1 Introduction

The goal of the paper is to prove the following result:

Theorem 1.1

Let \({\mathbb {M}}\) be a smooth, connected, oriented and \(n+m\) dimensional compact manifold. We assume that \({\mathbb {M}}\) is equipped with a Riemannian foliation \({\mathcal {F}}\) with bundle-like metric g and totally geodesic m-dimensional leaves. We also assume that the horizontal distribution \({\mathcal {H}}={\mathcal {F}}^\perp \) is bracket-generating and that there exists \(\varepsilon >0\) such that

$$\begin{aligned} (\nabla _v J)_w = - \frac{1}{2\varepsilon } [J_v, J_w] \end{aligned}$$
(1.1)

for any \(v,w \in T_x{\mathbb {M}}\), \(x \in {\mathbb {M}}\), where \(\nabla \) is the Bott connection of the foliation and J is the tensor defined in (2.2). Denoting \(\chi ({\mathbb {M}})\) the Euler characteristic of \({\mathbb {M}}\):

  • If n or m is odd, then \(\chi ({\mathbb {M}})=0\);

  • If n and m are both even, then

    $$\begin{aligned} \chi ({\mathbb {M}})= \int _{\mathbb {M}}\hat{\omega }_{\mathcal {H}}^\varepsilon \wedge \left[ \det \left( \frac{{\mathscr {T}}}{\sinh ({\mathscr {T}})}\right) ^{1/2}\right] _m . \end{aligned}$$

Notations are further explained in Sect. 4, but we point out that a remarkable feature of that result is that the density \( \hat{\omega }_{\mathcal {H}}^\varepsilon \wedge \left[ \det \left( \frac{{\mathscr {T}}}{\sinh ({\mathscr {T}})}\right) ^{1/2}\right] _m\) essentially only depends on horizontal curvature quantities. Therefore, the theorem illustrates further the fact already observed in [4] that topological properties of \({\mathbb {M}}\) might be obtained from horizontal curvature invariants only provided that the bracket-generating condition of the horizontal distribution is satisfied; thus, in essence, the theorem is a sub-Riemannian result. We also note that the condition (1.1) is satisfied in a large class of examples including the H-type foliations introduced in [5], see Example 2.4.

The proof of Theorem 1.1 is based on the study of the heat semigroup generated by the hypoelliptic sub-Laplacian on forms recently introduced in [4]. The heat equation approach to Chern–Gauss-Bonnet type formulas (or index formulas) that we are using is of course not new: It was suggested by Atiyah–Bott [1] and McKean-Singer [16] and first carried out by Patodi [18] and Gilkey [12] and is by now classical, see the book [9]. However, a difficulty in our setting is that the sub-Laplacian on forms we consider is only hypoelliptic but not elliptic. To carry out the required small-time asymptotics analysis to obtain the horizontal Chern–Gauss–Bonnet formula, we will make use of the probabilistic Brownian Chen series parametrix method first introduced in [3] and which is easy to adapt to hypoelliptic situations, see [2].

The paper is organized as follows. In Sect. 2, we introduce the horizontal Laplacian on forms \(\Delta _{{\mathcal {H}},\varepsilon }\) and prove that it is a self-adjoint operator if and only if the condition (1.1) is satisfied. In Sect. 3, we prove a McKean–Singer type formula for \(\Delta _{{\mathcal {H}},\varepsilon }\), namely that for every \(t > 0\),

$$\begin{aligned} \mathbf {Str} ( e^{t \Delta _{{\mathcal {H}},\varepsilon }}) =\chi ({\mathbb {M}}). \end{aligned}$$

Finally, in Sect. 4 we study the small-time asymptotics of \(\mathbf {Str} ( e^{t \Delta _{{\mathcal {H}},\varepsilon }}) \) and conclude the proof of Theorem 1.1.

2 Preliminaries

In this section, we first recall the framework and notations of Baudoin and Grong [4] and the references therein to which we refer for further details. We then prove a necessary and sufficient condition for the form horizontal Laplacian of a totally geodesic foliation to be a symmetric operator.

2.1 Totally geodesic foliations

Let \(({\mathbb {M}},g)\) be a smooth, oriented, connected, compact Riemannian manifold with dimension \(n+m\). We assume that \({\mathbb {M}}\) is equipped with a foliation \({\mathcal {F}}\) with m-dimensional leaves. The distribution \({\mathcal {V}}\) formed by vectors tangent to the leaves is referred to as the set of vertical directions (or vertical subbundle). Define the horizontal subbundle \({\mathcal {H}}= {\mathcal {V}}^\perp \) as its orthogonal complement. We will always assume in this paper that the horizontal distribution \({\mathcal {H}}\) is everywhere bracket-generating. The foliation is called Riemannian and totally geodesic if for any \(X \in \Gamma ({\mathcal {H}})\), \(Z \in \Gamma ({\mathcal {V}})\), the respective conditions are satisfied,

$$\begin{aligned} ({\mathcal {L}}_Z g)(X,X) =0, \quad ({\mathcal {L}}_X g)(Z,Z) =0. \end{aligned}$$

Equivalently, we can describe these conditions using the Bott connection. Write \(\pi _{{\mathcal {H}}}\) and \(\pi _{{\mathcal {V}}}\) for the respective orthogonal projections to \({\mathcal {H}}\) and \({\mathcal {V}}\). Let \(\nabla ^g\) be the Levi–Civita connection of g. Introduce a new connection \(\nabla \) on \(T{\mathbb {M}}\) according to the rules,

$$\begin{aligned} \nabla _X Y = \left\{ \begin{array}{ll} \pi _{\mathcal {H}}( \nabla ^g_X Y) &{} \text {for any }X, Y \in \Gamma ({\mathcal {H}}), \\ \pi _{\mathcal {H}}([X,Y] )&{} \text {for any }X \in \Gamma ({\mathcal {V}}), Y \in \Gamma ({\mathcal {H}}), \\ \pi _{\mathcal {V}}([X,Y] )&{} \text {for any }X \in \Gamma ({\mathcal {H}}), Y \in \Gamma ({\mathcal {V}}), \\ \pi _{\mathcal {V}}(\nabla ^g_X Y) &{} \text {for any }X, Y \in \Gamma ({\mathcal {V}}). \end{array}\right. \end{aligned}$$
(2.1)

We observe that \(\nabla \) preserves \({\mathcal {H}}\) and \({\mathcal {V}}\) under parallel transport. The foliation \({\mathcal {F}}\) is then both Riemannian and totally geodesic if and only if \(\nabla g = 0\). For the rest of the paper, we will assume that \(\nabla \) is indeed compatible with the metric g. The torsion T of \(\nabla \) is given by

$$\begin{aligned} T(X, Y) = - \pi _{{\mathcal {V}}} [\pi _{{\mathcal {H}}} X, \pi _{{\mathcal {H}}} Y ]. \end{aligned}$$

Define a corresponding endomorphism valued one-form \(Z \mapsto J_Z\) by

$$\begin{aligned} \langle J_Z X, Y \rangle _g = \langle Z, T(X,Y) \rangle _g, \qquad X,Y, Z \in \Gamma (T{\mathbb {M}}).\end{aligned}$$
(2.2)

Let \(g_{\mathcal {H}}\) and \(g_{\mathcal {V}}\) be the respective restrictions of g to \({\mathcal {H}}\) and \({\mathcal {V}}\). We then define the canonical variation g by \(g_\varepsilon = g_{\mathcal {H}}\oplus \frac{1}{\varepsilon } g_{\mathcal {V}}\), \(\varepsilon >0\), and make the following observations:

  1. (i)

    If \(({\mathbb {M}}, {\mathcal {F}}, g)\) is a Riemannian, totally geodesic foliation, then so is \(({\mathbb {M}}, {\mathcal {F}}, g_\varepsilon )\).

  2. (ii)

    Although the Levi-Civita connection \(\nabla ^{g_\varepsilon }\) of \(g_\varepsilon \) is different from the connection \(\nabla ^g\) of g, replacing \(\nabla ^g\) with \(\nabla ^{g_\varepsilon }\) in formula (2.1) will lead to exactly the same connection. In other words, when defining the Bott connection \(\nabla \), we obtain the same connection for any metric \(g_\varepsilon \) in the family of canonical variations.

