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

Relative heat content asymptotics for sub-Riemannian manifolds

Andrei Agrachev SISSA, Trieste, Italy Luca Rizzi SISSA, Trieste, Italy Univ. Grenoble Alpes, CNRS, Institut Fourier, Grenoble, France Tommaso Rossi SISSA, Trieste, Italy Univ. Grenoble Alpes, CNRS, Institut Fourier, Grenoble, France Institut für Angewandte Mathematik, Universität Bonn, Bonn, Germany
Abstract

The relative heat content associated with a subset ΩM\Omega\subset M of a sub-Riemannian manifold, is defined as the total amount of heat contained in Ω\Omega at time tt, with uniform initial condition on Ω\Omega, allowing the heat to flow outside the domain. In this work, we obtain a fourth-order asymptotic expansion in square root of tt of the relative heat content associated with relatively compact non-characteristic domains. Compared to the classical heat content that we studied in [RR21], several difficulties emerge due to the absence of Dirichlet conditions at the boundary of the domain. To overcome this lack of information, we combine a rough asymptotics for the temperature function at the boundary, coupled with stochastic completeness of the heat semi-group. Our technique applies to any (possibly rank-varying) sub-Riemannian manifold that is globally doubling and satisfies a global weak Poincaré inequality, including in particular sub-Riemannian structures on compact manifolds and Carnot groups.

1 Introduction

In this paper we study the asymptotics of the relative heat content in sub-Riemannian geometry. The latter is a vast generalization of Riemannian geometry, indeed a sub-Riemannian manifold MM is a smooth manifold where a metric is defined only on a subset of preferred directions 𝒟xTxM\mathcal{D}_{x}\subset T_{x}M at each point xMx\in M (called horizontal directions). For example, 𝒟\mathcal{D} can be a sub-bundle of the tangent bundle, but we will consider the most general case of rank-varying distributions. Moreover, we assume that 𝒟\mathcal{D} satisfies the so-called Hörmander condition, which ensures that MM is horizontally-path connected, and that the usual length-minimization procedure yields a well-defined metric.

Let MM be a sub-Riemannian manifold, equipped with a smooth measure ω\omega, let ΩM\Omega\subset M be an open relatively compact subset of MM, with smooth boundary, and consider the Cauchy problem for the heat equation in this setting:

(tΔ)u(t,x)\displaystyle\left(\partial_{t}-\Delta\right)u(t,x) =0,\displaystyle=0, (t,x)(0,)×M,\displaystyle\forall(t,x)\in(0,\infty)\times M, (1)
u(0,)\displaystyle u(0,\cdot) =𝟙Ω,\displaystyle=\mathds{1}_{\Omega}, in L2(M,ω),\displaystyle\text{in }L^{2}(M,\omega),

where 𝟙Ω\mathds{1}_{\Omega} is the indicator function of the set Ω\Omega, and Δ\Delta is the sub-Laplacian, defined with respect to ω\omega. By classical spectral theory, there exists a unique solution to (1),

u(t,x)=etΔ𝟙Ω(x),xM,t>0,u(t,x)=e^{t\Delta}\mathds{1}_{\Omega}(x),\qquad\forall\,x\in M,\ t>0, (2)

where etΔe^{t\Delta} denotes the heat semi-group in L2(M,ω)L^{2}(M,\omega), associated with Δ\Delta. The relative heat content is the function

HΩ(t)=Ωu(t,x)𝑑ω(x),t>0.H_{\Omega}(t)=\int_{\Omega}u(t,x)d\omega(x),\qquad\forall\,t>0. (3)

This quantity has been studied in connection with geometric properties of subsets of n\mathbb{R}^{n}, starting from the seminal work of De Giorgi [DG54], where he introduced the notion of perimeter of a set in n\mathbb{R}^{n} and proved a characterization of sets of finite perimeter in terms of the heat kernel. His result was subsequently refined, using techniques of functions of bounded variation: it was proven in [Led94] for balls in n\mathbb{R}^{n}, and in [MPPP07] for general subsets of n\mathbb{R}^{n}, that a borel set Ωn\Omega\subset\mathbb{R}^{n} with finite Lebesgue measure has finite perimeter à la De Giorgi if and only if

limt0πt(|Ω|HΩ(t))=P(Ω),\exists\lim_{t\to 0}\frac{\sqrt{\pi}}{\sqrt{t}}\big{(}|\Omega|-H_{\Omega}(t)\big{)}=P(\Omega), (4)

where |||\cdot| is the Lebesgue measure and PP is the perimeter measure in n\mathbb{R}^{n}. Notice that (4) is equivalent to a first-order111Here and throughout the paper, the notion of order is computed with respect to t\sqrt{t}. asymptotic expansion of HΩ(t)H_{\Omega}(t). A further development in this direction was then obtained in [AMM13], where the authors extended (4) to an asymptotic expansion of order 33 in t\sqrt{t}, assuming the boundary of Ωn\Omega\subset\mathbb{R}^{n} to be a C1,1C^{1,1} set. For simplicity, we state here the result of [AMM13, Thm. 1.1] assuming Ω\partial\Omega is smooth222The statement of Theorem 1.1 in [AMM13] differs from (5) by a sign in the third-order coefficient: the correct sign appears a few lines below the statement, in the expansion of the function Kt(E,Ec)K_{t}(E,E^{c}). :

HΩ(t)=|Ω|1πP(Ω)t1/2+(n1)212πΩ(HΩ2(x)+2(n1)2cΩ(x))𝑑n1(x)t3/2+o(t3/2),H_{\Omega}(t)=|\Omega|-\frac{1}{\sqrt{\pi}}P(\Omega)t^{1/2}\\ +\frac{(n-1)^{2}}{12\sqrt{\pi}}\int_{\partial\Omega}\left(H_{\partial\Omega}^{2}(x)+\frac{2}{(n-1)^{2}}c_{\partial\Omega}(x)\right)d\mathcal{H}^{n-1}(x)t^{3/2}+o(t^{3/2}), (5)

as t0t\to 0, where n1\mathcal{H}^{n-1} is the Hausdorff measure and, denoting by kiΩ(x)k_{i}^{\partial\Omega}(x) the principal curvatures of Ω\partial\Omega at the point xx,

HΩ(x)=1n1i=1n1kiΩ(x),cΩ(x)=i=1n1kiΩ(x)2,H_{\partial\Omega}(x)=\frac{1}{n-1}\sum_{i=1}^{n-1}k_{i}^{\partial\Omega}(x),\qquad c_{\partial\Omega}(x)=\sum_{i=1}^{n-1}k_{i}^{\partial\Omega}(x)^{2}, (6)

In the Riemannian setting, Van den Berg and Gilkey in [vdBG15] proved the existence of a complete asympotic expansion for HΩ(t)H_{\Omega}(t), generalizing (5), when Ω\partial\Omega is smooth. Moreover, they were able to compute explicitly the coefficients of the expansion up to order 44 in t\sqrt{t}. Their techniques are based on pseudo-differential calculus, and cannot be immediately adapted to the sub-Riemannian setting. In particular, what is missing is a global parametrix estimate for the heat kernel pt(x,y)p_{t}(x,y), cf. [vdBG15, Sec. 2.3]: for any kk\in\mathbb{N}, there exist Jk,Ck>0J_{k},C_{k}>0 such that

pt(x,y)j=0Jkptj(x,y)Ck(M×M)Cktk,as t0,\bigg{\|}p_{t}(x,y)-\sum_{j=0}^{J_{k}}p_{t}^{j}(x,y)\bigg{\|}_{C^{k}(M\times M)}\leq C_{k}t^{k},\qquad\text{as }t\to 0, (7)

where ptj(x,y)p_{t}^{j}(x,y) are suitable smooth functions, given explicitly in terms of the Euclidean heat kernel and iterated convolutions. The closest estimate analogue to (7) in the sub-Riemannian setting is the one proved recently in [CdVHT21, Thm. A] (see Theorem 2.9 for the precise statement), where the authors show an asymptotic expansion of the heat kernel in an asymptotic neighborhood of the diagonal, which is not enough to reproduce (7) and thus the argument of Van den Berg and Gilkey. Moreover, in this case, ptj(x,y)p_{t}^{j}(x,y) is expressed in terms of the heat kernel of the nilpotent approximation and iterated convolutions, thus posing technical difficulties for the explicit computations of the coefficients (which would be no longer “simple” gaussian-type integrals).

In this paper, under the assumption of not having characteristic points, we prove the existence of the asymptotic expansion of HΩ(t)H_{\Omega}(t), up to order 44 in t\sqrt{t}, as t0t\to 0. We remark that we include also the rank-varying case. In order to state our main results, let us introduce the following operator, acting on smooth functions compactly supported close to Ω\partial\Omega,

Nϕ=2g(ϕ,δ)+ϕΔδ,N\phi=2g(\nabla\phi,\nabla\delta)+\phi\Delta\delta, (8)

where δ:M\delta\colon M\rightarrow\mathbb{R} denotes the sub-Riemannian signed distance function from Ω\partial\Omega, see Section 4 for precise definitions.

Theorem 1.1.

Let MM be a compact sub-Riemannian manifold, equipped with a smooth measure ω\omega, and let ΩM\Omega\subset M be an open subset whose boundary is smooth and has no characteristic points. Then, as t0t\to 0,

HΩ(t)=ω(Ω)1πσ(Ω)t1/2112πΩ(N(Δδ)2(Δδ)2)𝑑σt3/2+o(t2),H_{\Omega}(t)=\omega(\Omega)-\frac{1}{\sqrt{\pi}}\sigma(\partial\Omega)t^{1/2}-\frac{1}{12\sqrt{\pi}}\int_{\partial\Omega}\left(N(\Delta\delta)-2(\Delta\delta)^{2}\right)d\sigma\,t^{3/2}+o(t^{2}), (9)

where σ\sigma denotes the sub-Riemannian perimeter measure.

Remark 1.2.

The compactness assumption in Theorem 1.1 is technical and can be relaxed by requiring, instead, global doubling of the measure and a global Poincaré inequality, see section 7 and in particular Theorem 7.3. Some notable examples satisfying these assumptions are:

  • MM is a Lie group with polynomial volume growth, the distribution is generated by a family of left-invariant vector fields satisfying the Hörmander condition and ω\omega is the Haar measure. This family includes also Carnot groups.

  • M=nM=\mathbb{R}^{n}, equipped with a sub-Riemannian structure induced by a family of vector fields {Y1,,YN}\{Y_{1},\ldots,Y_{N}\} with bounded coefficients together with their derivatives, and satisfying the Hörmander condition.

  • MM is a complete Riemannian manifold, equipped with the Riemannian measure, and with non-negative Ricci curvature.

See Section 7.1 for further details. In all these examples, Theorem 1.1 holds.

The strategy of the proof of Theorem 1.1 follows a similar strategy of [RR21], inspired by the method introduced in [Sav98], used for the classical heat content (11). However, as we are going to explain in Section 1.1, new technical difficulties arise, the main one being related to the fact that now u(t,)|Ω0u(t,\cdot)\rvert_{\partial\Omega}\neq 0. At order zero, we obtain the following result, see Section 2 for precise definitions.

Theorem 1.3.

Let MM be a sub-Riemannian manifold, equipped with a smooth measure ω\omega and let ΩM\Omega\subset M be an open relatively compact subset, whose boundary is smooth and has no characteristic points. Let xΩx\in\partial\Omega and consider a chart of privileged coordinates ψ:UVn\psi\colon U\rightarrow V\subset\mathbb{R}^{n} centered at xx, such that ψ(UΩ)=V{z1>0}\psi(U\cap\Omega)=V\cap\{z_{1}>0\}. Then,

limt0u(t,x)={z1>0}p^1x(0,z)𝑑ω^x(z)=12,xΩ,\lim_{t\to 0}u(t,x)=\int_{\{z_{1}>0\}}\hat{p}^{x}_{1}(0,z)d\hat{\omega}^{x}(z)=\frac{1}{2},\qquad\forall\,x\in\partial\Omega, (10)

where ω^x\hat{\omega}^{x} denotes the nilpotentization of ω\omega at xx and p^tx\hat{p}_{t}^{x} denotes the heat kernel associated with the nilpotent approximation of MM at xx and measure ω^x\hat{\omega}^{x}.

