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

Local Invariants and Geometry of
the sub-Laplacian on H-type foliations

Wolfram Bauer1 1[email protected] Irina Markina2 2[email protected] Abdellah Laaroussi3 3[email protected]  and  Gianmarco Vega-Molino4 4[email protected]
Abstract.

HH-type foliations (𝕄,,g)(\mathbb{M},\mathcal{H},g_{\mathcal{H}}) are studied in the framework of sub-Riemannian geometry with bracket generating distribution defined as the bundle transversal to the fibers. Equipping 𝕄\mathbb{M} with the Bott connection we consider the scalar horizontal curvature κ\kappa_{\mathcal{H}} as well as a new local invariant τ𝒱\tau_{\mathcal{V}} induced from the vertical distribution. We extend recent results on the small-time asymptotics of the sub-Riemannanian heat kernel on quaternion-contact (qc-)manifolds due to A. Laaroussi and we express the second heat invariant in sub-Riemannian geometry as a linear combination of κ\kappa_{\mathcal{H}} and τ𝒱\tau_{\mathcal{V}}. The use of an analog to normal coordinates in Riemannian geometry that are well-adapted to the geometric structure of HH-type foliations allows us to consider the pull-back of Korányi balls to 𝕄\mathbb{M}. We explicitly obtain the first three terms in the asymptotic expansion of their Popp volume for small radii. Finally, we address the question of when 𝕄\mathbb{M} is locally isometric as a sub-Riemannian manifold to its HH-type tangent group.

Key words and phrases:
sub-Riemannian geometry, Bott connection and curvature, Heisenberg type group
2020 Mathematics Subject Classification:
53C17,58J35
The work of the second author was partially supported by the project Pure Mathematics in Norway, funded by Trond Mohn Foundation and TromsøResearch Foundation.
The fourth author is partially supported by the Trond Mohn Foundation - Grant TMS2021STG02 (GeoProCo).

1. Introduction

Sub-Riemannian geometry and the analysis of intrinsically associated differential equations, such as the sub-Riemannian heat or sub-Riemannian wave equation, have attracted increasing interest during the last decades [1, 2, 5, 9, 13, 14, 15, 17, 24, 34, 40]. Sub-Riemannian geometry is the study of bracket generating metric distributions inside the tangent bundle of a given manifold. On the one hand, such geometries allow for the definition of a metric space structure, of tangent groups or geodesic curves. On the other hand, on a regular sub-Riemannian manifold an analytic object, namely the intrinsic hypoelliptic sub-Laplacian, can be defined in form of a second order differential operator [2, 6] and generalizes the Beltrami-Laplace operator in Riemannian geometry. The sub-Riemannian heat kernel KsubK_{\mathrm{sub}} is the fundamental solution of the corresponding heat operator and it is expected to encode specific geometric data of the sub-Riemannian structure. With varying degrees of generality the existence of the small time asymptotic expansion of KsubK_{\mathrm{sub}} or of the sub-Riemannian heat trace on compact manifolds has been obtained by analytical or stochastic methods [14, 17, 23, 24, 28, 40].

In some cases even explicit formulas for the sub-Riemannian heat kernel are available [7, 8, 11, 12, 14] and are of great use in the asymptotic analysis. However, due to a large variety of sub-Riemannian structures (e.g. see [36] for many examples) it seems that a geometric interpretation of the coefficients (heat invariants) in the small time asymptotic expansion of KsubK_{\mathrm{sub}} requires a case-by-case analysis. We mention the results for unimodular Lie groups [40], particular sub-Riemannian manifolds with symmetries [5], sub-Riemannian structures on the sphere 𝕊7\mathbb{S}^{7} [13], and on quaternionic contact (qc) manifolds [34]. In the present paper we study the geometric interpretation of the second heat invariant for the intrinsic sub-Laplacian on HH-type foliations. Introduced in [10], HH-type foliations form good model spaces inside the family of sub-Riemannian manifolds with step two bracket generating distribution. An HH-type foliation 𝕄\mathbb{M} carries an horizontal distribution \mathcal{H} of constant rank with a totally geodesic and integrable complement 𝒱\mathcal{V}. Moreover, 𝕄\mathbb{M} is the total space of a Clifford bundle adapted to the splitting 𝒱\mathcal{H}\oplus\mathcal{V} of the tangent space (see Section 2 for a precise definition). HH-type foliations generalize previously studied examples such as Sasakian manifolds, Twistor spaces or 3-Sasakian manifolds and they carry a canonical connection (Bott-connection) which induces useful notions of horizontal and vertical curvature. We mention that the class of HH-type foliations has a non-empty intersection with qc-manifolds, similar aspects of which have been recently studied by the third author in [34].

The local models (tangent groups) of an H-type foliation 𝕄\mathbb{M} are H(eisenberg)-type Lie groups GG which were introduced by A. Kaplan in the 80th in his study of fundamental solutions to hypoelliptic PDE, cf. [31]. The corresponding HH-type Lie algebras are in one-to-one correspondence with representations of Clifford algebras [22]. Such algebras are nilpotent of step two and the complement of their center (first layer) is naturally identified with a left-invariant bracket generating distribution which we call horizontal. Equipped with an inner product, which is induced by the metric on 𝕄\mathbb{M}, the (tangent) HH-type Lie algebras define the first algebraic and metric invariants of 𝕄\mathbb{M} as a sub-Riemannian manifold (see [37]).

HH-type groups endowed with a left-invariant metric have the structure of a sub-Riemannian manifold themselves and subsequently have become an important class of examples in PDE [20, 25, 38], geometry and analysis [18, 32] or geometric measure theory, [3, 26]; see also the survey paper [41]. In analogy to known results in Riemannian geometry or the analysis in [34] we expect that for sufficiently well-behaved examples the second heat invariant of the sub-Laplacian can be expressed using a suitable concept of curvature. We remark that the horizontal distribution on a step two nilpotent Lie group has vanishing curvature as well as vanishing second heat invariant. The class of HH-type foliations, which we are considering in this paper, is more interesting when studying the effect of curvature terms on the heat kernel analysis of the sub-Laplacian. Roughly speaking, such manifolds may be thought of as “curved versions” of the HH-type groups.

The task of expressing the second heat invariant of the sub-Laplacian on an HH-type foliation in terms of curvature tensors requires the choice of a connection which is suitably adapted to the sub-Riemannian and Clifford bundle structure of 𝕄\mathbb{M}. In this paper we choose the Bott connection in [45, Chapter 5], see also [27] which - different from the Levi-Civita connection in Riemannian geometry - has non-vanishing torsion. In case of a contact manifold or for pseudo-Hermitian or strictly pseudo-convex CR manifolds a natural connection is given by the Tanno connection [44] and the Tanaka-Webster connection [43, 46], respectively. On qc-manifolds we may as well use the Biquard connection [19, 34].

We mention that all these examples have non-empty intersections with the class of HH-type foliations. In these intersections the Bott connection coincides with the Tanno and the Tanaka-Webster connection, respectively, and it differs from the Biquard connection by torsion properties. Different choices of a connection may lead to different representations of the second heat invariant [19].

Our heat kernel analysis is based on an appropriate choice of privileged coordinates on 𝕄\mathbb{M} constructed by Kunkel in [26] (see also [30]) combined with recent methods from [24]. In fact, an inspection of the proofs in [24] shows that the second heat invariant can be obtained as a convolution integral. The integrand is built from the heat kernel of the sub-Laplacian of the local nilpotent approximation (see [14] for an explicit formula) and an operator which, in particular, depends on the curvature tensor, see Eq. 5.3.

Our main result concerning the geometric interpretation of the second heat invariant for an HH-type foliation is stated below as Theorem A and will be explained more in detail next.

Let q𝕄q\in\mathbb{M} and t>0t>0. As is well-known the heat kernel KsubK_{\textup{sub}} of the intrinsic sub-Laplace operator Δsub\Delta_{\textup{sub}} on 𝕄\mathbb{M} has a short time asymptotic expansion of the form

(1.1) Ksub(t;q,q)=tQ2[c0(q)+c1(q)t++cN(q)tN+O(tN+1)],as t0.K_{\textup{sub}}(t;q,q)=t^{-\frac{Q}{2}}\Big{[}c_{0}(q)+c_{1}(q)t+\ldots+c_{N}(q)t^{N}+O(t^{N+1})\Big{]},\hskip 17.22217pt\mbox{\it as }\>t\downarrow 0.

Here QQ denotes the Hausdorff dimension of 𝕄\mathbb{M} as a metric space equipped with the Carnot-Carathéodory metric. The function cj(q)c_{j}(q) with j=0,1,2j=0,1,2\ldots is referred to as the (j+1)(j+1)-th heat invariant of the sub-Riemannian structure.

Let \nabla denote the Bott connection on 𝕄\mathbb{M} with torsion TT. Moreover, κ\kappa_{\mathcal{H}} and τ\tau are the horizontal scalar curvature and an invariant defined through the vertical bundle, respectively. We note that an HH-type foliation 𝕄\mathbb{M} is said to have horizontally parallel torsion when T=0\nabla_{\mathcal{H}}T=0.

Theorem A.

Let (𝕄,,g)(\mathbb{M},\mathcal{H},g_{\mathcal{H}}) be an H-type foliation with horizontally parallel torsion. Then the second heat invariant c1(q)c_{1}(q) is a linear combination of the local geometric invariants κ\kappa_{\mathcal{H}} and τ\tau:

(1.2) c1=C1κ+C2τ,c_{1}=C_{1}\kappa_{\mathcal{H}}+C_{2}\tau,

where C1C_{1} and C2C_{2} are universal constants depending only on nn and mm, the ranks of the horizontal and vertical distribution, respectively.

We address two further problems in connection with HH-type foliations which so far seemed to not have been settled in the literature. The first question concerns the asymptotic behaviour of the Popp volume vol of small balls on 𝕄\mathbb{M}. Let r>0r>0 and consider the preimage B(q,r)𝕄B(q,r)\subset\mathbb{M} around q𝕄q\in\mathbb{M} of a Korányi ball in n+m\mathbb{R}^{n+m} of radius rr under (parabolic) privileged coordinates. In Section 4 it is shown:

Theorem B.

Let q𝕄q\in\mathbb{M}. As r0r\to 0 it holds:

rQvol(B(q,r))=an,mbn,mκ(q)r2+O(r3),r^{-Q}\cdot\mathrm{vol}(B(q,r))=a_{n,m}-b_{n,m}\kappa_{\mathcal{H}}(q)r^{2}+O(r^{3}),

where an,ma_{n,m} and bn,mb_{n,m} are suitable positive constants (which have explicit integral expressions).

Finally, we give a sufficient and necessary condition for an HH-type foliation with horizontally parallel torsion to be locally isometric as a sub-Riemannian manifold to its tangent group (HH-type group).

Theorem C.

Let (𝕄,,g)(\mathbb{M},\mathcal{H},g_{\mathcal{H}}) be an H-type foliation with horizontally parallel torsion. Then, 𝕄\mathbb{M} is locally isometric as a sub-Riemannian manifold to its tangent group if and only if the curvature tensor of the Bott connection vanishes, i.e. R0R\equiv 0.

The structure of the paper is as follows: In Section 2 we recall some basic notions in sub-Riemannian geometry and we introduce HH-type foliations 𝕄\mathbb{M} which form a model class of sub-Riemannian manifolds with a Clifford bundle structure, [10].

We recall the construction of privileged coordinates [33] in Section 3. Throughout the paper all local calculations will be performed within such coordinates which are well-adapted to the sub-Riemannian structure on 𝕄\mathbb{M}. In particular, the coordinate system obeys linear scaling in horizontal directions much like Riemannian normal coordinates, but obeys parabolic scaling in vertical directions reflecting the bracket structure. An essential ingredient to our analysis is the calculation of the homogeneous parts of the associated co-frame and connection one-forms in 3.3. Finally, C (3.7) is achieved.

In Section 4 we calculate the Popp volume 𝒫\mathcal{P} on an HH-type foliation which is used in the definition of the intrinsic sub-Laplacian Δsub\Delta_{\mathrm{sub}}. A local expression of Δsub\Delta_{\mathrm{sub}} is then given in 4.4. Section 4 contains a proof of B (4.3) as its main result.

Based on our calculations in local privileged coordinates and methods from [24] we introduce a new invariant τ\tau in Section 5 and derive a representation of the second heat invariant as a linear combination of τ\tau and the horizontal scalar curvature κ\kappa_{\mathcal{H}} with respect to the Bott connection. This result can be further refined when assuming horizontally parallel Clifford structure. The section concludes by proving 5.12 (A).

The paper ends with an appendix which collects various technical calculations.

2. Sub-Riemannian geometry and H-type foliations

In the setting of sub-Riemannian geometry, one works with a smooth manifold 𝕄\mathbb{M} of dimension n+mn+m equipped with a pair (,g)(\mathcal{H},g_{\mathcal{H}}) where \mathcal{H} is a rank nn subbundle of T𝕄T\mathbb{M} and gg_{\mathcal{H}} is a symmetric, positive-definite (2,0)-tensor on \mathcal{H}. We insist that the bracket-generating condition holds, which we say is satisfied if at every point p𝕄p\in\mathbb{M} one can generate all of Tp𝕄T_{p}\mathbb{M} by taking sufficiently many Lie brackets of vector fields in Γ()\Gamma(\mathcal{H}) at pp. Notably, this is equivalent to Hörmander’s condition in PDEs [36]. One can define the sub-Riemannian (Carnot-Carathéodory) distance dcc(p,q)d_{cc}(p,q) for two points p,q𝕄p,q\in\mathbb{M} by the usual infimum formula taken over the space of smooth paths C,p,qC_{\mathcal{H},p,q} connecting pp to qq such that γ˙(t)γ(t)\dot{\gamma}(t)\in\mathcal{H}_{\gamma(t)} at almost every point along γ\gamma. The now famous theorem of Chow [21] and Rashevsky [39] tells us that (𝕄,dcc)(\mathbb{M},d_{cc}) is a metric space when \mathcal{H} is bracket-generating.

Such a triple (𝕄,,g)(\mathbb{M},\mathcal{H},g_{\mathcal{H}}) is called a sub-Riemannian manifold. These objects arise naturally in myriad settings and have been extensively studied in recent decades with increasing interest, see [1, 36, 42] for an overview.

An established approach is to study sub-Riemannian manifolds by equipping a Riemannian penalty metric

gε=g1εg𝒱g_{\varepsilon}=g_{\mathcal{H}}\oplus\frac{1}{\varepsilon}g_{\mathcal{V}}

parameterized by ε>0\varepsilon>0 on an appropriately chosen complementary distribution 𝒱\mathcal{V}.

Notation 2.1.

In the sequel we will write ,𝒱\mathcal{H},\mathcal{V} as subscripts to denote appropriate projections of tensor fields.

In some key examples there is a natural choice of a complement, but generically this is not the case. However, when such a choice can be made one can then attempt to recover purely sub-Riemannian results by working with Riemannian tools on (𝕄,gε)(\mathbb{M},g_{\varepsilon}) and considering the limit ε0+\varepsilon\rightarrow 0^{+}, however one should understand that Riemannian definitions of curvature do not make sense in this limit.

2.1. H-type foliations

We now review the notion of H-type foliations introduced in [10] and motivated by earlier results of [9]. In particular, there is a sub-class of H-type foliations associated in an essential way to qc-manifolds equipped with a totally-geodesic foliation, as appear in [34] which motivates the present paper.

Let (𝕄,g)(\mathbb{M},g) be a smooth, oriented, connected Riemannian manifold of dimension n+mn+m. To keep formulas shorter we will also use the notation g(,)=,g(\cdot,\cdot)=\langle\cdot,\cdot\rangle for the Riemannian metric and occasionally we write X2=g(X,X)\|X\|^{2}=g(X,X). We assume that 𝕄\mathbb{M} is equipped with a Riemannian foliation with bundle-like complete metric and totally geodesic mm-dimensional leaves. The subbundle 𝒱\mathcal{V} formed by vectors tangent to the leaves is referred to as the vertical distribution. The subbundle \mathcal{H} orthogonal to 𝒱\mathcal{V} will be called the horizontal distribution.

We recall that the foliation is called totally-geodesic if the leaves of the foliation are totally-geodesic submanifolds, and said to have bundle-like metric if geodesics that are anywhere tangent to \mathcal{H} remain always tangent to \mathcal{H}. These conditions have the (respective) characterizations with respect to the Lie derivative [45]

(2.1) (Zg)(X,X)=0,(Xg)(Z,Z)=0(\mathcal{L}_{Z}g)(X,X)=0,\qquad(\mathcal{L}_{X}g)(Z,Z)=0

for vector fields XΓ(),ZΓ(𝒱)X\in\Gamma(\mathcal{H}),Z\in\Gamma(\mathcal{V}).

On such a manifold there is a canonical connection \nabla that preserves the metric and the foliation structure, the Bott connection, see [10, 45] and also a generalization to equiregular higher step sub-Riemannian manifolds, the Hladky connection [27].

Definition 2.2.

The Bott connection on a totally-geodesic foliation with bundle-like metric is uniquely characterized by the following properties:

  1. (1)

    (Metric) g=0\nabla g=0,

  2. (2)

    (Compatible) For XΓ(T𝕄)X\in\Gamma(T\mathbb{M}), X and X𝒱𝒱.\nabla_{X}\mathcal{H}\subseteq\mathcal{H}\text{ and }\nabla_{X}\mathcal{V}\subseteq\mathcal{V}.

  3. (3)

    (Torsion) The torsion T(X,Y):=XYYX[X,Y]T(X,Y):=\nabla_{X}Y-\nabla_{Y}X-[X,Y] satisfies

    1. (a)

      T(,)𝒱,T(\mathcal{H},\mathcal{H})\subseteq\mathcal{V},

    2. (b)

      T(,𝒱)=T(𝒱,𝒱)=0.T(\mathcal{H},\mathcal{V})=T(\mathcal{V},\mathcal{V})=0.

For every ZΓ(𝒱)Z\in\Gamma(\mathcal{V}) define a bundle endomorphism JZ:J_{Z}:\mathcal{H}\rightarrow\mathcal{H} by

(2.2) g(JZX,Y)=g(Z,T(X,Y)).g(J_{Z}X,Y)=g(Z,T(X,Y)).
Lemma 2.3.

Suppose the H-type condition

(2.3) JZ2=Z2IdJ_{Z}^{2}=-\|Z\|^{2}\mathrm{Id}_{\mathcal{H}}

is satisfied. Then the tangent space Tq𝕄T_{q}\mathbb{M} at any point q𝕄q\in\mathbb{M} is generated by [X,]q[X,\mathcal{H}]_{q} and q\mathcal{H}_{q} for every XΓ()X\in\Gamma(\mathcal{H}) such that Xq0X_{q}\neq 0.

Such a distribution is also called strongly bracket generating or fat, cf. [35, 42].

Proof.

Observe that the torsion of the Bott connection is given by

(2.4) T(X,Y)=[X,Y]𝒱.T(X,Y)=-[X_{\mathcal{H}},Y_{\mathcal{H}}]_{\mathcal{V}}.

Letting XΓ(),ZΓ(𝒱)X\in\Gamma(\mathcal{H}),Z\in\Gamma(\mathcal{V}), we have the following relation for the bundle map JZJ_{Z}

(2.5) g(Z,T(X,JZX))=g(JZX,JZX)=g(X,JZ2X)=g(Z,X2Z).g\left(Z,T(X,J_{Z}X)\right)=g\left(J_{Z}X,J_{Z}X\right)=-g\left(X,J^{2}_{Z}X\right)=g\left(Z,\|X\|^{2}Z\right).

Combining the last 2 equalities we obtain

(2.6) [X,JZX]𝒱=X2Z[X,J_{Z}X]_{\mathcal{V}}=-\|X\|^{2}Z

and the claim follows. ∎

Definition 2.4.