  3. (iii)

    For any fixed \(\varepsilon >0\), define a connection

    $$\begin{aligned} {\hat{\nabla }}^\varepsilon _XY = \nabla _X Y + \frac{1}{\varepsilon } J_X Y.\end{aligned}$$
    (2.3)

    This connection preserves \({\mathcal {H}}\) and \({\mathcal {V}}\) under parallel transport and is compatible with \(g_{\varepsilon '}\) for any \(\varepsilon ' >0\). Furthermore, its torsion

    $$\begin{aligned} {\hat{T}}^\varepsilon (X,Y) = T(X,Y) + \frac{1}{\varepsilon } J_X Y - \frac{1}{\varepsilon } J_Y X, \end{aligned}$$

    is skew-symmetric with respect to \(g_\varepsilon \). Hence, if we consider its adjoint connection

    $$\begin{aligned} \nabla _X^\varepsilon Y = {\hat{\nabla }}_X^\varepsilon Y - {\hat{T}}^\varepsilon (X,Y) = \nabla _X Y - T(X,Y) + \frac{1}{\varepsilon } J_Y X,\end{aligned}$$
    (2.4)

    it will also be compatible with \(g_\varepsilon \). However, \({\mathcal {H}}\) and \({\mathcal {V}}\) are not parallel with respect to \(\nabla ^\varepsilon \).

2.2 Horizontal Laplacian on forms

For the totally geodesic Riemannian foliation \(({\mathbb {M}}, {\mathcal {F}}, g)\), define its horizontal Laplacian on functions \( f \in C^\infty ({\mathbb {M}})\) by

$$\begin{aligned} \Delta _{{\mathcal {H}}} f = {{\,\mathrm{tr}\,}}_{{\mathcal {H}}} \nabla _{\times }df(\times ).\end{aligned}$$
(2.5)

We note that since \({\mathcal {H}}\) is assumed to be bracket-generating, from Hörmander’s theorem, \(\Delta _{{\mathcal {H}}}\) is a subelliptic operator. We also note that since \(g_{\mathcal {H}}\) and the Bott connection are independent of \(\varepsilon >0\), the horizontal Laplacian is as well; that is, the choice of any metric \(g_\varepsilon \) in the canonical variation family will not change \(g_{\mathcal {H}}\), the Bott connection, or the horizontal Laplacian.

Consider now the totally geodesic Riemannian foliation \(({\mathbb {M}}, {\mathcal {F}}, g_\varepsilon )\) for some fixed \(\varepsilon >0\). We want to extend the horizontal Laplacian on functions (2.5) to a differential operator on forms \(\Delta _{{\mathcal {H}},\varepsilon }\) satisfying the following requirements:

  1. (I)

    \(\Delta _{{\mathcal {H}},\varepsilon } f = \Delta _{{\mathcal {H}}}f\) for any smooth function f;

  2. (II)

    The operator \(\Delta _{{\mathcal {H}}, \varepsilon }\) is of Weitzenböck type, i.e., \(\Delta _{{\mathcal {H}}, \varepsilon } = L_{{\mathcal {H}}, \varepsilon } + {\mathscr {R}}_\varepsilon \) where \({\mathscr {R}}_\varepsilon \) is a zero-order differential operator and

    $$\begin{aligned} L_{{\mathcal {H}},\varepsilon } ={{\,\mathrm{tr}\,}}_{{\mathcal {H}}} {\tilde{\nabla }}_{\times , \times }^2,\end{aligned}$$
    (2.6)

    is the connection horizontal Laplacian of some connection \({\tilde{\nabla }}\) compatible with \(g_\varepsilon \);

  3. (III)

    If d is the exterior differential, then

    $$\begin{aligned}{}[\Delta _{{\mathcal {H}},\varepsilon }, d] =0. \end{aligned}$$

Given these requirements, there is an essentially unique extension of \(\Delta _{{\mathcal {H}}}\) to forms, see [4, 15] for details. We call \(\Delta _{{\mathcal {H}},\varepsilon }\) the \(\varepsilon \)-horizontal Laplacian on forms. This operator can described as follows.

Proposition 2.1

(Horizontal Laplacian on forms, see [4]) Consider the \(\varepsilon \)-horizontal divergence operator defined by

$$\begin{aligned} \delta _{{\mathcal {H}},\varepsilon } \eta = - {{\,\mathrm{tr}\,}}_{{\mathcal {H}}} (\nabla _\times ^\varepsilon \eta )(\times , \cdot ). \end{aligned}$$

The operator

$$\begin{aligned} \Delta _{{\mathcal {H}},\varepsilon } = - \delta _{{\mathcal {H}},\varepsilon } d - d\delta _{{\mathcal {H}},\varepsilon } \end{aligned}$$

is called the \(\varepsilon \)-horizontal Laplacian on forms, and it satisfies the requirements \(\mathrm {(I)}, \mathrm {(II)}, \mathrm {(III)}\). In particular, this operator has Weitzenböck decomposition \(\Delta _{{\mathcal {H}}, \varepsilon } = L_{{\mathcal {H}}, \varepsilon } + {\mathscr {R}}_\varepsilon \) where \(L_{{\mathcal {H}},\varepsilon }\) is defined as in (2.6) relative to \(\nabla ^\varepsilon \).

We can describe the zero order operator \({\mathscr {R}}_\varepsilon \) can be made explicit, see [4]. For later use, we will prefer to write the operators using Fermion calculus, see Appendix A.1. Let \(X_1, \dots , X_n\) and \(Z_1, \dots , Z_m\) be local orthonormal bases of, respectively, \({\mathcal {H}}\) and \({\mathcal {V}}\). Define \(a_i = \iota _{X_i}\) and \(b_r = \iota _{Z_r}\) for the corresponding annihilation operators, with the dual operators \(a^*_i = X^*_i \wedge \) and \(b_r^* = Z_r^* \wedge \) acting by wedge products. The dual are here relative to the \(L^2\) inner product with respect to the fixed metric g. Relative to the curvature tensor \({\hat{R}}^\varepsilon \) of \({\hat{\nabla }}^\varepsilon \), write

$$\begin{aligned} {\hat{R}}^{\varepsilon ,l}_{ijk} = \langle {\hat{R}}^\varepsilon (X_i, X_j) X_k, X_l \rangle _g,\end{aligned}$$
(2.7)

and use similar notation for other tensors with indices ijkl denoting evaluations with respect to the basis of \({\mathcal {H}}\), indices rs with respect to the basis of \({\mathcal {V}}\). We emphasize that these indices are always defined relative to the fixed metric g. Then, \({\mathscr {R}}_\varepsilon \) is given by

$$\begin{aligned} {\mathscr {R}}_\varepsilon&= \sum _{i,j,k=1}^n {\hat{R}}_{ijk}^{\varepsilon ,i} a_k^* a_j + \sum _{i,k=1}^n \sum _{r=1}^m {\hat{R}}_{irk}^{\varepsilon ,i} a_k^* b_r + \frac{1}{2} \sum _{i,j,k,l=1}^n {\hat{R}}_{ijk}^{\varepsilon ,l} a_k^* a_l^* a_j a_i \nonumber \\&\quad + \,\sum _{i,j,k=1}^n \sum _{r=1}^m {\hat{R}}_{irk}^{\varepsilon ,l} a_k^* a_l ^*b_r a_i + \frac{1}{2} \sum _{i,j=1}^n \sum _{r,s=1}^m {\hat{R}}_{rsi}^{\varepsilon ,j} a_i^* a_j^* b_r b_s. \end{aligned}$$
(2.8)