This result can be seen as a partial generalization of [CCSGM13, Prop. 3], where the authors proved an asymptotic expansion of u(t,x)u(t,x) up to order 11 in t\sqrt{t} for xΩx\in\partial\Omega, for a special class of non-characteristic domains in Carnot groups.

Remark 1.4.

Our proof of Theorem 1.3 does not yield an asymptotic series for u(t,)|Ωu(t,\cdot)\rvert_{\partial\Omega} at order higher than 0. Indeed a complete asymptotic series of this quantity seems difficult to achieve, cf. Section 6.

Remark 1.5.

When Ω\partial\Omega has no characteristic points, the conormal bundle 𝒜(Ω):={λTMλ,Tπ(λ)Ω=0}\mathcal{A}(\partial\Omega):=\{\lambda\in T^{*}M\mid\langle\lambda,T_{\pi(\lambda)}\partial\Omega\rangle=0\} does not intersect the characteristic set and, as a consequence, the principal symbol of the sub-Laplacian is elliptic near 𝒜(Ω)\mathcal{A}(\partial\Omega). Thus, it is likely that microlocal analysis techniques in the spirit of [CdVHT18] could yield the existence of a complete asymptotic expansion of the relative heat content (but not an explicit expression and geometric interpretation of the coefficients). We thank Yves Colin de Verdière and the anonymous referee for pointing out this fact.

1.1 Strategy of the proof of Theorem 1.1

To better understand the new technical difficulties in the study of the relative heat content HΩ(t)H_{\Omega}(t), let us compare it with the classical heat content QΩ(t)Q_{\Omega}(t) and illustrate the strategy of the proof of Theorem 1.1.

The classical heat content.

We highlight the differences between the relative heat content HΩ(t)H_{\Omega}(t) and the classical one QΩ(t)Q_{\Omega}(t): let ΩM\Omega\subset M an open set in MM, then for all t>0t>0, we have

HΩ(t)=Ωu(t,x)𝑑ω(x),QΩ(t)=Ωu0(t,x)𝑑ω(x),H_{\Omega}(t)=\int_{\Omega}u(t,x)d\omega(x),\qquad Q_{\Omega}(t)=\int_{\Omega}u_{0}(t,x)d\omega(x), (11)

where u(t,x)u(t,x) is the solution to (1) and u0(t,x)u_{0}(t,x) is the solution to the Dirichlet problem for the heat equation, associated with Ω\Omega, i.e.

(tΔ)u0(t,x)\displaystyle(\partial_{t}-\Delta)u_{0}(t,x) =0,\displaystyle=0, (t,x)(0,)×Ω,\displaystyle\forall(t,x)\in(0,\infty)\times\Omega, (12)
u0(t,x)\displaystyle u_{0}(t,x) =0,\displaystyle=0, (t,x)(0,)×Ω,\displaystyle\forall(t,x)\in(0,\infty)\times\partial\Omega,
u0(0,x)\displaystyle u_{0}(0,x) =1,\displaystyle=1, xΩ,\displaystyle\forall x\in\Omega,

The crucial difference is that u0(t,)|Ω=0u_{0}(t,\cdot)\rvert_{\partial\Omega}=0, for any t>0t>0, whereas u(t,)|Ω0u(t,\cdot)\rvert_{\partial\Omega}\neq 0 in general. Thus, there is no a priori relation between HΩ(t)H_{\Omega}(t) and QΩ(t)Q_{\Omega}(t): the only relevant information is given by domain monotonicity, which implies that:

QΩ(t)HΩ(t),t>0,Q_{\Omega}(t)\leq H_{\Omega}(t),\qquad\forall\,t>0, (13)

and clearly this does not give the asymptotics of the latter. See also [vdB13] for other comparison results in the Euclidean setting.

Failure of Duhamel’s principle.

In [RR21], we established a complete asymptotic expansion of QΩ(t)Q_{\Omega}(t), as t0t\to 0, provided that Ω\partial\Omega has no characteristic points. The proof of this result relied on an iterated application of the Duhamel’s principle and the fact that u0(t,x)|Ω=0u_{0}(t,x)\rvert_{\partial\Omega}=0. Following the same strategy, we apply Duhamel’s principle to a localized version of HΩ(t)H_{\Omega}(t): fix a function ϕCc(M)\phi\in C_{c}^{\infty}(M), compactly supported in a tubular neighborhood around Ω\partial\Omega and such that 0ϕ10\leq\phi\leq 1 and ϕ\phi is identically 11, close to Ω\partial\Omega. Then, using off-diagonal estimates for the heat kernel, one can prove that:

ω(Ω)HΩ(t)=Iϕ(t,0)+O(t),as t0,\omega(\Omega)-H_{\Omega}(t)=I\phi(t,0)+O(t^{\infty}),\qquad\text{as }t\to 0, (14)

where Iϕ(t,r)I\phi(t,r) is defined for t>0t>0 and r0r\geq 0 as

Iϕ(t,r)=Ωr(1u(t,x))ϕ(x)𝑑ω(x),I\phi(t,r)=\int_{\Omega_{r}}(1-u(t,x))\phi(x)d\omega(x), (15)

here Ωr={xΩδ(x)>r}\Omega_{r}=\{x\in\Omega\mid\delta(x)>r\}, with δ:Ω\delta\colon\Omega\rightarrow\mathbb{R} denoting the distance function from the boundary. Hence, the small-time behavior of HΩ(t)H_{\Omega}(t) is captured by Iϕ(t,0)I\phi(t,0). By Duhamel’s principle and the sub-Riemannian mean value lemma, cf. Section 4 for details, we obtain the following:

Iϕ(t,0)=1π0tΩ(1u(τ,y))ϕ(y)𝑑σ(y)(tτ)1/2𝑑τ+O(t),as t0.I\phi(t,0)=\frac{1}{\sqrt{\pi}}\int_{0}^{t}\int_{\partial\Omega}\left(1-u(\tau,y)\right)\phi(y)d\sigma(y)(t-\tau)^{-1/2}d\tau+O(t),\qquad\text{as }t\to 0. (16)

For the classical heat content, u0u_{0} satisfies Dirichlet boundary condition, cf. (12), hence (16) would give the first-order asymptotics (and then one could iterate). On the contrary, in this case, we do not have prior knowledge of u(t,y)u(t,y) as yΩy\in\partial\Omega and t0t\to 0. Thus, already for the first-order asymptotics, Duhamel’s principle alone is not enough, and we need some information on the asymptotic behavior of u(t,)|Ωu(t,\cdot)\rvert_{\partial\Omega}.

First-order asymptotics.

We study the asymptotics of u(t,)|Ωu(t,\cdot)\rvert_{\partial\Omega}. Using the notion of nilpotent approximation of a sub-Riemannian manifold, cf. Section 2.3, we deduce the zero-order asymptotic expansion of u(t,)|Ωu(t,\cdot)\rvert_{\partial\Omega} as t0t\to 0, proving Theorem 1.3. This is enough to infer the first-order expansion of HΩ(t)H_{\Omega}(t), by means of (16). At this point, we iterate the Duhamel’s principle to obtain the higher-order terms of the expansion of HΩ(t)H_{\Omega}(t). However, already at the first iteration, we obtain the following formula for IϕI\phi:

Iϕ(t,0)=1π0tΩ(1u(τ,))ϕ𝑑σ(tτ)1/2𝑑τ+12π0t0τΩ(1u(τ^,))Nϕ𝑑σ((ττ^)(tτ))1/2𝑑τ^𝑑τ+O(t3/2),I\phi(t,0)=\frac{1}{\sqrt{\pi}}\int_{0}^{t}\int_{\partial\Omega}(1-u(\tau,\cdot))\phi\hskip 0.59998ptd\sigma(t-\tau)^{-1/2}d\tau\\ +\frac{1}{2\pi}\int_{0}^{t}\int_{0}^{\tau}\int_{\partial\Omega}(1-u(\hat{\tau},\cdot))N\phi\hskip 0.59998ptd\sigma((\tau-\hat{\tau})(t-\tau))^{-1/2}d\hat{\tau}\hskip 0.59998ptd\tau+O(t^{3/2}), (17)

as t0t\to 0. Therefore, the zero-order asymptotic expansion of u(t,)|Ωu(t,\cdot)\rvert_{\partial\Omega} no longer suffices for obtaining the second-order asymptotics of HΩ(t)H_{\Omega}(t).

The outside contribution IcϕI^{c}\phi.

We mentioned that the crucial difference between HΩ(t)H_{\Omega}(t) and QΩ(t)Q_{\Omega}(t), defined in (11), is related to the fact that u(t,)|Ω0u(t,\cdot)\rvert_{\partial\Omega}\neq 0, whereas u0(t,)|Ω=0u_{0}(t,\cdot)\rvert_{\partial\Omega}=0, for any t>0t>0. From a physical viewpoint, this distinction comes from the fact that, since the boundary Ω\partial\Omega is no longer insulated, the heat governed by the Cauchy problem u(t,x)u(t,x), solution to (1), can flow also outside of Ω\Omega, whereas u0(t,x)u_{0}(t,x), solution to the Dirichlet problem (12), is confined in Ω\Omega, and the external temperature is 0. Hence, we can imagine that the asymptotic expansion of HΩ(t)H_{\Omega}(t) is affected by the boundary, both from the inside and from the outside of Ω\Omega.

Interpreting IϕI\phi as the inside contribution to the asymptotics of HΩH_{\Omega}, we are going to formalize the physical intuition of having heat flowing outside of Ω\Omega, defining an outside contribution, IcϕI^{c}\phi to the asymptotics333The notation “superscript cc” stands for complement. Indeed the outside contribution is the inside contribution of the complement of Ω\Omega, see Section 5.1.. The starting observation is the following simple relation: setting

KΩ(t)=MΩu(t,x)𝑑ω(x),t>0,K_{\Omega}(t)=\int_{M\setminus\Omega}u(t,x)d\omega(x),\qquad\forall\,t>0, (18)

we have, by divergence theorem,

HΩ(t)+KΩ(t)=ω(Ω),t>0.H_{\Omega}(t)+K_{\Omega}(t)=\omega(\Omega),\qquad\forall\,t>0. (19)

Similarly to (15), for a suitable smooth function ϕ\phi, one may define a localized version of KΩ(t)K_{\Omega}(t), which we call Icϕ(t,r)I^{c}\phi(t,r), so that

KΩ(t)=Icϕ(t,0)+O(t),as t0,K_{\Omega}(t)=I^{c}\phi(t,0)+O(t^{\infty}),\qquad\text{as }t\to 0, (20)

see Section 5.1 for precise definitions. Using (14), (19) and (20), we show the following relation:

Iϕ(t,0)Icϕ(t,0)=O(t),as t0,I\phi(t,0)-I^{c}\phi(t,0)=O(t^{\infty}),\qquad\text{as }t\to 0, (21)

for a suitable smooth function ϕ\phi. On the other hand, for the localized quantity Iϕ(t,0)Icϕ(t,0)I\phi(t,0)-I^{c}\phi(t,0) we have a Duhamel’s principle, thanks to which we are able to study the asymptotic expansion, up to order 33, of the integral of u(t,x)u(t,x) over Ω\partial\Omega, cf. Theorem 5.4. The limitation to the order 33 of the asymptotics is technical and seems difficult to overcome, cf. Remark 5.5. Inserting this asymptotics in (17), we obtain the asymptotics up to order 33 of the expansion of HΩ(t)H_{\Omega}(t), as t0t\to 0.

Fourth-order asymptotics.