If the H-type condition is satisfied then the sub-Riemannian manifold (𝕄,,g)(\mathbb{M},\mathcal{H},g_{\mathcal{H}}) is called an H-type foliation.

Remark 2.5.

We observe that it can be at first counter-intuitive to call (𝕄,,g)(\mathbb{M},\mathcal{H},g_{\mathcal{H}}) a foliation as \mathcal{H} is bracket generating, however we emphasize that the foliation tangent to 𝒱\mathcal{V} uniquely determines its gg-complement \mathcal{H}; since we focus on these objects’ structure as sub-Riemannian manifolds, we will continue to emphasize the sub-Riemannian distribution. See also the related [10, Remark 2.3].

Note that due to polarization the H-type condition is equivalent to

(2.7) JZJW+JWJZ=2g(Z,W)Id,Z,WΓ(𝒱).J_{Z}J_{W}+J_{W}J_{Z}=-2g(Z,W)\mathrm{Id}_{\mathcal{H}},\qquad Z,W\in\Gamma(\mathcal{V}).

2.2. Curvature and frames

Let us denote by RR the curvature tensor of the Bott connection defined by

(2.8) R(X,Y)Z:=XYZYXZ[X,Y]Z.\displaystyle R(X,Y)Z:=\nabla_{X}\nabla_{Y}Z-\nabla_{Y}\nabla_{X}Z-\nabla_{[X,Y]}Z.
Notation 2.6.

In the following we denote by {θ1,,θn}\left\{\theta^{1},\cdots,\theta^{n}\right\} (resp. {η1,,ηm}\left\{\eta^{1},\cdots,\eta^{m}\right\}) the metric dual frame of a horizontal frame {X1,,Xn}\left\{X_{1},\cdots,X_{n}\right\} of \mathcal{H} (resp. vertical frame {Z1,,Zm}\left\{Z_{1},\cdots,Z_{m}\right\} of 𝒱\mathcal{V}).

It is occasionally convenient to have a notation for the entire frame and coframe. In these cases we may write {Y1,,Yn+m}\{Y_{1},\cdots,Y_{n+m}\} and {ν1,,νn+m}\{\nu^{1},\cdots,\nu^{n+m}\} and one should interpret Ya=XaY_{a}=X_{a} for a{1,,n},Yn+a=Zaa\in\{1,\cdots,n\},Y_{n+a}=Z_{a} for a{1,,m}a\in\{1,\cdots,m\}, and analogously for the νa\nu^{a}. Furthermore, in order to have a consistent index notation we will use different letters for different ranges of indices as follows:

(2.9) a,b,c,d{1,,n+m},α,β,γ,δ{1,,n},i,j,k{1,,m}.a,b,c,d\in\{1,\cdots,n+m\},\hskip 5.69054pt\alpha,\beta,\gamma,\delta\in\{1,\cdots,n\},\quad i,j,k\in\{1,\cdots,m\}.

With this convention we set

(2.10) Jαβi:=g(JZiXα,Xβ),Tαβi:=ηi(T(Xα,Xβ)) and Rabcd:=νd(R(Ya,Yb)Yc)J_{\alpha\beta}^{i}:=g(J_{Z_{i}}X_{\alpha},X_{\beta}),\quad T_{\alpha\beta}^{i}:=\eta^{i}\left(T(X_{\alpha},X_{\beta})\right)\text{ and }R_{abc}^{d}:=\nu^{d}\left(R(Y_{a},Y_{b})Y_{c}\right)

and define the horizontal scalar curvature induced by the Bott connection

(2.11) κ=α,β=1nRαββα.\kappa_{\mathcal{H}}=\sum_{\alpha,\beta=1}^{n}R_{\alpha\beta\beta}^{\alpha}.

Throughout the text we frequently apply the first Bianchi identity for a connection with torsion,

(2.12) R(X,Y)Z=XT(Y,Z)+T(T(Y,Z),X),\circlearrowright R(X,Y)Z=\circlearrowright\nabla_{X}T(Y,Z)+\circlearrowright T(T(Y,Z),X),

which in case of the Bott connection the torsion properties imply that this takes the simpler form

(2.13) R(X,Y)Z=XT(Y,Z)Γ(𝒱)\circlearrowright R(X,Y)Z=\circlearrowright\nabla_{X}T(Y,Z)\in\Gamma(\mathcal{V})

with \circlearrowright denoting the cyclic sum.

Notation 2.7.

In the sequel we use a nonstandard summation convention: whenever any index is repeated twice it must be summed. When unclear, we will make the summations explicit. For example, (2.11) becomes

(2.14) κ=Rαββα.\kappa_{\mathcal{H}}=R_{\alpha\beta\beta}^{\alpha}.

in this notation.

Take care to note that this is not the standard Einstein summation convention, wherein repeated indices are only summed when they appear as both a superscript and subscript.

3. Construction of privileged coordinates

Consider n+m=nm\mathbb{R}^{n+m}=\mathbb{R}^{n}\oplus\mathbb{R}^{m} with coordinates (x,z)=(x1,,xn,z1,,zm)(x,z)=(x^{1},\ldots,x^{n},z^{1},\ldots,z^{m}) and the natural dilation δt:n+mn+m\delta_{t}\colon\mathbb{R}^{n+m}\rightarrow\mathbb{R}^{n+m} defined for t>0t>0 by

(3.1) δt(x,z)=(tx,t2z).\delta_{t}(x,z)=(tx,t^{2}z).

We associate the following weights ww for the generators 11, xαx^{\alpha}, ziz^{i}, xα\frac{\partial}{\partial x^{\alpha}}, zi\frac{\partial}{\partial z^{i}} of polynomial vector fields:

(3.2) w(1)=0,w(xα)=1,w(zi)=2,w(xα)=1,w(zi)=2.w(1)=0,\quad w(x^{\alpha})=1,\quad w(z^{i})=2,\quad w\left(\frac{\partial}{\partial x^{\alpha}}\right)=-1,\quad w\left(\frac{\partial}{\partial z^{i}}\right)=-2.

A formal monomial is a product of the generators, the order of which is the sum of weights of each term. A polynomial vector field on n+m\mathbb{R}^{n+m} (considered as a differential operator of first order) is said to be homogeneous if it is a sum of monomials of the same weight. For any smooth vector field, applying the Taylor expansion for the coefficients, one can rearrange the terms into a sum of homogeneous terms

(3.3) Xl=2+X(l),X\sim\sum_{l=-2}^{+\infty}X^{(l)},

where X(l)X^{(l)} is a homogeneous vector field of order ll. Written in this form, we will call this a parabolic Taylor expansion.

We say that a system of coordinates Ψ:Un+m\Psi\colon U\to\mathbb{R}^{n+m} centered at qU𝕄q\in U\subseteq\mathbb{M} is privileged if for any horizontal vector field XΓ()X\in\Gamma(\mathcal{H}) the terms of the homogeneous order 2-2 vanish, see more details in [1, Chapter 10.4] or [16].

Now we will construct a system of privileged coordinates that is convenient for the further calculations; it is reminiscent of normal Euclidean coordinates and normal coordinates for CR manifolds [30, 33]. Let q𝕄q\in\mathbb{M} be fixed. Consider an arbitrary tangent vector X+Zq𝒱q=Tq𝕄X+Z\in\mathcal{H}_{q}\oplus\mathcal{V}_{q}=T_{q}\mathbb{M} and define γX+Z\gamma_{X+Z} to be the curve starting at qq with

γX+Z(0)=q,γ˙X+Z(0)=X,Dtγ˙X+Z|t=0=Z and Dt2γ˙X+Z=0,\gamma_{X+Z}(0)=q,\hskip 5.69054pt\dot{\gamma}_{X+Z}(0)=X,\hskip 5.69054pt\left.D_{t}\dot{\gamma}_{X+Z}\right|_{t=0}=Z\text{ and }D_{t}^{2}\dot{\gamma}_{X+Z}=0,

where DtD_{t} denotes the covariant derivative along γX+Z\gamma_{X+Z}. According to [33] (see also [30]), there are neighborhoods 0OTq𝕄0\in O\subset T_{q}\mathbb{M} and qU𝕄q\in U\subset\mathbb{M} so that the map φ:OU\varphi:O\longrightarrow U defined by

φ(X+Z):=γX+Z(1)\varphi(X+Z):=\gamma_{X+Z}(1)

is a diffeomorphism, and satisfies the parabolic scaling φ(tX+t2Z)=γX+Z(t)\varphi(tX+t^{2}Z)=\gamma_{X+Z}(t) whenever either side is defined. Such a curve γX+Z\gamma_{X+Z} will be called a parabolic geodesic.

Fix a horizontal (resp. vertical) orthonormal frame {X1(q),,Xn(q)}\{X_{1}(q),\dots,X_{n}(q)\} (resp. {Z1(q),,Zm(q)}\{Z_{1}(q),\dots,Z_{m}(q)\}) at qq and extend these vectors to be parallel along parabolic geodesics starting at qq. It is known (cf. [33]) that such extensions necessarily are smooth in a neighbourhood of q𝕄q\in\mathbb{M}. In this way, we obtain a smooth local frame {X1,,Xn,Z1,,Zm}\{X_{1},\dots,X_{n},Z_{1},\dots,Z_{m}\} for T𝕄=𝒱T\mathbb{M}=\mathcal{H}\oplus\mathcal{V}. Note that the corresponding dual frame {θ1,,θn,η1,,ηm}\{\theta^{1},\dots,\theta^{n},\eta^{1},\dots,\eta^{m}\} is also parallel along parabolic geodesics. In the following we call such a frame a special frame.

We define a privileged coordinate system (x,z)(x,z) on UU by composing the inverse of φ\varphi with the map λ:Tq𝕄n+m\lambda:T_{q}\mathbb{M}\longrightarrow\mathbb{R}^{n+m} defined by λ(W):=(θα(W),ηi(W))\lambda(W):=(\theta^{\alpha}(W),\eta^{i}(W)) (cf. [1, Example 10.31 (ii)]). Note that if we consider another horizontal (resp. vertical) frame at qq, then these frames are related via an orthogonal transformation. Hence, the resulting coordinates are related via a block diagonal matrix with orthogonal entries, i.e. a matrix of the form

(3.4) (A00B)𝐎(n+m)\begin{pmatrix}A&0\\ 0&B\end{pmatrix}\in{\bf O}(n+m)

with A𝐎(n)A\in{\bf O}(n) and B𝐎(m)B\in{\bf O}(m), where 𝐎(k){\bf O}(k) with kk\in\mathbb{N} denotes the group of orthogonal k×kk\times k real matrices.

From now on we will work with this special frame and co-frame and frequently write

(3.5) Ji:=JZi for i=1,,m.J_{i}:=J_{Z_{i}}\text{ for }i=1,\cdots,m.

The infinitesimal generator P~\tilde{P} of the +\mathbb{R}^{+}-action δt\delta_{t} (Eq. 3.1) on n+m\mathbb{R}^{n+m} is given in these coordinates by

(3.6) P~(x,z):=α=1nxαxα+2i=1mzizi.\tilde{P}_{(x,z)}:=\sum_{\alpha=1}^{n}x^{\alpha}\frac{\partial}{\partial x^{\alpha}}+2\sum_{i=1}^{m}z^{i}\frac{\partial}{\partial z^{i}}.

We adapt the notion of homogeneity to tensor fields on 𝕄\mathbb{M}. We will denote by PP the pullback of P~\tilde{P} to Γ(T𝕄)\Gamma(T\mathbb{M}) and say that a tensor field Θ\Theta is homogeneous of order l=ord(Θ)l=\mathrm{ord}(\Theta)\in\mathbb{Z} with respect to the above dilations if

(3.7) P(Θ)=lΘ,\mathcal{L}_{P}(\Theta)=l\Theta,
Proposition 3.1.

Whenever they make sense, the following properties hold.

  1. (1)

    Let XX and YY be homogeneous vector fields. Then [X,Y][X,Y] is homogeneous of order ord(X)+ord(Y)\mathrm{ord}(X)+\mathrm{ord}(Y).

  2. (2)

    Let XX (resp. ν\nu) be a homogeneous vector field (resp. 11-form). Then ν(X)\nu(X) is homogeneous of order ord(ν)+ord(X)\mathrm{ord}(\nu)+\mathrm{ord}(X).

  3. (3)

    Let ff (resp. XX) be a homogeneous function (resp. vector field). Then fXfX is homogeneous of order ord(f)+ord(X)\mathrm{ord}(f)+\mathrm{ord}(X).

We now prove a lemma crucial for our asymptotic analysis on H-type foliations, (cf. [30, 33]):

Lemma 3.2.

In the above privileged coordinates, it holds:

(3.8) θα(P)=xα,ηi(P),=zi and ωba(P)=0.\theta^{\alpha}(P)=x^{\alpha},\eta^{i}(P),=z^{i}\text{ and }\omega^{a}_{b}(P)=0.

Here ωba\omega^{a}_{b} denotes the connection 11-forms w.r.t. the constructed special frame, i.e.

(3.9) Xb=Xaωba.\nabla X_{b}=X_{a}\otimes\omega^{a}_{b}.

In particular, at p𝕄p\in\mathbb{M} with the coordinates (x,z)=(x1,,xn,z1,,zm)n+m(x,z)=(x^{1},\dots,x^{n},z^{1},\dots,z^{m})\in\mathbb{R}^{n+m} we can write

Pp=α=1nxαXα+i=1mziZi.P_{p}=\sum_{\alpha=1}^{n}x^{\alpha}X_{\alpha}+\sum_{i=1}^{m}z^{i}Z_{i}.
Proof.

It will suffice to prove the statement along parabolic geodesics. Construct a privileged coordinate system on a neighborhood UU around q𝕄q\in\mathbb{M} as explained above, and denote the coordinate map by Ψ:=λφ1:Un+m\Psi:=\lambda\circ\varphi^{-1}\colon U\rightarrow\mathbb{R}^{n+m}.

Let

W:=xαXα(q)+ziZi(q)Tq𝕄,W:=\sum x^{\alpha}X_{\alpha}(q)+\sum z^{i}Z_{i}(q)\in T_{q}\mathbb{M},

be the vector written in the special frame {X1,,Xn,Z1,,Zm}\{X_{1},\cdots,X_{n},Z_{1},\cdots,Z_{m}\}, where (x,z)=Ψ(p)n+m(x,z)=\Psi(p)\in\mathbb{R}^{n+m}. The parabolic geodesic generated by WW can be written in local coordinates as

(3.10) Ψ(γW(t))=λ(tX(q)+t2Z(q))=(tx,t2z)=δt(x,z).\Psi(\gamma_{W}(t))=\lambda(tX(q)+t^{2}Z(q))=(tx,t^{2}z)=\delta_{t}(x,z).

By definition

(3.11) P~Ψ(γW(t))=txαxα+2t2zizi\tilde{P}_{\Psi(\gamma_{W}(t))}=\sum tx^{\alpha}\frac{\partial}{\partial x^{\alpha}}+\sum 2t^{2}z^{i}\frac{\partial}{\partial z^{i}}

and so

(3.12) γ˙W(t)=ddt(Ψ1δt)(x,z)=Ψ(xαxα+2tzizi)Ψ(γW(t))=t1(ΨP~)Ψ(γW(t))=t1PγW(t).\begin{split}\dot{\gamma}_{W}(t)&=\frac{d}{dt}(\Psi^{-1}\circ\delta_{t})(x,z)\\ &=\Psi^{*}\left(\sum x^{\alpha}\frac{\partial}{\partial x^{\alpha}}+\sum 2tz^{i}\frac{\partial}{\partial z^{i}}\right)_{\Psi(\gamma_{W}(t))}\\ &=t^{-1}(\Psi^{*}\tilde{P})_{\Psi(\gamma_{W}(t))}\\ &=t^{-1}P_{\gamma_{W}(t)}.\end{split}

We now observe that for any YY along the geodesic γW(t)\gamma_{W}(t) in the special frame one has

(3.13) ddtDtγ˙W,Y=(γ˙Wg)(Dtγ˙W,Y)+Dt2γ˙W,Y+Dtγ˙W,DtY=0\frac{d}{dt}\langle D_{t}\dot{\gamma}_{W},Y\rangle=(\nabla_{\dot{\gamma}_{W}}g)(D_{t}\dot{\gamma}_{W},Y)+\langle D^{2}_{t}\dot{\gamma}_{W},Y\rangle+\langle D_{t}\dot{\gamma}_{W},D_{t}Y\rangle=0

and so it follows from the initial condition that Dtγ˙W=ZD_{t}\dot{\gamma}_{W}=Z is parallel, or equivalently the vector field ΨZ\Psi_{*}Z has constant coefficients in n+m\mathbb{R}^{n+m}. For any 1-form ν\nu parallel along γW\gamma_{W} we have that

(3.14) ddt(ν(γ˙W(t)))=ν(Dtγ˙W(t))=ν(Z).\frac{d}{dt}(\nu(\dot{\gamma}_{W}(t)))=\nu(D_{t}\dot{\gamma}_{W}(t))=\nu(Z).

We recover a system of equations for the special co-frame {θα,ηi}\{\theta^{\alpha},\eta^{i}\}

(3.15) {ddt(θα(γ˙W(t)))=0θα(γ˙W(0))=xα{ddt(ηi(γ˙W(t)))=ziηi(γ˙W(0))=0\begin{cases}\frac{d}{dt}(\theta^{\alpha}(\dot{\gamma}_{W}(t)))&=0\\ \theta^{\alpha}(\dot{\gamma}_{W}(0))&=x^{\alpha}\end{cases}\qquad\begin{cases}\frac{d}{dt}(\eta^{i}(\dot{\gamma}_{W}(t)))&=z^{i}\\ \eta^{i}(\dot{\gamma}_{W}(0))&=0\end{cases}

from which we obtain the expressions

(3.16) θα(γ˙W(t))=xα,ηi(γ˙W(t))=tzi.\theta^{\alpha}(\dot{\gamma}_{W}(t))=x^{\alpha},\qquad\eta^{i}(\dot{\gamma}_{W}(t))=tz^{i}.

Applying this with (3.12),

(3.17) θα(PγW(t))=txα,ηi(PγW(t))=t2zi,\theta^{\alpha}(P_{\gamma_{W}(t)})=tx^{\alpha},\qquad\eta^{i}(P_{\gamma_{W}(t)})=t^{2}z^{i},

and so we can conclude

(3.18) PγW(t)=α=1ntxαXα+i=1mt2ziZi.P_{\gamma_{W}(t)}=\sum_{\alpha=1}^{n}tx^{\alpha}X_{\alpha}+\sum_{i=1}^{m}t^{2}z^{i}Z_{i}.

In particular, for pUp\in U with parabolic coordinates (x,z)(x,z), we have

(3.19) Pp=PγW(1)=α=1nxαXα+i=1mziZiP_{p}=P_{\gamma_{W}(1)}=\sum_{\alpha=1}^{n}x^{\alpha}X_{\alpha}+\sum_{i=1}^{m}z^{i}Z_{i}

as desired.

For the connection forms, observe that

ωba(P)=g(Xb,PXa)=tg(Xb,DtXa)\omega^{a}_{b}(P)=g(X_{b},\nabla_{P}X_{a})=tg(X_{b},D_{t}X_{a})

follows from the fact P=tγ˙WP=t\dot{\gamma}_{W} for some t>0t>0 with initial vector WW. Since the frame is parallel along any parabolic geodesic γW\gamma_{W}, we obtain that DtXa=0D_{t}X_{a}=0 which concludes the proof. ∎

As mentioned in [30], if ν\nu is a differential form, then the homogeneous part ν(l)\nu^{(l)} of degree l>0l>0 of its parabolic Taylor expansion can be computed by the formula

(3.20) ν(l)=1l(P(ν))(l)=1l(Pdν+d(Pν))(l) for l1.\nu^{(l)}=\frac{1}{l}\left(\mathcal{L}_{P}(\nu)\right)^{(l)}=\frac{1}{l}\left(P\lrcorner\hskip 1.42262ptd\nu+d(P\lrcorner\hskip 1.42262pt\nu)\right)^{(l)}\text{ for }l\geq 1.