We want to give a formula for this operator that shows the dependence of \(\varepsilon \) explicitly. Let T and R be the curvature of the Bott connection \(\nabla \) and use indices after semi-colons to denote covariant derivatives with respect to this connection. Using Lemma A.2, Appendix, we can write

$$\begin{aligned} {\mathscr {R}}_\varepsilon&= \sum _{i,j,k=1}^n \left( R_{kji}^k + \frac{1}{\varepsilon } \sum _{r=1}^m T_{ik}^r T^r_{jk} \right) a_i^* a_j - \sum _{i,j=1}^n \sum _{r=1}^m T_{ij;i}^r a_j^* b_r \nonumber \\&\quad +\, \frac{1}{2} \sum _{i,j,k,l=1}^n \left( R_{kli}^j +\frac{1}{\varepsilon } \sum _{r=1}^m T_{kl}^r T_{ij}^r \right) a_i^* a_j^* a_l a_k + \sum _{i,j,k=1}^n \sum _{r=1}^m \frac{1}{\varepsilon } T_{ij;k}^r a_i^* a_j^* b_r a_k\nonumber \\&\quad + \,\frac{1}{2} \sum _{i,j=1}^n \sum _{r,s=1}^m \left( \frac{2}{\varepsilon } T_{ij;r}^s + \frac{1}{\varepsilon ^2} \sum _{k=1}^n (T_{kj}^r T_{ik}^s - T_{kj}^s T_{ik}^r) \right) a_i^* a_j^* b_s b_r. \end{aligned}$$
(2.9)

2.3 Symmetry of the horizontal Laplacian

Consider the exterior algebra

$$\begin{aligned} \Omega = \Omega ({\mathbb {M}}) = \bigoplus _{k=0}^{\dim {\mathbb {M}}} \Omega ^k, \end{aligned}$$

with the \(L^2\)-inner product from \(g_\varepsilon \). When restricted to elements in \(\Omega ^0 \oplus \Omega ^1\), the operator \(\Delta _{{\mathcal {H}},\varepsilon }\) is symmetric if and only if \({\mathcal {H}}\) satisfies the Yang–Mills condition, i.e., if

$$\begin{aligned} \sum _{i=1}^n T_{ij;i}^r =0,\qquad \text {for any}\quad j=1, \dots , n, r= 1, \dots , m. \end{aligned}$$

see [6]. Considering all forms, we have the following result.

Proposition 2.2

The operator \(\Delta _{{\mathcal {H}},\varepsilon }\) is symmetric with respect to the \(L^2\)-inner product of \(g_\varepsilon \) if and only if

$$\begin{aligned} (\nabla _v J)_w = - \frac{1}{2\varepsilon } [J_v, J_w],\end{aligned}$$
(2.10)

for any \(v,w \in T_xM\), \(x \in M\). In particular, \(\nabla _v J = 0\) for any \(v \in {\mathcal {H}}\).

We note that under the above condition, the expression of \({\mathscr {R}}_\varepsilon \) reduces to

$$\begin{aligned} {\mathscr {R}}_\varepsilon&= \sum _{i,j,k=1}^n \left( R_{kji}^k + \frac{1}{\varepsilon } \sum _{r=1}^m T_{ik}^r T^r_{jk} \right) a_i^* a_j + \frac{1}{2} \sum _{i,j,k,l=1}^n \left( R_{kli}^j +\frac{1}{\varepsilon } \sum _{r=1}^m T_{kl}^r T_{ij}^r \right) a_i^* a_j^* a_l a_k. \end{aligned}$$
(2.11)

Proof

\(L_{{\mathcal {H}},\varepsilon }\) is symmetric by Grong and Thalmaier [15, Lemma A.1], so we only need to determine when \({\mathscr {R}}_\varepsilon \) is symmetric. We choose a local bases \(X_1, \dots , X_n\) and \(Z_1, \dots , Z_m\) of, respectively, \({\mathcal {H}}\) and \({\mathcal {V}}\). We consider the representation of \({\mathscr {R}}_\varepsilon \) as in (2.9). Then, for \({\mathscr {R}}_\varepsilon \) to be symmetric, we must have

$$\begin{aligned} 0&= \langle {\mathscr {R}}_\varepsilon X_k^* \wedge Z_r^*, X_i^* \wedge X^j \rangle _{\varepsilon } - \langle {\mathscr {R}}_\varepsilon X_i^* \wedge X_j, X_k^* \wedge Z_r^* \rangle _{\varepsilon } = \frac{1}{\varepsilon } T_{ij;k}^r, \\ 0&= \langle {\mathscr {R}}_\varepsilon Z_r^* \wedge Z_s^*, X_i^* \wedge X^j \rangle _{\varepsilon } - \langle {\mathscr {R}}_\varepsilon X_i^* \wedge X_j, Z_r^* \wedge Z_s^* \rangle _{\varepsilon } \\&= \frac{2}{\varepsilon } T_{ij;r}^s + \frac{1}{\varepsilon ^2} \sum _{k=1}^n (T_{kj}^r T_{ik}^s - T_{kj}^s T_{ik}^r). \end{aligned}$$

These equations are clearly equivalent to (2.10). If these hold, then \({\mathscr {R}}_\varepsilon \) reduces to the expression (2.11), which is symmetric by Lemma A.3 (i). \(\square \)

Remark 2.3

If we assume that \(m=1\) (i.e., the leaves are one-dimensional), then it is immediate from the previous result that the following are equivalent:

  1. (i)

    \(\Delta _{{\mathcal {H}}, \varepsilon }\) is symmetric for some \(\varepsilon >0\).

  2. (ii)

    \(\Delta _{{\mathcal {H}}, \varepsilon }\) is symmetric for all \(\varepsilon >0\).

  3. (iii)

    \(\nabla J=0\).

Recall that the statement \(\nabla J = 0\) is equivalent to \(\nabla T = 0\). For \(m > 1\), the above statement remains true if we replace (i) by the following assumption

  1. (i’)

    \(\Delta _{{\mathcal {H}}, \varepsilon }\) is symmetric at least two values \(\varepsilon >0\) and \(\varepsilon ' >0\).

Example 2.4

(H-type foliations) Following definitions given in [5], we say that a foliated Riemannian manifold \(({\mathbb {M}}, {\mathcal {F}}, g)\) is of H-type if for every \(Z \in \Gamma ({\mathcal {V}})\), we have \(J_Z^2 = - \Vert Z \Vert _{{\mathcal {V}}}^2 \pi _{{\mathcal {H}}}\). Expand the definition of J from taking values from \({\mathcal {V}}\) to its Clifford algebra \(\mathbf {Cl}({\mathcal {V}})\) by the rule \(J_1 = \pi _{{\mathcal {H}}}\) and iteratively \(J_{u \cdot v} = J_u J_v\), \(u, v \in \mathbf {Cl}({\mathcal {V}})\). We then further say that the foliation is of horizontally parallel Clifford type if \(\nabla _X J =0\) for any horizontal vector fields \(X \in \Gamma ({\mathcal {H}})\) and while for \(u,v \in {\mathcal {V}}\).

$$\begin{aligned} (\nabla _{u}J)_v \in J_{\mathbf {Cl}({\mathcal {V}})}. \end{aligned}$$

It then turns out that for some \(\kappa \in {\mathbb {R}}\),

$$\begin{aligned} (\nabla _{u}J)_v = - \kappa J_{u \cdot v + \langle u, v \rangle } = - \frac{\kappa }{2} [J_u, J_v]. \end{aligned}$$

The number \(\kappa \) determines the Ricci curvature of \(\nabla \), see [5, Theorem 3.16]. We see that if we have an H-type Riemannian foliation \(({\mathbb {M}}, {\mathcal {F}}, g)\) of horizontally parallel Clifford type, then \(\Delta _{{\mathcal {H}},\varepsilon }\) is symmetric with respect to \(g_\varepsilon \) for \(\varepsilon = \frac{1}{\kappa }\).

Finally, to conclude the section we point out the following result. For the definition of the Carnot–Carathéodory metric \(d_{cc}\) of the sub-Riemannian manifold \(({\mathbb {M}}, {\mathcal {H}}, g_{{\mathcal {H}}})\) and the tangent cone of a metric space, see, e.g., [13].

Corollary 2.5

Assume that \(\Delta _{{\mathcal {H}},\varepsilon }\) is symmetric on forms for some fixed \(\varepsilon >0\). Then, the following holds:

  1. (a)

    The horizontal bundle \({\mathcal {H}}\) has step 2, that is \({\mathcal {H}}+ [{\mathcal {H}}, {\mathcal {H}}] = T{\mathbb {M}}\). In particular, the torsion T of the Bott connection \(\nabla \) will be surjective on \({\mathcal {V}}\).