Since we have at disposal only the asymptotics of the integral of u(t,x)u(t,x) over Ω\partial\Omega, up to order 33, we need a finer argument to obtain the fourth-order asymptotics of HΩ(t)H_{\Omega}(t). The simple but compelling relation is based once again on (14), (19) and (20), thanks to which we can write:

ω(Ω)HΩ(t)=12(Iϕ(t,0)+Icϕ(t,0))+O(t),as t0.\omega(\Omega)-H_{\Omega}(t)=\frac{1}{2}\left(I\phi(t,0)+I^{c}\phi(t,0)\right)+O(t^{\infty}),\qquad\text{as }t\to 0. (22)

Now for the sum of the contributions Iϕ(t,0)+Icϕ(t,0)I\phi(t,0)+I^{c}\phi(t,0), the Duhamel’s principle implies the following:

Iϕ(t,0)+Icϕ(t,0)=2πσ(Ω)t1/2+12π0t0τΩ(12u(τ^,x))Nϕ(y)𝑑σ(y)((ττ^)(tτ))1/2𝑑τ^𝑑τ+o(t).I\phi(t,0)+I^{c}\phi(t,0)=\frac{2}{\sqrt{\pi}}\sigma(\partial\Omega)t^{1/2}\\ +\frac{1}{2\pi}\int_{0}^{t}\int_{0}^{\tau}\int_{\partial\Omega}(1-2u(\hat{\tau},x))N\phi(y)d\sigma(y)\left((\tau-\hat{\tau})(t-\tau)\right)^{-1/2}d\hat{\tau}\hskip 0.59998ptd\tau+o(t). (23)

This time notice how the integral of u(t,x)u(t,x) over Ω\partial\Omega appears in a first-order term (as opposed to what happened in (16) or (17)), thus its asymptotic expansion up to order 33 implies a fourth-order expansion for HΩ(t)H_{\Omega}(t), concluding the proof of Theorem 1.1.

1.2 From the heat kernel asymptotics to the relative heat content asymptotics

In [CdVHT21, Thm. A], the authors proved the existence of small-time asymptotics of the hypoelliptic heat kernel, pt(x,y)p_{t}(x,y), see Theorem 2.9 below for the precise statement. In Theorem 1.3 we are able to exploit this result to obtain the zero-order asymptotics of the function

u(t,x)=etΔ𝟙Ω(x)=Ωpt(x,y)𝑑ω(y),t>0,xΩ.u(t,x)=e^{t\Delta}\mathds{1}_{\Omega}(x)=\int_{\Omega}p_{t}(x,y)d\omega(y),\qquad\forall\,t>0,\quad x\in\partial\Omega. (24)

However, we are not able to extend Theorem 1.3 to higher-order asymptotics since, roughly speaking, the remainder terms in Theorem 2.9 are not uniform as t0t\to 0. If we had a better control on the remainders, we could indeed integrate (in a suitable way) the small-time heat kernel asymptotics to obtain the corresponding expansion for u(t,x)u(t,x). Finally, from such an expansion, the relative heat content asymptotics would follow from the localization principle (14) and the (iterated) Duhamel’s principle (16). This is done in Section 6.

1.3 Characteristic points

In order to prove our main results, we need the non-characteristic assumption on the domain Ω\Omega. We recall that for a subset ΩM\Omega\subset M with smooth boundary, xΩx\in\partial\Omega is a characteristic point if 𝒟xTx(Ω)\mathcal{D}_{x}\subset T_{x}(\partial\Omega). As was the case for the classical heat content, cf. [RR21], the non-characteristic assumption is crucial to follow our strategy, since it guarantees the smoothness of the signed distance function close to Ω\partial\Omega, cf. Theorem 4.1. Nevertheless, one might ask whether Theorem 1.1 holds for domains with characteristic points, at least formally.

On the one hand, the coefficients, up to order 22, are well-defined even in presence of characteristic points, cf. [Bal03]. While, for what concerns the integrand of the third-order coefficient, its integrability, with respect to the sub-Riemannian induced measure σ\sigma, is related to integrability of the sub-Riemannian mean curvature \mathcal{H}, with respect to the Riemannian induced measure. The latter is a non-trivial property, which has been studied in [DGN12], and holds in the Heisenberg group, for surfaces with mildly-degenerate characteristic points in the sense of [Ros23].

On the other hand, differently from what happens in the case of the Dirichlet problem, the heat kernel pt(x,y)p_{t}(x,y) associated with (1) is smooth at the boundary of Ω\Omega, for positive times, even in presence of characteristic points. Thus, in principle, there is no obstacle in obtaining an asymptotic expansion of HΩ(t)H_{\Omega}(t) also in that case. Moreover, in Carnot groups of step 22, a similar result to (4) holds, cf. [BMP12, GT20]. In particular, the characterization of sets of finite horizontal perimeter in Carnot groups of step 22 is independent of the presence of characteristic points, indicating that an asymptotic expansion such as (9) may still hold, dropping the non-characteristic assumption.

1.4   Notation

Throughout the article, for a set UMU\subset M, we will use the notation Cc(U)C_{c}^{\infty}(U), even in the compact case, so that all the statements need not be modified in the non-compact case, when the generalization is possible, cf. Theorem 7.3. Moreover, in the non-compact and complete case, the set ΩM\Omega\subset M is assumed to be open and bounded.

Acknowledgments.

This work was supported by the Grant ANR-18-CE40-0012 of the ANR, by the Project VINCI 2019 ref. c2-1212 and by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program (grant agreement No. 945655).
We thank Yves Colin de Verdière, Luc Hillairet and Emmanuel Trélat for stimulating discussion regarding the small-time asymptotics of hypoelliptic heat kernels. We also thank the anonymous referee for valuable comments and remarks.

2 Preliminaries

We recall some essential facts in sub-Riemannian geometry, following [ABB20].

2.1 Sub-Riemannian geometry

Let MM be a smooth, connected finite-dimensional manifold. A sub-Riemannian structure on MM is defined by a set of NN global smooth vector fields X1,,XNX_{1},\ldots,X_{N}, called a generating frame. The generating frame defines a distribution of subspaces of the tangent spaces at each point xMx\in M, given by

𝒟x=span{X1(x),,XN(x)}TxM.\mathcal{D}_{x}=\mathrm{span}\{X_{1}(x),\ldots,X_{N}(x)\}\subseteq T_{x}M. (25)

We assume that the distribution satisfies the Hörmander condition, i.e. the Lie algebra of smooth vector fields generated by X1,,XNX_{1},\dots,X_{N}, evaluated at the point xx, coincides with TxMT_{x}M, for all xMx\in M. The generating frame induces a norm on the distribution at xx, namely

vg=inf{i=1Nui2i=1NuiXi(x)=v},v𝒟x,\|v\|_{g}=\inf\left\{\sum_{i=1}^{N}u_{i}^{2}\mid\sum_{i=1}^{N}u_{i}X_{i}(x)=v\right\},\qquad\forall\,v\in\mathcal{D}_{x}, (26)

which, in turn, defines an inner product on 𝒟x\mathcal{D}_{x} by polarization, which we denote by gx(v,v)g_{x}(v,v). Let T>0T>0. We say that γ:[0,T]M\gamma\colon[0,T]\to M is a horizontal curve, if it is absolutely continuous and

γ˙(t)𝒟γ(t),for a.e.t[0,T].\dot{\gamma}(t)\in\mathcal{D}_{\gamma(t)},\qquad\text{for a.e.}\,t\in[0,T]. (27)

This implies that there exists u:[0,T]Nu:[0,T]\to\mathbb{R}^{N}, such that

γ˙(t)=i=1Nui(t)Xi(γ(t)),for a.e.t[0,T].\dot{\gamma}(t)=\sum_{i=1}^{N}u_{i}(t)X_{i}(\gamma(t)),\qquad\text{for a.e.}\,t\in[0,T]. (28)

Moreover, we require that uL2([0,T],N)u\in L^{2}([0,T],\mathbb{R}^{N}). If γ\gamma is a horizontal curve, then the map tγ˙(t)gt\mapsto\|\dot{\gamma}(t)\|_{g} is integrable on [0,T][0,T]. We define the length of a horizontal curve as follows:

(γ)=0Tγ˙(t)g𝑑t.\ell(\gamma)=\int_{0}^{T}\|\dot{\gamma}(t)\|_{g}dt. (29)

The sub-Riemannian distance is defined, for any x,yMx,y\in M, by

dSR(x,y)=inf{(γ)γ horizontal curve between x and y}.d_{\mathrm{SR}}(x,y)=\inf\{\ell(\gamma)\mid\gamma\text{ horizontal curve between $x$ and $y$}\}. (30)

By the Chow-Rashevsky Theorem, the distance dSR:M×Md_{\mathrm{SR}}\colon M\times M\to\mathbb{R} is finite and continuous. Furthermore it induces the same topology as the manifold one.

Remark 2.1.

The above definition includes all classical constant-rank sub-Riemannian structures as in [Mon02, Rif14] (where 𝒟\mathcal{D} is a vector distribution and gg a symmetric and positive tensor on 𝒟\mathcal{D}), but also general rank-varying sub-Riemannian structures. Moreover, the same sub-Riemannian structure can arise from different generating families.

2.2 The relative heat content

Let MM be a sub-Riemannian manifold. Let ω\omega be a smooth measure on MM, i.e. by a positive tensor density. The divergence of a smooth vector field is defined by

divω(X)ω=Xω,XΓ(TM),\mathrm{div}_{\omega}(X)\omega=\mathcal{L}_{X}\omega,\qquad\forall\,X\in\Gamma(TM), (31)

where X\mathcal{L}_{X} denotes the Lie derivative in the direction of XX. The horizontal gradient of a function fC(M)f\in C^{\infty}(M), denoted by f\nabla f, is defined as the horizontal vector field (i.e. tangent to the distribution at each point), such that

gx(f(x),v)=v(f)(x),v𝒟x,g_{x}(\nabla f(x),v)=v(f)(x),\qquad\forall\,v\in\mathcal{D}_{x}, (32)

where vv acts as a derivation on ff. In terms of a generating frame as in (25), one has

f=i=1NXi(f)Xi,fC(M).\nabla f=\sum_{i=1}^{N}X_{i}(f)X_{i},\qquad\forall\,f\in C^{\infty}(M). (33)

We recall the divergence theorem (we stress that MM is not required to be orientable): let ΩM\Omega\subset M be open with smooth boundary, then

Ω(fdivωX+g(f,X))𝑑ω=Ωfg(X,ν)𝑑σ,\int_{\Omega}\left(f\mathrm{div}_{\omega}X+g(\nabla f,X)\right)d\omega=-\int_{\partial\Omega}fg(X,\nu)d\sigma, (34)

for any smooth function ff and vector field XX, such that the vector field fXfX is compactly supported. In (34), ν\nu is the inward-pointing normal vector field to Ω\Omega and σ\sigma is the induced sub-Riemannian measure on Ω\partial\Omega (i.e. the one whose density is σ=|iνω|Ω\sigma=|i_{\nu}\omega|_{\partial\Omega}).