Furthermore, using the structure equations (see 7.1 in the Appendix) of the Riemannian foliation 𝕄\mathbb{M} (w.r.t. the special frame and co-frame):

(3.21) {dθα=θβωβα,dηi=12Jαβiθαθβ+ηjωji,dωab=12Rαβabθαθβ+Rαiabθαηi+12Rjkabηjηk+ωacωcb.\begin{cases}d\theta^{\alpha}&=\theta^{\beta}\wedge\omega_{\beta}^{\alpha},\\ d\eta^{i}&=\frac{1}{2}J^{i}_{\alpha\beta}\theta^{\alpha}\wedge\theta^{\beta}+\eta^{j}\wedge\omega_{j}^{i},\\ d\omega_{a}^{b}&=\frac{1}{2}R_{\alpha\beta a}^{b}\theta^{\alpha}\wedge\theta^{\beta}+R_{\alpha ia}^{b}\theta^{\alpha}\wedge\eta^{i}+\frac{1}{2}R_{jka}^{b}\eta^{j}\wedge\eta^{k}+\omega_{a}^{c}\wedge\omega_{c}^{b}.\\ \end{cases}

We then obtain exact expressions for the homogeneous parts of the special co-frame and connection 11-forms.

Proposition 3.3.

In privileged coordinates (xα,zi)(x^{\alpha},z^{i}), we have θα(0)=ηi(0)=0\theta^{\alpha(0)}=\eta^{i(0)}=0 and the higher order components and connection 11-forms are given for l1l\geq 1 by

(3.22) {θα(l)=1l(dxα+xβωβα)(l),ηi(l)=1l(dzi+zjωji+Jαβixαθβ)(l),ωab(l)=1l(Rαβabxαθβ+Rαiabxαηi+Riαabziθα+Rjkabzjηk)(l).\begin{cases}\theta^{\alpha(l)}&=\frac{1}{l}\left(dx^{\alpha}+x^{\beta}\omega_{\beta}^{\alpha}\right)^{(l)},\\ \eta^{i(l)}&=\frac{1}{l}\left(dz^{i}+z^{j}\omega_{j}^{i}+J_{\alpha\beta}^{i}x^{\alpha}\theta^{\beta}\right)^{(l)},\\ \omega_{a}^{b(l)}&=\frac{1}{l}\left(R_{\alpha\beta a}^{b}x^{\alpha}\theta^{\beta}+R_{\alpha ia}^{b}x^{\alpha}\eta^{i}+R_{i\alpha a}^{b}z^{i}\theta^{\alpha}+R_{jka}^{b}z^{j}\eta^{k}\right)^{(l)}.\\ \end{cases}
Proof.

The identity θα(0)=ηi(0)=0\theta^{\alpha(0)}=\eta^{i(0)}=0 follows from the observation that the one forms θα\theta^{\alpha} and ηi\eta^{i} are linear combinations of dxαdx^{\alpha} and dzidz^{i} which are of homogeneous order one and two, respectively. The remaining identities follow from Eq. 3.21 and Eq. 3.20. ∎

Applying the proposition, we recover the particular expressions.

ll θα(l)\theta^{\alpha(l)} ηi(l)\eta^{i(l)} ωab(l)\omega^{b(l)}_{a}
1 dxαdx^{\alpha} 0 0
2 0 12(dzi+Jαβi(q)xαdxβ)\frac{1}{2}(dz^{i}+J_{\alpha\beta}^{i}(q)x^{\alpha}dx^{\beta}) 12Rαβab(q)xαdxβ\frac{1}{2}R_{\alpha\beta a}^{b}(q)x^{\alpha}dx^{\beta}
3 16Rγδβα(q)xβxγdxδ\frac{1}{6}R_{\gamma\delta\beta}^{\alpha}(q)x^{\beta}x^{\gamma}dx^{\delta} 13Jαβi(1)xαdxβ\frac{1}{3}J_{\alpha\beta}^{i(1)}x^{\alpha}dx^{\beta}
4 14(zjωji(2)+Jαβi(q)xαθβ(3)+Jαβi(2)xαdxβ)\frac{1}{4}(z^{j}\omega_{j}^{i(2)}+J_{\alpha\beta}^{i}(q)x^{\alpha}\theta^{\beta(3)}+J_{\alpha\beta}^{i(2)}x^{\alpha}dx^{\beta})

The following properties will be useful in the sequel.

Lemma 3.4.

In privileged coordinates (xα,zi)(x^{\alpha},z^{i}), it holds:

  1. (1)

    The homogeneous terms of JαβiJ_{\alpha\beta}^{i} of order one and two are given by

    (3.23) Jαβi(1)\displaystyle J_{\alpha\beta}^{i(1)} =xγXγ(Jαβi)(q),\displaystyle=x^{\gamma}X_{\gamma}(J_{\alpha\beta}^{i})(q),
    (3.24) Jαβi(2)\displaystyle J_{\alpha\beta}^{i(2)} =12(xγxδXγXδ(Jαβi)(q)+zjZj(Jαβi)(q)).\displaystyle=\frac{1}{2}\left(x^{\gamma}x^{\delta}X_{\gamma}X_{\delta}(J_{\alpha\beta}^{i})(q)+z^{j}Z_{j}(J_{\alpha\beta}^{i})(q)\right).
  2. (2)

    The connection 11-forms ωab\omega_{a}^{b} vanish at the point qq.

  3. (3)

    The horizontal derivatives of JαβiJ_{\alpha\beta}^{i} at qq are related to the torsion TT by

    Xγ(Jαβi)(q)=gq(Zi,(XγT)(Xα,Xβ)).X_{\gamma}(J_{\alpha\beta}^{i})(q)=g_{q}\big{(}Z_{i},(\nabla_{X_{\gamma}}T)(X_{\alpha},X_{\beta})\big{)}.
Proof.

(1): Let ff be a smooth function near q𝕄q\in\mathbb{M} and consider its parabolic Taylor expansion around qq in privileged coordinates (xα,zi)(x^{\alpha},z^{i}), namely:

f=f(q)+l>0f(l),f=f(q)+\sum_{l>0}f^{(l)},

where f(l)f^{(l)} denotes the ll-homogeneous part of ff. Applying the Lie derivative P\mathcal{L}_{P} to both sides of this expansion we find

f(l)=1l(P(f))(l), for l1.f^{(l)}=\frac{1}{l}\big{(}P(f)\big{)}^{(l)},\text{ for }l\geq 1.

According to 3.2 and choosing f=Jαβif=J_{\alpha\beta}^{i} we obtain

Jαβi(l)=1l(P(Jαβi))(l).J_{\alpha\beta}^{i(l)}=\frac{1}{l}\left(P(J_{\alpha\beta}^{i})\right)^{(l)}.

Hence, it follows that

(3.25) Jαβi(1)=(P(Jαβi))(1)=(xγXγ(Jαβi)+zjZj(Jαβi))(1)=xγXγ(Jαβi)(q)\displaystyle J_{\alpha\beta}^{i(1)}=\left(P(J_{\alpha\beta}^{i})\right)^{(1)}=\big{(}x^{\gamma}X_{\gamma}(J_{\alpha\beta}^{i})+z^{j}Z_{j}(J_{\alpha\beta}^{i})\big{)}^{(1)}=x^{\gamma}X_{\gamma}(J_{\alpha\beta}^{i})(q)

and similarly

(3.26) Jαβi(2)\displaystyle J_{\alpha\beta}^{i(2)} =12(xγXγ(Jαβi)+zjZj(Jαβi))(2)\displaystyle=\frac{1}{2}\big{(}x^{\gamma}X_{\gamma}(J_{\alpha\beta}^{i})+z^{j}Z_{j}(J_{\alpha\beta}^{i})\big{)}^{(2)}
(3.27) =12(xγ(Xγ(Jαβi))(1)+zjZj(Jαβi)(q))\displaystyle=\frac{1}{2}\big{(}x^{\gamma}\big{(}X_{\gamma}(J_{\alpha\beta}^{i})\big{)}^{(1)}+z^{j}Z_{j}(J_{\alpha\beta}^{i})(q)\big{)}
(3.28) =12(xγxδXγXδ(Jαβi)(q)+zjZj(Jαβi)(q)).\displaystyle=\frac{1}{2}\big{(}x^{\gamma}x^{\delta}X_{\gamma}X_{\delta}(J_{\alpha\beta}^{i})(q)+z^{j}Z_{j}(J_{\alpha\beta}^{i})(q)\big{)}.

(2): In privileged coordinates (xα,zi)(x^{\alpha},z^{i}) we write ωab=Caαbdxα+Daibdzi\omega_{a}^{b}=C_{a\alpha}^{b}dx^{\alpha}+D_{ai}^{b}dz^{i} with some smooth functions CaαbC_{a\alpha}^{b} and DaibD_{ai}^{b}. Then, we obtain

ωab(1)=Caαb(q)dxα and ωab(2)=Caαb(1)dxα+Daib(q)dzi.\omega_{a}^{b(1)}=C_{a\alpha}^{b}(q)dx^{\alpha}\hskip 8.61108pt\text{ and }\hskip 8.61108pt\omega_{a}^{b(2)}=C_{a\alpha}^{b(1)}dx^{\alpha}+D_{ai}^{b}(q)dz^{i}.

However, according to 3.3 we have ωab(1)=0\omega_{a}^{b(1)}=0 and ωab(2)=12Rαβab(q)xαdxβ.\omega_{a}^{b(2)}=\frac{1}{2}R_{\alpha\beta a}^{b}(q)x^{\alpha}dx^{\beta}. In particular, this shows that Caαb(q)=Daib(q)=0,C_{a\alpha}^{b}(q)=D_{ai}^{b}(q)=0, i.e. ωab(q)=0\omega_{a}^{b}(q)=0.

(3): Since \nabla is a metric connection, we have:

(3.29) Xγ(Jαβi)\displaystyle X_{\gamma}(J_{\alpha\beta}^{i}) =Xγg(JZiXα,Xβ)=Xγg(Zi,T(Xα,Xβ))\displaystyle=X_{\gamma}g\big{(}J_{Z_{i}}X_{\alpha},X_{\beta}\big{)}=X_{\gamma}g\big{(}Z_{i},T(X_{\alpha},X_{\beta})\big{)}
(3.30) =g(XγZi,T(Xα,Xβ))+g(Zi,Xγ(T(Xα,Xβ))).\displaystyle=g\big{(}\nabla_{X_{\gamma}}Z_{i},T(X_{\alpha},X_{\beta})\big{)}+g\big{(}Z_{i},\nabla_{X_{\gamma}}(T(X_{\alpha},X_{\beta}))\big{)}.

At the point q𝕄q\in\mathbb{M} we have ωab(q)=0\omega_{a}^{b}(q)=0 according to (2) and therefore,

(3.31) (XγXα)q=(XγXβ)q=(XγZi)q=0(\nabla_{X_{\gamma}}X_{\alpha})_{q}=(\nabla_{X_{\gamma}}X_{\beta})_{q}=(\nabla_{X_{\gamma}}Z_{i})_{q}=0

showing the assertion. ∎

Now we compute the low order homogeneous parts in the parabolic Taylor series of the special frame {X1,,Xn,Z1,,Zm}\{X_{1},\ldots,X_{n},Z_{1},\ldots,Z_{m}\}. For this, we first define the following linearly independent vector fields on n+m\mathbb{R}^{n+m}:

(3.32) X^α:=xα+Jαβi(q)xβzi and Z^i:=2zi.\hat{X}_{\alpha}:=\frac{\partial}{\partial x^{\alpha}}+J_{\alpha\beta}^{i}(q)x^{\beta}\frac{\partial}{\partial z^{i}}\text{ and }\hat{Z}_{i}:=2\frac{\partial}{\partial z^{i}}.

Note that X^α\hat{X}_{\alpha} (resp. Z^i\hat{Z}_{i}) is homogeneous of degree 1-1 (resp. 2-2).

Remark 3.5.

The advantage of the chosen privileged coordinates is that the nilpotent Lie algebra (that depends on the choice of privileged coordinates) will coincide with the HH-type Lie algebra, which is the first algebraic and metric invariant at each point q𝕄q\in\mathbb{M} as mentioned in the introduction. Moreover, the vector fields X^α\hat{X}_{\alpha} and Z^i\hat{Z}_{i} are left invariant vector fields on the corresponding (HH-type) Lie group 𝔾(q)\mathbb{G}(q) with group law “\star” as in e.g. [16].

Lemma 3.6.

In privileged coordinates (xα,zi)(x^{\alpha},z^{i}), for the horizontal frame we can express the low order homogeneous parts as

(3.33) Xα(1)=X^α,Xα(0)=13Jαβi(1)xβZ^i and Xα(1)=fαβX^β+hαiZ^i,X_{\alpha}^{(-1)}=\hat{X}_{\alpha},\hskip 5.69054ptX_{\alpha}^{(0)}=\frac{1}{3}J_{\alpha\beta}^{i(1)}x^{\beta}\hat{Z}_{i}\;\text{ and }\;X_{\alpha}^{(1)}=f^{\beta}_{\alpha}\hat{X}_{\beta}+h^{i}_{\alpha}\hat{Z}_{i},

where

(3.34) fαβ\displaystyle f^{\beta}_{\alpha} =16xγxδRαγδβ(q),\displaystyle=\frac{1}{6}x^{\gamma}x^{\delta}R_{\alpha\gamma\delta}^{\beta}(q),
(3.35) hαi\displaystyle h^{i}_{\alpha} =18Rαβji(q)zjxβ+124Jβγi(q)Rαβγγ(q)xβxβxγ+14xβJαβi(2).\displaystyle=\frac{1}{8}R_{\alpha\beta j}^{i}(q)z^{j}x^{\beta}+\frac{1}{24}J_{\beta\gamma}^{i}(q)R_{\alpha\beta^{\prime}\gamma^{\prime}}^{\gamma}(q)x^{\beta}x^{\beta^{\prime}}x^{\gamma^{\prime}}+\frac{1}{4}x^{\beta}J_{\alpha\beta}^{i(2)}.

Similarly, for the vertical frame it holds that

(3.37) Zi(2)=Z^i and Zi(1)=Zi(0)=0.Z_{i}^{(-2)}=\hat{Z}_{i}\text{ and }Z_{i}^{(-1)}=Z_{i}^{(0)}=0.
Proof.

Locally near 0n+m0\in\mathbb{R}^{n+m} we consider the expansion:

(3.38) Xα=sαβX^β+rαjZ^j,X_{\alpha}=s_{\alpha}^{\beta}\hat{X}_{\beta}+r_{\alpha}^{j}\hat{Z}_{j},

where sαβs_{\alpha}^{\beta} and rαjr_{\alpha}^{j} are smooth functions. The parabolic Taylor expansion of sαβs_{\alpha}^{\beta} and rαjr_{\alpha}^{j} around 0 have the form:

(3.39) sαβ\displaystyle s_{\alpha}^{\beta} =sαβ(0)+sαβ(1)+sαβ(2)+\displaystyle=s^{\beta(0)}_{\alpha}+s^{\beta(1)}_{\alpha}+s^{\beta(2)}_{\alpha}+\cdots
(3.40) rαj\displaystyle r_{\alpha}^{j} =rαj(0)+rαj(1)+rαj(2)+.\displaystyle=r^{j(0)}_{\alpha}+r^{j(1)}_{\alpha}+r^{j(2)}_{\alpha}+\cdots.

Here sαβ(l)s^{\beta(l)}_{\alpha} (resp. rαj(l)r^{j(l)}_{\alpha}) (for l0l\geq 0) denotes the homogeneous part of order ll in the expansion. Applying θγ\theta^{\gamma} (resp. ηi\eta^{i}) to both sides of Eq. 3.38 and taking the homogeneous parts of order ll, we obtain recursive formulas for sαγ(l)s^{\gamma(l)}_{\alpha} and rαi(l)r^{i(l)}_{\alpha} (for l1l\geq 1):

(3.41) sαγ(l)\displaystyle s^{\gamma(l)}_{\alpha} =m=0l1sαβ(m)θγ(lm+1)(X^β)m=0lrαj(m)θγ(lm+2)(Z^j),\displaystyle=-\sum_{m=0}^{l-1}s^{\beta(m)}_{\alpha}\theta^{\gamma(l-m+1)}(\hat{X}_{\beta})-\sum_{m=0}^{l}r^{j(m)}_{\alpha}\theta^{\gamma(l-m+2)}(\hat{Z}_{j}),
(3.42) rαi(l)\displaystyle r^{i(l)}_{\alpha} =m=0l1sαβ(m)ηi(lm+1)(X^β)m=0l1rαj(m)ηi(lm+2)(Z^j)\displaystyle=-\sum_{m=0}^{l-1}s^{\beta(m)}_{\alpha}\eta^{i(l-m+1)}(\hat{X}_{\beta})-\sum_{m=0}^{l-1}r^{j(m)}_{\alpha}\eta^{i(l-m+2)}(\hat{Z}_{j})

with initial conditions sαγ(0)=δαγs^{\gamma(0)}_{\alpha}=\delta_{\alpha\gamma} and rαi(0)=0r^{i(0)}_{\alpha}=0. Combining these identities with the expressions for the homogeneous components of θγ\theta^{\gamma} and ηi\eta^{i} in 3.3 we find:

(3.43) sαβ(1)=0,rαj(0)=rαj(1)=0, and rαj(2)=13Jαβj(1)xβ.s^{\beta(1)}_{\alpha}=0,\quad r^{j(0)}_{\alpha}=r^{j(1)}_{\alpha}=0,\text{ and }r^{j(2)}_{\alpha}=\frac{1}{3}J_{\alpha\beta}^{j(1)}x^{\beta}.

Hence, it follows that

(3.44) sαβ(2)=θβ(3)(X^α)=16Rαγδβ(q)xγxδ,s^{\beta(2)}_{\alpha}=-\theta^{\beta(3)}(\hat{X}_{\alpha})=\frac{1}{6}R_{\alpha\gamma\delta}^{\beta}(q)x^{\gamma}x^{\delta},

and

(3.45) rαj(3)\displaystyle r^{j(3)}_{\alpha} =ηj(4)(X^α)\displaystyle=-\eta^{j(4)}(\hat{X}_{\alpha})
(3.46) =18Rαβij(q)zixβ+124Jβγj(q)Rαβγγ(q)xβxβxγ+14xβJαβj(2).\displaystyle=\frac{1}{8}R_{\alpha\beta i}^{j}(q)z^{i}x^{\beta}+\frac{1}{24}J_{\beta\gamma}^{j}(q)R_{\alpha\beta^{\prime}\gamma^{\prime}}^{\gamma}(q)x^{\beta}x^{\beta^{\prime}}x^{\gamma^{\prime}}+\frac{1}{4}x^{\beta}J_{\alpha\beta}^{j(2)}.

Now the assumption follows from the formula (l1l\geq 1)

(3.47) Xα(l)=sαβ(l+1)X^β+rαj(l+2)Z^j.X_{\alpha}^{(l)}=s_{\alpha}^{\beta(l+1)}\hat{X}_{\beta}+r_{\alpha}^{j(l+2)}\hat{Z}_{j}.

The identities Zi(2)=Z^iZ_{i}^{(-2)}=\hat{Z}_{i} and Zi(1)=Zi(0)=0Z_{i}^{(-1)}=Z_{i}^{(0)}=0 follow by a similar calculation. ∎

We constructed a graded step two nilpotent Lie algebra generated by the vector fields X1(1),,Xn(1)X_{1}^{(-1)},\cdots,X_{n}^{(-1)} on n+m\mathbb{R}^{n+m}; as was mentioned, this Lie algebra coincides with an HH-type Lie algebra. The corresponding connected, simply connected Lie group is the metric tangent group (or cone), see [16] or [1, Proposition 10.77].

The next theorem shows that the curvature tensor RR of the Bott connection is exactly the obstruction of the HH-type foliation (with horizontal parallel torsion, that is XT=0\nabla_{X}T=0 for all XΓ()X\in\Gamma(\mathcal{H})) to be locally isometric to an HH-type group in the sense of sub-Riemannian manifolds. We recall the definition (see [6]):

Let (𝕄j,j,,j)(\mathbb{M}_{j},\mathcal{H}_{j},\langle\cdot,\cdot\rangle_{j}) for j=1,2j=1,2 be sub-Riemannian manifolds. We call a smooth map

φ:𝕄1𝕄2\varphi:\mathbb{M}_{1}\rightarrow\mathbb{M}_{2}

horizontal if φ(1)2\varphi_{*}(\mathcal{H}_{1})\subset\mathcal{H}_{2}.

Moreover, φ\varphi is called a (local) sub-Riemannian isometry, if it is a horizontal (local) diffeomorphism such that ϕ:(1,,1)(2,,2)\phi_{*}:(\mathcal{H}_{1},\langle\cdot,\cdot\rangle_{1})\longrightarrow(\mathcal{H}_{2},\langle\cdot,\cdot\rangle_{2}) is an isometry.

Theorem 3.7.

Let 𝕄\mathbb{M} be an H-type foliation with horizontally parallel torsion. Then 𝕄\mathbb{M} is locally isometric (as a sub-Riemannian manifold) to its tangent group if and only if the curvature tensor of the Bott connection vanishes, i.e. R=0R=0.

Proof.

It suffices to construct a local isometry between 𝕄\mathbb{M} and its tangent group under the condition R0R\equiv 0. Assume that the curvature tensor RR is identically zero. Then, due to the properties of the Bott connection we can construct near every fixed point q𝕄q\in\mathbb{M}, a parallel orthonormal frame {X1,,Xn,Z1,,Zm}\{X_{1},\cdots,X_{n},Z_{1},\cdots,Z_{m}\}, where X1,,XnX_{1},\cdots,X_{n} (resp. Z1,,ZmZ_{1},\cdots,Z_{m}) are horizontal (resp. vertical) vector fields. Next, we look at the Lie bracket relations of this frame. Then it holds that

(3.48) [Xα,Zi]\displaystyle\textstyle[X_{\alpha},Z_{i}] =T(Xα,Zi)=0,\displaystyle=-T(X_{\alpha},Z_{i})=0,
(3.49) [Zi,Zj]\displaystyle[Z_{i},Z_{j}] =T(Zi,Zj)=0,\displaystyle=-T(Z_{i},Z_{j})=0,
(3.50) [Xα,Xβ]\displaystyle[X_{\alpha},X_{\beta}] =T(Xα,Xβ)=k=1mJαβkZk.\displaystyle=-T(X_{\alpha},X_{\beta})=-\textstyle\sum_{k=1}^{m}J_{\alpha\beta}^{k}Z_{k}.

Since the torsion is horizontally parallel, it follows that the functions JαβkJ_{\alpha\beta}^{k} are in fact constant. Now we consider the structure equations (3.21) with respect to this frame and the associated co-frame θ1,,θn,η1,,ηm\theta^{1},\cdots,\theta^{n},\eta^{1},\cdots,\eta^{m}:

(3.51) dθα=0 and dηi=12Jαβiθαθβ.d\theta^{\alpha}=0\text{ and }d\eta^{i}=\frac{1}{2}J^{i}_{\alpha\beta}\theta^{\alpha}\wedge\theta^{\beta}.

Then, for every α\alpha we can find a locally defined smooth function xαx^{\alpha} such that θα=dxα\theta^{\alpha}=dx^{\alpha}. Inserting this in Eq. 3.51 we can find a smooth function ziz_{i} such that

(3.52) ηi=dzi+12Jαβixαdxβ.\eta^{i}=dz^{i}+\frac{1}{2}J^{i}_{\alpha\beta}x^{\alpha}dx^{\beta}.

Finally, the map p(x1(p),,xn(p),z1(p),,zm(p))p\mapsto(x^{1}(p),\cdots,x^{n}(p),z^{1}(p),\cdots,z^{m}(p)) defines a local sub-Riemannian isometry between 𝕄\mathbb{M} and its tangent HH-type group. ∎

4. Popp measure and volume of small parabolic geodesic balls

Let {X1,,Xn,Z1,,Zm}\{X_{1},\cdots,X_{n},Z_{1},\cdots,Z_{m}\} be a special frame near q𝕄q\in\mathbb{M} in the sense of Section 3. We observe that this frame is also an adapted frame for the sub-Riemannian structure of the H-type Riemannian foliation 𝕄\mathbb{M} in the sense of [6]; that is, {X1,,Xn}\{X_{1},\cdots,X_{n}\} spans the horizontal distribution while {Z1,,Zm}\{Z_{1},\cdots,Z_{m}\} spans the vertical distribution. It follows from the results therein that the Popp measure 𝒫\mathcal{P} associated to the corresponding equiregular sub-Riemannian structure can be expressed locally in the form

(4.1) 𝒫=1detBθ1θnη1ηm.\mathcal{P}=\frac{1}{\sqrt{\det{B}}}\theta^{1}\wedge\cdots\wedge\theta^{n}\wedge\eta^{1}\wedge\cdots\wedge\eta^{m}.

Here B=(Bij)ijB=(B_{ij})_{ij} is the (m×m)(m\times m)-matrix function locally defined near qq with coefficients given by

(4.2) Bij:=α,β=1nbαβibαβj,B_{ij}:=\sum_{\alpha,\beta=1}^{n}b_{\alpha\beta}^{i}b_{\alpha\beta}^{j},

where bαβib_{\alpha\beta}^{i} are defined for α,β=1,,n\alpha,\beta=1,\cdots,n and i=1,,mi=1,\cdots,m by

(4.3) bαβi:=g([Xα,Xβ],Zi)=g(JiXα,Xβ).b_{\alpha\beta}^{i}:=g\big{(}[X_{\alpha},X_{\beta}],Z_{i}\big{)}=-g(J_{i}X_{\alpha},X_{\beta}).

Hence it follows that

(4.4) Bij=α,β=1ng(JiXα,Xβ)g(JjXα,Xβ)=α=1ng(JiXα,JjXα)=nδij.B_{ij}=\sum_{\alpha,\beta=1}^{n}g(J_{i}X_{\alpha},X_{\beta})g(J_{j}X_{\alpha},X_{\beta})=\sum_{\alpha=1}^{n}g(J_{i}X_{\alpha},J_{j}X_{\alpha})=n\delta_{ij}.

In the last equality we used the skew-symmetry of the almost complex structures JiJ_{i} together with the HH-type condition. This shows that BB is a diagonal matrix:

(4.5) B=nIdm×mB=n\cdot\mathrm{Id}\in\mathbb{R}^{m\times m}

and hence we obtain a formula for the Popp measure:

Lemma 4.1.

The Popp measure 𝒫\mathcal{P} for the H-type Riemannian foliation (𝕄,,g)(\mathbb{M},\mathcal{H},g_{\mathcal{H}}) has the form

(4.6) 𝒫=nm/2ω,\mathcal{P}=n^{-m/2}\omega,

where ω:=θ1θnη1ηm\omega:=\theta^{1}\wedge\cdots\wedge\theta^{n}\wedge\eta^{1}\wedge\cdots\wedge\eta^{m}.

Let us denote by Ψ:qU𝕄n+m\Psi\colon q\in U\subseteq\mathbb{M}\rightarrow\mathbb{R}^{n+m} the (privileged) coordinate map and define the parabolic ball with center qq and (small) radius r>0r>0 by

(4.7) B(q,r):=Ψ1(B^(0,r)),B(q,r):=\Psi^{-1}\left(\hat{B}(0,r)\right),

where

(4.8) B^(0,r):={(x,z)n+m:(x,z)h:=(|x|4+|z|2)1/4r}\hat{B}(0,r):=\left\{(x,z)\in\mathbb{R}^{n+m}\colon\|(x,z)\|_{h}:=(|x|^{4}+|z|^{2})^{1/4}\leq r\right\}

is the Korányi ball of radius r>0r>0 centered at 0n+m0\in\mathbb{R}^{n+m}. Here |||\cdot| denotes the Euclidean inner product on n\mathbb{R}^{n} and m\mathbb{R}^{m}.

Remark 4.2.

Note that a parabolic ball is independent of the choice of an orthonormal frame at qq and that according to the ball-box theorem (see [36, Theorem 2.10]) the sub-Riemannian distance dccd_{cc} in the privileged coordinates (x,z)(x,z) is equivalent to the homogeneous distance dhd_{h} defined by

(4.9) dh((x,z),(x,z)):=(x,z)1(x,z)hd_{h}\left((x,z),(x^{\prime},z^{\prime})\right):=\|(x,z)^{-1}\star(x^{\prime},z^{\prime})\|_{h}

for (x,z),(x,z)n+m𝔾(q)(x,z),(x^{\prime},z^{\prime})\in\mathbb{R}^{n+m}\simeq\mathbb{G}(q). Here \star denotes the group law on 𝔾(q)\mathbb{G}(q), as in 3.5.

The next lemma provides a geometric interpretation of the scalar curvature induced by the Bott connection:

Theorem 4.3.

Let q𝕄q\in\mathbb{M}. As r0r\to 0 it holds:

(4.10) rQvol(B(q,r))=an,mbn,mκ(q)r2+O(r3),r^{-Q}\cdot\mathrm{vol}(B(q,r))=a_{n,m}-b_{n,m}\kappa_{\mathcal{H}}(q)r^{2}+O(r^{3}),

where an,ma_{n,m} and bn,mb_{n,m} are suitable positive constants (which will be given in the proof below) and Q=n+2mQ=n+2m is the Hausdorff dimension of the metric space (𝕄,dcc)(\mathbb{M},d_{cc}).

Proof.

We write

(4.11) vol(B(q,r))=1nm/2B(q,r)ω=1nm/2B^(0,1)ω^r,\mathrm{vol}(B(q,r))=\frac{1}{n^{m/2}}\int_{B(q,r)}\omega=\frac{1}{n^{m/2}}\int_{\hat{B}(0,1)}{{\hat{\omega}}_{r}},

where we denote the pushforward to n+m\mathbb{R}^{n+m} of ω\omega as ω^=Ψω{\hat{\omega}}=\Psi_{*}\omega and its pullback by δr\delta_{r} as ω^r=δrω^{{\hat{\omega}}_{r}}=\delta_{r}^{*}{\hat{\omega}}.

By 3.3, the low order homogeneous parts in the expansion of ω^r{{\hat{\omega}}_{r}} as r0+r\rightarrow 0^{+} start at order Q=n+2mQ=n+2m and can be explicitly computed as

(4.12) ω^r=rQω^(Q)+rQ+1ω^(Q+1)+rQ+2ω^(Q+2)+O(rQ+3), as r0+.{{\hat{\omega}}_{r}}=r^{Q}{\hat{\omega}}^{(Q)}+r^{Q+1}{\hat{\omega}}^{(Q+1)}+r^{Q+2}{\hat{\omega}}^{(Q+2)}+O(r^{Q+3}),\text{ as }r\rightarrow 0^{+}.

We calculate the homogeneous terms ω^(j){\hat{\omega}}^{(j)} for j=Q,Q+1,Q+2j=Q,Q+1,Q+2.

For j=Qj=Q,

(4.13) ω^(Q)\displaystyle{\hat{\omega}}^{(Q)} =θ1(1)θn(1)η1(2)ηm(2)\displaystyle=\theta^{1(1)}\wedge\cdots\wedge\theta^{n(1)}\wedge\eta^{1(2)}\wedge\cdots\wedge\eta^{m(2)}
(4.14) =12mdx1dxndz1dzm.\displaystyle=\frac{1}{2^{m}}dx^{1}\wedge\cdots\wedge dx^{n}\wedge dz^{1}\wedge\cdots\wedge dz^{m}.

It holds that ω^(Q+1)=0{\hat{\omega}}^{(Q+1)}=0, since θi(2)=0\theta^{i(2)}=0 and ηi(3)\eta^{i(3)} is a linear combination of dxjdx^{j}. In fact, a similar argument will hold in general for jj odd.

Finally, for j=Q+2j=Q+2 we have:

(4.15) ω^(Q+2)=162mRγαβα(q)xβxγdx1dxndz1dzm.{\hat{\omega}}^{(Q+2)}=\frac{1}{6\cdot 2^{m}}R_{\gamma\alpha\beta}^{\alpha}(q)x^{\beta}x^{\gamma}dx^{1}\wedge\cdots\wedge dx^{n}\wedge dz^{1}\wedge\cdots\wedge dz^{m}.

It follows that an,ma_{n,m} is given by

(4.16) an,m:=1(4m)m/2B^(0,1)𝑑x1𝑑xn𝑑z1𝑑zma_{n,m}:=\frac{1}{(4m)^{m/2}}\int_{\hat{B}(0,1)}dx^{1}\cdots dx^{n}dz^{1}\cdots dz^{m}

and the coefficient of the r2r^{2} term has the value

(4.17) 16(4n)m/2B^(0,1)Rγαβα(q)xβxγ𝑑x1𝑑xn𝑑z1𝑑zm.\frac{1}{6(4n)^{m/2}}\int_{\hat{B}(0,1)}R_{\gamma\alpha\beta}^{\alpha}(q)x^{\beta}x^{\gamma}dx^{1}\cdots dx^{n}dz^{1}\cdots dz^{m}.

Since the ball B^(0,1){{\hat{B}(0,1)}} is invariant under the reflection xγxγx^{\gamma}\rightarrow-x^{\gamma} we conclude that:

(4.18) B^(0,1)xβxγ𝑑x1𝑑xn𝑑z1𝑑zm\int_{\hat{B}(0,1)}x^{\beta}x^{\gamma}\ dx^{1}\cdots dx^{n}dz^{1}\cdots dz^{m}

is non-zero if and only if β=γ\beta=\gamma and moreover the integral does not depend on this value. Since Rβαβα=RβααβR_{\beta\alpha\beta}^{\alpha}=-R_{\beta\alpha\alpha}^{\beta} it follows that the coefficient in front of r2r^{2} will be of the form bn,mκ(q)-b_{n,m}\kappa_{\mathcal{H}}(q) with

(4.19) bn,m:=16(4m)m/2B^(0,1)(xα)2𝑑x1𝑑xn𝑑z1𝑑zm,b_{n,m}:=\frac{1}{6(4m)^{m/2}}\int_{\hat{B}(0,1)}(x^{\alpha})^{2}\ dx^{1}\cdots dx^{n}dz^{1}\cdots dz^{m},

where the index α\alpha can be chosen to be any of 1,2,,n1,2,\ldots,n. ∎

Being an equiregular sub-Riemannian manifold, (𝕄,,g)(\mathbb{M},\mathcal{H},g_{\mathcal{H}}) carries an intrinsic sub-Laplacian Δsub\Delta_{\mathrm{sub}} induced by the Popp measure 𝒫\mathcal{P}, (see [2, 6]). More precisely, in terms of the special frame {X1,,Xn,Z1,,Zm}\{X_{1},\cdots,X_{n},Z_{1},\cdots,Z_{m}\}, this second order, positive and hypoelliptic differential operator Δsub\Delta_{\mathrm{sub}} can be expressed explicitly in the following form (see [6]):

(4.20) Δsub=(α=1nXα2+div𝒫(Xα)Xα).\Delta_{\mathrm{sub}}=-\left(\sum_{\alpha=1}^{n}X_{\alpha}^{2}+\text{div}_{\mathcal{P}}(X_{\alpha})X_{\alpha}\right).

Here the divergence operator is defined by

(4.21) div𝒫(Xα)𝒫=Xα(𝒫).\mathrm{div}_{\mathcal{P}}(X_{\alpha})\mathcal{P}=\mathcal{L}_{X_{\alpha}}(\mathcal{P}).

We now calculate the divergence in terms of the geometric data. Applying the Leibniz law for the Lie derivative,

(4.22) Xα(𝒫)=1nm/2a=1n+mν1νa1Xα(νa)νa+1νn+m.\mathcal{L}_{X_{\alpha}}(\mathcal{P})=\frac{1}{n^{m/2}}\sum_{a=1}^{n+m}\nu^{1}\wedge\cdots\wedge\nu^{a-1}\wedge\mathcal{L}_{X_{\alpha}}(\nu^{a})\wedge\nu^{a+1}\wedge\cdots\wedge\nu^{n+m}.

Since {ν1,,νn+m}\{\nu^{1},\cdots,\nu^{n+m}\} is a coframe, we write for α{1,,n}\alpha\in\{1,\cdots,n\} and a{1,,n+m}a\in\{1,\cdots,n+m\}:

(4.23) Xα(νa)=fαbaνb.\mathcal{L}_{X_{\alpha}}(\nu^{a})=f_{\alpha b}^{a}\nu^{b}.

with the functions fαbaf_{\alpha b}^{a} given by fαba=Xα(νa)(Xb)=νa([Xb,Xα]).f_{\alpha b}^{a}=\mathcal{L}_{X_{\alpha}}(\nu^{a})(X_{b})=\nu^{a}([X_{b},X_{\alpha}]). Hence we obtain

(4.24) div𝒫(Xα)=θβ([Xβ,Xα])+ηi([Zi,Xα])\mathrm{div}_{\mathcal{P}}(X_{\alpha})=\theta^{\beta}([X_{\beta},X_{\alpha}])+\eta^{i}([Z_{i},X_{\alpha}])

Now, using the properties of the Bott connection, a calculation shows that ωα:=θβ([Xβ,Xα])=ωαβ(Xβ)\omega_{\alpha}:=\theta^{\beta}([X_{\beta},X_{\alpha}])=\omega_{\alpha}^{\beta}(X_{\beta}) and ηi([Zi,Xα])=0\eta^{i}([Z_{i},X_{\alpha}])=0; summarizing the above calculation gives us that

Lemma 4.4.

The intrinsic sub-Laplacian Δsub\Delta_{\mathrm{sub}} has the expression

(4.25) Δsub=(α=1nXα2+ωαXα).\Delta_{\mathrm{sub}}=-\left(\sum_{\alpha=1}^{n}X_{\alpha}^{2}+\omega_{\alpha}X_{\alpha}\right).

By definition the sub-Laplacian Δsub\Delta_{\mathrm{sub}} is positive and it coincides up to a constant factor with the (negative) horizontal Laplacian defined in [10, Remark 2.19].

5. Heat invariants

Using 3.3 and 4.4, we compute the low order homogeneous terms in the decomposition of the sub-Laplacian into a sum of homogeneous differential operators with polynomial coefficients:

Δsub=(Δ^sub+𝒜(1)+𝒜(0)+)withΔ^sub:=αX^α2.\Delta_{\mathrm{sub}}=-\left(\hat{\Delta}_{\mathrm{sub}}+\mathcal{A}^{(-1)}+\mathcal{A}^{(0)}+\cdots\right)\hskip 8.61108pt\text{with}\hskip 8.61108pt\hat{\Delta}_{\mathrm{sub}}:=\sum_{\alpha}\hat{X}_{\alpha}^{2}.