  2. (b)

    The tangent cones of the metric space \(({\mathbb {M}}, d_{cc})\) at any pair of points \(x, y \in {\mathbb {M}}\) are isometric.

Proof

  1. (a)

    Recall that if \(\Delta _{{\mathcal {H}},\varepsilon }\) is symmetric on forms for some \(\varepsilon >0\), then in particular \(\nabla _v J =0\) for any \(v \in {\mathcal {H}}\). We can rewrite it as \(\nabla _v T = 0\) for any \(v \in {\mathcal {H}}\) since \(\nabla \) is compatible with g. Define \({\mathcal {H}}^2 = {\mathcal {H}}+ [{\mathcal {H}}, {\mathcal {H}}]\) and let \(X_1, X_2, X_3 \in \Gamma ({\mathcal {H}})\) be arbitrary. We first see that

    $$\begin{aligned} T(X_2, X_3) = \nabla _{X_2} X_3 - \nabla _{X_3} X_2 - [X_2,X_3] =0 \mod {\mathcal {H}}^2, \end{aligned}$$

    since \(\nabla \) preserves \({\mathcal {H}}\). Furthermore, by the definition of the Bott connection

    $$\begin{aligned}{}[X_1, [X_2, X_3]]&= - [X_1, T(X_2, X_3)] \mod {\mathcal {H}}^2 = - \nabla _{X_1} T(X_2, X_3) \mod {\mathcal {H}}^2 \\&\quad = - T(\nabla _{X_1} X_2, X_3) - T(X_2, \nabla _{X_1} X_3) \mod {\mathcal {H}}^2 = 0 \mod {\mathcal {H}}^2. \end{aligned}$$

    It follows that \({\mathcal {H}}\) only generates \({\mathcal {H}}^2\). As we assumed that \({\mathcal {H}}\) is bracket generating, we have \({\mathcal {H}}^2 = TM\).

  2. (b)

    Since both \({\mathcal {H}}\) and \({\mathcal {H}}^2 = {\mathcal {H}}+ [{\mathcal {H}}, {\mathcal {H}}] = TM\) have constant rank, it follows by Mitchell [17] and Bellaïche [8] that the tangent cone at a point x is a Carnot group \(G_x\). Its Lie algebra \({\mathfrak {g}}_x\) is given by

    $$\begin{aligned} {\mathfrak {g}}_{x} = {\mathfrak {g}}_{x,1} \oplus {\mathfrak {g}}_{x,2} = {\mathcal {H}}_x \oplus T_xM/{\mathcal {H}}_x, \end{aligned}$$

    where \(TM/{\mathcal {H}}_x\) is the center, and for \(X_x, Y_x \in {\mathcal {H}}_x = {\mathfrak {g}}_{x,1}\) the Lie bracket is defined as

    $$\begin{aligned} {{\,\mathrm{[ \! [}\,}}X_x, Y_x {{\,\mathrm{] \! ]}\,}}= [X,Y] |_x \mod {\mathcal {H}}_x. \end{aligned}$$

    where XY are any pair of vector fields extending this vectors. The Carnot group \(G_x\) is then the corresponding simply connected Lie group of \({\mathfrak {g}}_x\) with the sub-Riemannian structure given by left translation of \({\mathfrak {g}}_x = {\mathcal {H}}_x\) and its inner product.

    If identify \({\mathfrak {g}}_x = {\mathcal {H}}_x \oplus T_xM /{\mathcal {H}}_x\) with \(T_xM = {\mathcal {H}}_x \oplus {\mathcal {V}}_x\) through the map \(v \mod {\mathcal {H}}_x \mapsto \pi _{{\mathcal {V}}_x}(v)\), \(v \in T_xM\), then the Lie bracket becomes,

    $$\begin{aligned} {{\,\mathrm{[ \! [}\,}}v, w {{\,\mathrm{] \! ]}\,}}= - T(v,w), \qquad v, w\in T_xM. \end{aligned}$$

    Let now y be any other point and let \(\gamma :[0,1] \rightarrow {\mathbb {M}}\) be any horizontal curve from x to y, which exists form our assumption that \({\mathcal {H}}\) satisfies the bracket-generating condition. Then, \(\nabla _{{\dot{\gamma }}(t)} T =0\) for any \(t \in [0,1]\), so if we write

    $$\begin{aligned} {{\,\mathrm{/\! \! /}\,}}_{\gamma ,t} = {{\,\mathrm{/\! \! /}\,}}_t : T_{x} {\mathbb {M}}\rightarrow T_{\gamma (t)} {\mathbb {M}}, \end{aligned}$$

    for the parallel transport map along \(\gamma \), then this satisfies

    $$\begin{aligned} {{\,\mathrm{/\! \! /}\,}}_t T(u, v) = T({{\,\mathrm{/\! \! /}\,}}_t u,{{\,\mathrm{/\! \! /}\,}}_t v), \qquad v,w \in T_x{\mathbb {M}}. \end{aligned}$$

    As a consequence, \({{\,\mathrm{/\! \! /}\,}}_1:{\mathfrak {g}}_x= T_x {\mathbb {M}}\rightarrow {\mathfrak {g}}_y = T_y {\mathbb {M}}\) is a Lie algebra isomorphism, which can be integrated to a Lie group isomorphism from \(G_x\) to \(G_y\). Since the parallel transport \({{\,\mathrm{/\! \! /}\,}}_1\) also maps \({\mathcal {H}}_x\) onto \({\mathcal {H}}_y\) isometrically, the induced map on Carnot groups is in fact a sub-Riemannian isometry.

\(\square \)

3 Horizontal McKean–Singer theorem

We work on a totally geodesic foliation \(({\mathbb {M}},{\mathcal {F}},g)\) and assume that there is some \(0< \varepsilon < +\infty \) such that horizontal Laplacian \(\Delta _{{\mathcal {H}},\varepsilon }\), is symmetric. From Proposition 2.2, this assumption is equivalent to the fact that

$$\begin{aligned} (\nabla _v J)_w = -\frac{1}{2\varepsilon } [J_v, J_w]. \end{aligned}$$

Since \(\Delta _{{\mathcal {H}},\varepsilon }\) commutes with d on smooth forms and is symmetric, it also commutes on smooth forms with the coderivative \(\delta _\varepsilon \), and thus, it also commutes with the Hodge–de Rham operator \(\Delta _\varepsilon := -d\delta _\varepsilon -\delta _\varepsilon d\) on smooth forms. From Hodge theorem, the operator \(\Delta _\varepsilon \) is elliptic with a compact resolvent and the space of \(L^2\)-forms can be decomposed as \(\oplus _{k=0}^{+\infty } E_{\lambda _k}\) where the \(E_{\lambda _k}\)’s are the eigenspaces of \(\Delta _\varepsilon \). Those eigenspaces only contain smooth forms, therefore \(\Delta _{{\mathcal {H}},\varepsilon }(E_{\lambda _k}) \subset E_{\lambda _k}\). This implies that \(\Delta _{{\mathcal {H}},\varepsilon }\) is essentially self-adjoint and generates the semigroup:

$$\begin{aligned} e^{t \Delta _{{\mathcal {H}},\varepsilon }}=\oplus _{k=0}^{+\infty } e^{t \Delta _{{\mathcal {H}},\varepsilon }\mid E_{\lambda _k}} \end{aligned}$$
(3.1)

By hypoellipticity (see [4, Lemma 4.9]), this semigroup has a smooth kernel \(p_{{\mathcal {H}}, \varepsilon } (t, x,y)\) and is a bounded trace class operator in \( L_\mu ^2 (\wedge ^\cdot {\mathbb {M}}, g_\varepsilon )\). Let us denote by \(E^+_0(\Delta _{{\mathcal {H}},\varepsilon })\) (resp. \(E^-_0(\Delta _{{\mathcal {H}},\varepsilon })\)) the space of harmonic even forms for \(\Delta _{{\mathcal {H}},\varepsilon }\) (resp. the space of harmonic odd forms for \(\Delta _{{\mathcal {H}},\varepsilon }\)).