The sub-Laplacian is the operator Δ=divω\Delta=\mathrm{div}_{\omega}\circ\nabla, acting on C(M)C^{\infty}(M). Again, we may write its expression with respect to a generating frame (25), obtaining

Δf=i=1N{Xi2(f)+Xi(f)divω(Xi)},fC(M).\Delta f=\sum_{i=1}^{N}\left\{X^{2}_{i}(f)+X_{i}(f)\mathrm{div}_{\omega}(X_{i})\right\},\qquad\forall\,f\in C^{\infty}(M). (35)

We denote by L2(M,ω)L^{2}(M,\omega), or simply by L2L^{2}, the space of real functions on MM which are square-integrable with respect to the measure ω\omega. Let ΩM\Omega\subset M be an open relatively compact set with smooth boundary. This means that the closure Ω¯\bar{\Omega} is a compact manifold with smooth boundary. We consider the Cauchy problem for the heat equation on Ω\Omega, that is we look for functions uu such that

(tΔ)u(t,x)\displaystyle\left(\partial_{t}-\Delta\right)u(t,x) =0,\displaystyle=0, (t,x)(0,)×M,\displaystyle\forall(t,x)\in(0,\infty)\times M, (36)
u(0,)\displaystyle u(0,\cdot) =𝟙Ω,\displaystyle=\mathds{1}_{\Omega}, in L2(M,ω),\displaystyle\text{in }L^{2}(M,\omega),

where u(0,)u(0,\cdot) is a shorthand notation for the L2L^{2}-limit of u(t,x)u(t,x) as t0t\to 0. Notice that Δ\Delta is symmetric with respect to the L2L^{2}-scalar product and negative, moreover, if (M,dSR)(M,d_{\mathrm{SR}}) is complete as a metric space, it is essentially self-adjoint, see [Str86]. Thus, there exists a unique solution to (36), and it can be represented as

u(t,x)=etΔ𝟙Ω(x),xM,t>0,u(t,x)=e^{t\Delta}\mathds{1}_{\Omega}(x),\qquad\forall\,x\in M,\ t>0, (37)

where etΔ:L2L2e^{t\Delta}\colon L^{2}\rightarrow L^{2} denotes the heat semi-group, associated with Δ\Delta. We remark that for all φL2\varphi\in L^{2}, the function etΔφe^{t\Delta}\varphi is smooth for all (t,x)(0,)×M(t,x)\in(0,\infty)\times M, by hypoellipticity of the heat operator and there exists a heat kernel associated with (36), i.e. a positive function pt(x,y)C((0,+)×M×M)p_{t}(x,y)\in C^{\infty}((0,+\infty)\times M\times M) such that:

u(t,x)=Mpt(x,y)𝟙Ω(y)𝑑ω(y)=Ωpt(x,y)𝑑ω(y).u(t,x)=\int_{M}p_{t}(x,y)\mathds{1}_{\Omega}(y)d\omega(y)=\int_{\Omega}p_{t}(x,y)d\omega(y). (38)
Definition 2.2 (Relative heat content).

Let u(t,x)u(t,x) be the solution to (36). We define the relative heat content, associated with Ω\Omega, as

HΩ(t)=Ωu(t,x)𝑑ω(x),t>0.H_{\Omega}(t)=\int_{\Omega}u(t,x)d\omega(x),\qquad\forall\,t>0. (39)
Remark 2.3.

If we consider, instead of Ω\Omega, a set which is the closure of an open set, then the Cauchy problem (36) has a unique solution and relative heat content is still well-defined.

We recall here a property of the solution to (36): it satisfies a weak maximum principle, meaning that

0u(t,x)1,xΩ,t>0.0\leq u(t,x)\leq 1,\qquad\forall\,x\in\Omega,\ \forall\,t>0. (40)

This can be proven following the blueprint of the Riemannian proof (see [Gri09, Thm. 5.11]).

Definition 2.4 (Characteristic point).

We say that xΩx\in\partial\Omega is a characteristic point, or tangency point, if the distribution is tangent to Ω\partial\Omega at xx, that is

𝒟xTx(Ω).\mathcal{D}_{x}\subseteq T_{x}(\partial\Omega). (41)

We will assume that Ω\partial\Omega has no characteristic points. We say in this case that Ω\Omega is a non-characteristic domain.

2.3 Nilpotent approximation of MM

We introduce the notion of nilpotent approximation of a sub-Riemannian manifold, see [Jea14, Bel96] for details. This will be used only in Sections 3 and 6.

Sub-Riemannian flag.

Let MM be an nn-dimensional sub-Riemannian manifold with distribution 𝒟\mathcal{D}. We define the flag of 𝒟\mathcal{D} as the sequence of subsheafs 𝒟kTM\mathcal{D}^{k}\subset TM such that

𝒟1=𝒟,𝒟k+1=𝒟k+[𝒟,𝒟k],k1,\mathcal{D}^{1}=\mathcal{D},\qquad\mathcal{D}^{k+1}=\mathcal{D}^{k}+[\mathcal{D},\mathcal{D}^{k}],\qquad\forall\,k\geq 1, (42)

with the convention that 𝒟0={0}\mathcal{D}^{0}=\{0\}. Under the Hörmander condition, the flag of the distribution defines an exhaustion of TxMT_{x}M, for any point xMx\in M, i.e. there exists r(x)r(x)\in\mathbb{N} such that:

{0}=𝒟x0𝒟x1𝒟xr(x)1𝒟xr(x)=TxM.\{0\}=\mathcal{D}^{0}_{x}\subset\mathcal{D}^{1}_{x}\subset\ldots\subset\mathcal{D}^{r(x)-1}_{x}\subsetneq\mathcal{D}^{r(x)}_{x}=T_{x}M. (43)

The number r(x)r(x) is called degree of nonholonomy at xx. We set nk(x)=dim𝒟xkn_{k}(x)=\dim\mathcal{D}^{k}_{x}, for any k0k\geq 0, then the collection of r(x)r(x) integers

(n1(x),,nr(x)(x))\left(n_{1}(x),\ldots,n_{r(x)}(x)\right) (44)

is called growth vector at xx, and we have nr(x)(x)=n=dimMn_{r(x)}(x)=n=\dim M. Associated with the growth vector, we can define the sub-Riemannian weights wi(x)w_{i}(x) at xx, setting for any i{1,,n}i\in\{1,\ldots,n\},

wi(x)=j,if and only ifnj1(x)+1inj(x).w_{i}(x)=j,\qquad\text{if and only if}\qquad n_{j-1}(x)+1\leq i\leq n_{j}(x). (45)

A point xMx\in M is said to be regular if the growth vector is constant in a neighborhood of xx, and singular otherwise. The sub-Riemannian structure on MM is said to be equiregular if all points of MM are regular. In this case, the weights are constant as well on MM. Finally, given any xMx\in M, we define the homogeneous dimension of MM at xx as

𝒬(x)=i=1r(x)i(ni(x)ni1(x))=i=1nwi(x).\mathcal{Q}(x)=\sum_{i=1}^{r(x)}i(n_{i}(x)-n_{i-1}(x))=\sum_{i=1}^{n}w_{i}(x). (46)

We recall that, if xx is regular, then 𝒬(x)\mathcal{Q}(x) coincides with the Hausdorff dimension of (M,dSR)(M,d_{\mathrm{SR}}) at xx, cf. [Mit85]. Moreover, 𝒬(x)>n\mathcal{Q}(x)>n, for any xMx\in M such that 𝒟xTxM\mathcal{D}_{x}\subsetneq T_{x}M.

Privileged coordinates.

Let MM be a sub-Riemannian manifold with generating frame (25) and ff be the germ of a smooth function ff at xMx\in M. We call nonholonomic derivative of order kk\in\mathbb{N} of ff, the quantity

Xj1Xjkf(x),X_{j_{1}}\cdots X_{j_{k}}f(x), (47)

for any family of indexes {j1,,jk}{1,,N}\{j_{1},\ldots,j_{k}\}\subset\{1,\ldots,N\}. Then, the nonholonomic order of ff at the point xx is

ordx(f)=min{k{j1,,jk}{1,,N} s.t. Xj1Xjkf(x)0}.\mathrm{ord}_{x}(f)=\min\left\{k\in\mathbb{N}\mid\exists\{j_{1},\ldots,j_{k}\}\subset\{1,\ldots,N\}\text{ s.t. }X_{j_{1}}\cdots X_{j_{k}}f(x)\neq 0\right\}. (48)
Definition 2.5 (Privileged coordinates).

Let MM be a nn-dimensional sub-Riemannian manifold and xMx\in M. A system of local coordinates (z1,,zn)(z_{1},\ldots,z_{n}) centered at xx is said to be privileged at xx if

ordx(zj)=wj(x),j=1,,n.\mathrm{ord}_{x}(z_{j})=w_{j}(x),\qquad\forall\,j=1,\ldots,n. (49)

Notice that privileged coordinates (z1,,zn)(z_{1},\ldots,z_{n}) at xx satisfy the following property

zi|x𝒟xwi,zi|x𝒟xwi1,i=1,,n.{\partial_{z_{i}}}_{\rvert x}\in\mathcal{D}^{w_{i}}_{x},\qquad{\partial_{z_{i}}}_{\rvert x}\notin\mathcal{D}^{w_{i}-1}_{x},\qquad\forall i=1,\ldots,n. (50)

A local frame of TMTM consisting of nn vector fields {Z1,,Zn}\{Z_{1},\ldots,Z_{n}\} and satisfying (50) is said to be adapted to the flag (43) at xx. Thus, privileged coordinates are always adapted to the flag. In addition, given a local frame adapted to the sub-Riemannian flag at xx, say {Z1,,Zn}\{Z_{1},\ldots,Z_{n}\}, we can define a set of privileged coordinates at xx, starting from {Z1,,Zn}\{Z_{1},\ldots,Z_{n}\}, i.e.

n(z1,,zn)ez1Z1eznZn(x).\mathbb{R}^{n}\ni(z_{1},\ldots,z_{n})\mapsto e^{z_{1}Z_{1}}\circ\cdots\circ e^{z_{n}Z_{n}}(x). (51)

Moreover, in these coordinates, the vector field Z1Z_{1} is exactly z1\partial_{z_{1}}.

Nilpotent approximation.

Let MM be a sub-Riemannian manifold and let xMx\in M with weights as in (45). Consider ψ=(z1,,zn):UV\psi=(z_{1},\ldots,z_{n})\colon U\rightarrow V a chart of privileged coordinates at xx, where UMU\subset M is a relatively compact neighborhood of xx and VnV\subset\mathbb{R}^{n} is a neighborhood of 0. Then, for any ε\varepsilon\in\mathbb{R}, we can define the dilation at xx as

δε:nn;δε(z)=(εw1(x)z1,,εwn(x)zn).\delta_{\varepsilon}\colon\mathbb{R}^{n}\rightarrow\mathbb{R}^{n};\qquad\delta_{\varepsilon}(z)=\left(\varepsilon^{w_{1}(x)}z_{1},\ldots,\varepsilon^{w_{n}(x)}z_{n}\right). (52)

Using such dilations, we obtain the nilpotent (or first-order) approximation of the generating frame (25), indeed setting Yi=ψXiY_{i}=\psi_{*}X_{i}, for any i=1,Ni=1\ldots,N, define

X^ix=limε0εδ1ε(Yi),i=1,N,\widehat{X}_{i}^{x}=\lim_{\varepsilon\to 0}\varepsilon\delta_{\frac{1}{\varepsilon}*}(Y_{i}),\qquad\forall\,i=1\ldots,N, (53)

where the limit is taken in the CC^{\infty}-topology of n\mathbb{R}^{n}. Notice that the vector field X^ix\widehat{X}_{i}^{x} is defined on the whole n\mathbb{R}^{n}, even though YiY_{i} was defined only on VnV\subset\mathbb{R}^{n}.

Theorem 2.6.

Let MM be a nn-dimensional sub-Riemannian manifold with generating frame {X1,,XN}\{X_{1},\ldots,X_{N}\} and consider its first-order approximation at xx as in (53). Then, the frame {X^1x,,X^Nx}\{\widehat{X}_{1}^{x},\ldots,\widehat{X}_{N}^{x}\} of vector fields on n\mathbb{R}^{n} generates a nilpotent Lie algebra of step r(x)=wn(x)r(x)=w_{n}(x) and satisfies the Hörmander condition.

The proof of this theorem can be found in [Jea14]. Recall that a Lie algebra is said to be nilpotent of step ss if ss is the smallest integer such that all the brackets of length greater than ss are zero.

Definition 2.7 (Nilpotent approximation).

Let MM be a sub-Riemannian manifold and let xMx\in M. Then, Theorem 2.6 implies that the frame {X^1x,,X^Nx}\{\widehat{X}_{1}^{x},\ldots,\widehat{X}_{N}^{x}\} is a generating frame for a sub-Riemannian structure on n\mathbb{R}^{n}: we denote the resulting sub-Riemannian manifold M^x\widehat{M}^{x}. This is the so-called nilpotent approximation of MM at the point xx.

Notice that the sub-Riemannian distance of M^x\widehat{M}^{x}, denoted by d^x\hat{d}^{x}, is 11-homogeneous with respect to the dilations (52).

Remark 2.8.

Up to isometries, the nilpotent approximation of MM at xx coincides with the Gromov-Hausdorff metric tangent space of (M,dSR)(M,d_{\mathrm{SR}}) at xx. Moreover, M^x\widehat{M}^{x} is isometric to a quotient of a Carnot group. See [Gro96, Bel96, Mon02] for further details.