Note that Δ^sub\hat{\Delta}_{\mathrm{sub}} is the intrinsic sub-Laplacian on the HH-type group 𝔾(q)\mathbb{G}(q) (cf. Remark 3.5). The operators 𝒜(1)\mathcal{A}^{(-1)} and 𝒜(0)\mathcal{A}^{(0)} are given by:

(5.1) 𝒜(1):=α(X^αXα(0)+Xα(0)X^α),\displaystyle\mathcal{A}^{(-1)}:=\sum_{\alpha}\Big{(}\hat{X}_{\alpha}X_{\alpha}^{(0)}+X_{\alpha}^{(0)}\hat{X}_{\alpha}\Big{)},
(5.2) 𝒜(0):=α(X^αXα(1)+Xα(1)X^α+(Xα(0))2+12Rγβαβ(q)xγX^α).\displaystyle\mathcal{A}^{(0)}:=\sum_{\alpha}\Big{(}\hat{X}_{\alpha}X_{\alpha}^{(1)}+X_{\alpha}^{(1)}\hat{X}_{\alpha}+\left(X_{\alpha}^{(0)}\right)^{2}+\frac{1}{2}R_{\gamma\beta\alpha}^{\beta}(q)x^{\gamma}\hat{X}_{\alpha}\Big{)}.

According to the results in [24, p. 28,46] and based on Duhamel’s formula the second heat invariant c1(q)c_{1}(q) at qq is given by the Schwartz kernel K1(1,0,0)K_{1}(1,0,0) (at time t=1t=1) of the operator

C1(t):=0te(ts)Δ^sub(𝒜(0)(q)esΔ^sub+𝒜(1)(q)C0(s))𝑑s,C_{1}(t):=\int_{0}^{t}e^{(t-s)\hat{\Delta}_{\mathrm{sub}}}\left(\mathcal{A}^{(0)}(q)e^{s\hat{\Delta}_{\mathrm{sub}}}+\mathcal{A}^{(-1)}(q)C_{0}(s)\right)ds,

where

C0(t)=0te(ts)Δ^sub𝒜(1)(q)esΔ^sub𝑑s.C_{0}(t)=\int_{0}^{t}e^{(t-s)\hat{\Delta}_{\mathrm{sub}}}\mathcal{A}^{(-1)}(q)e^{s\hat{\Delta}_{\mathrm{sub}}}ds.

From now on and in order to simplify the formulas we assume that the torsion is horizontally parallel, i.e. T=0\nabla_{\mathcal{H}}T=0. According to 3.4, (3) this implies that all horizontal derivatives of JαβiJ_{\alpha\beta}^{i} at the point qq vanish. As a consequence, at the point qq, we have Jαβi(1)=0J_{\alpha\beta}^{i(1)}=0 by 3.4, (1) and hence, Xα(0)=0X_{\alpha}^{(0)}=0 for all α\alpha by 3.6. With this in mind, it follows that 𝒜(1)=0\mathcal{A}^{(-1)}=0 under the condition T=0\nabla_{\mathcal{H}}T=0, and the second heat invariant c1(q)c_{1}(q) is obtained from the Schwartz kernel K1(1,0,0)K_{1}(1,0,0) of

C1(t)=0te(ts)Δ^sub𝒜(0)(q)esΔ^sub𝑑sC_{1}(t)=\int_{0}^{t}e^{(t-s)\hat{\Delta}_{\mathrm{sub}}}\mathcal{A}^{(0)}(q)e^{s\hat{\Delta}_{\mathrm{sub}}}ds

at (0,0)(0,0) and for time t=1t=1. More precisely, we have

(5.3) c1(q)=01n+mK^2(s,0,ξ)(𝒜(0)(q)K^1)(1s,ξ,0)𝑑ξ𝑑s,c_{1}(q)=\int_{0}^{1}\int_{\mathbb{R}^{n+m}}\hat{K}_{2}(s,0,\xi)(\mathcal{A}^{(0)}(q)\hat{K}_{1})(1-s,\xi,0)d\xi ds,

where K^1\hat{K}_{1} denotes the heat kernel associated with Δ^sub\hat{\Delta}_{\mathrm{sub}} on 𝔾(q)n+m\mathbb{G}(q)\cong\mathbb{R}^{n+m} equipped with the nilpotentization of the Popp measure at the point qq, and the operator 𝒜(0)\mathcal{A}^{(0)} acts on the second component of K^1\hat{K}_{1}.

Lemma 5.1.

Assume that T=0\nabla_{\mathcal{H}}T=0. Then the second heat invariant c1(q)c_{1}(q) is a linear combination of the following components of tensors at qq:

(5.4) Rαβγδ,Rαβij,JαλiRαβγδ,JδλiRαβγδ,JαγiRαβij,\displaystyle R_{\alpha\beta\gamma}^{\delta},\hskip 5.69054ptR_{\alpha\beta i}^{j},\hskip 5.69054ptJ_{\alpha\lambda}^{i}R_{\alpha\beta\gamma}^{\delta},\hskip 5.69054ptJ_{\delta\lambda}^{i}R_{\alpha\beta\gamma}^{\delta},\hskip 5.69054ptJ_{\alpha\gamma}^{i}R_{\alpha\beta i}^{j},
(5.5) JααiJββjRαγδβ,XαXβ(Jγδi),JααiXγXδ(Jαβj),\displaystyle J_{\alpha\alpha^{\prime}}^{i}J_{\beta\beta^{\prime}}^{j}R_{\alpha\gamma\delta}^{\beta},\hskip 5.69054ptX_{\alpha}X_{\beta}(J_{\gamma\delta}^{i}),\hskip 5.69054ptJ_{\alpha\alpha^{\prime}}^{i}X_{\gamma}X_{\delta}(J_{\alpha\beta}^{j}),
(5.6) Zi(Jαβj),JαβiZj(Jαγk).\displaystyle Z_{i}(J_{\alpha\beta}^{j}),\hskip 5.69054ptJ_{\alpha\beta}^{i}Z_{j}(J_{\alpha\gamma}^{k}).
Proof.

According to the data and notation in 3.6 and using Xα(0)=0X_{\alpha}^{(0)}=0 for all α\alpha we can express the operator 𝒜(0)\mathcal{A}^{(0)} in the form

𝒜(0)=(fαβ+fβα)X^αX^β+X^α(fβα)X^β+(X^α(hαi)Z^i+2hαiX^αZ^i+12Rγβαβ(q)xγX^α).\mathcal{A}^{(0)}=(f_{\alpha}^{\beta}+f_{\beta}^{\alpha})\hat{X}_{\alpha}\hat{X}_{\beta}+\hat{X}_{\alpha}(f_{\beta}^{\alpha})\hat{X}_{\beta}+\Big{(}\hat{X}_{\alpha}(h^{i}_{\alpha})\hat{Z}_{i}+2h^{i}_{\alpha}\hat{X}_{\alpha}\hat{Z}_{i}+\frac{1}{2}R_{\gamma\beta\alpha}^{\beta}(q)x^{\gamma}\hat{X}_{\alpha}\Big{)}.

Then the statement directly follows from Eq. 5.3. ∎

By 5.1, we know that the second heat invariant c1(q)c_{1}(q) can be expressed in terms of components of certain tensors with respect to the orthonormal basis {Xα,Zi}\{X_{\alpha},Z_{i}\} at qq. In the following lemmas we will simplify these expressions.

Remark 5.2.

The key observation is that c1(q)c_{1}(q) is independent of the choice of such an orthonormal basis; it follows that the linear combinations occurring in 5.1 must be invariant under the action of the group 𝐎(n)×𝐎(m){\bf O}(n)\times{\bf O}(m) and therefore we can use techniques from the classical invariance theory of the orthogonal group (cf. [4, 47]) to obtain a more precise expression of the second heat invariant in terms of components of the curvature and torsion tensors.

Lemma 5.3.

The second heat invariant c1(q)c_{1}(q) is a linear combination of the following traces:

RαββαR_{\alpha\beta\beta}^{\alpha} JαγiJβδiRαγδβJ_{\alpha\gamma}^{i}J_{\beta\delta}^{i}R_{\alpha\gamma\delta}^{\beta} JαγiJβδiRαδγβJ_{\alpha\gamma}^{i}J_{\beta\delta}^{i}R_{\alpha\delta\gamma}^{\beta}
JαβiXγXγ(Jαβi)J_{\alpha\beta}^{i}X_{\gamma}X_{\gamma}(J_{\alpha\beta}^{i}) JαβiXβXγ(Jαγi)J_{\alpha\beta}^{i}X_{\beta}X_{\gamma}(J_{\alpha\gamma}^{i}) JαβiXγXβ(Jαγi)J_{\alpha\beta}^{i}X_{\gamma}X_{\beta}(J_{\alpha\gamma}^{i})

where we emphasize that we use the summation rule described in 2.7.

Proof.

First we will write the coefficients of the tensors in their covariant form by contracting with the metric tensor. For instance gσδRαβγδ=Rαβγσ.g_{\sigma\delta}R_{\alpha\beta\gamma}^{\delta}=R_{\alpha\beta\gamma\sigma}. The tensor with the coefficients RαβγσR_{\alpha\beta\gamma\sigma} will be considered as an element of the vector space

(5.7) ((n))4=(n)(n)(n)(n).((\mathbb{R}^{n})^{*})^{\otimes 4}=(\mathbb{R}^{n})^{*}\otimes(\mathbb{R}^{n})^{*}\otimes(\mathbb{R}^{n})^{*}\otimes(\mathbb{R}^{n})^{*}.

More precisely, we denote V=(n)V=(\mathbb{R}^{n})^{*} and write {e1,,en}\{e^{1},\ldots,e^{n}\} for the standard dual basis in VV. Then

(5.8) {Rαβγδ:α,β,γ,δ{1,,n}}RαβγδeαeβeγeδV4.\big{\{}R_{\alpha\beta\gamma\delta}\colon\alpha,\beta,\gamma,\delta\in\{1,\ldots,n\}\big{\}}\mapsto R_{\alpha\beta\gamma\delta}e^{\alpha}\otimes e^{\beta}\otimes e^{\gamma}\otimes e^{\delta}\in V^{\otimes 4}.

The vector space VV can be considered as an 𝐎(n){\bf O}(n) module under the standard action of the orthogonal group. This action is extended to V4V^{\otimes 4} componentwise.

Having this example in mind, we proceed like in the CR case [15]. We consider the space V=(n)V=(\mathbb{R}^{n})^{*} (resp. W=(m)W=(\mathbb{R}^{m})^{*}) with the standard dual basis {e1,,en}\{e^{1},\cdots,e^{n}\} (resp. {f1,,fm}\{f^{1},\cdots,f^{m}\}) as an 𝐎(n){\bf O}(n) (resp. 𝐎(m){\bf O}(m)) module. These actions extend naturally to the tensor products

(5.9) VrWs:=VVr timesWWs times.V^{\otimes r}\otimes W^{\otimes s}:=\underbrace{V\otimes\cdots\otimes V}_{r\text{ times}}\otimes\underbrace{W\otimes\cdots\otimes W}_{s\text{ times}}.

We write all the tensors occurring in 5.1 in the covariant form and consider them as elements of the corresponding vector space E:=VrWsE:=V^{\otimes r}\otimes W^{\otimes s} for appropriate integers rr and ss.

Note that this tensor representation is equivariant under the action of the group 𝐎(n)×𝐎(m){\bf O}(n)\times{\bf O}(m); that is, the 𝐎(n)×𝐎(m){\bf O}(n)\times{\bf O}(m) group action on the chosen orthonormal frame {X1(q),,Xn(q),Z1(q),,Zm(q)}\{X_{1}(q),\ldots,X_{n}(q),Z_{1}(q),\ldots,Z_{m}(q)\} at q𝕄q\in\mathbb{M} commutes with the representation map. Let us denote by ΘE\Theta\in E one of the tensors from 5.1, written in the covariant form in the fixed orthonormal frame at q𝕄q\in\mathbb{M}. By the invariance theory [4, 47] there is a linear functional f:Ef\colon E\rightarrow\mathbb{R}, such that

(5.10) f(UΘ)=c1(q) for all U𝐎(n)×𝐎(m).f(U\Theta)=c_{1}(q)\hskip 8.61108pt\text{{\it{ for all }}}\hskip 8.61108ptU\in{\bf{O}}(n)\times{\bf{O}}(m).

By replacing the functional ff by its average over 𝐎(n)×𝐎(m){\bf{O}}(n)\times{\bf{O}}(m), we assume that ff is 𝐎(n)×𝐎(m){\bf{O}}(n)\times{\bf{O}}(m)-invariant. It is known that an 𝐎(n)×𝐎(m){\bf{O}}(n)\times{\bf{O}}(m)-invariant linear functional on E=VrWsE=V^{\otimes r}\otimes W^{\otimes s} is non-zero only in the case where rr and ss are both even and moreover such a functional must be a complete contraction (up to a constant multiple). Hence, the second heat invariant c1(q)=f(Θ)c_{1}(q)=f(\Theta) must be a linear combination of

RαββαR_{\alpha\beta\beta}^{\alpha} RααββR_{\alpha\alpha\beta}^{\beta} RαβαβR_{\alpha\beta\alpha}^{\beta} JαγiJβδiRαγδβJ_{\alpha\gamma}^{i}J_{\beta\delta}^{i}R_{\alpha\gamma\delta}^{\beta}
JαγiJβδiRαδγβJ_{\alpha\gamma}^{i}J_{\beta\delta}^{i}R_{\alpha\delta\gamma}^{\beta} JαβiXγXγ(Jαβi)J_{\alpha\beta}^{i}X_{\gamma}X_{\gamma}(J_{\alpha\beta}^{i}) JαβiXβXγ(Jαγi)J_{\alpha\beta}^{i}X_{\beta}X_{\gamma}(J_{\alpha\gamma}^{i}) JαβiXγXβ(Jαγi).J_{\alpha\beta}^{i}X_{\gamma}X_{\beta}(J_{\alpha\gamma}^{i}).

5.3 follows now from the fact that the curvature tensor RR is skew-symmetric in the first two lower components so that Rααββ=0R_{\alpha\alpha\beta}^{\beta}=0 and Rαβαβ=RαββαR_{\alpha\beta\alpha}^{\beta}=-R_{\alpha\beta\beta}^{\alpha}. (Note that in this final equality summations are not necessary despite the convention 2.7.) ∎

The following result can be found in [10] and will be useful in the sequel:

Lemma 5.4.

Let (𝕄,,g)(\mathbb{M},\mathcal{H},g_{\mathcal{H}}) be an H-type foliation with horizontally parallel torsion, i.e. T=0\nabla_{\mathcal{H}}T=0. Then (1) - (3) hold for XΓ()X\in\Gamma(\mathcal{H}) and Z,WΓ(𝒱)Z,W\in\Gamma(\mathcal{V}):

  1. (1)

    (ZJ)W=(WJ)Z(\nabla_{Z}J)_{W}=-(\nabla_{W}J)_{Z}.

  2. (2)

    (ZJ)W(\nabla_{Z}J)_{W} is skew-symmetric and anti-commutes with JWJ_{W}.

  3. (3)

    (ZJ)WX2=R(Z,W)W,ZX2\|(\nabla_{Z}J)_{W}X\|^{2}=\langle R(Z,W)W,Z\rangle\|X\|^{2}.

Proof.

See Lemmas 2.62.6, 2.182.18 and 3.53.5 in [10]. ∎

To study the traces from 5.3 we introduce the following operator.

Definition 5.5.

We define the bundle like operator M(Z,W):qqM(Z,W)\colon\mathcal{H}_{q}\to\mathcal{H}_{q} for q𝕄q\in\mathbb{M} by

(5.11) XM(Z,W)X=JWJZ(ZJ)WXX\mapsto M(Z,W)X=J_{W}J_{Z}(\nabla_{Z}J)_{W}X

for XΓ(),Z,WΓ(𝒱)X\in\Gamma(\mathcal{H}),Z,W\in\Gamma(\mathcal{V}).

5.4 immediately implies that M(Z,W)=0M(Z,W)=0 for linearly dependent vectors Z,W𝒱qZ,W\in\mathcal{V}_{q}, q𝕄q\in\mathbb{M}.

Lemma 5.6.

Under the assumptions of the preceding lemma, the following holds for Z,W𝒱Z,W\in\mathcal{V}:

  1. (1)

    The symmetric part of the operator M(Z,W)M(Z,W) is given by

    N(Z,W):=M(Z,W)+Z,W(ZJ)W.N(Z,W):=M(Z,W)+\langle Z,W\rangle(\nabla_{Z}J)_{W}.

    In particular, M(Z,W)M(Z,W) is symmetric if ZZ and WW are orthogonal Z,W=0\langle Z,W\rangle=0.

  2. (2)

    The eigenvalues of the operator N(Z,W)N(Z,W) are given by

    ±(Z2W2Z,W2)R(Z,W)W,Z.\pm\sqrt{(\|Z\|^{2}\|W\|^{2}-\langle Z,W\rangle^{2})\langle R(Z,W)W,Z\rangle}.

    In particular,

    (5.12) tr(M(Z,W))\displaystyle\textup{tr}\left(M(Z,W)\right) =tr(N(Z,W))\displaystyle=\textup{tr}\left(N(Z,W)\right)
    (5.13) =σ(Z,W)(Z2W2Z,W2)R(Z,W)W,Z,\displaystyle=\sigma(Z,W)\sqrt{(\|Z\|^{2}\|W\|^{2}-\langle Z,W\rangle^{2})\langle R(Z,W)W,Z\rangle},

    where σ(Z,W)\sigma(Z,W) denotes the number of positive eigenvalues minus the number of negative eigenvalues of N(Z,W)N(Z,W).

  3. (3)

    The operator N(Z,W)N(Z,W) is zero if and only if R(Z,W)W,Z=0\langle R(Z,W)W,Z\rangle=0, i.e. the sectional curvature of the vertical plane generated by ZZ and WW vanishes. Otherwise, N(Z,W)N(Z,W) is invertible.

Proof.

It is sufficient to prove the statements (1) and (2).
(1): 5.4 and the HH-type condition Eq. 2.7 show:

(5.14) N(Z,W)\displaystyle N(Z,W) =12(M(Z,W)+M(W,Z))\displaystyle=\frac{1}{2}\Big{(}M(Z,W)+M(W,Z)\Big{)}
(5.15) =12(M(Z,W)+JZJW(WJ)Z)\displaystyle=\frac{1}{2}\Big{(}M(Z,W)+J_{Z}J_{W}(\nabla_{W}J)_{Z}\Big{)}
(5.16) =12(M(Z,W)(JWJZ2Z,W)(ZJ)W)\displaystyle=\frac{1}{2}\Big{(}M(Z,W)-(-J_{W}J_{Z}-2\langle Z,W\rangle)(\nabla_{Z}J)_{W}\Big{)}
(5.17) =M(Z,W)+Z,W(ZJ)W.\displaystyle=M(Z,W)+\langle Z,W\rangle(\nabla_{Z}J)_{W}.

(2): Note that M(Z,W)M(Z,W) and (ZJ)W(\nabla_{Z}J)_{W} commute according to 5.4. Let XX be any eigenvector of N(Z,W)N(Z,W), then we obtain:

(5.18) N(Z,W)X2\displaystyle\|N(Z,W)X\|^{2} =(M(Z,W)+Z,W(ZJ)W)2X,X\displaystyle=\big{\langle}(M(Z,W)+\langle Z,W\rangle(\nabla_{Z}J)_{W})^{2}X,X\big{\rangle}
=M(Z,W)2X,X+Z,W2(ZJ)WX2.\displaystyle=\big{\langle}M(Z,W)^{2}X,X\big{\rangle}+\langle Z,W\rangle^{2}\|(\nabla_{Z}J)_{W}X\|^{2}.

In the last equality we have used Eq. 2.7 and the skew-symmetry of (ZJ)W(\nabla_{Z}J)_{W} which show that:

(5.19) M(Z,W)(ZJ)WX,X\displaystyle\big{\langle}M(Z,W)(\nabla_{Z}J)_{W}X,X\big{\rangle} =JWJZ(ZJ)WX,(ZJ)WX\displaystyle=-\big{\langle}J_{W}J_{Z}(\nabla_{Z}J)_{W}X,(\nabla_{Z}J)_{W}X\big{\rangle}
(5.20) =Z,W(ZJ)WX2.\displaystyle=\langle Z,W\rangle\|(\nabla_{Z}J)_{W}X\|^{2}.

As for the first summand:

M(Z,W)2X,X\displaystyle\big{\langle}M(Z,W)^{2}X,X\big{\rangle} =JWJZ(ZJ)WX,JZJW(ZJ)WX\displaystyle=-\big{\langle}J_{W}J_{Z}(\nabla_{Z}J)_{W}X,J_{Z}J_{W}(\nabla_{Z}J)_{W}X\big{\rangle}
=JWJZ(ZJ)WX2+2Z,WJWJZ(ZJ)W,(ZJ)WX\displaystyle=\big{\|}J_{W}J_{Z}(\nabla_{Z}J)_{W}X\big{\|}^{2}+2\langle Z,W\rangle\big{\langle}J_{W}J_{Z}(\nabla_{Z}J)_{W},(\nabla_{Z}J)_{W}X\big{\rangle}
(5.21) =(Z2W22Z,W2)(ZJ)WX2.\displaystyle=\Big{(}\|Z\|^{2}\|W\|^{2}-2\langle Z,W\rangle^{2}\Big{)}\big{\|}(\nabla_{Z}J)_{W}X\big{\|}^{2}.

Combining Eq. 5.18 and Eq. 5.21 with 5.4, this shows

(5.22) N(Z,W)X2=(Z2W2Z,W2)R(Z,W)W,ZX2,\big{\|}N(Z,W)X\|^{2}=\Big{(}\|Z\|^{2}\|W\|^{2}-\langle Z,W\rangle^{2}\Big{)}\big{\langle}R(Z,W)W,Z\big{\rangle}\|X\|^{2},

which completes the proof. ∎

We define an important invariant

(5.23) τ𝒱:=i,j=1mtr(M(Zi,Zj)),\tau_{\mathcal{V}}:=\sum_{i,j=1}^{m}\mathrm{tr}\hskip 2.84526pt\big{(}M(Z_{i},Z_{j})\big{)},

where {Z1,,Zm}\{Z_{1},\cdots,Z_{m}\} is a local orthonormal frame of the vertical distribution 𝒱\mathcal{V} and tr\mathrm{tr} denotes the pointwise matrix trace of M(Zi,Zj)End()M(Z_{i},Z_{j})\in\mathrm{End}(\mathcal{H}). Note that the definition of τ𝒱\tau_{\mathcal{V}} does not depend on the choice of an orthonormal frame of 𝒱\mathcal{V}.

Remark 5.7.

If Z1,,ZmZ_{1},\cdots,Z_{m} is an orthonormal frame of 𝒱\mathcal{V}, then by 5.6, property (2)(2) above we can write

τ𝒱=ijσ(Zi,Zj)Rijji,\tau_{\mathcal{V}}=\sum_{ij}\sigma(Z_{i},Z_{j})\sqrt{R_{ijj}^{i}},

Since τ𝒱\tau_{\mathcal{V}} is independent of the choice of an orthonormal frame Z1,,ZmZ_{1},\cdots,Z_{m}, it is natural to ask whether one can use this invariance to further simplify this expression. As we will see below, under further assumption on the leaves of our foliation the invariant τ𝒱\tau_{\mathcal{V}} will be everywhere constant.

In the next proposition we need the notion of horizontally parallel Clifford structure on an HH-type foliation. We start with a definition (see [10, Definition 3.1]).

Definition 5.8.

An HH-type foliation (𝕄,,g)(\mathbb{M},\mathcal{H},g_{\mathcal{H}}) with horizontally parallel torsion TT is said to have horizontally parallel Clifford structure if there exists a smooth bundle map Ψ:𝒱×𝒱C2(𝒱)\Psi:\mathcal{V}\times\mathcal{V}\longrightarrow C\ell_{2}(\mathcal{V}) (Clifford bundle) such that for all Z1,Z2Γ(𝒱)Z_{1},Z_{2}\in\Gamma(\mathcal{V}):

(5.24) (Z1J)Z2=JΨ(Z1,Z2).(\nabla_{Z_{1}}J)_{Z_{2}}=J_{\Psi(Z_{1},Z_{2})}.

We refer to [10] for examples and further properties. With this we can prove: —

Proposition 5.9.

Suppose m2m\geq 2 and assume that the torsion is horizontally parallel. Then

  1. (1)

    Let Z,W𝒱Z,W\in\mathcal{V} be orthonormal, and suppose that XΓ()X\in\Gamma(\mathcal{H}) is a λ\lambda-eigenvector for M(Z,W)=N(Z,W)=JWJZ(ZJ)WM(Z,W)=N(Z,W)=J_{W}J_{Z}(\nabla_{Z}J)_{W}. The vectors JZX,JWX,JZJWXJ_{Z}X,J_{W}X,J_{Z}J_{W}X are all λ\lambda-eigenvectors for N(Z,W)N(Z,W).

  2. (2)

    σ=4k\sigma=4k, kk\in\mathbb{N}.

Assume moreover that the sectional curvature κ𝒱\kappa_{\mathcal{V}} of the leaves is a positive constant; that is, for any Z,W𝒱Z,W\in\mathcal{V}

(5.25) R(Z,W)W,Z=κ𝒱(Z2W2Z,W2).\langle R(Z,W)W,Z\rangle=\kappa_{\mathcal{V}}\left(\|Z\|^{2}\|W\|^{2}-\langle Z,W\rangle^{2}\right).

It then follows that

  1. (3)

    σ(Z,W)=σ\sigma(Z,W)=\sigma is independent of the choice of vectors Z,W𝒱Z,W\in\mathcal{V} and it holds:

    τ𝒱=m(m1)σκ𝒱.\tau_{\mathcal{V}}=m(m-1)\sigma\sqrt{\kappa_{\mathcal{V}}}.
  2. (4)

    If 𝕄\mathbb{M} also has an horizontally parallel Clifford structure (see [10, Theorem 3.6]) then σ\sigma and τ𝒱\tau_{\mathcal{V}} are constant; specifically

    (5.26) σ=n and τ𝒱=m(m1)nκ𝒱.\sigma=-n\text{ and }\tau_{\mathcal{V}}=-m(m-1)n\sqrt{\kappa_{\mathcal{V}}}.
Proof.

(1) It follows from 5.4 that JZJ_{Z} and JWJ_{W} anti-commute with (ZJ)W(\nabla_{Z}J)_{W}, and from the HH-type condition that they also anti-commute with JWJZJ_{W}J_{Z}; together these imply that N(Z,W)N(Z,W) commutes with each of JZ,JW,J_{Z},J_{W}, and JZJWJ_{Z}J_{W}, completing the proof of (1). Lemma 5.6 (2) shows that

λ=±R(Z,W)W,Z\lambda=\pm\sqrt{\langle R(Z,W)W,Z\rangle}

and the vector space span{X,JZX,JWX,JZJWX}{\rm span}\{X,J_{Z}X,J_{W}X,J_{Z}J_{W}X\} is an eigenspace with eigenvalue λ\lambda. At the point q𝕄q\in\mathbb{M} at the level of the tangent group it will form a subgroup isomorphic to the complexified Heisenberg group.

(2) N(Z,W)N(Z,W) is symmetric, therefore diagonalizable. Partition q\mathcal{H}_{q} with q𝕄q\in\mathbb{M} into eigenspaces spanned by X1,(JZX)q,(JWX)q,(JVJWX)qX_{1},(J_{Z}X)_{q},(J_{W}X)_{q},(J_{V}J_{W}X)_{q}, which is always possible by (1); this immediately implies σ=4k\sigma=4k for some kk\in\mathbb{Z}.

(3) Let Z,W𝒱Z,W\in\mathcal{V} be linearly independent. By 5.6, we know that

(5.27) tr(N(Z,W))=σ(Z,W)(Z2W2Z,W2)R(Z,W)W,Z.\mathrm{tr}(N(Z,W))=\sigma(Z,W)\sqrt{(\|Z\|^{2}\|W\|^{2}-\langle Z,W\rangle^{2})\langle R(Z,W)W,Z\rangle}.

The idea is to consider σ\sigma as a function defined on the Grassmann 22-plane bundle G2(𝒱)G_{2}(\mathcal{V}) as follows:

(5.28) σ(Z,W)=tr(N(Z,W))(Z2W2Z,W2)R(Z,W)W,Z.\sigma(Z,W)=\frac{\textup{tr}(N(Z,W))}{\sqrt{(\|Z\|^{2}\|W\|^{2}-\langle Z,W\rangle^{2})\langle R(Z,W)W,Z\rangle}}.

Note that the right hand side of this equation depends only on the 22-plane generated by ZZ and WW. In that way, we consider σ:G2(𝒱)\sigma\colon\mathrm{G}_{2}(\mathcal{V})\rightarrow\mathbb{R} which is a smooth function with discrete values in \mathbb{Z}. Since G2(𝒱)\mathrm{G}_{2}(\mathcal{V}) is connected, it follows that σ\sigma is a constant function.

Hence

(5.29) τ𝒱=ij,i,j=1mtr(M(Zi,Zj))=ij,i,j=1mσκ𝒱=m(m1)σκ𝒱\tau_{\mathcal{V}}=\sum_{i\neq j,\ i,j=1}^{m}\mathrm{tr}(M(Z_{i},Z_{j}))=\sum_{i\neq j,\ i,j=1}^{m}\sigma\sqrt{\kappa_{\mathcal{V}}}=m(m-1)\sigma\sqrt{\kappa_{\mathcal{V}}}

by tr(M(Z,W))=tr(N(Z,W))\mathrm{tr}(M(Z,W))=\mathrm{tr}(N(Z,W)). This shows that τ𝒱\tau_{\mathcal{V}} is constant on 𝕄\mathbb{M} with value

(5.30) τ𝒱=m(m1)σκ𝒱.\tau_{\mathcal{V}}=m(m-1)\sigma\sqrt{\kappa_{\mathcal{V}}}.

(4) According to [10, Theorem 3.6] the assumption of having horizontally parallel Clifford structure implies that the sectional curvature κ𝒱\kappa_{\mathcal{V}} of the leaves of the foliation associated to 𝒱\mathcal{V} is constant and

(5.31) (ZiJ)Zj=κ𝒱JZiJZj(\nabla_{Z_{i}}J)_{Z_{j}}=-\sqrt{\kappa_{\mathcal{V}}}J_{Z_{i}}J_{Z_{j}}

Then all eigenvalues are the same λ=κ𝒱\lambda=-\sqrt{\kappa_{\mathcal{V}}}. The claim follows immediately. ∎

Proposition 5.10.

Assume that the torsion is horizontally parallel and recall from (2.11) the scalar curvature κ\kappa_{\mathcal{H}}. Then at q𝕄q\in\mathbb{M}:

  1. (1)

    JαγiJβδiRαδγβ=R(Xα,JZiXβ)JZiXα,Xβ=κ+2τ𝒱,J_{\alpha\gamma}^{i}J_{\beta\delta}^{i}R_{\alpha\delta\gamma}^{\beta}=\langle R(X_{\alpha},J_{Z_{i}}X_{\beta})J_{Z_{i}}X_{\alpha},X_{\beta}\rangle=\kappa_{\mathcal{H}}+2\tau_{\mathcal{V}},

  2. (2)

    JαγiJβδiRαγδβ=R(Xα,JZiXα)JZiXβ,Xβ=2κ+4τ𝒱,J_{\alpha\gamma}^{i}J_{\beta\delta}^{i}R_{\alpha\gamma\delta}^{\beta}=\langle R(X_{\alpha},J_{Z_{i}}X_{\alpha})J_{Z_{i}}X_{\beta},X_{\beta}\rangle=2\kappa_{\mathcal{H}}+4\tau_{\mathcal{V}},

  3. (3)

    JαβiXγXγ(Jαβi)=0,J_{\alpha\beta}^{i}X_{\gamma}X_{\gamma}(J_{\alpha\beta}^{i})=0,

  4. (4)

    JαβiXβXγ(Jαγi)=12τ𝒱,J_{\alpha\beta}^{i}X_{\beta}X_{\gamma}(J_{\alpha\gamma}^{i})=\frac{1}{2}\tau_{\mathcal{V}},

  5. (5)

    JαβiXγXβ(Jαγi)=12τ𝒱.J_{\alpha\beta}^{i}X_{\gamma}X_{\beta}(J_{\alpha\gamma}^{i})=-\frac{1}{2}\tau_{\mathcal{V}}.

We recall that the summations are implied, per 2.7.

Proof.

(1) We recall the following formula for the commutator of R(,)R(\cdot,\cdot) and JJ on horizontal vectors. A proof of Eq. 5.32 can be found in [10, Lemma 3.18].

(5.32) [R(X,Y),JZ]=(T(X,Y)J)Z+J(ZT)(X,Y)[R(X,Y),J_{Z}]=\left(\nabla_{T(X,Y)}J\right)_{Z}+J_{(\nabla_{Z}T)(X,Y)}

for X,YΓ()X,Y\in\Gamma(\mathcal{H}) and ZΓ(𝒱)Z\in\Gamma(\mathcal{V}). Hence, we can write

(5.33) JαγiJβδiRαδγβ\displaystyle J_{\alpha\gamma}^{i}J_{\beta\delta}^{i}R_{\alpha\delta\gamma}^{\beta} =R(Xα,JZiXβ)JZiXα,Xβ\displaystyle=\langle R(X_{\alpha},J_{Z_{i}}X_{\beta})J_{Z_{i}}X_{\alpha},X_{\beta}\rangle
(5.34) =JZiR(Xα,JZiXβ)Xα,Xβ\displaystyle=\langle J_{Z_{i}}R(X_{\alpha},J_{Z_{i}}X_{\beta})X_{\alpha},X_{\beta}\rangle
(5.35) +(T(Xα,JZiXβ)J)ZiXα,Xβ+J(ZiT)(Xα,JZiXβ))Xα,Xβ.\displaystyle+\left\langle\left(\nabla_{T(X_{\alpha},J_{Z_{i}}X_{\beta})}J\right)_{Z_{i}}X_{\alpha},X_{\beta}\right\rangle+\left\langle J_{\left(\nabla_{Z_{i}}T)(X_{\alpha},J_{Z_{i}}X_{\beta})\right)}X_{\alpha},X_{\beta}\right\rangle.

Since JZiJ_{Z_{i}} is an isometry, writing JZiXβ=XγJ_{Z_{i}}X_{\beta}=X_{\gamma} and Xβ=JZiXγX_{\beta}=-J_{Z_{i}}X_{\gamma} the first term on the right hand side JZiR(Xα,JZiXβ)Xα,Xβ\langle J_{Z_{i}}R(X_{\alpha},J_{Z_{i}}X_{\beta})X_{\alpha},X_{\beta}\rangle can be interpreted as the scalar curvature κ\kappa_{\mathcal{H}}. The second term can be rewritten in the form

(5.36) (T(Xα,JZiXβ)J)ZiXα,Xβ\displaystyle\left\langle\left(\nabla_{T(X_{\alpha},J_{Z_{i}}X_{\beta})}J\right)_{Z_{i}}X_{\alpha},X_{\beta}\right\rangle =(T(Xα,Xβ)J)ZiXα,JZiXβ\displaystyle=-\left\langle\left(\nabla_{T(X_{\alpha},X_{\beta})}J\right)_{Z_{i}}X_{\alpha},J_{Z_{i}}X_{\beta}\right\rangle
(5.37) =JZjXα,Xβ(ZjJ)ZiXα,JZiXβ\displaystyle=-\langle J_{Z_{j}}X_{\alpha},X_{\beta}\rangle\cdot\left\langle\left(\nabla_{Z_{j}}J\right)_{Z_{i}}X_{\alpha},J_{Z_{i}}X_{\beta}\right\rangle
(5.38) =JZjXα,JZi(ZjJ)ZiXα\displaystyle=\left\langle J_{Z_{j}}X_{\alpha},J_{Z_{i}}\left(\nabla_{Z_{j}}J\right)_{Z_{i}}X_{\alpha}\right\rangle
(5.39) =Xα,JZjJZi(ZiJ)ZjXα\displaystyle=\left\langle X_{\alpha},J_{Z_{j}}J_{Z_{i}}\left(\nabla_{Z_{i}}J\right)_{Z_{j}}X_{\alpha}\right\rangle
(5.40) =tr(JZjJZi(ZiJ)Zj).\displaystyle=\mathrm{tr}\left(J_{Z_{j}}J_{Z_{i}}\left(\nabla_{Z_{i}}J\right)_{Z_{j}}\right).

Finally, the last term can be expressed as

(5.41) J(ZiT)(Xα,JZiXβ))Xα,Xβ\displaystyle\Big{\langle}J_{\left(\nabla_{Z_{i}}T)(X_{\alpha},J_{Z_{i}}X_{\beta})\right)}X_{\alpha},X_{\beta}\Big{\rangle} =J(ZiT)(Xα,Xβ))Xα,JZiXβ\displaystyle=-\Big{\langle}J_{\left(\nabla_{Z_{i}}T)(X_{\alpha},X_{\beta})\right)}X_{\alpha},J_{Z_{i}}X_{\beta}\Big{\rangle}
(5.42) =(ZiT)(Xα,Xβ),ZjJZjXα,JZiXβ\displaystyle=-\langle\left(\nabla_{Z_{i}}T\right)(X_{\alpha},X_{\beta}),Z_{j}\rangle\cdot\langle J_{Z_{j}}X_{\alpha},J_{Z_{i}}X_{\beta}\rangle
(5.43) =(ZiJ)ZjXα,XβJZiJZjXα,Xβ\displaystyle=\langle\left(\nabla_{Z_{i}}J\right)_{Z_{j}}X_{\alpha},X_{\beta}\rangle\cdot\langle J_{Z_{i}}J_{Z_{j}}X_{\alpha},X_{\beta}\rangle
(5.44) =JZjJZi(ZiJ)ZjXα,Xα\displaystyle=\langle J_{Z_{j}}J_{Z_{i}}\left(\nabla_{Z_{i}}J\right)_{Z_{j}}X_{\alpha},X_{\alpha}\rangle
(5.45) =tr(JZjJZi(ZiJ)Zj).\displaystyle=\text{tr}\left(J_{Z_{j}}J_{Z_{i}}\left(\nabla_{Z_{i}}J\right)_{Z_{j}}\right).

Here in the second line we used that (ZiT)(Xα,Xβ)(\nabla_{Z_{i}}T)(X_{\alpha},X_{\beta}) has components only in the direction of 𝒱\mathcal{V}. For the third line we applied the formula

(5.46) (WJ)ZX,Y=Z,(WT)(X,Y),X,YΓ(),Z,WΓ(𝒱)\langle\left(\nabla_{W}J\right)_{Z}X,Y\rangle=\langle Z,\left(\nabla_{W}T\right)(X,Y)\rangle,\quad X,Y\in\Gamma(\mathcal{H}),\quad Z,W\in\Gamma(\mathcal{V})

that is the consequence of the differentiation

(5.47) W(JZX,Y)=W(Z,T(X,Y)).W\left(\langle J_{Z}X,Y\rangle\right)=W\left(\langle Z,T(X,Y)\rangle\right).

(2) Using the first Bianchi identity Eq. 2.13 and the fact that T(,)𝒱T(\mathcal{H},\mathcal{H})\subset\mathcal{V}, T(,𝒱)=0T(\mathcal{H},\mathcal{V})=0, we obtain:

(5.48) R(Xα,JZiXα)JZiXβ,Xβ++R(JZiXα,JZiXβ)Xα,Xβ+R(JZiXβ,Xα)JZiXα,Xβ=0.\langle R(X_{\alpha},J_{Z_{i}}X_{\alpha})J_{Z_{i}}X_{\beta},X_{\beta}\rangle+\\ +\langle R(J_{Z_{i}}X_{\alpha},J_{Z_{i}}X_{\beta})X_{\alpha},X_{\beta}\rangle+\langle R(J_{Z_{i}}X_{\beta},X_{\alpha})J_{Z_{i}}X_{\alpha},X_{\beta}\rangle=0.

By using JZiXα=XγJ_{Z_{i}}X_{\alpha}=X_{\gamma}, Xα=JZiXγX_{\alpha}=-J_{Z_{i}}X_{\gamma} for the second term and skew symmetry of the curvature tensor with respect to the two first vectors, see [27, Lemma 3.7], we obtain the desired result.

(3) By the H-type condition JZi2=IdJ_{Z_{i}}^{2}=-\mathrm{Id}_{\mathcal{H}}, it follows that JαβiJαβi=1J_{\alpha\beta}^{i}J_{\alpha\beta}^{i}=1 for all α\alpha. Taking the first derivative along XγX_{\gamma}, we obtain JαβiXγ(Jαβi)=0J_{\alpha\beta}^{i}X_{\gamma}\left(J_{\alpha\beta}^{i}\right)=0 for all α,β,i\alpha,\beta,i, and γ\gamma. Taking again the derivative along XγX_{\gamma}, we obtain

(5.49) Xγ(Jαβi)Xγ(Jαβi)+JαβiXγXγ(Jαβi)=0.X_{\gamma}\left(J_{\alpha\beta}^{i}\right)X_{\gamma}\left(J_{\alpha\beta}^{i}\right)+J_{\alpha\beta}^{i}X_{\gamma}X_{\gamma}\left(J_{\alpha\beta}^{i}\right)=0.

Since the torsion is horizontally parallel and T(,)𝒱T(\mathcal{H},\mathcal{H})\subset\mathcal{V} it follows that Xγ(Jαβi)(q)X_{\gamma}\left(J_{\alpha\beta}^{i}\right)(q) vanishes at q𝕄q\in\mathbb{M} and hence

(5.50) JαβiXγXγ(Jαβi)=0.J_{\alpha\beta}^{i}X_{\gamma}X_{\gamma}\left(J_{\alpha\beta}^{i}\right)=0.

(4) Using the H-type condition JZi2=IdJ_{Z_{i}}^{2}=-\mathrm{Id}_{\mathcal{H}}, it follows that JαβiJαγi=δβγJ_{\alpha\beta}^{i}J_{\alpha\gamma}^{i}=\delta_{\beta\gamma} for all β\beta and γ\gamma. Hence, using the assumption that the torsion is horizontally parallel and 3.4 (3), we can write at q𝕄q\in\mathbb{M}

(5.51) JαβiXβXγ(Jαγi)+XβXγ(Jαβi)Jαγi=0.J_{\alpha\beta}^{i}X_{\beta}X_{\gamma}\left(J_{\alpha\gamma}^{i}\right)+X_{\beta}X_{\gamma}\left(J_{\alpha\beta}^{i}\right)J_{\alpha\gamma}^{i}=0.

This shows that at qq

(5.52) JαβiXβXγ(Jαγi)=JαβiXγXβ(Jαγi).J_{\alpha\beta}^{i}X_{\beta}X_{\gamma}\left(J_{\alpha\gamma}^{i}\right)=-J_{\alpha\beta}^{i}X_{\gamma}X_{\beta}\left(J_{\alpha\gamma}^{i}\right).

Now, we write at qq

(5.53) JαβiXβXγ(Jαγi)\displaystyle J_{\alpha\beta}^{i}X_{\beta}X_{\gamma}\left(J_{\alpha\gamma}^{i}\right) =12Jαβi[Xβ,Xγ](Jαγi)\displaystyle=\frac{1}{2}J_{\alpha\beta}^{i}[X_{\beta},X_{\gamma}]\left(J_{\alpha\gamma}^{i}\right)
(5.54) =12JαβiJβγjZj(Jαγi)\displaystyle=\frac{1}{2}J_{\alpha\beta}^{i}J_{\beta\gamma}^{j}Z_{j}\left(J_{\alpha\gamma}^{i}\right)
(5.55) =12JZiXα,JZjXγ(ZjJ)ZiXα,Xγ\displaystyle=-\frac{1}{2}\langle J_{Z_{i}}X_{\alpha},J_{Z_{j}}X_{\gamma}\rangle\langle\left(\nabla_{Z_{j}}J\right)_{Z_{i}}X_{\alpha},X_{\gamma}\rangle
(5.56) =12τ𝒱(q).\displaystyle=\frac{1}{2}\tau_{\mathcal{V}}(q).

(5) The last statement can be shown by similar calculation. ∎

Remark 5.11.

Note that the traces in (1)(1) and (2)(2) from the preceding proposition were computed in [29] for quaternionic contact manifolds endowed with the Biquard connection. Therein it was shown that these traces are of the form CκC\kappa_{\mathcal{H}} where CC is a dimensional constant and κ\kappa_{\mathcal{H}} is the scalar curvature with respect to the Biquard connection. Furthermore, any 33-Sasakian manifold can be considered both as an H-type foliation (endowed with the Bott connection) with horizontally parallel torsion, or as a quaternionic contact manifold (endowed with the Biquard connection). In that setting, our result about the traces in (1)(1) and (2)(2) coincides with the result from [29], because both local invariants κ\kappa_{\mathcal{H}} and τ𝒱\tau_{\mathcal{V}} are global constants and related to each other by some dimensional constant.

Now, we summarize the preceding computations in the following main theorem:

Theorem 5.12.

Let (𝕄,,g)(\mathbb{M},\mathcal{H},g_{\mathcal{H}}) be an H-type foliation with horizontally parallel torsion. Then the second heat invariant c1c_{1} is a linear combination of the local invariants κ\kappa_{\mathcal{H}} and τ𝒱\tau_{\mathcal{V}}:

(5.57) c1=C1κ+C2τ𝒱,c_{1}=C_{1}\kappa_{\mathcal{H}}+C_{2}\tau_{\mathcal{V}},

where C1C_{1} and C2C_{2} are universal constants depending only on nn and mm.

6. Open Problems

We present a list of open problems that seem closely related to our analysis:

Let us consider the open subset 𝒰\mathcal{U} of the Grassmann 2-plane bundle G2(𝒱)\mathrm{G}_{2}(\mathcal{V}) defined by

(6.1) 𝒰:={ZWG2(𝒱):S(ZW)0},\mathcal{U}:=\{Z\wedge W\in\mathrm{G}_{2}(\mathcal{V}):S(Z\wedge W)\neq 0\},

where S(ZW)S(Z\wedge W) denotes the sectional curvature of the plane spanned by Z,WZ,W and the map σ:𝒰\sigma:\mathcal{U}\longrightarrow\mathbb{Z}, which assigns to ZWZ\wedge W the difference of the numbers of positive and negative eigenvalues of N(Z,W)N(Z,W).

Note that 𝒰=\mathcal{U}=\emptyset if and only if the torsion TT is completely parallel. Assume now that 𝒰\mathcal{U}\neq\emptyset, then it is natural to ask:

  1. (a)

    Is the map σ\sigma constant on 𝒰\mathcal{U}?

Let {Z1,,Zm}\{Z_{1},\ldots,Z_{m}\} be an orthonormal frame of 𝒱\mathcal{V}. Then, the invariant τ\tau can be expressed as (see 5.7 and 5.6, (2)):

(6.2) τ=i,jσijRijji,\tau=\sum_{i,j}\sigma_{ij}\sqrt{R_{ijj}^{i}},

where σij:=σ(ZiZj).\sigma_{ij}:=\sigma(Z_{i}\wedge Z_{j}).

  1. (b)

    Are the σij\sigma_{ij}’s independent of ii and jj? Are the sectional curvatures RijjiR_{ijj}^{i} independent of ii and jj?

Note that τ\tau is independent of the choice of an orthonormal basis of the vertical space, which suggests that the answer may be “yes”.

7. Appendix

In this appendix we present detailed proofs of technical lemmas which were applied in the previous sections.

Proposition 7.1.

With the notation in Section 3 and 5.4 we have the identities:

  • (a)

    dθα=θβωβαd\theta^{\alpha}=\theta^{\beta}\wedge\omega_{\beta}^{\alpha}

  • (b)

    dηi=12Jβγiθβθγ+ηkωkid\eta^{i}=\frac{1}{2}J^{i}_{\beta\gamma}\theta^{\beta}\wedge\theta^{\gamma}+\eta^{k}\wedge\omega_{k}^{i}

  • (c)

    dωab=12Rβγabθβθγ+Rβjabθβηj+12Rjkabηjηk+ωacωcbd\omega_{a}^{b}=\frac{1}{2}R_{\beta\gamma a}^{b}\theta^{\beta}\wedge\theta^{\gamma}+R_{\beta ja}^{b}\theta^{\beta}\wedge\eta^{j}+\frac{1}{2}R_{jka}^{b}\eta^{j}\wedge\eta^{k}+\omega_{a}^{c}\wedge\omega_{c}^{b}.

Proof.

(a): Let α{1,,n}\alpha\in\{1,\ldots,n\} and consider an expansion:

(7.1) dθα\displaystyle d\theta^{\alpha} =β<γaβγαθβθγ+β,ibβiαθβηi+i<jcijαηiηj.\displaystyle=\sum_{\beta<\gamma}a_{\beta\gamma}^{\alpha}\theta^{\beta}\wedge\theta^{\gamma}+\sum_{\beta,i}b_{\beta i}^{\alpha}\theta^{\beta}\wedge\eta^{i}+\sum_{i<j}c_{ij}^{\alpha}\eta^{i}\wedge\eta^{j}.

Since θβ\theta^{\beta} is dual to XβX_{\beta} and the torsion of horizontal vectors is vertical we conclude that the coefficients aβγαa_{\beta\gamma}^{\alpha} are given by:

(7.2) aβγα\displaystyle a_{\beta\gamma}^{\alpha} =2dθα(Xβ,Xγ)=Xβθα(Xγ)=0Xγθα(Xβ)=0θα([Xβ,Xγ])\displaystyle=2d\theta^{\alpha}(X_{\beta},X_{\gamma})=\underbrace{X_{\beta}\theta^{\alpha}(X_{\gamma})}_{=0}-\underbrace{X_{\gamma}\theta^{\alpha}(X_{\beta})}_{=0}-\theta^{\alpha}\big{(}[X_{\beta},X_{\gamma}]\big{)}
(7.3) =g(Xα,[Xβ,Xγ])\displaystyle=-g\big{(}X_{\alpha},\big{[}X_{\beta},X_{\gamma}\big{]}\big{)}
(7.4) =g(Xα,XβXγ)+g(Xα,XγXβ)+g(Xα,T(Xβ,Xγ))=0.\displaystyle=-g\big{(}X_{\alpha},\nabla_{X_{\beta}}X_{\gamma}\big{)}+g\big{(}X_{\alpha},\nabla_{X_{\gamma}}X_{\beta}\big{)}+\underbrace{g\big{(}X_{\alpha},T(X_{\beta},X_{\gamma})\big{)}}_{=0}.

Recall that ωαc\omega_{\alpha}^{c} denote the connection one-forms, i.e. Xα=Xcωαc\nabla X_{\alpha}=X_{c}\otimes\omega_{\alpha}^{c} and therefore:

(7.5) aβγα\displaystyle a_{\beta\gamma}^{\alpha} =g(Xα,Xcωγc(Xβ))+g(Xα,Xcωβc(Xγ))=ωγα(Xβ)+ωβα(Xγ).\displaystyle=-g\big{(}X_{\alpha},X_{c}\otimes\omega_{\gamma}^{c}(X_{\beta})\big{)}+g\big{(}X_{\alpha},X_{c}\otimes\omega_{\beta}^{c}(X_{\gamma})\big{)}=-\omega_{\gamma}^{\alpha}(X_{\beta})+\omega_{\beta}^{\alpha}(X_{\gamma}).

Hence, we find:

(7.6) β<γaβγαθβθγ\displaystyle\sum_{\beta<\gamma}a_{\beta\gamma}^{\alpha}\theta^{\beta}\wedge\theta^{\gamma} =β<γωγα(Xβ)θβθγ+β<γωβα(Xγ)θβθγ\displaystyle=-\sum_{\beta<\gamma}\omega_{\gamma}^{\alpha}(X_{\beta})\theta^{\beta}\wedge\theta^{\gamma}+\sum_{\beta<\gamma}\omega_{\beta}^{\alpha}(X_{\gamma})\theta^{\beta}\wedge\theta^{\gamma}
=ωβα(Xγ)θβθγ.\displaystyle=\omega_{\beta}^{\alpha}(X_{\gamma})\theta^{\beta}\wedge\theta^{\gamma}.

Similarly, we calculate bβiα=g(Xα,[Xβ,Zi])b_{\beta i}^{\alpha}=-g(X_{\alpha},[X_{\beta},Z_{i}]). Since T(Xβ,Zi)=0T(X_{\beta},Z_{i})=0 according to property (3) of the connection we find:

(7.7) bβiα\displaystyle b_{\beta i}^{\alpha} =g(Xα,XβZiΓ(𝒱))+g(Xα,ZiXβ)=g(Xα,ZiXβ)=ωβα(Zi).\displaystyle=-g\big{(}X_{\alpha},\underbrace{\nabla_{X_{\beta}}Z_{i}}_{\in\Gamma(\mathcal{V})}\big{)}+g\big{(}X_{\alpha},\nabla_{Z_{i}}X_{\beta}\big{)}=g\big{(}X_{\alpha},\nabla_{Z_{i}}X_{\beta}\big{)}=\omega_{\beta}^{\alpha}(Z_{i}).

Therefore:

(7.8) β,ibβiαθβηi=β,iωβα(Zi)θβηi.\sum_{\beta,i}b_{\beta i}^{\alpha}\theta^{\beta}\wedge\eta^{i}=\sum_{\beta,i}\omega_{\beta}^{\alpha}(Z_{i})\theta^{\beta}\wedge\eta^{i}.

Finally, since T(Zi,Zj)=0T(Z_{i},Z_{j})=0 according to (3):

(7.9) cijα\displaystyle c_{ij}^{\alpha} =2dθα(Zi,Zj)=g(Xα,[Zi,Zj])=g(Xα,ZiZjΓ(𝒱)ZjZiΓ(𝒱))=0.\displaystyle=2d\theta^{\alpha}(Z_{i},Z_{j})=-g\big{(}X_{\alpha},\big{[}Z_{i},Z_{j}\big{]}\big{)}=-g\big{(}X_{\alpha},\underbrace{\nabla_{Z_{i}}Z_{j}}_{\in\Gamma(\mathcal{V})}-\underbrace{\nabla_{Z_{j}}Z_{i}}_{\in\Gamma(\mathcal{V})}\big{)}=0.

Combining Eq. 7.6, Eq. 7.8 and Eq. 7.9 together with ωαβ=ωαβ(Xc)θc\omega_{\alpha}^{\beta}=\omega_{\alpha}^{\beta}(X_{c})\theta^{c} yields (a):

(7.10) dθα=θβ[ωβα(Xγ)θγ+ωβα(Zi)ηi]=θβωβα.d\theta^{\alpha}=\theta^{\beta}\wedge\big{[}\omega_{\beta}^{\alpha}(X_{\gamma})\theta^{\gamma}+\omega_{\beta}^{\alpha}(Z_{i})\eta^{i}\big{]}=\theta^{\beta}\wedge\omega_{\beta}^{\alpha}.

(b): Let i{1,,m}i\in\{1,\ldots,m\} and consider an expansion of the left hand side:

(7.11) dηi\displaystyle d\eta^{i} =β<γdβγiθβθγ+β,jeβjiθβηj+j<kfjkiηjηk.\displaystyle=\sum_{\beta<\gamma}d_{\beta\gamma}^{i}\theta^{\beta}\wedge\theta^{\gamma}+\sum_{\beta,j}e_{\beta j}^{i}\theta^{\beta}\wedge\eta^{j}+\sum_{j<k}f_{jk}^{i}\eta^{j}\wedge\eta^{k}.

We calculate the coefficients dβγid_{\beta\gamma}^{i}, eβjie_{\beta j}^{i} and fjkif_{jk}^{i}. First, note that:

(7.12) dβγi\displaystyle d_{\beta\gamma}^{i} =2dηi(Xβ,Xγ)=ηi([Xβ,Xγ])=g(Zi,[Xβ,Xγ])\displaystyle=2d\eta^{i}(X_{\beta},X_{\gamma})=-\eta^{i}\big{(}[X_{\beta},X_{\gamma}]\big{)}=-g\big{(}Z_{i},[X_{\beta},X_{\gamma}]\big{)}
(7.13) =g(Zi,XβXγΓ()XγXβΓ()T(Xβ,Xγ))\displaystyle=-g\big{(}Z_{i},\underbrace{\nabla_{X_{\beta}}X_{\gamma}}_{\in\Gamma(\mathcal{H})}-\underbrace{\nabla_{X_{\gamma}}X_{\beta}}_{\in\Gamma(\mathcal{H})}-T(X_{\beta},X_{\gamma})\big{)}
(7.14) =g(Zi,T(Xβ,Xγ))\displaystyle=g\big{(}Z_{i},T(X_{\beta},X_{\gamma})\big{)}
(7.15) =g(JZiXβ,Xγ)=Jβγi.\displaystyle=g\big{(}J_{Z_{i}}X_{\beta},X_{\gamma}\big{)}=J_{\beta\gamma}^{i}.

Similarly, for eβjie_{\beta j}^{i} and fjkif_{jk}^{i}:

(7.16) eβji\displaystyle e_{\beta j}^{i} =g(Zi,XβZjΓ(𝒱)ZjXβΓ()T(Xβ,Zj)=0)\displaystyle=-g\big{(}Z_{i},\underbrace{\nabla_{X_{\beta}}Z_{j}}_{\in\Gamma(\mathcal{V})}-\underbrace{\nabla_{Z_{j}}X_{\beta}}_{\in\Gamma(\mathcal{H})}-\underbrace{T(X_{\beta},Z_{j})}_{=0}\big{)}
(7.17) =g(Zi,XβZj)=ωji(Xβ),\displaystyle=-g\big{(}Z_{i},\nabla_{X_{\beta}}Z_{j}\big{)}=-\omega_{j}^{i}(X_{\beta}),
(7.18) fjki\displaystyle f_{jk}^{i} =g(Zi,ZjZkΓ(𝒱)ZkZjΓ(𝒱)T(Zj,Zk)=0)\displaystyle=-g\big{(}Z_{i},\underbrace{\nabla_{Z_{j}}Z_{k}}_{\in\Gamma(\mathcal{V})}-\underbrace{\nabla_{Z_{k}}Z_{j}}_{\in\Gamma(\mathcal{V})}-\underbrace{T(Z_{j},Z_{k})}_{=0}\big{)}
(7.19) =ωki(Zj)+ωji(Zk).\displaystyle=-\omega_{k}^{i}(Z_{j})+\omega_{j}^{i}(Z_{k}).

Combining the identities and using the skew-symmetry Jβγi=JγβiJ_{\beta\gamma}^{i}=-J_{\gamma\beta}^{i} gives:

(7.20) dηi\displaystyle d\eta^{i} =β<γJβγiθβθγβ,jωji(Xβ)θβηjj<k(ωki(Zj)ωji(Zk))ηjηk\displaystyle=\sum_{\beta<\gamma}J_{\beta\gamma}^{i}\theta^{\beta}\wedge\theta^{\gamma}-\sum_{\beta,j}\omega_{j}^{i}(X_{\beta})\theta^{\beta}\wedge\eta^{j}-\sum_{j<k}\Big{(}\omega_{k}^{i}(Z_{j})-\omega_{j}^{i}(Z_{k})\Big{)}\eta^{j}\wedge\eta^{k}
(7.21) =12Jβγiθβθγ+β,kωki(Xβ)ηkθβ+j,kωki(Zj)ηkηj\displaystyle=\frac{1}{2}J_{\beta\gamma}^{i}\theta^{\beta}\wedge\theta^{\gamma}+\sum_{\beta,k}\omega_{k}^{i}(X_{\beta})\eta^{k}\wedge\theta^{\beta}+\sum_{j,k}\omega_{k}^{i}(Z_{j})\eta^{k}\wedge\eta^{j}
(7.22) =12Jβγiθβθγ+kηk(βωki(Xβ)θβ+jωki(Zj)ηj)\displaystyle=\frac{1}{2}J_{\beta\gamma}^{i}\theta^{\beta}\wedge\theta^{\gamma}+\sum_{k}\eta^{k}\wedge\left(\sum_{\beta}\omega_{k}^{i}(X_{\beta})\theta^{\beta}+\sum_{j}\omega_{k}^{i}(Z_{j})\eta^{j}\right)
(7.23) =12Jβγiθβθγ+ηkωki.\displaystyle=\frac{1}{2}J_{\beta\gamma}^{i}\theta^{\beta}\wedge\theta^{\gamma}+\eta^{k}\wedge\omega_{k}^{i}.

(c): Let a,b{1,,n+m}a,b\in\{1,\ldots,n+m\} and consider an expansion of the left hand side:

(7.24) dωab\displaystyle d\omega_{a}^{b} =β<γnaβγbθβθγ+β,jhaβjbθβηj+j<kmajkbηjηk.\displaystyle=\sum_{\beta<\gamma}n_{a\beta\gamma}^{b}\theta^{\beta}\wedge\theta^{\gamma}+\sum_{\beta,j}h_{a\beta j}^{b}\theta^{\beta}\wedge\eta^{j}+\sum_{j<k}m_{ajk}^{b}\eta^{j}\wedge\eta^{k}.

Again, we calculate the coefficients in the expansion.

(7.25) naβγb\displaystyle n_{a\beta\gamma}^{b} =2dωab(Xβ,Xγ)=Xβωab(Xγ)Xγωab(Xβ)ωab([Xβ,Xγ]).\displaystyle=2\,d\omega_{a}^{b}(X_{\beta},X_{\gamma})=X_{\beta}\omega_{a}^{b}(X_{\gamma})-X_{\gamma}\omega_{a}^{b}(X_{\beta})-\omega_{a}^{b}\big{(}[X_{\beta},X_{\gamma}]\big{)}.

Recall that XγXa=Xcωac(Xγ)\nabla_{X_{\gamma}}X_{a}=X_{c}\otimes\omega_{a}^{c}(X_{\gamma}) and therefore ωab(Xγ)=g(Xb,XγXa)\omega_{a}^{b}(X_{\gamma})=g\big{(}X_{b},\nabla_{X_{\gamma}}X_{a}\big{)}. According to (1) the connection \nabla is metric and hence:

(7.26) Xβωab(Xγ)\displaystyle X_{\beta}\omega_{a}^{b}(X_{\gamma}) =Xβg(Xb,XγXa)=g(XβXb,XγXa)+g(Xb,XβXγXa)\displaystyle=X_{\beta}g\big{(}X_{b},\nabla_{X_{\gamma}}X_{a}\big{)}=g\big{(}\nabla_{X_{\beta}}X_{b},\nabla_{X_{\gamma}}X_{a}\big{)}+g\big{(}X_{b},\nabla_{X_{\beta}}\nabla_{X_{\gamma}}X_{a}\big{)}
(7.27) Xγωab(Xβ)\displaystyle X_{\gamma}\omega_{a}^{b}(X_{\beta}) =g(XγXb,XβXa)+g(Xb,XγXβXa).\displaystyle=g\big{(}\nabla_{X_{\gamma}}X_{b},\nabla_{X_{\beta}}X_{a}\big{)}+g\big{(}X_{b},\nabla_{X_{\gamma}}\nabla_{X_{\beta}}X_{a}\big{)}.

Moreover, one has:

(7.28) ωab([Xβ,Xγ])\displaystyle\omega_{a}^{b}\big{(}[X_{\beta},X_{\gamma}\big{]}\big{)} =g(Xb,[Xβ,Xγ]Xa)\displaystyle=g\big{(}X_{b},\nabla_{[X_{\beta},X_{\gamma}]}X_{a}\big{)}
(7.29) g(XβXb,XγXa)\displaystyle g\big{(}\nabla_{X_{\beta}}X_{b},\nabla_{X_{\gamma}}X_{a}\big{)} =g(Xρ,XβXb)g(Xρ,XγXa)=ωρb(Xβ)ωaρ(Xγ),\displaystyle=g\big{(}X_{\rho},\nabla_{X_{\beta}}X_{b}\big{)}g\big{(}X_{\rho},\nabla_{X_{\gamma}}X_{a}\big{)}=-\omega^{b}_{\rho}(X_{\beta})\omega_{a}^{\rho}(X_{\gamma}),

where we used the property that for all a,b,ca,b,c:

(7.30) ωac(Xb)=g(Xc,XbXa)=g(XbXc,Xa)=ωca(Xb),\omega_{a}^{c}(X_{b})=g(X_{c},\nabla_{X_{b}}X_{a})=-g(\nabla_{X_{b}}X_{c},X_{a})=-\omega_{c}^{a}(X_{b}),

based on the fact that the Bott connection \nabla is metric and that the basis is orthonormal. It also shows the skew-symmetry of the connection form; that is ωac=ωca\omega_{a}^{c}=-\omega_{c}^{a}.

Inserting these identities into the equations above yields:

(7.31) naβγb\displaystyle n_{a\beta\gamma}^{b} =g(Xb,R(Xβ,Xγ)Xa)+ωaρ(Xβ)ωρb(Xγ)ωaρ(Xγ)ωρb(Xβ)\displaystyle=g\big{(}X_{b},R(X_{\beta},X_{\gamma})X_{a}\big{)}+\omega_{a}^{\rho}(X_{\beta})\omega^{b}_{\rho}(X_{\gamma})-\omega_{a}^{\rho}(X_{\gamma})\omega^{b}_{\rho}(X_{\beta})
(7.32) =Rβγab+ωaρ(Xβ)ωρb(Xγ)ωaρ(Xγ)ωρb(Xβ).\displaystyle=R_{\beta\gamma a}^{b}+\omega_{a}^{\rho}(X_{\beta})\omega^{b}_{\rho}(X_{\gamma})-\omega_{a}^{\rho}(X_{\gamma})\omega^{b}_{\rho}(X_{\beta}).

A similar calculation shows

(7.33) haβjb\displaystyle h_{a\beta j}^{b} =2dωab(Xβ,Zj)=Xβ(ωab(Zj))Zj(ωab(Xβ))ωab([Xβ,Zj])\displaystyle=2\,d\omega_{a}^{b}(X_{\beta},Z_{j})=X_{\beta}\big{(}\omega_{a}^{b}(Z_{j})\big{)}-Z_{j}\big{(}\omega_{a}^{b}(X_{\beta})\big{)}-\omega_{a}^{b}\big{(}[X_{\beta},Z_{j}]\big{)}
(7.34) =Rβjab+ωaρ(Xβ)ωρb(Zj)ωaρ(Zj)ωρb(Xβ),\displaystyle=R_{\beta ja}^{b}+\omega_{a}^{\rho}(X_{\beta})\omega^{b}_{\rho}(Z_{j})-\omega_{a}^{\rho}(Z_{j})\omega^{b}_{\rho}(X_{\beta}),

and

(7.35) majkb\displaystyle m_{ajk}^{b} =2dωab(Zj,Zk))=Zjωab(Zk)Zkωab(Zj)ωab([Zj,Zk])\displaystyle=2\,d\omega_{a}^{b}(Z_{j},Z_{k}))=Z_{j}\omega_{a}^{b}(Z_{k})-Z_{k}\omega_{a}^{b}(Z_{j})-\omega_{a}^{b}\big{(}[Z_{j},Z_{k}]\big{)}
(7.36) =Rjkab+ωaρ(Zj)ωρb(Zk)ωaρ(Zk)ωρb(Zj).\displaystyle=R_{jka}^{b}+\omega_{a}^{\rho}(Z_{j})\omega^{b}_{\rho}(Z_{k})-\omega_{a}^{\rho}(Z_{k})\omega^{b}_{\rho}(Z_{j}).

Observe that for a,b,c,d{1,,n+m}a,b,c,d\in\{1,\ldots,n+m\},

(7.37) ωacωcb=ωac(Xd)ωcb(Xl)θdθl.\omega_{a}^{c}\wedge\omega^{b}_{c}=\omega_{a}^{c}(X_{d})\omega^{b}_{c}(X_{l})\theta^{d}\wedge\theta^{l}.

Using the identities in the expansion Eq. 7.24 together with the skew-symmetry of the coefficients naβγb=naγβbn_{a\beta\gamma}^{b}=-n_{a\gamma\beta}^{b} and majkb=makjbm_{ajk}^{b}=-m_{akj}^{b} defined above implies:

(7.38) dωab=\displaystyle d\omega_{a}^{b}= (12Rβγab+ωaρ(Xβ)ωρb(Xγ))θβθγ\displaystyle\Big{(}\frac{1}{2}R_{\beta\gamma a}^{b}+\omega_{a}^{\rho}(X_{\beta})\omega^{b}_{\rho}(X_{\gamma})\Big{)}\theta^{\beta}\wedge\theta^{\gamma}
(7.39) +Rβjabθβηj+ωaρ(Xβ)ωρb(Zj)θβηjωaρ(Zj)ωρb(Xβ)θβηj\displaystyle+R_{\beta ja}^{b}\theta^{\beta}\wedge\eta^{j}+\omega_{a}^{\rho}(X_{\beta})\omega^{b}_{\rho}(Z_{j})\theta^{\beta}\wedge\eta^{j}-\omega_{a}^{\rho}(Z_{j})\omega^{b}_{\rho}(X_{\beta})\theta^{\beta}\wedge\eta^{j}
(7.40) +(12Rjkab+ωaρ(Zj)ωρb(Zk))ηjηk\displaystyle+\Big{(}\frac{1}{2}R_{jka}^{b}+\omega_{a}^{\rho}(Z_{j})\omega^{b}_{\rho}(Z_{k})\Big{)}\eta^{j}\wedge\eta^{k}
(7.41) =12Rβγabθβθγ+Rβjabθβηj+12Rjkabηjηk+ωacωcb.\displaystyle=\frac{1}{2}R_{\beta\gamma a}^{b}\theta^{\beta}\wedge\theta^{\gamma}+R_{\beta ja}^{b}\theta^{\beta}\wedge\eta^{j}+\frac{1}{2}R_{jka}^{b}\eta^{j}\wedge\eta^{k}+\omega_{a}^{c}\wedge\omega^{b}_{c}.

We have used ωbc(Xβ)=0\omega_{b}^{c}(X_{\beta})=0 if β{1,,n}\beta\in\{1,\ldots,n\} and c>nc>n. ∎

References

  • [1] A. Agrachev, D. Barilari, and U. Boscain. A comprehensive introduction to sub-Riemannian geometry, volume 181 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2020. From the Hamiltonian viewpoint, With an appendix by Igor Zelenko.
  • [2] A. Agrachev, U. Boscain, J.-P. Gauthier, and F. Rossi. The intrinsic hypoelliptic Laplacian and its heat kernel on unimodular Lie groups. J. Funct. Anal., 256(8):2621–2655, 2009.
  • [3] L. Ambrosio, B. Kleiner, and E. Le Donne. Rectifiability of sets of finite perimeter in Carnot groups: existence of a tangent hyperplane. J. Geom. Anal., 19(3):509–540, 2009.
  • [4] M. Atiyah, R. Bott, and V. K. Patodi. On the heat equation and the index theorem. Invent. Math., 19:279–330, 1973.
  • [5] D. Barilari, U. Boscain, and R. W. Neel. Heat kernel asymptotics on sub-Riemannian manifolds with symmetries and applications to the bi-Heisenberg group. Ann. Fac. Sci. Toulouse Math. (6), 28(4):707–732, 2019.
  • [6] D. Barilari and L. Rizzi. A formula for Popp’s volume in sub-Riemannian geometry. Anal. Geom. Metr. Spaces, 1(2013):42–57, 2013.
  • [7] F. Baudoin and M. Bonnefont. The subelliptic heat kernel on SU(2): Representations, Asymptotics and Gradient bounds, Mar. 2008.
  • [8] F. Baudoin and G. Cho. The subelliptic heat kernel of the octonionic Hopf fibration. Potential Anal., 55(2):211–228, 2021.
  • [9] F. Baudoin and N. Garofalo. Curvature-dimension inequalities and Ricci lower bounds for sub-Riemannian manifolds with transverse symmetries. J. Eur. Math. Soc., 19(1):151–219, 2017.
  • [10] F. Baudoin, E. Grong, L. Rizzi, and G. Vega-Molino. H-type foliations. to appear: Differential Geom. Appl., 2022.
  • [11] F. Baudoin and J. Wang. The subelliptic heat kernel on the CR sphere. Math. Z., 275(1):135–150, 2013.
  • [12] F. Baudoin and J. Wang. The subelliptic heat kernels of the quaternionic Hopf fibration. Potential Anal., 41(3):959–982, 2014.
  • [13] W. Bauer and A. Laaroussi. Trivializable and Quaternionic Subriemannian Structures on 𝕊7\mathbb{S}^{7} and Subelliptic Heat Kernel. J. Geom. Anal., 32(8):1–37, 2022.
  • [14] R. Beals, B. Gaveau, and P. Greiner. The Green function of model step two hypoelliptic operators and the analysis of certain tangential Cauchy Riemann complexes. Adv. Math., 121(2):288–345, 1996.
  • [15] R. Beals, P. Greiner, and N. Stanton. The heat equation on a CR manifold. J. Differential Geom., 20(2):343–387, 1984.
  • [16] A. Bellaïche. The tangent space in sub-Riemannian geometry. In Sub-Riemannian geometry, volume 144 of Progr. Math., pages 1–78. Birkhäuser, Basel, 1996.
  • [17] G. Ben Arous. Développement asymptotique du noyau de la chaleur hypoelliptique sur la diagonale. Ann. Inst. Fourier (Grenoble), 39(1):73–99, 1989.
  • [18] E. Binz and S. Pods. The geometry of Heisenberg groups, volume 151 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2008. With applications in signal theory, optics, quantization, and field quantization, With an appendix by Serge Preston.
  • [19] O. Biquard. Métriques d’Einstein asymptotiquement symétriques. Asterisque, (265):vi+109, 2000.
  • [20] A. Bonfiglioli, E. Lanconelli, and F. Uguzzoni. Stratified Lie groups and potential theory for their sub-Laplacians. Springer Monographs in Mathematics. Springer, Berlin, 2007.
  • [21] W.-L. Chow. Über Systeme von linearen partiellen Differentialgleichungen erster Ordnung. Math. Ann., 117:98–105, 1939.
  • [22] M. Cowling, A. H. Dooley, A. Korányi, and F. Ricci. HH-type groups and Iwasawa decompositions. Adv. Math., 87(1):1–41, 1991.
  • [23] S. Dave and S. Haller. The heat asymptotics on filtered manifolds. J. Geom. Anal., 30(1):337–389, 2020.
  • [24] Y. C. de Verdière, L. Hillairet, and E. Trélat. Small-time asymptotics of hypoelliptic heat kernels near the diagonal, nilpotentization and related results. Ann. H. Lebesgue, 4:897–971, 2021.
  • [25] G. B. Folland. A fundamental solution for a subelliptic operator. Bull. Amer. Math. Soc., 79:373–376, 1973.
  • [26] B. Franchi, R. Serapioni, and F. Serra Cassano. Rectifiability and perimeter in the Heisenberg group. Math. Ann., 321(3):479–531, 2001.
  • [27] R. K. Hladky. Connections and curvature in sub-Riemannian geometry. Houston J. Math., 38(4):1107–1134, 2012.
  • [28] Y. Inahama and S. Taniguchi. Heat trace asymptotics on equiregular sub-Riemannian manifolds. J. Math. Soc. Japan, 72(4):1049–1096, 2020.
  • [29] S. Ivanov, I. Minchev, and D. Vassilev. Quaternionic contact Einstein structures and the quaternionic contact Yamabe problem. Mem. Amer. Math. Soc., 231, 2014.
  • [30] D. Jerison and J. M. Lee. Intrinsic CR normal coordinates and the CR Yamabe problem. J. Differential Geom., 29(2):303–343, 1989.
  • [31] A. Kaplan. Fundamental solutions for a class of hypoelliptic PDE generated by composition of quadratic forms. Trans. Amer. Math. Soc., 258(1):147–153, 1980.
  • [32] A. Korányi and H. M. Reimann. Foundations for the theory of quasiconformal mappings on the Heisenberg group. Adv. Math., 111(1):1–87, 1995.
  • [33] C. S. Kunkel. Quaternionic contact pseudohermitian normal coordinates. PhD thesis, 2008. Thesis (Ph.D.)–University of Washington.
  • [34] A. Laaroussi. Heat kernel asymptotics for quaternionic contact manifolds. arXiv preprint 2103.00892, Mar. 2021.
  • [35] G. Métivier. Fonction spectrale et valeurs propres d’une classe d’opérateurs non elliptiques. Comm. Partial Differential Equations, 1(5):467–519, 1976.
  • [36] R. Montgomery. A tour of subriemannian geometries, their geodesics and applications, volume 91 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2002.
  • [37] T. Morimoto. Cartan connection associated with a subriemannian structure. Differential Geom. Appl., 26(1):75–78, 2008.
  • [38] D. Müller and A. Seeger. Sharp LpL^{p} bounds for the wave equation on groups of Heisenberg type. Anal. PDE, 8(5):1051–1100, 2015.
  • [39] P. K. Rashevsky. Any two points of a totally nonholonomic space may be connected by an admissible line. Uch. Zap. Ped. Inst. im. Liebknechta, 2:83–94, 1938.
  • [40] C. Séguin and A. Mansouri. Short-time asymptotics of heat kernels of hypoelliptic Laplacians on unimodular Lie groups. J. Funct. Anal., 262(9):3891–3928, 2012.
  • [41] S. Semmes. An introduction to Heisenberg groups in analysis and geometry. Notices Amer. Math. Soc., 50(6):640–646, 2003.
  • [42] R. S. Strichartz. Sub-Riemannian geometry. J. Differential Geom., 24(2):221–263, 1986.
  • [43] N. Tanaka. On non-degenerate real hypersurfaces, graded Lie algebras and Cartan connections. Jpn. J. Math., 2(1):131–190, 1976.
  • [44] S. Tanno. Variational problems on contact Riemannian manifolds. Trans. Amer. Math. Soc., 314(1):349–379, 1989.
  • [45] P. Tondeur. Foliations on Riemannian manifolds. Universitext. Springer-Verlag, New York, 1988.
  • [46] S. M. Webster. Pseudohermitian structures on a real hypersurface. J. Differential Geom., 13(1):25–41, 1978.
  • [47] H. Weyl. The classical groups. Princeton Landmarks in Mathematics. Princeton University Press, Princeton, NJ, 1997. Their invariants and representations, Fifteenth printing, Princeton Paperbacks.