The goal of the section is to prove the following theorem, which is an analogue for our horizontal Laplacian of the classical McKean–Singer formula found in [16] :

Theorem 3.1

(Horizontal McKean-Singer formula) For every \(t >0\),

$$\begin{aligned} \mathbf {Str} ( e^{t \Delta _{{\mathcal {H}},\varepsilon }}):&= \int _{\mathbb {M}}\mathbf {Tr} (p^{+}_{{\mathcal {H}},\varepsilon } (t,x,x) ) d\mu (x) - \int _{\mathbb {M}}\mathbf {Tr} (p^{-}_{{\mathcal {H}},\varepsilon } (t,x,x) ) d\mu (x) \\&=\dim E^+_0(\Delta _{{\mathcal {H}},\varepsilon }) -\dim E^-_0(\Delta _{{\mathcal {H}},\varepsilon }) \\&=\chi ({\mathbb {M}}) \end{aligned}$$

where \(\chi ({\mathbb {M}})\) is the Euler characteristic of \({\mathbb {M}}\).

We turn to the proof of Theorem 3.1. We denote by

$$\begin{aligned} {\mathbf {D}}_\varepsilon =d+\delta _\varepsilon \end{aligned}$$

the Dirac operator of the metric \(g_\varepsilon \). Observe that \({\mathbf {D}}_\varepsilon \) commutes with \(\Delta _{{\mathcal {H}},\varepsilon }\) since both d and \(\delta _\varepsilon \) commute with it. The main idea to prove Theorem 3.1 is to introduce a deformation of \(\Delta _{{\mathcal {H}},\varepsilon }\) as follows:

$$\begin{aligned} \square _{\varepsilon , \theta }= (1-\theta )\Delta _{{\mathcal {H}},\varepsilon }-\theta {\mathbf {D}}_\varepsilon ^2 , \quad \theta \in [0,1]. \end{aligned}$$

A first lemma is the following:

Lemma 3.2

Let \(\lambda \) be a nonzero eigenvalue of \(\square _{\varepsilon , \theta }\). Then, \({\mathbf {D}}_\varepsilon : E^+_\lambda (\square _{\varepsilon , \theta }) \rightarrow E^-_\lambda (\square _{\varepsilon , \theta })\) is an isomorphism. Therefore, \(\dim E^+_\lambda (\square _{\varepsilon , \theta }) =\dim E^-_\lambda (\square _{\varepsilon , \theta })\).

Proof

Let \(\lambda \) be a nonzero eigenvalue of \(\square _{\varepsilon , \theta }\). The corresponding eigenspace \(E_\lambda (\square _{\varepsilon , \theta })\) is finite-dimensional since \(e^{t \square _{\varepsilon , \theta } }\) is a compact operator for \(t>0\). Moreover, since \({\mathbf {D}}_\varepsilon \) commutes with \(\square _{\varepsilon , \theta }\), \({\mathbf {D}}_\varepsilon : E^+_\lambda (\square _{\varepsilon , \theta }) \rightarrow E^-_\lambda (\square _{\varepsilon , \theta })\) is well defined. Let now \(\alpha \in E^+_\lambda (\square _{\varepsilon , \theta }) \) such that \({\mathbf {D}}_{\varepsilon } \alpha =0\). One has then

$$\begin{aligned} d\alpha =-\delta _\varepsilon \alpha . \end{aligned}$$

This implies that

$$\begin{aligned} \Vert d\alpha \Vert ^2_{ L^2 (\wedge ^\cdot {\mathbb {M}}, g_\varepsilon )}= -\langle d\alpha , \delta _\varepsilon \alpha \rangle _{ L^2 (\wedge ^\cdot {\mathbb {M}}, g_\varepsilon )}=0, \end{aligned}$$

so \(d\alpha =0\). Similarly, one has \(\Vert \delta _\varepsilon \alpha \Vert ^2_{ L^2 (\wedge ^\cdot {\mathbb {M}}, g_\varepsilon )}=0,\) so \(\delta _\varepsilon \alpha =0\). Therefore,

$$\begin{aligned} \alpha =\frac{1-\theta }{\lambda } \Delta _{{\mathcal {H}},\varepsilon }\alpha =-\frac{1-\theta }{\lambda } (d \delta _{{\mathcal {H}},\varepsilon } + \delta _{{\mathcal {H}},\varepsilon } d) \alpha =-\frac{1-\theta }{\lambda } d \delta _{{\mathcal {H}},\varepsilon }\alpha . \end{aligned}$$

One deduces

$$\begin{aligned} \Vert \alpha \Vert ^2_{ L^2 (\wedge ^\cdot {\mathbb {M}}, g_\varepsilon )}=-\frac{1-\theta }{\lambda } \langle \alpha , d \delta _{{\mathcal {H}},\varepsilon }\alpha \rangle _{ L^2 (\wedge ^\cdot {\mathbb {M}}, g_\varepsilon )}=-\frac{1-\theta }{\lambda } \langle \delta _\varepsilon \alpha , \delta _{{\mathcal {H}},\varepsilon }\alpha \rangle _{ L^2 (\wedge ^\cdot {\mathbb {M}}, g_\varepsilon )}=0. \end{aligned}$$

As a consequence, \({\mathbf {D}}_\varepsilon : E^+_\lambda (\square _{\varepsilon , \theta }) \rightarrow E^-_\lambda (\square _{\varepsilon , \theta })\) is injective. Let us now prove that it is surjective. Let \(\alpha \in E^-_\lambda (\square _{\varepsilon , \theta })\) which is orthogonal to the space \({\mathbf {D}}_\varepsilon E^+_\lambda (\square _{\varepsilon , \theta })\). For every \(\omega \in E^+_\lambda (\square _{\varepsilon , \theta })\), one has

$$\begin{aligned} 0=\langle \alpha , {\mathbf {D}}_\varepsilon \omega \rangle _{ L^2 (\wedge ^\cdot {\mathbb {M}}, g_\varepsilon )}=\langle {\mathbf {D}}_\varepsilon \alpha , \omega \rangle _{ L^2 (\wedge ^\cdot {\mathbb {M}}, g_\varepsilon )}. \end{aligned}$$

Thus, \({\mathbf {D}}_\varepsilon \alpha =0\) and from the first part of the proof, we deduce that \(\alpha =0\). We conclude that \({\mathbf {D}}_\varepsilon : E^+_\lambda (\square _{\varepsilon , \theta }) \rightarrow E^-_\lambda (\square _{\varepsilon , \theta })\) is indeed an isomorphism. \(\square \)

A second lemma is the following:

Lemma 3.3

For every \(t>0\), the map \(\theta \rightarrow \mathbf {Str} ( e^{t \square _{\varepsilon , \theta } })\) is continuous on [0, 1].

Proof

Let \(q_{\varepsilon ,\theta }(t,x,y)\) be the heat kernel of \(\square _{\varepsilon , \theta }=(1-\theta ) \Delta _{{\mathcal {H}},\varepsilon }- \theta {\mathbf {D}}^2_\varepsilon \), \(p_{{\mathcal {H}},\varepsilon } (t,x,y)\) be the heat kernel of \(\Delta _{{\mathcal {H}},\varepsilon }\) and \(p_\varepsilon (t,x,y)\) be the heat kernel of \(-{\mathbf {D}}^2_\varepsilon \). Since \(-{\mathbf {D}}^2_\varepsilon \) and \(\Delta _{{\mathcal {H}},\varepsilon }\) commute, we have

$$\begin{aligned} e^{t\square _{\varepsilon , \theta } }= e^{t(1-\theta ) \Delta _{{\mathcal {H}},\varepsilon }} e^{-t\theta {\mathbf {D}}^2_\varepsilon }. \end{aligned}$$

Therefore:

$$\begin{aligned} q_{\varepsilon ,\theta }(t,x,y) =\int _{\mathbb {M}}p_{{\mathcal {H}},\varepsilon } (t(1-\theta ),x,z) p_\varepsilon (t \theta ,z,y) dz \end{aligned}$$

and the result easily follows since

$$\begin{aligned} \mathbf {Str} ( e^{t \square _{\varepsilon , \theta } })=\int _{\mathbb {M}}q_{\varepsilon ,\theta }(t,x,x) dx. \end{aligned}$$

\(\square \)

We are now ready for the proof of Theorem  3.1.

Proof

From the first lemma:

$$\begin{aligned}&\mathbf {Str} ( e^{t \square _{\varepsilon , \theta } }) \\&\quad = \dim E^+_0(\square _{\varepsilon , \theta }) -\dim E^-_0(\square _{\varepsilon , \theta })+ \sum _{ \lambda \ne 0}( \dim E^+_\lambda (\square _{\varepsilon , \theta }) -\dim E^-_\lambda (\square _{\varepsilon , \theta })) e^{\lambda t} \\&\quad = \dim E^+_0(\square _{\varepsilon , \theta }) -\dim E^-_0(\square _{\varepsilon , \theta }). \end{aligned}$$

Therefore, \(\mathbf {Str} ( e^{t \square _{\varepsilon , \theta } }) \in {\mathbb {Z}}\). From the second lemma, \(\theta \rightarrow \mathbf {Str} ( e^{t \square _{\varepsilon , \theta } })\) is continuous, thus constant. We deduce

$$\begin{aligned} \mathbf {Str} ( e^{t \square _{\varepsilon , 0} }) =\mathbf {Str} ( e^{t \square _{\varepsilon , 1} }). \end{aligned}$$

Since \(\square _{{\varepsilon ^*}, 1} = -{\mathbf {D}}_\varepsilon ^2\) is the Hodge–de Rham Laplacian of the Riemannian manifold \(({\mathbb {M}}, g_\varepsilon )\), from the usual Riemannian Hodge theory (see [16]), we have

$$\begin{aligned} \mathbf {Str} ( e^{t \square _{\varepsilon , 1} })=\chi ({\mathbb {M}}), \end{aligned}$$

4 which concludes the proof. \(\square \)

Remark 3.4

(Dependence on the symmetry condition) It would obviously be beneficial to prove the above statement without the assumption of symmetry on \(\Delta _{{\mathcal {H}},\varepsilon }\). A semigroup approach to non-symmetric horizontal Laplacians has been used, see [15, Appendix A]. In the above proof, however, we really rely on the fact that \(\Delta _{{\mathcal {H}},\varepsilon }\) commutes with the codifferential \(\delta _\varepsilon \), and with the Laplace–Beltrami operator \(- {\mathbf {D}}_\varepsilon ^2\). We can no longer use these properties if we remove the symmetry assumption.

4 Horizontal Chern–Gauss–Bonnet formula

As before, we consider the horizontal Laplacian

$$\begin{aligned} \Delta _{{\mathcal {H}},\varepsilon }=-d \delta _{{\mathcal {H}},\varepsilon }-\delta _{{\mathcal {H}},\varepsilon } d, \end{aligned}$$

and assume that it is symmetric for a fixed \(\varepsilon \). As seen earlier, \(\Delta _{{\mathcal {H}},\varepsilon }\) satisfies the Weitzenböck identity

$$\begin{aligned} \Delta _{{\mathcal {H}},\varepsilon }= L_{{\mathcal {H}},\varepsilon } - {\mathscr {R}}_\varepsilon =-(\nabla _{\mathcal {H}}^\varepsilon )^* \nabla _{\mathcal {H}}^\varepsilon -{\mathscr {R}}_\varepsilon . \end{aligned}$$
(4.1)

where the later equality follows from [15, Lemma 2.1]. The goal of the section is to compute the pointwise limit

$$\begin{aligned} \lim _{t \rightarrow 0} \mathbf {Str}\text { } (p_{{\mathcal {H}},\varepsilon } (t,x,x)) \end{aligned}$$

and deduce from it our horizontal Chern–Gauss–Bonnet formula. The computation of that limit will be based on the probabilist method of Brownian Chen series (see [3, 7]) which has the advantage of being easily adapted to subelliptic operators like \(\Delta _{{\mathcal {H}},\varepsilon }\), see [2]. For convenience and to introduce notation, we include in Appendix A.2 the main elements of that theory.

A first step to implement the method in [2] is to study the small-time heat kernel asymptotics of a diffusion tangent to the scalar horizontal Laplacian \(\Delta _{{\mathcal {H}}}\) . Since we assume that \(\Delta _{{\mathcal {H}},\varepsilon }\) is symmetric, from Corollary 2.5 one has \(T{\mathbb {M}}={\mathcal {H}} + [{\mathcal {H}},{\mathcal {H}}]\), and thus the tangent diffusion will take its values in a two-step Carnot group [the so-called tangent cone, see Corollary 2.5(b)] for which an explicit formula for the heat kernel is known (see [10, 11]). In a local horizontal frame \(\{ X_1,\ldots ,X_n \}\) around \(x_0\) write

$$\begin{aligned}V_t(x_0) = \sum _{i=1}^n \sqrt{2} X_i (x_0) B^i_t + \sum _{1 \le i <j \le n} \pi _{{\mathcal {V}}}( [X_i,X_j] (x_0)) \int _0^t B^i_s dB^j_s -B^j_s dB^i_s,\end{aligned}$$

where \((B_t)_{t \ge 0}\) is a Brownian motion in \({\mathbb {R}}^n\). We note that \(V_t(x_0)\) can be written in a basis free way as

$$\begin{aligned} \sqrt{2} B_t(x_0) - \int _0^t T(B_s(x_0), dB_s(x_0)) \end{aligned}$$

where \(B_t(x_0)= \sum _{i=1}^n X_i (x_0) B^i_t \) is a standard Brownian motion in \({\mathcal {H}}_{x_0}\).

Lemma 4.1

Let \(x_0 \in {\mathbb {M}}\). For \(t>0\), let \(d_t (x_0)\) be the density at 0 of the \(T_{x_0} {\mathbb {M}}\) valued random variable \(V_t(x_0)\). Then, when \(t \rightarrow 0\),

$$\begin{aligned} d_t (x_0) \sim \frac{2^m}{(4\pi t)^{\frac{n}{2}+m} } \int _{{\mathcal {V}}_{x_0}} \det \left( \frac{\sqrt{J_z^* J_z}}{\sinh \sqrt{J_z^* J_z}}\right) ^{1/2} dz. \end{aligned}$$

Proof

The process \((V_t(x_0))_{t \ge 0}\) is the horizontal Brownian motion in the tangent cone \(G_{x_0}\) which is a 2-step Carnot group when it is identified with \(T_{x_0}{\mathbb {M}}\) using the group exponential map. The heat kernel of the horizontal Laplacian is known explicitly in 2-step Carnot groups (see [10, 11]) which yields the small-time asymptotics. \(\square \)

Remark 4.2

We note that \(d_t(x_0)\) is independent of \(x_0\) because of Corollary 2.5(b).

In the sequel, we will use the notation \({\mathcal {F}}_I\) (defined with respect to the connection \(D=\nabla ^{\varepsilon }\)) and \(\Lambda _I(B)_t\), as introduced and discussed in Appendix A.2.

Corollary 4.3

It will hold that as \(t \rightarrow 0\)

$$\begin{aligned} \mathbf {Str} ( p_{{\mathcal {H}},\varepsilon }(t,x_0,x_0)) \sim d_t(x_0) {\mathbb {E}}\left( \left. \mathbf {Str} \left( \exp \left( \sum _{I,d(I) \le n+2m} \Lambda _I(B)_t {\mathcal {F}}_I\right) (x_0) \right) \right| B_1= 0 \right) \end{aligned}$$

where \(d_t(x_0)\) is the density at 0 of \(V_t(x)\), as in Lemma 4.1.

Proof

Since \({\mathcal {H}}\) is two-step bracket generating, the homogeneous dimension is \(Q = \dim {\mathcal {H}} + 2\dim {\mathcal {V}} = n + 2m\). Taking \(N = n + 2m\) in Theorem A.1, and applying similar arguments as in the proof of Proposition 4.2 in [3], the corollary follows by recognizing that for \(|I| > 2, X_I\) is a linear combination of \(X_i , [X_j,X_k]\) so that when \(t \rightarrow 0\) the density at 0 of

$$\begin{aligned}\sum _{I, d(I) \le n+2m} \Lambda _I(B)_t X_I \end{aligned}$$

is equivalent to \(d_t(x_0)\) from the previous lemma. \(\square \)

Applying the previous results, we are now able to compute \(\lim _{t \rightarrow 0} \mathbf {Str} (p_{{\mathcal {H}},\varepsilon } (t,x_0,x_0))\). Choose local orthonormal bases \(X_1, \dots , X_n\) and \(Z_1, \dots , Z_m\) of, respectively, \({\mathcal {H}}\) and \({\mathcal {V}}\).

Lemma 4.4

The integral

$$\begin{aligned} {\mathcal {J}} = {\mathcal {J}}(x_0)=\frac{2^m}{(2\pi )^{\frac{n}{2}+m} } \int _{{\mathcal {V}}_{x_0}} \det \left( \frac{\sqrt{J_z^* J_z}}{\sinh \sqrt{J_z^* J_z}}\right) ^{1/2} dz, \end{aligned}$$

is a constant, so independent of the point \(x_0 \in {\mathbb {M}}\) chosen. Furthermore, it holds that

$$\begin{aligned} \lim _{t \rightarrow 0} \mathbf {Str}(p_{{\mathcal {H}},\varepsilon } (t,x_0,x_0) ) = {\left\{ \begin{array}{ll} \frac{{\mathcal {J}}}{\left( \frac{n}{2}+m\right) ! }{\mathbb {E}} \left( \left. \mathbf {Str} \left[ A_{x_0}^{\frac{n}{2}+m} \right] \right| B_1=0\right) , \quad \text {if n is even}\\ \\ 0, \quad \text {if n is odd}. \end{array}\right. } \end{aligned}$$

where the random variable \(A_{x_0}\) is given by

$$\begin{aligned} A_{x_0}= & {} - \frac{1}{2} \sum _{i,j,k,l=1}^n \left( R_{kli}^j +\frac{1}{\varepsilon } \sum _{r=1}^m T_{kl}^r T_{ij}^r \right) a_i^* a_j^* a_l a_k\nonumber \\&\quad , \sum _{1 \le i<j \le n} \sum _{r,s=1}^m T_{ij;r}^s b_r^* b_s \int _0^1 B^i_t dB^j_t -B^j_t dB^i_t. \end{aligned}$$
(4.2)

Proof

First, observe that

$$\begin{aligned}{\mathcal {J}}(x_0) = (2t)^{\frac{n}{2}+m} d_t(x_0),\end{aligned}$$

and so the independence of \({\mathcal {J}}(x_0)\) from \(x_0\) follows from Corollary 2.5(b) as in Remark 4.2.

Consider the expansion

$$\begin{aligned}\mathbf {Str} \left[ {\text {exp}}\left( \sum _{I,d(I) \le n+2m} \Lambda _I(B)_t {\mathcal {F}}_I\right) (x_0)\right] \ = \sum _{k \ge 0}\frac{1}{k!}\mathbf {Str} \left[ \left( \sum _{I,d(I) \le n+2m} \Lambda _I(B)_t {\mathcal {F}}_I\right) ^k(x_0)\right] . \end{aligned}$$

From the Weitzenböck identity (4.1), we have for \(i,j \in \{1,\dots ,n+m\}\) that

$$\begin{aligned}{\mathcal {F}}_0 = -{\mathscr {R}}_\varepsilon , \quad {\mathcal {F}}_i = 0, \quad {\mathcal {F}}_{(i,j)} = \hat{R}^\varepsilon (Y_i,Y_j) \end{aligned}$$

where \(\{Y_1, \dots , Y_{n+m}\}\) form a local orthonormal frame and the \(\{c_i,c^*_i\}_{i=1}^{n+m}\) form the associated Fermion calculus of \(T{\mathbb {M}}\). Equation (2.11) allows us to write

$$\begin{aligned}{\mathscr {R}}_\varepsilon = \sum _{i,j ,k=1}^n \langle {\hat{R}}^\varepsilon (X_i, X_k) X_j, X_i \rangle _g a_k^* a_i + \sum _{i,j,k,l} \langle {\hat{R}}^\varepsilon (X_i, X_j) X_k, X_l \rangle _g a_i^* a_j^* a_l a_k\end{aligned}$$

where \(\{a_i,a^*_i\}\) form the Fermion calculus for \({\mathcal {H}}\).

Recalling equation (A.1) in the appendix, we see that the supertrace will vanish for any term that is not of full degree; from our expressions for \({\mathcal {F}}_I\), it is thus clear that for \(k < \frac{n}{2}+m\)

$$\begin{aligned}\mathbf {Str} \left[ \left( \sum _{I, d(I) \le n+2m} \Lambda _I(B)_t {\mathcal {F}}_I \right) ^k(x_0)\right] = 0.\end{aligned}$$

Let us assume that n is even. Applying the scaling property of Brownian motion, when \(t \rightarrow 0\) the term \(k = \frac{n}{2} + m\) will be dominant. More precisely,

$$\begin{aligned}&\textstyle {\mathbb {E}}\textstyle \left( \left. \mathbf {Str} \left[ {\text {exp}}\left( \sum _{I,d(I) \le n+2m} \Lambda _I(B)_t {\mathcal {F}}_I\right) (x_0)\right] \right| B_1=0 \right) \nonumber \\&\quad \textstyle = \frac{1}{\left( \frac{n}{2}+m\right) !}{\mathbb {E}}\left( \left. \mathbf {Str}\left[ \left( \sum _{I, d(I) \le n+2m} \Lambda _I(B)_t {\mathcal {F}}_I \right) ^{\frac{n}{2}+m}(x_0)\right] \right| B_1=0 \right) + O\left( t^{\frac{n}{2}+m+\frac{1}{2}}\right) . \end{aligned}$$
(4.3)

Then, we have,

$$\begin{aligned}&\scriptstyle {\mathbb {E}} \scriptstyle \left( \left. \mathbf {Str}\left[ \left( \sum _{I, d(I) \le n+2m} \Lambda _I(B)_t {\mathcal {F}}_I \right) ^{\frac{n}{2}+m}(x_0)\right] \right| B_1=0 \right) \nonumber \\&\quad \scriptstyle = {\mathbb {E}}\left( \left. \mathbf {Str}\left[ \left( - t {\mathscr {R}}_\varepsilon (x_0)+ \sum _{1 \le i<j \le n} \sum _{r,s=1}^s \hat{R}_{iir}^{\varepsilon ,s} b_r^* b_s \int _0^t B^i_u dB^j_u -B^j_u dB^i_u \right) ^{\frac{n}{2}+m} \right] \right| B_1=0 \right) + O\left( t^{\frac{n}{2}+m+\frac{1}{2}}\right) . \end{aligned}$$
(4.4)

We can further simplify this expression using that by Lemma A.2, Appendix, we know that \({\hat{R}}_{ijr}^{\varepsilon ,s} = R_{ijr}^s = T_{ij;r}^s\). We also use (2.11) and the fact that only the last term in \({\mathscr {R}}_\varepsilon \) contributes to the supertrace. Combining Lemma 4.1, Corollary 4.3, and Eqs. (4.3) and (4.4), we apply the scaling property of Brownian motion again to find

$$\begin{aligned}&\mathbf {Str} (p_{{\mathcal {H}},\varepsilon } (t,x_0,x_0) ) =\frac{{\mathcal {J}}}{\left( \frac{n}{2}+m\right) ! }{\mathbb {E}} \left( \mathbf {Str} \left[ \left. A_{x_0}^{\frac{n}{2}+m} \right] \right| \right. \\&\left. B_1 = 0 \right) + O\left( t^{\frac{1}{2}}\right) .\end{aligned}$$

If n is odd, we get by similar arguments that

$$\begin{aligned}\mathbf {Str} (p_{{\mathcal {H}},\varepsilon } (t,x_0,x_0) ) = O\left( t^{\frac{1}{2}}\right) .\end{aligned}$$

completing the proof. \(\square \)

In what follows, we will introduce the tensor \({\mathscr {T}}\) by

$$\begin{aligned} {\mathscr {T}}(Y_1, Y_2) = {\hat{R}}^\varepsilon (\pi _{{\mathcal {H}}} Y_1, Y_2) \pi _{{\mathcal {V}}} = \pi _{{\mathcal {V}}} {\hat{R}}^\varepsilon (\pi _{{\mathcal {H}}} Y_1, Y_2). \end{aligned}$$

We observe that for any \(X_1, X_2 \in \Gamma ({\mathcal {H}})\) and \(Z \in {\mathcal {V}}\),

$$\begin{aligned} {\mathscr {T}}(X_1, X_2) Z= (\nabla _Z T)(X_1, X_2) = \frac{1}{2\varepsilon } \left( T(J_Z X_1, X_2) + T(X_1, J_Z X_2) \right) , \end{aligned}$$

where the latter equality follows from the symmetry condition of \(\Delta _{{\mathcal {H}},\varepsilon }\).

Example 4.5

(H-type foliation) We again consider the case of the of H-type foliations as in Example 2.4. We recall that in this case, we have that \(\Delta _{{\mathcal {H}},\varepsilon }\) for \(\varepsilon = \frac{1}{\kappa }\). Let \(x \in {\mathbb {M}}\) be a fixed point and let \(\mathbf {Cl}({\mathcal {V}}_x)\) be the Clifford algebra of the vertical space. We remark that in this case, for any \(u,v \in {\mathcal {H}}_x\) with \(v \in ({{\,\mathrm{span}\,}}_{\zeta \in \mathbf {Cl}({\mathcal {V}}_x)} J_{\zeta } u)^\perp \), we have \({\mathscr {T}}(u,v) = 0\). On the other hand, if \(v = J_\zeta u\), then for any \(z \in {\mathcal {V}}_x\),

$$\begin{aligned} {\mathscr {T}}(u, J_\zeta u)z = \kappa \pi _{{\mathcal {V}}_x} (z \cdot \zeta ^{\mathrm {odd}}), \end{aligned}$$

where \(\zeta ^{\mathrm {odd}}\) is the odd part of \(\zeta \) and \(\pi _{{\mathcal {V}}_x} \mathrm {Cl}({\mathcal {V}}_x) \rightarrow {\mathcal {V}}_x\) is the projection to the first-order part.

We can use the above definition and the previous lemma to prove the following.

Proposition 4.6

Assume that n or m is odd, then

$$\begin{aligned} \lim _{t \rightarrow 0} \mathbf {Str}\text { } (p_{{\mathcal {H}},\varepsilon } (t,x,x) ) \, dx = 0 \end{aligned}$$

Assume that both n and m are even, then

$$\begin{aligned}\lim _{t \rightarrow 0} \mathbf {Str}\text { } (p_{{\mathcal {H}},\varepsilon } (t,x,x) ) \, dx = \hat{\omega }_{\mathcal {H}}^\varepsilon \wedge \left[ \det \left( \frac{{\mathscr {T}}}{\sinh ({\mathscr {T}})}\right) ^{1/2}\right] _m \end{aligned}$$

where \(\left[ \cdot \right] _m \) denotes the m-form part and \(\hat{\omega }_{\mathcal {H}}^\varepsilon \) is the horizontal Euler form, locally defined as

$$\begin{aligned} \hat{\omega }_{\mathcal {H}}^\varepsilon =\frac{(-1)^{n/2} m!}{ 2^{n/2} \left( \frac{n}{2}+m \right) !} {\mathcal {J}} \sum _{\sigma ,\tau \in {\mathfrak {S}}_n} \epsilon (\sigma ) \epsilon (\tau ) \prod _{i=1}^{n-1} \hat{R}^{\varepsilon , \tau (i+1)}_{\sigma (i) \sigma (i+1) \tau (i) }dx_{\mathcal {H}}, \end{aligned}$$

In the above formula, \({\mathfrak {S}}_n\) is the set of the permutations of the indices \(\{1,...,n\}\), \(\epsilon \) the signature of a permutation, \(\hat{R}^{\varepsilon ,l}_{ijk}\) is as in (2.7) and \(dx_{\mathcal {H}}\) the n-form \(X_1^* \wedge \cdots \wedge X_n^*\).

Proof

We first assume that both n and m are even. It remains to compute \({\mathbb {E}} \left( \mathbf {Str} \left[ \left. A_{x_0}^{\frac{n}{2}+m} \right] \right| B_1 = 0 \right) \). Looking at (4.2), we have

$$\begin{aligned}&{\scriptstyle {\mathbb {E}} \left( \mathbf {Str} \left[ \left. A_{x_0}^{\frac{n}{2}+m} \right] \right| B_1 = 0 \right) } \\&\quad ={\scriptstyle \mathbf {Str} \left[ \left( -\sum _{i,j,k,l} \langle {\hat{R}}^\varepsilon (X_i, X_j) X_k, X_l \rangle _g a_i^* a_j^* a_l a_k\right) ^{n/2} {\mathbb {E}}\left[ \left. \left( \sum _{1 \le i<j \le n} {\mathscr {T}}(X_i, X_j)(x_0) \int _0^1 B^i_s dB^j_s -B^j_s dB^i_s \right) ^{m} \right| B_1=0 \right] \right] } \end{aligned}$$

The term \(\left( \sum _{i,j,k,l} \langle {\hat{R}}^\varepsilon (X_i, X_j) X_k, X_l \rangle _g a_i^* a_j^* a_l a_k\right) ^{n/2}\) is then analyzed as in the proof of Proposition 5.6 in [7] (see also Lemma 2.35 in [19]) and up to constant yields the horizontal Euler form \(\hat{\omega }_{\mathcal {H}}^\varepsilon \). On the other hand, using again the formula for the supertrace, the term

$$\begin{aligned} {\mathbb {E}}\left[ \left. \left( \sum _{1 \le i<j \le n} {\mathscr {T}}(X_i, X_j)(x_0) \int _0^1 B^i_s dB^j_s -B^j_s dB^i_s \right) ^{m} \right| B_1=0 \right] \end{aligned}$$

can be replaced with

$$\begin{aligned} m! \, {\mathbb {E}} \left[ \left. \exp \left( \sum _{1 \le i<j \le n} {\mathscr {T}}(X_i, X_j)(x_0) \int _0^1 B^i_s dB^j_s -B^j_s dB^i_s \right) \right| B_1=0\right] \end{aligned}$$

and is analyzed using the Lévy area formula as in the proof of Theorem 4.3 in [3]: it yields the top degree Fermionic piece of \(\det \left( \frac{{\mathscr {T}}}{\sinh ({\mathscr {T}})}\right) ^{1/2} (x_0) \in \mathbf {End} \left( \wedge {\mathcal {V}}_{x_0}^*\right) \) (Fermionic calculus is done here on \({\mathcal {V}}_{x_0}\)).

If n is even and m is odd, a similar analysis shows that

$$\begin{aligned} {\mathbb {E}} \left( \left. \mathbf {Str} \left[ A_{x_0}^{\frac{n}{2}+m} \right] \right| B_1=0\right) =0. \end{aligned}$$

\(\square \)

Combining Theorem 3.1 and Proposition 4.6 finally yields our main theorem:

Theorem 4.7

Assume that both n and m are even, then

$$\begin{aligned} \chi ({\mathbb {M}})= \int _{\mathbb {M}}\hat{\omega }_{\mathcal {H}}^\varepsilon \wedge \left[ \det \left( \frac{{\mathscr {T}}}{\sinh {\mathscr {T}}}\right) ^{1/2}\right] _m . \end{aligned}$$

Assume that n or m is odd, then \(\chi ({\mathbb {M}})=0\).

As a corollary, since \(\nabla J = 0\) implies \({\mathscr {T}}= 0\), we obtain the following result:

Corollary 4.8

Assume that \(\nabla J=0\), then \(\chi ({\mathbb {M}})=0\).