Nilpotentized sub-Laplacian.

Let MM be a sub-Riemannian manifold, equipped with a smooth measure ω\omega, and let (z1,,zn)(z_{1},\ldots,z_{n}) be a set of privileged coordinates at xMx\in M. We will use the same symbol ω\omega to denote measure in coordinates. The nilpotentization ω^x\hat{\omega}^{x} of ω\omega at xx is defined as follows:

ω^x,f=limε01|ε|𝒬(x)δεω,f,fCc(n).\langle\hat{\omega}^{x},f\rangle=\lim_{\varepsilon\to 0}\frac{1}{\left|\varepsilon\right|^{\mathcal{Q}(x)}}\langle\delta_{\varepsilon}^{*}\omega,f\rangle,\qquad\forall\,f\in C_{c}^{\infty}(\mathbb{R}^{n}). (54)

Notice that, denoting by dz=dz1dzndz=dz_{1}\ldots dz_{n} the Lebesgue measure on n\mathbb{R}^{n}, we have

δε(dz)=|ε|𝒬(x)dz,ε0,\delta_{\varepsilon}^{*}(dz)=\left|\varepsilon\right|^{\mathcal{Q}(x)}dz,\qquad\forall\,\varepsilon\neq 0, (55)

thus, the limit in (54) exists. Finally, we can define the nilpotentized sub-Laplacian according to (35), acting on C(n)C^{\infty}(\mathbb{R}^{n}), i.e.

Δ^x=divω^x(^x)=i=1N(X^ix)2.\widehat{\Delta}^{x}=\mathrm{div}_{\hat{\omega}^{x}}\left(\widehat{\nabla}^{x}\right)=\sum_{i=1}^{N}(\widehat{X}^{x}_{i})^{2}. (56)

We remark that in (56) there is no divergence term, since

divω^x(X^ix)=0i{1,,N}.\mathrm{div}_{\hat{\omega}^{x}}(\widehat{X}^{x}_{i})=0\qquad\forall i\in\{1,\ldots,N\}. (57)

As in the general sub-Riemannian context, in the nilpotent approximation M^x\widehat{M}^{x}, we may consider the Cauchy heat problem (36) in L2(n,ω^x)L^{2}(\mathbb{R}^{n},\hat{\omega}^{x}). We will the denote the associated heat kernel as

p^tx(z,z)C((0,+)×n×n).\hat{p}_{t}^{x}(z,z^{\prime})\in C^{\infty}((0,+\infty)\times\mathbb{R}^{n}\times\mathbb{R}^{n}). (58)

Heat kernel asymptotics.

Let MM be a sub-Riemannian manifold, equipped with a smooth measure ω\omega and denote by pt(x,y)p_{t}(x,y) the heat kernel (38). We have the following result.

Theorem 2.9 ([CdVHT21, Thm. A]).

Let MM be a sub-Riemannian manifold and let ψ:UV\psi\colon U\rightarrow V be a chart of privileged coordinates at xMx\in M. Then, for any mm\in\mathbb{N},

|ε|𝒬(x)pε2τ(δε(z),δε(z))=p^τx(z,z)+i=1mεifix(τ,z,z)+o(|ε|m),as ε0,|\varepsilon|^{\mathcal{Q}(x)}p_{\varepsilon^{2}\tau}(\delta_{\varepsilon}(z),\delta_{\varepsilon}(z^{\prime}))=\hat{p}_{\tau}^{x}(z,z^{\prime})+\sum_{i=1}^{m}\varepsilon^{i}f_{i}^{x}(\tau,z,z^{\prime})+o(|\varepsilon|^{m}),\qquad\text{as }\varepsilon\to 0, (59)

in the CC^{\infty}-topology of (0,)×V×V(0,\infty)\times V\times V, where fixf_{i}^{x}’s are smooth functions satisfying the following homogeneity property: for i=0,,mi=0,\ldots,m

|ε|𝒬(x)εifix(ε2τ,δε(z),δε(z))=fix(τ,z,z),(τ,z,z)(0,)×n×n,|\varepsilon|^{\mathcal{Q}(x)}\varepsilon^{-i}f_{i}^{x}(\varepsilon^{2}\tau,\delta_{\varepsilon}(z),\delta_{\varepsilon}(z^{\prime}))=f_{i}^{x}(\tau,z,z^{\prime}),\qquad\forall\,(\tau,z,z^{\prime})\in(0,\infty)\times\mathbb{R}^{n}\times\mathbb{R}^{n}, (60)

where, for i=0i=0, we set f0x(τ,z,z)=p^τx(z,z)f_{0}^{x}(\tau,z,z^{\prime})=\hat{p}^{x}_{\tau}(z,z^{\prime}). In (59), we are considering the heat kernel ptp_{t} in coordinates, with a little abuse of notation.

Remark 2.10.

We will drop the dependence on the center of the privileged coordinates if there is no confusion.

3 Small-time asymptotics of u(t,x)u(t,x) at the boundary

We prove here Theorem 1.3, regarding the zero-order asymptotics of u(t,)|Ωu(t,\cdot)\rvert_{\partial\Omega} as t0t\to 0.

Theorem 3.1.

Let MM be a compact sub-Riemannian manifold, equipped with a smooth measure ω\omega and let ΩM\Omega\subset M be an open subset, whose boundary is smooth and has no characteristic points. Let xΩx\in\partial\Omega and consider a chart of privileged coordinates ψ:UVn\psi\colon U\rightarrow V\subset\mathbb{R}^{n} centered at xx, such that ψ(UΩ)=V{z1>0}\psi(U\cap\Omega)=V\cap\{z_{1}>0\}. Then,

limt0u(t,x)={z1>0}p^1x(0,z)𝑑ω^x(z)=12,xΩ,\lim_{t\to 0}u(t,x)=\int_{\{z_{1}>0\}}\hat{p}^{x}_{1}(0,z)d\hat{\omega}^{x}(z)=\frac{1}{2},\qquad\forall\,x\in\partial\Omega, (61)

where ω^x\hat{\omega}^{x} denotes the nilpotentization of ω\omega at xx and p^tx\hat{p}_{t}^{x} denotes the heat kernel associated with the nilpotent approximation of MM at xx and measure ω^x\hat{\omega}^{x}.

Remark 3.2.

A chart of privileged coordinates, such that ψ(UΩ)=V{z1>0}\psi(U\cap\Omega)=V\cap\{z_{1}>0\} always exists, provided that Ω\partial\Omega has no characteristic points. Indeed, in this case, there exists a tubular neighborhood of the boundary, cf. Theorem 4.1, which is built through the flow of δ\nabla\delta, namely

G:(r0,r0)×ΩΩr0r0;G(t,q)=etδ(q),G\colon(-r_{0},r_{0})\times\partial\Omega\rightarrow\Omega_{-r_{0}}^{r_{0}};\qquad G(t,q)=e^{t\nabla\delta}(q), (62)

is a diffeomorphism such that Gt=δG_{*}\partial_{t}=\nabla\delta and δ(G(t,q))=t\delta(G(t,q))=t. Here δ:M\delta\colon M\rightarrow\mathbb{R} is the signed distance function444We warn the reader that δ\delta without a subscript always denotes the signed distance function and should not be confused with dilations δε\delta_{\varepsilon}. from Ω\partial\Omega and Ωr0r0={r0<δ<r0}\Omega_{-r_{0}}^{r_{0}}=\{-r_{0}<\delta<r_{0}\}, see Section 4.1 for precise definitions. Therefore, choosing an adapted frame for the distribution at xx, say {Z1,,Zn}\{Z_{1},\ldots,Z_{n}\} where Z1=δZ_{1}=\nabla\delta, we can define a set of privileged coordinates as in (51):

n(z1,,zn)ez1Z1ez2Z2eznZn(x)φ(z2,,zn)=G(z1,φ(z2,,zn)).\mathbb{R}^{n}\ni(z_{1},\ldots,z_{n})\mapsto e^{z_{1}Z_{1}}\circ\underbrace{e^{z_{2}Z_{2}}\circ\cdots\circ e^{z_{n}Z_{n}}(x)}_{\varphi(z_{2},\ldots,z_{n})}=G(z_{1},\varphi(z_{2},\ldots,z_{n})). (63)

The resulting set of coordinates ψ\psi satisfies ψ(δ)=z1\psi_{*}(\nabla\delta)=\partial_{z_{1}} and, denoting by VV the neighborhood of 0 in n\mathbb{R}^{n} where ψ\psi is invertible, ψ(UΩ)={z1>0}V\psi(U\cap\Omega)=\{z_{1}>0\}\cap V. Here, esX(q)e^{sX}(q) denotes the flow of the vector field XX, starting at qq, evaluated at time ss.

Proof of Theorem 3.1.

Let pt(x,y)p_{t}(x,y) be the heat kernel of MM, then we may write

u(t,x)=Ωpt(x,y)𝑑ω(y),xM.u(t,x)=\int_{\Omega}p_{t}(x,y)d\omega(y),\qquad\forall\,x\in M. (64)

For a fixed xMx\in M, denoting by UU any relatively compact neighborhood of xx, we have

u(t,x)=UΩpt(x,y)𝑑ω(y)+ΩUpt(x,y)𝑑ω(y)=UΩpt(x,y)𝑑ω(y)+O(t),\begin{split}u(t,x)&=\int_{U\cap\Omega}p_{t}(x,y)d\omega(y)+\int_{\Omega\setminus U}p_{t}(x,y)d\omega(y)\\ &=\int_{U\cap\Omega}p_{t}(x,y)d\omega(y)+O(t^{\infty}),\end{split} (65)

as t0t\to 0. Indeed, since the heat kernel is exponentially decaying outside the diagonal, cf. [JSC86, Prop. 3],

ΩUpt(x,y)𝑑ω(y)ω(ΩU)CUecUt=O(t),as t0.\int_{\Omega\setminus U}p_{t}(x,y)d\omega(y)\leq\omega(\Omega\setminus U)C_{U}e^{-\frac{c_{U}}{t}}=O(t^{\infty}),\qquad\text{as }t\to 0. (66)

as t0t\to 0. Now, for xΩx\in\partial\Omega, fix the set of privileged coordinates ψ:UVn\psi\colon U\rightarrow V\subset\mathbb{R}^{n}, defined as in the statement and assume without loss of generality that δε(V)V\delta_{\varepsilon}(V)\subset V for any |ε|1|\varepsilon|\leq 1, where δε\delta_{\varepsilon} is the dilation (52) of the nilpotent approximation of MM. Also set

Vε=δε(V{z1>0}),|ε|1.V_{\varepsilon}=\delta_{\varepsilon}\left(V\cap\{z_{1}>0\}\right),\qquad\forall\,|\varepsilon|\leq 1. (67)

when the limits exist, we have:

limt0u(t,x)=limt0UΩpt(x,y)𝑑ω(y)=limt0V1pt(0,z)𝑑ω(z),\lim_{t\to 0}u(t,x)=\lim_{t\to 0}\int_{U\cap\Omega}p_{t}(x,y)d\omega(y)=\lim_{t\to 0}\int_{V_{1}}p_{t}(0,z)d\omega(z), (68)

where, in the last equation, we are considering the expression of the heat kernel and the measure in coordinates. We want to apply (59) at order 11 in ε\varepsilon, so let us rephrase the statement as follows: for any compact set KVK\subset V,

|ε|𝒬pε2τ(0,δε(z))=p^τ(0,z)+εR(ε,τ,z),as ε0,|\varepsilon|^{\mathcal{Q}}p_{\varepsilon^{2}\tau}(0,\delta_{\varepsilon}(z))=\hat{p}_{\tau}(0,z)+\varepsilon R(\varepsilon,\tau,z),\qquad\text{as }\varepsilon\to 0, (69)

where RR is a smooth function such that

supε[1,1]zK|R(ε,τ,z)|C(τ,K),\sup_{\begin{subarray}{c}\varepsilon\in[-1,1]\\ z\in K\end{subarray}}\left|R(\varepsilon,\tau,z)\right|\leq C(\tau,K), (70)

with C(τ,K)>0C(\tau,K)>0. Notice that (70) is not uniform in τ\tau, in the sense that τC(τ,K)\tau\mapsto C(\tau,K) can explode as τ0\tau\to 0, in general. Moreover, without loss of generality and, up to restrictions of UU, we can assume that (70) holds globally on V¯1\overline{V}_{1}. For a fixed parameter L>1L>1, we set τ=1/L\tau=1/L and ε2=tL\varepsilon^{2}=tL in (69), obtaining:

|tL|𝒬/2pt(0,δtL(z))=p^1/L(0,z)+tLR(tL,1/L,z),as t0,|tL|^{\mathcal{Q}/2}p_{t}(0,\delta_{\sqrt{tL}}(z))=\hat{p}_{1/L}(0,z)+\sqrt{tL}R(\sqrt{tL},1/L,z),\qquad\text{as }t\to 0, (71)

where the remainder RR is bounded as t0t\to 0 on the compact sets of VV, but with a constant depending on LL. Inserting the above expansion in (68), and writing the measure in coordinates dω(z)=ω(z)dzd\omega(z)=\omega(z)dz with ω()C(V1)\omega(\cdot)\in C^{\infty}(V_{1}), we have:

u(t,x)\displaystyle u(t,x) =V1pt(0,z)ω(z)𝑑z+O(t)\displaystyle=\int_{V_{1}}p_{t}(0,z)\omega(z)dz+O(t^{\infty}) (72)
=VtLpt(0,z)ω(z)𝑑z+V1VtLpt(0,z)𝑑ω(z)+O(t)\displaystyle=\int_{V_{\sqrt{tL}}}p_{t}(0,z)\omega(z)dz+\int_{V_{1}\setminus V_{\sqrt{tL}}}p_{t}(0,z)d\omega(z)+O(t^{\infty}) (73)
=V1|tL|𝒬/2pt(0,δtL(z))ω(δtL(z))𝑑z+V1VtLpt(0,z)𝑑ω(z)+O(t)\displaystyle=\int_{V_{1}}|tL|^{\mathcal{Q}/2}p_{t}(0,\delta_{\sqrt{tL}}(z))\omega(\delta_{\sqrt{tL}}(z))dz+\int_{V_{1}\setminus V_{\sqrt{tL}}}p_{t}(0,z)d\omega(z)+O(t^{\infty}) (74)
=V1(p^1/L(0,z)+tLR(tL,1/L,z))ω(δtL(z))𝑑z\displaystyle=\int_{V_{1}}\left(\hat{p}_{1/L}(0,z)+\sqrt{tL}R(\sqrt{tL},1/L,z)\right)\omega(\delta_{\sqrt{tL}}(z))dz (75)
+V1VtLpt(0,z)𝑑ω(z)+O(t),\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\,+\int_{V_{1}\setminus V_{\sqrt{tL}}}p_{t}(0,z)d\omega(z)+O(t^{\infty}), (76)

where in the third equality we performed the change of variable zδ1/tL(z)z\mapsto\delta_{1/\sqrt{tL}}(z) in the first integral. Let us discuss the terms appearing in (75) and (76). First of all, for any L>1L>1, by definition of the nilpotentization of ω\omega given in (54), we get

limt0V1p^1/L(0,z)ω(δtL(z))𝑑z=V1p^1/L(0,z)𝑑ω^(z).\lim_{t\to 0}\int_{V_{1}}\hat{p}_{1/L}(0,z)\omega(\delta_{\sqrt{tL}}(z))dz=\int_{V_{1}}\hat{p}_{1/L}(0,z)d\hat{\omega}(z). (77)

Moreover, for a fixed L>1L>1, the integral of RR is bounded as t0t\to 0, therefore, using (70), we have:

|tLV1R(tL,1/L,z)𝑑ω(z)|CLt,t1,\left|\sqrt{tL}\int_{V_{1}}R(\sqrt{tL},1/L,z)d\omega(z)\right|\leq C_{L}\sqrt{t},\qquad\forall\,t\leq 1, (78)

where CL>0C_{L}>0 is a constant depending on the fixed LL. Secondly, by an upper Gaussian bound for the heat kernel in compact sub-Riemannian manifold, [JSC86, Thm. 2], we obtain the following estimate for (76):

V1VtLpt(0,z)𝑑ω(z)V1VtLC1eβdSR2(0,z)tt𝒬/2𝑑ω(z),\int_{V_{1}\setminus V_{\sqrt{tL}}}p_{t}(0,z)d\omega(z)\leq\int_{V_{1}\setminus V_{\sqrt{tL}}}\frac{C_{1}e^{-\frac{\beta d_{\mathrm{SR}}^{2}(0,z)}{t}}}{t^{\mathcal{Q}/2}}d\omega(z), (79)

where C1,β>0C_{1},\beta>0 are positive constants. Now, by the Ball-Box Theorem [Jea14, Thm. 2.1], the sub-Riemannian distance function at the origin is comparable with the sub-Riemannian distance of M^x\widehat{M}^{x}, denoted by d^\hat{d}. In particular, there exists a constant c>0c>0 such that

dSR2(0,z)cd^2(0,z),zV.d_{\mathrm{SR}}^{2}(0,z)\geq c\,\hat{d}^{2}(0,z),\qquad\forall\,z\in V. (80)

Since in (79) we are integrating over the set V1VtLV_{1}\setminus V_{\sqrt{tL}} and d^\hat{d} is 1-homogeneous with respect to δε\delta_{\varepsilon}, we conclude that

dSR2(0,z)ctL,zV1VtL.d_{\mathrm{SR}}^{2}(0,z)\geq c\,tL,\qquad\forall\,z\in V_{1}\setminus V_{\sqrt{tL}}. (81)

Therefore, using also (80), the term (79) can be estimated as follows:

V1VtLpt(0,z)𝑑ω(z)C1ecβL2V1eβdSR2(0,z)2tt𝒬/2𝑑ω(z)C1ecβL2V1eβcd^2(0,z)2tt𝒬/2ω(z)𝑑zC~ecβL2,\int_{V_{1}\setminus V_{\sqrt{tL}}}p_{t}(0,z)d\omega(z)\leq C_{1}e^{-\frac{c\beta L}{2}}\int_{V_{1}}\frac{e^{-\frac{\beta d_{\mathrm{SR}}^{2}(0,z)}{2t}}}{t^{\mathcal{Q}/2}}d\omega(z)\\ \leq C_{1}e^{-\frac{c\beta L}{2}}\int_{V_{1}}\frac{e^{-\frac{\beta c\,\hat{d}^{2}(0,z)}{2t}}}{t^{\mathcal{Q}/2}}\omega(z)dz\leq\tilde{C}e^{-\frac{c\beta L}{2}}, (82)

where C~>0\tilde{C}>0 is independent of tt and LL. The last inequality in (82) follows from the fact that, after a change of variable zδ1/t(z)z\mapsto\delta_{1/\sqrt{t}}(z), the integral

V1eβcd^2(0,z)2tt𝒬/2ω(z)𝑑z<+,\int_{V_{1}}\frac{e^{-\frac{\beta c\,\hat{d}^{2}(0,z)}{2t}}}{t^{\mathcal{Q}/2}}\omega(z)dz<+\infty, (83)

is uniformly bounded with respect to t[0,)t\in[0,\infty).

Therefore, for any L>1L>1, we obtain the following estimates for the limit of uu:

lim supt0u(t,x)V1p^1/L(0,z)𝑑ω^(z)+C~ecβL2,lim inft0u(t,x)V1p^1/L(0,z)𝑑ω^(z)C~ecβL2.\begin{split}\limsup_{t\to 0}u(t,x)&\leq\int_{V_{1}}\hat{p}_{1/L}(0,z)d\hat{\omega}(z)+\tilde{C}e^{-\frac{c\beta L}{2}},\\ \liminf_{t\to 0}u(t,x)&\geq\int_{V_{1}}\hat{p}_{1/L}(0,z)d\hat{\omega}(z)-\tilde{C}e^{-\frac{c\beta L}{2}}.\end{split} (84)

In order to evaluate the limits in (84), let us firstly notice that, since p^\hat{p} enjoys upper and lower Gaussian bounds (see for example [CdVHT21, App. C]), reasoning as we did for (82), we can prove the following:

V1p^1/L(0,z)𝑑ω^(z)={z1>0}p^1/L(0,z)𝑑ω^(z)+O(eβL).\int_{V_{1}}\hat{p}_{1/L}(0,z)d\hat{\omega}(z)=\int_{\{z_{1}>0\}}\hat{p}_{1/L}(0,z)d\hat{\omega}(z)+O\left(e^{-\beta^{\prime}L}\right). (85)

Secondly, thanks to (60) for p^\hat{p}, we have the following parity property

p^t(0,z)=p^t(0,δ1(z)),t>0,zn,\hat{p}_{t}(0,z)=\hat{p}_{t}(0,\delta_{-1}(z)),\qquad\forall\,t>0,\ z\in\mathbb{R}^{n}, (86)

and, by the choice of privileged coordinates, δ1({z1>0})={z1<0}\delta_{-1}(\{z_{1}>0\})=\{z_{1}<0\}. Thus, using also the stochastic completeness of the nilpotent approximation, we obtain for any t0t\geq 0,

1\displaystyle 1 =np^t(0,z)𝑑ω^(z)={z1>0}p^t(0,z)𝑑ω^(z)+{z1<0}p^t(0,z)𝑑ω^(z)\displaystyle=\int_{\mathbb{R}^{n}}\hat{p}_{t}(0,z)d\hat{\omega}(z)=\int_{\{z_{1}>0\}}\hat{p}_{t}(0,z)d\hat{\omega}(z)+\int_{\{z_{1}<0\}}\hat{p}_{t}(0,z)d\hat{\omega}(z) (87)
=2{z1>0}p^t(0,z)𝑑ω^(z),\displaystyle=2\int_{\{z_{1}>0\}}\hat{p}_{t}(0,z)d\hat{\omega}(z), (88)

having performed the change of variables zδ1(z)z\mapsto\delta_{-1}(z) in the last equality. Hence, the integral in (85) is

V1p^1/L(0,z)𝑑ω^(z)=12+O(eβL).\int_{V_{1}}\hat{p}_{1/L}(0,z)d\hat{\omega}(z)=\frac{1}{2}+O\left(e^{-\beta^{\prime}L}\right). (89)

Finally, we optimize the inequalities (84) with respect to LL, taking LL\to\infty and concluding the proof. ∎

Remark 3.3.

In the non-compact case, if MM is globally doubling and supports a global Poincaré inequality, the proof above is still valid, cf. Theorem 7.3. Otherwise, a different proof is needed, see [Ros21, App. D] for details.

4 First-order asymptotic expansion of HΩ(t)H_{\Omega}(t)

In this section, we introduce the technical tools that allow us to prove the first-order asymptotic expansion of the relative heat content starting from Theorem 3.1. The new ingredient is a definition of an operator IΩI_{\Omega}, which depends on the base set Ω\Omega.

4.1 A mean value lemma

Define δ:M\delta\colon M\rightarrow\mathbb{R} to be the signed distance function from Ω\partial\Omega, i.e.

δ(x)={dSR(x,Ω)xΩ,dSR(x,Ω)xMΩ,\delta(x)=\begin{cases}d_{\mathrm{SR}}(x,\partial\Omega)&x\in\Omega,\\ -d_{\mathrm{SR}}(x,\partial\Omega)&x\in M\setminus\Omega,\end{cases} (90)

where dSR(,Ω):M[0,+)d_{\mathrm{SR}}(\cdot,\partial\Omega)\colon M\rightarrow[0,+\infty) denotes the usual distance function from Ω\partial\Omega. Let us introduce the following notation: for any a,ba,b\in\mathbb{R}, with a<ba<b, we set

Ωab={xMa<δ(x)<b},\Omega_{a}^{b}=\{x\in M\mid a<\delta(x)<b\}, (91)

with the understanding that if bb (or aa) is omitted, it is assumed to be ++\infty (or -\infty), for example555Notice that the set Ω+=M\Omega_{-\infty}^{+\infty}=M, thus omitting both indexes can create confusion. We will never do that and Ω\Omega will always denote the starting subset of MM.

Ωr=Ωr+={xMr<δ(x)}.\Omega_{r}=\Omega_{r}^{+\infty}=\{x\in M\mid r<\delta(x)\}. (92)

In the non-characteristic case, [FPR20, Prop. 3.1] can be extended without difficulties to the signed distance function.

Theorem 4.1 (Double-sided tubular neighborhood).

Let MM be a sub-Riemannian manifold and let ΩM\Omega\subset M be an open relatively compact subset of MM whose boundary is smooth and has no characteristic points. Let δ:M\delta:M\to\mathbb{R} be the signed distance function from Ω\partial\Omega. Then, we have:

  • i)

    δ\delta is Lipschitz with respect to the sub-Riemannian distance and δg1\|\nabla\delta\|_{g}\leq 1 a.e.;

  • ii)

    there exists r0>0r_{0}>0 such that δ:Ωr0r0\delta\colon\Omega_{-r_{0}}^{r_{0}}\to\mathbb{R} is smooth;

  • iii)

    there exists a smooth diffeomorphism G:(r0,r0)×ΩΩr0r0G:(-r_{0},r_{0})\times\partial\Omega\to\Omega_{-r_{0}}^{r_{0}}, such that

    δ(G(t,y))=tandGt=δ,(t,y)(r0,r0)×Ω.\delta(G(t,y))=t\quad\text{and}\quad G_{*}\partial_{t}=\nabla\delta,\qquad\forall\,(t,y)\in(-r_{0},r_{0})\times\partial\Omega. (93)

    Moreover, δg1\|\nabla\delta\|_{g}\equiv 1 on Ωr0r0\Omega_{-r_{0}}^{r_{0}}.

In particular, the following co-area formula for the signed distance function holds

Ω0rv(x)𝑑ω(x)=0rΩsv(s,y)𝑑σ(y)𝑑s,r0,\int_{\Omega_{0}^{r}}v(x)d\omega(x)=\int_{0}^{r}\int_{\partial\Omega_{s}}v(s,y)d\sigma(y)ds,\qquad\forall\,r\geq 0, (94)

where σ\sigma is the induced measure on Ωs\partial\Omega_{s}, namely the positive measure with density |iδω||Ωs|i_{\nabla\delta}\omega|_{|\partial\Omega_{s}}. From (94), we deduce the sub-Riemannian mean value lemma, see [RR21, Thm. 4.1] for a proof.

Proposition 4.2.

Let MM be a compact sub-Riemannian manifold, equipped with a smooth measure ω\omega, let ΩM\Omega\subset M be an open subset of MM with smooth boundary and no characteristic points and let δ:M\delta\colon M\to\mathbb{R} be the signed distance function from Ω\partial\Omega. Fix a smooth function vC(M)v\in C^{\infty}(M) and define

F(r)=Ωrv(x)𝑑ω(x),r0.F(r)=\int_{\Omega_{r}}v(x)d\omega(x),\qquad\forall\,r\geq 0. (95)

Then there exists r0>0r_{0}>0 such that the function FF is smooth on [0,r0)[0,r_{0}) and, for 0r<r00\leq r<r_{0}:

F′′(r)=ΩrΔv(x)𝑑ω(x)Ωrv(y)divω(ν(y))𝑑σ(y),F^{\prime\prime}(r)=\int_{\Omega_{r}}{\Delta v(x)d\omega(x)}-\int_{\partial\Omega_{r}}{v(y)\mathrm{div}_{\omega}\left(\nu(y)\right)d\sigma(y)}, (96)

where ν\nu is the inward-pointing unit normal to Ωr\Omega_{r}, σ\sigma is the induced measure on Ωr\partial\Omega_{r}.

Remark 4.3.

If vCc(M)v\in C_{c}^{\infty}(M), then neither MM nor Ω\Omega is required to be compact for Proposition 4.2 to be true, indeed its proof relies on (94), which continues to hold, and the divergence theorem (34), which applies if supp(v)\mathrm{supp}(v) is compact. Moreover, we remark that νr\nu_{r} is equal to δ\nabla\delta up to sign. We prefer to keep νr\nu_{r} in (95), since we are going to apply it when the integral is performed over Ωr\Omega_{r} or its complement.

If we choose the function vv in the definition of FF to be 1u(t,x)1-u(t,x), where u(t,)=etΔ𝟙Ωu(t,\cdot)=e^{t\Delta}\mathds{1}_{\Omega}, then, FF satisfies a non-homogeneous one-dimensional heat equation.

Corollary 4.4.

Under the hypotheses of Proposition 4.2, the function

F(t,r)=Ωr(1u(t,x))𝑑ω(x),t>0,r0,F(t,r)=\int_{\Omega_{r}}\left(1-u(t,x)\right)d\omega(x),\qquad\forall\,t>0,\quad r\geq 0, (97)

where u(t,x)=etΔ𝟙Ω(x)u(t,x)=e^{t\Delta}\mathds{1}_{\Omega}(x), satisfies the following non-homogeneous one-dimensional heat equation:

(tr2)F(t,r)=Ωr(1u(t,))divω(ν)𝑑σ,t>0,r[0,r0).(\partial_{t}-\partial_{r}^{2})F(t,r)=\int_{\partial\Omega_{r}}{\left(1-u(t,\cdot)\right)\mathrm{div}_{\omega}(\nu)d\sigma},\qquad t>0,\quad r\in[0,r_{0}). (98)

Here ν\nu is the inward-pointing unit normal to Ωr\Omega_{r}, σ\sigma is the induced measure on Ωr\partial\Omega_{r}.

Corollary 4.4 holds only for rr0r\leq r_{0}, however we would like to extend it to the whole positive half-line, in order to apply a Duhamel’s principle. This can be done up to an error which is exponentially small.

4.2 Localization principle

Proposition 4.5.

Let MM be a compact sub-Riemannian manifold, equipped with a smooth measure ω\omega, and let ΩM\Omega\subset M be an open subset of MM, with smooth boundary. Moreover, let KMK\subset M be a closed set such that KΩ=K\cap\partial\Omega=\emptyset. Then

𝟙Ω(x)u(t,x)=O(t),uniformly for xK,\mathds{1}_{\Omega}(x)-u(t,x)=O(t^{\infty}),\qquad\text{uniformly for }x\in K, (99)

where u(t,x)=etΔ𝟙Ω(x)u(t,x)=e^{t\Delta}\mathds{1}_{\Omega}(x).

Proof.

The statement is a direct consequence of the off-diagonal estimate for the heat kernel in compact sub-Riemannian manifold (see [JSC86, Prop. 3]):

pt(x,y)Caeca/t,x,y with d(x,y)a,t<1,p_{t}(x,y)\leq C_{a}e^{-c_{a}/t},\qquad\forall\,x,y\text{ with }d(x,y)\geq a,\quad t<1, (100)

for suitable constants Ca,ca>0C_{a},c_{a}>0, depending only on aa. Now, since KΩ=K\cap\partial\Omega=\emptyset, we can write KK as a disjoint union

K=K1K2with K1Ω,K2MΩ.K=K_{1}\sqcup K_{2}\qquad\text{with }K_{1}\subset\Omega,\quad K_{2}\subset M\setminus\Omega. (101)

At this point, for i=1,2i=1,2, set ai=dSR(Ki,Ω)>0a_{i}=d_{\mathrm{SR}}(K_{i},\partial\Omega)>0 by hypothesis, and let xK1x\in K_{1}. Then, using the stochastic completeness of MM, we have:

|𝟙Ω(x)u(t,x)|=1u(t,x)=MΩpt(x,y)𝑑ω(y)C1ec1/tω(MΩ),\left|\mathds{1}_{\Omega}(x)-u(t,x)\right|=1-u(t,x)=\int_{M\setminus\Omega}p_{t}(x,y)d\omega(y)\leq C_{1}e^{-c_{1}/t}\omega(M\setminus\Omega), (102)

which is exponentially decaying, uniformly in K1K_{1}. Analogously, if xK2x\in K_{2}, we have

|𝟙Ω(x)u(t,x)|=u(t,x)=Mpt(x,y)𝟙Ω(y)𝑑ω(y)=Ωpt(x,y)𝑑ω(y)C2ec2/tω(Ω),\left|\mathds{1}_{\Omega}(x)-u(t,x)\right|=u(t,x)=\int_{M}p_{t}(x,y)\mathds{1}_{\Omega}(y)d\omega(y)=\int_{\Omega}p_{t}(x,y)d\omega(y)\leq C_{2}e^{-c_{2}/t}\omega(\Omega), (103)

uniformly in K2K_{2}. ∎

Remark 4.6.

In the non-compact case, Proposition 4.5 may fail. Indeed, on the one hand the off-diagonal estimate (100) is not always available, on the other hand the measure of MΩM\setminus\Omega appearing in (102) is infinite. Under additional assumption on MM, we are able to recover a localization principle, see Section 7.

Let MM be compact. Thanks to Proposition 4.5, we can extend the function FF defined in (97), to a solution to a non-homogeneous heat equation such as (98) on the whole half-line. More precisely, let ϕ,ηCc(M)\phi,\eta\in C_{c}^{\infty}(M) such that

ϕ+η1,supp(ϕ)Ωr0r0,supp(η)Ωr0/2Ωr0/2,\phi+\eta\equiv 1,\qquad\mathrm{supp}(\phi)\subset\Omega_{-r_{0}}^{r_{0}},\qquad\mathrm{supp}(\eta)\subset\Omega^{-r_{0}/2}\cup\Omega_{r_{0}/2}, (104)

where r0r_{0} is defined in Proposition 4.2. We have then, for r[0,r0)r\in[0,r_{0}),

F(t,r)\displaystyle F(t,r) =Ωr(1u(t,x))ϕ(x)𝑑ω(x)+Ωr(1u(t,x))η(x)𝑑ω(x)\displaystyle=\int_{\Omega_{r}}\left(1-u(t,x)\right)\phi(x)d\omega(x)+\int_{\Omega_{r}}\left(1-u(t,x)\right)\eta(x)d\omega(x) (105)
=Ωr(1u(t,x))ϕ(x)𝑑ω(x)+supp(η)Ωr(1u(t,x))η(x)𝑑ω(x)\displaystyle=\int_{\Omega_{r}}\left(1-u(t,x)\right)\phi(x)d\omega(x)+\int_{\mathrm{supp}(\eta)\cap\Omega_{r}}\left(1-u(t,x)\right)\eta(x)d\omega(x) (106)
=Ωr(1u(t,x))ϕ(x)𝑑ω(x)+O(t),\displaystyle=\int_{\Omega_{r}}\left(1-u(t,x)\right)\phi(x)d\omega(x)+O(t^{\infty}), (107)

where we used Proposition 4.5 to deal with the second term, having set K=supp(η)ΩrK=\mathrm{supp}(\eta)\cap\Omega_{r}. For this reason, we may focus on the first term in (107).

Definition 4.7.

For all t>0t>0 and r0r\geq 0, we define the operators IΩ,ΛΩ:Cc(Ωr0r0)C((0,)×[0,))I_{\Omega},\Lambda_{\Omega}:C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}})\to C^{\infty}((0,\infty)\times[0,\infty)), associated with Ω\Omega, by

IΩϕ(t,r)\displaystyle I_{\Omega}\phi(t,r) =Ωr(1u(t,x))ϕ(x)𝑑ω(x),\displaystyle=\int_{\Omega_{r}}{\left(1-u(t,x)\right)\phi(x)d\omega(x)}, (108)
ΛΩϕ(t,r)\displaystyle\Lambda_{\Omega}\phi(t,r) =rIΩϕ(t,r)=Ωr(1u(t,y))ϕ(y)𝑑σ(y),\displaystyle=-\partial_{r}I_{\Omega}\phi(t,r)=-\int_{\partial\Omega_{r}}{\left(1-u(t,y)\right)\phi(y)d\sigma(y)}, (109)

for any ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}), and where σ\sigma denotes the induced measure on Ωr\partial\Omega_{r} and u(t,)=etΔ𝟙Ω()u(t,\cdot)=e^{t\Delta}\mathds{1}_{\Omega}(\cdot).

Remark 4.8.

We stress that, for every ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}), IΩϕI_{\Omega}\phi, ΛΩϕ\Lambda_{\Omega}\phi are indeed smooth in both variables thanks to the choice of the parameter r0>0r_{0}>0 as in Proposition 4.2, together with the smoothness of the solution to the heat equation. Moreover, ΛΩϕ\Lambda_{\Omega}\phi is compactly supported in the rr-variable.

Thanks to the localization principle, we can improve Corollary 4.4, obtaining a better result for IΩϕ(t,r)I_{\Omega}\phi(t,r).

Lemma 4.9.

Let L=tr2L=\partial_{t}-\partial_{r}^{2} be the one-dimensional heat operator. Then, for any ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}),

L(IΩϕ(t,r))=IΩΔϕ(t,r)+ΛΩNΩϕ(t,r),t>0,r0,L\left(I_{\Omega}\phi(t,r)\right)=I_{\Omega}\Delta\phi(t,r)+\Lambda_{\Omega}N_{\Omega}\phi(t,r),\qquad\forall\,t>0,\quad r\geq 0, (110)

where NΩN_{\Omega} is the operator defined by:

NΩϕ=2g(ϕ,ν)+ϕdivω(ν),ϕCc(Ωr0r0),N_{\Omega}\phi=2g\left(\nabla\phi,\nu\right)+\phi\,\mathrm{div}_{\omega}(\nu),\qquad\forall\,\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}), (111)

and ν\nu is the inward-pointing unit normal to Ω\Omega.

4.3 Duhamel’s principle for IΩϕI_{\Omega}\phi

We recall for the convenience of the reader a one-dimensional version of the Duhamel’s principle, see [RR21, Lem. 5.4].

Lemma 4.10 (Duhamel’s principle).

Let fC((0,)×[0,))f\in C((0,\infty)\times[0,\infty)), v0,v1C([0,))v_{0},v_{1}\in C([0,\infty)), such that f(t,)f(t,\cdot) and v0v_{0} are compactly supported and assume that

limt0f(t,r),r0.\exists\lim_{t\to 0}f(t,r),\qquad\forall r\geq 0. (112)

Consider the non-homogeneous heat equation on the half-line:

Lv(t,r)\displaystyle Lv(t,r) =f(t,r),\displaystyle=f(t,r), for t>0,r>0,\displaystyle\text{for }t>0,\ r>0, (113)
v(0,r)\displaystyle v(0,r) =v0(r),\displaystyle=v_{0}(r), for r>0,\displaystyle\text{for }r>0,
rv(t,0)\displaystyle\partial_{r}v(t,0) =v1(t),\displaystyle=v_{1}(t), for t>0,\displaystyle\text{for }t>0,

where L=tr2L=\partial_{t}-\partial_{r}^{2}. Then, for t>0t>0 and r0r\geq 0, the solution to (113) is given by

v(t,r)=0e(t,r,s)v0(s)𝑑s+0t0e(tτ,r,s)f(τ,s)𝑑s𝑑τ0te(tτ,r,0)v1(τ)𝑑τ,v(t,r)=\int_{0}^{\infty}e(t,r,s)v_{0}(s)ds+\int_{0}^{t}\int_{0}^{\infty}e(t-\tau,r,s)f(\tau,s)ds\hskip 0.59998ptd\tau\\ -\int_{0}^{t}e(t-\tau,r,0)v_{1}(\tau)d\tau, (114)

where e(t,r,s)e(t,r,s) is the Neumann heat kernel on the half-line, that is

e(t,r,s)=14πt(e(rs)2/4t+e(r+s)2/4t).e(t,r,s)=\frac{1}{\sqrt{4\pi t}}\left(e^{-(r-s)^{2}/4t}+e^{-(r+s)^{2}/4t}\right). (115)

Finally, we apply Lemma 4.10 to obtain an asymptotic equality for IΩϕI_{\Omega}\phi. The main difference with the result of [RR21, Thm. 5.6] is that the former will not be a true first-order asymptotic expansion.

Corollary 4.11.

Let MM be a compact sub-Riemannian manifold, equipped with a smooth measure ω\omega, and let ΩM\Omega\subset M be an open subset whose boundary is smooth and has no characteristic points. Then, for any function ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}),

IΩϕ(t,0)=1π0tΩ(1u(τ,y))ϕ(y)𝑑σ(y)(tτ)1/2𝑑τ+O(t),I_{\Omega}\phi(t,0)=\frac{1}{\sqrt{\pi}}\int_{0}^{t}\int_{\partial\Omega}\left(1-u(\tau,y)\right)\phi(y)d\sigma(y)(t-\tau)^{-1/2}d\tau+O(t), (116)

as t0t\to 0, where u(t,)=etΔ𝟙Ω()u(t,\cdot)=e^{t\Delta}\mathds{1}_{\Omega}(\cdot).

Proof.

By Lemma 4.9, the function IΩϕ(t,r)I_{\Omega}\phi(t,r) satisfies the following Neumann problem on the half-line:

LIΩϕ(t,r)\displaystyle LI_{\Omega}\phi(t,r) =f(t,r),\displaystyle=f(t,r), for t>0,r>0,\displaystyle\text{for }t>0,\ r>0, (117)
IΩϕ(0,r)\displaystyle I_{\Omega}\phi(0,r) =0,\displaystyle=0, for r>0,\displaystyle\text{for }r>0,
rIΩϕ(t,0)\displaystyle\partial_{r}I_{\Omega}\phi(t,0) =ΛΩϕ(t,0),\displaystyle=-\Lambda_{\Omega}\phi(t,0), for t>0,\displaystyle\text{for }t>0,

where the source is given by f(t,r)=IΩΔϕ(t,r)+ΛΩNΩϕ(t,r)f(t,r)=I_{\Omega}\Delta\phi(t,r)+\Lambda_{\Omega}N_{\Omega}\phi(t,r). Thus, applying Duhamel’s formula (114), we have:

IΩϕ(t,0)=0t0+e(tτ,0,s)f(τ,s)𝑑s𝑑τ+1π0t1tτΛΩϕ(t,0)𝑑τ.I_{\Omega}\phi(t,0)=\int_{0}^{t}\int_{0}^{+\infty}e(t-\tau,0,s)f(\tau,s)ds\hskip 0.59998ptd\tau+\frac{1}{\sqrt{\pi}}\int_{0}^{t}\frac{1}{\sqrt{t-\tau}}\Lambda_{\Omega}\phi(t,0)d\tau. (118)

Since the source is uniformly bounded by the weak maximum principle (40), the first integral is a remainder of order tt, as t0t\to 0, concluding the proof. ∎

Remark 4.12.

We mention that a relevant role in the sequel will be played by the operators IΩI_{\Omega}, cf. Definition 4.7, associated with either Ω\Omega or its complement Ωc\Omega^{c}.

4.4 First-order asymptotics

In this section we prove the first-order asymptotic expansion of HΩ(t)H_{\Omega}(t), cf. Theorem 1.1 at order 11. We will use Corollary 4.11, for the inside contribution:

Iϕ(t,r)=Ωr(1u(t,x))ϕ(x)𝑑ω(x),t>0,r0,I\phi(t,r)=\int_{\Omega_{r}}{\left(1-u(t,x)\right)\phi(x)d\omega(x)},\qquad\forall\,t>0,\quad r\geq 0, (119)

for any ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}), and where σ\sigma denotes the induced measure on Ωr\partial\Omega_{r} and u(t,)=etΔ𝟙Ω()u(t,\cdot)=e^{t\Delta}\mathds{1}_{\Omega}(\cdot) is the solution to (36). The quantity (119) is just Definition 4.7, applied with base set ΩM\Omega\subset M.

Theorem 4.13.

Let MM be a compact sub-Riemannian manifold, equipped with a smooth measure ω\omega, and let ΩM\Omega\subset M be an open subset whose boundary is smooth and has no characteristic points. Then,

HΩ(t)=ω(Ω)1πσ(Ω)t1/2+o(t1/2),as t0.H_{\Omega}(t)=\omega(\Omega)-\frac{1}{\sqrt{\pi}}\sigma(\partial\Omega)t^{1/2}+o(t^{1/2}),\qquad\text{as }t\to 0. (120)
Proof.

Let ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}) be as in (104), namely

0ϕ1,andϕ1inΩr0/2r0/2.0\leq\phi\leq 1,\qquad\text{and}\qquad\phi\equiv 1\quad\text{in}\quad\Omega_{-r_{0}/2}^{r_{0}/2}. (121)

Then, by the localization principle, cf. (107), we have that

ω(Ω)HΩ(t)=Iϕ(t,0)+O(t),as t0.\omega(\Omega)-H_{\Omega}(t)=I\phi(t,0)+O(t^{\infty}),\qquad\text{as }t\to 0. (122)

Applying Corollary 4.11, we have:

Iϕ(t,0)=1π0tΩ(1u(τ,y))ϕ(y)𝑑σ(y)(tτ)1/2𝑑τ+O(t),as t0.I\phi(t,0)=\frac{1}{\sqrt{\pi}}\int_{0}^{t}\int_{\partial\Omega}\left(1-u(\tau,y)\right)\phi(y)d\sigma(y)(t-\tau)^{-1/2}d\tau+O(t),\qquad\text{as }t\to 0. (123)

Thus, to infer the first-order term of the asymptotic expansion we have to compute the following limit:

limt0Iϕ(t,0)t1/2=limt01t1/2π0tΩ(1u(τ,y))ϕ(y)𝑑σ(y)(tτ)1/2𝑑τ.\lim_{t\to 0}\frac{I\phi(t,0)}{t^{1/2}}=\lim_{t\to 0}\frac{1}{t^{1/2}\sqrt{\pi}}\int_{0}^{t}\int_{\partial\Omega}\left(1-u(\tau,y)\right)\phi(y)d\sigma(y)(t-\tau)^{-1/2}d\tau. (124)

Firstly, by the change of variable in the integral τtτ\tau\mapsto t\tau, we rewrite the argument of the limit (124) as

1π01Ω(1u(tτ,y))ϕ(y)𝑑σ(y)(1τ)1/2𝑑τ.\frac{1}{\sqrt{\pi}}\int_{0}^{1}\int_{\partial\Omega}\left(1-u(t\tau,y)\right)\phi(y)d\sigma(y)(1-\tau)^{-1/2}d\tau. (125)

Secondly, we apply the dominated convergence theorem. Indeed, on the one hand, by Theorem 3.1 we have point-wise convergence

(1u(tτ,y))ϕ(y)t012ϕ(y),yΩ,τ(0,1),\left(1-u(t\tau,y)\right)\phi(y)\xrightarrow{t\to 0}\frac{1}{2}\phi(y),\qquad\forall\,y\in\partial\Omega,\quad\tau\in(0,1), (126)

and on the other hand, by the maximum principle

|Ω(1u(tτ,y))ϕ(y)𝑑σ(y)(1τ)1/2|Ω|ϕ|𝑑σ(1τ)1/2L1(0,1),\left|\int_{\partial\Omega}\left(1-u(t\tau,y)\right)\phi(y)d\sigma(y)(1-\tau)^{-1/2}\right|\leq\int_{\partial\Omega}|\phi|d\sigma(1-\tau)^{-1/2}\in L^{1}(0,1), (127)

for any t>0t>0. Therefore, we finally obtain that:

Iϕ(t,0)=tπΩϕ(y)𝑑σ(y)+o(t1/2),as t0.I\phi(t,0)=\sqrt{\frac{t}{\pi}}\int_{\partial\Omega}\phi(y)d\sigma(y)+o(t^{1/2}),\qquad\text{as }t\to 0. (128)

Recalling that ϕ|Ω1\phi_{|\partial\Omega}\equiv 1, we conclude the proof. ∎

Remark 4.14.

The above technique used to evaluate the first-order coefficient causes a loss of precision in the remainder, with respect to the expression (123), where the remainder is O(t)O(t). This loss comes from the application of Theorem 3.1, which does not contain any remainder estimate.

5 Higher-order asymptotic expansion of HