}^{-1}X(Y+c_{1})&-c_{4}+c_{0}^{-1}X(X-c_{2})&0\\ -c_{5}-c_{0}^{-1}Y(Y+c_{1})&-T-c_{6}+Yc_{0}^{-1}(X-c_{2})&0\end{array}\right),
Mat(D(2))\displaystyle{\rm Mat}(D^{(2)}) =(0Yc1Xc2).\displaystyle=\left(\begin{array}[]{ccc}0&-Y-c_{1}&X-c_{2}\end{array}\right).

The zeros in the matrices above illustrate DD acting on E0E_{0} while being trivial on F0F_{0}.

Description of ~\tilde{\Box}, P~\tilde{P}, L~\tilde{L}

For k=0,3k=0,3, we have ~(k)=0\tilde{\Box}^{(k)}=0, while for k=1,2k=1,2

Mat(~(1))=\displaystyle{\rm Mat}\left(\tilde{\Box}^{(1)}\right)= Mat(d0(1,t))Mat(d~(1))=c0(000000c1c2c0),\displaystyle{\rm Mat}\big{(}d_{0}^{(1,t)}\big{)}{\rm Mat}\big{(}\tilde{d}^{(1)}\big{)}=c_{0}\left(\begin{array}[]{ccc}0&0&0\\ 0&0&0\\ c_{1}&c_{2}&c_{0}\end{array}\right)\ ,
Mat(~(2))=\displaystyle{\rm Mat}\big{(}\tilde{\Box}^{(2)}\big{)}= Mat(d~(1))Mat(d0(1,t))=diag(c02,0,0).\displaystyle{\rm Mat}\big{(}\tilde{d}^{(1)}\big{)}{\rm Mat}\big{(}d_{0}^{(1,t)}\big{)}={\rm diag}\big{(}c_{0}^{2},0,0\big{)}\ .

Consequently, P~(k)\tilde{P}^{(k)} is the identity on Ωk(M)\Omega^{k}(M) for k=0,3k=0,3. Since ((z0)(0~))j=0\big{(}(z-\Box_{0})(\Box_{0}-\tilde{\Box})\big{)}^{j}=0 for j>1j>1, the formulae for Cauchy residues and the von Neumann series simplify into

P~=Π0+Π0(0~)c02prc02+c02prc02(0~)Π0,\displaystyle\tilde{P}=\Pi_{0}+\Pi_{0}(\Box_{0}-\tilde{\Box})c_{0}^{-2}{\rm pr}_{c_{0}^{2}}+c_{0}^{-2}{\rm pr}_{c_{0}^{2}}(\Box_{0}-\tilde{\Box})\Pi_{0}\ ,

where prc02{\rm pr}_{c_{0}^{2}} denotes the projection onto the c02c_{0}^{2}-eigenspace of 0\Box_{0}. Hence, for k=1,2k=1,2,

Mat(P~(1))=\displaystyle{\rm Mat}\big{(}\tilde{P}^{(1)}\big{)}= diag(1,1,0)c01(000000c1c20)\displaystyle{\rm diag}(1,1,0)-c_{0}^{-1}\left(\begin{array}[]{ccc}0&0&0\\ 0&0&0\\ c_{1}&c_{2}&0\end{array}\right)
Mat(P~(2))=\displaystyle{\rm Mat}\big{(}\tilde{P}^{(2)}\big{)}= diag(0,1,1).\displaystyle{\rm diag}(0,1,1)\ .

We now describe

L~=P~Π0+(IP~)(IΠ0)=I+(P~Π0)(I+2Π0).\displaystyle\tilde{L}=\tilde{P}\Pi_{0}+(\text{\rm I}-\tilde{P})(\text{\rm I}-\Pi_{0})=\text{\rm I}+(\tilde{P}-\Pi_{0})(-\text{\rm I}+2\Pi_{0})\ .

For k=0,3k=0,3, L~(k)\tilde{L}^{(k)} is the identity on Ωk(M)\Omega^{k}(M) while for k=1,2k=1,2,

Mat(L~(1))=I3c01(000000c1c20),Mat(L~(2))=I3,\displaystyle{\rm Mat}\big{(}\tilde{L}^{(1)}\big{)}=\text{\rm I}_{3}-c_{0}^{-1}\left(\begin{array}[]{ccc}0&0&0\\ 0&0&0\\ c_{1}&c_{2}&0\end{array}\right)\ ,\ {\rm Mat}\big{(}\tilde{L}^{(2)}\big{)}=\text{\rm I}_{3}\ ,

where I3=diag(1,1,1)\text{\rm I}_{3}={\rm diag}(1,1,1) denotes the 3-by-3 identity matrix.

Computation of D~\tilde{D}

We now describe

D~(k)=(L(k+1))1d~(k)L~(k)Π0(k)\displaystyle\tilde{D}^{(k)}=\big{(}L^{(k+1)}\big{)}^{-1}\tilde{d}^{(k)}\tilde{L}^{(k)}\Pi_{0}^{(k)}

for k=0,1,2k=0,1,2.

Mat(D~(0))=(000),Mat(D~(1))=(000c3c40c5c60),{\rm Mat}\big{(}\tilde{D}^{(0)}\big{)}=\left(\begin{array}[]{c}0\\ 0\\ 0\end{array}\right),\quad{\rm Mat}\big{(}\tilde{D}^{(1)}\big{)}=\left(\begin{array}[]{ccc}0&0&0\\ -c_{3}&-c_{4}&0\\ -c_{5}&-c_{6}&0\end{array}\right),
Mat(D~(2))=(0c1c2).{\rm Mat}\big{(}\tilde{D}^{(2)}\big{)}=\left(\begin{array}[]{ccc}0&-c_{1}&-c_{2}\end{array}\right).

Appendix B Examples

In this appendix, we give two examples of a differential operator D:Ω(M)Ω(M)D\colon\Omega^{\bullet}(M)\to\Omega^{\bullet}(M) that respects the filtration (2.11) whose transpose DtD^{t} does not necessarily respect the filtration.

Example B.1.

We consider the 3-dimensional Heisenberg group \mathbb{H}, and the canonical basis X,Y,TX,Y,T of its Lie algebra 𝔥\mathfrak{h}. The only non-trivial bracket is [X,Y]=T[X,Y]=T. Identifying 𝔥\mathfrak{h} with the space of left-invariant vector fields on \mathbb{H}, we obtain a filtration of the tangent bundle of \mathbb{H} in subbundles:

{0}=H0H1:=span{X,Y}H2:=span{X,Y,T}=T1.\{0\}=H^{0}\subset H^{1}:={\rm span}_{\mathbb{R}}\{X,Y\}\subset H^{2}:={\rm span}_{\mathbb{R}}\{X,Y,T\}=T\mathbb{H}^{1}\,.

The dual of the left-invariant frame 𝕏:=(X,Y,T)\mathbb{X}:=(X,Y,T) of TT\mathbb{H} yields the global coframe Θ:=(X,Y,T)\Theta:=(X^{\ast},Y^{\ast},T^{\ast}) (as well as the natural linear isomorphism between Ω()\Omega^{\bullet}(\mathbb{H}) and C()𝔥C^{\infty}(\mathbb{H})\otimes\wedge^{\bullet}\mathfrak{h}^{*}).

Let us now consider the de Rham differential d:Ω()Ω()d\colon\Omega^{\bullet}(\mathbb{H})\to\Omega^{\bullet}(\mathbb{H}) acting on the 1-form fTΩ1,2()f\,T^{\ast}\in\Omega^{1,\geq 2}(\mathbb{H}) with fC()f\in C^{\infty}(\mathbb{H}):

d(fT)=dfT+fdT=Xf(XT)+Yf(YT)f(XY)Ω2,2().\displaystyle d(fT^{\ast})=df\wedge T^{\ast}+f\,dT^{\ast}=Xf\,(X^{\ast}\wedge T^{\ast})+Yf\,(Y^{\ast}\wedge T^{\ast})-f\,(X^{\ast}\wedge Y^{\ast})\in\Omega^{2,\geq 2}(\mathbb{H})\,.

If we now consider its formal transpose dt=(1)k3+1d:Ωk+1()Ωk()d^{t}=(-1)^{k\cdot 3+1}\star d\star\colon\Omega^{k+1}(\mathbb{H})\to\Omega^{k}(\mathbb{H}) acting on the 2-form g(XY)Ω2,2()g\,(X^{\ast}\wedge Y^{\ast})\ \in\ \Omega^{2,\geq 2}(\mathbb{H}) with gC()g\in C^{\infty}(\mathbb{H}):

dt(g(XY))=\displaystyle d^{t}(g\,(X^{\ast}\wedge Y^{\ast}))= d(gT)=(g(XY)+Xg(XT)+Yg(YT))\displaystyle\star d(g\,T^{\ast})=\star(-g\,(X^{\ast}\wedge Y^{\ast})+Xg\,(X^{\ast}\wedge T^{\ast})+Yg\,(Y^{\ast}\wedge T^{\ast}))
=\displaystyle= gTXgY+YgXΩ1,1()\displaystyle-g\,T^{\ast}-Xg\,Y^{\ast}+Yg\,X^{\ast}\in\Omega^{1,\geq 1}(\mathbb{H})

and so dt(Ω2,2())Ω1,1()d^{t}(\Omega^{2,\geq 2}(\mathbb{H}))\subset\Omega^{1,\geq 1}(\mathbb{H}), but not Ω1,2()\Omega^{1,\geq 2}(\mathbb{H}).

Applying similar computations, one can easily check that the transpose of the operator gr(d)Φ{\rm gr}(d)^{\Phi} instead respects the filtration.

A situation similar to Example B.1 also holds for the algebraic part of the de Rham differential d~:Ω(𝔾)Ω(𝔾)\tilde{d}\colon\Omega^{\bullet}(\mathbb{G})\to\Omega^{\bullet}(\mathbb{G}), when considering a nilpotent Lie group 𝔾\mathbb{G} with a filtration on its Lie algebra 𝔤\mathfrak{g} that is not coming from a homogeneous structure as in the following example:

Example B.2.

Let us consider the 4-dimensional Engel group 𝔾\mathbb{G}, that is, the connected simply connected nilpotent Lie group with Lie algebra 𝔤=span{X1,X2,X3,X4}\mathfrak{g}={\rm span}_{\mathbb{R}}\{X_{1},X_{2},X_{3},X_{4}\} and [X1,Xi]=Xi+1[X_{1},X_{i}]=X_{i+1} for i=2,3i=2,3. We identify 𝔤\mathfrak{g} with the space of left-invariant vector fields, and we consider the following left-invariant filtration of T𝔾T\mathbb{G}

{0}=H0H1=span{X1,X2}H2=span{X1,X2,X3,X4}=T𝔾.\displaystyle\{0\}=H^{0}\subset H^{1}={\rm span}_{\mathbb{R}}\{X_{1},X_{2}\}\subset H^{2}={\rm span}_{\mathbb{R}}\{X_{1},X_{2},X_{3},X_{4}\}=T\mathbb{G}\,.

The coframe of the canonical frame 𝕏=(X1,,X2,X3,X4)\mathbb{X}=(X_{1},,X_{2},X_{3},X_{4}) is denoted by Θ=(θ1,θ2,θ3,θ4)\Theta=(\theta^{1},\theta^{2},\theta^{3},\theta^{4}) Then, for an arbitrary form f1θ3+f2θ4f_{1}\,\theta^{3}+f_{2}\,\theta^{4} in Ω1,2(𝔾)\Omega^{1,\geq 2}(\mathbb{G}), with f1,f2C(𝔾)f_{1},f_{2}\in C^{\infty}(\mathbb{G}), we have

d~(f1θ3+f2θ4)=\displaystyle\tilde{d}(f_{1}\,\theta^{3}+f_{2}\,\theta^{4})= f1(θ1θ2)f2(θ1θ3)Ω2,2(𝔾).\displaystyle-f_{1}\,(\theta^{1}\wedge\theta^{2})-f_{2}\,(\theta^{1}\wedge\theta^{3})\ \in\Omega^{2,\geq 2}(\mathbb{G})\,.

However, given a form g(θ1θ3)Ω2,3(𝔾)g\,(\theta^{1}\wedge\theta^{3})\ \in\Omega^{2,\geq 3}(\mathbb{G}) with gC(𝔾)g\in C^{\infty}(\mathbb{G}), we obtain d~t(g(θ1θ3))=gθ4Ω1,2(𝔾)\tilde{d}^{t}(g\,(\theta^{1}\wedge\theta^{3}))=-g\,\theta^{4}\in\Omega^{1,\geq 2}(\mathbb{G}).

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. KK-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 n\mathbb{H}^{n}. 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 KK-theory for the group SU(n,1){\rm SU}(n,1). 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 L2L^{2} des espaces symétriques de rang 11. C. R. Acad. Sci. Paris Sér. I Math., 320(4), 1995.
  • [Jul19] Pierre Julg. How to prove the Baum-Connes conjecture for the groups Sp(n,1)Sp(n,1)? 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 q,p\ell^{q,p} cohomology of Carnot groups. Ann. H. Lebesgue, 1:267–295, 2018.
  • [PT19] Pierre Pansu and Francesca Tripaldi. Averages and the q,1\ell^{q,1} 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.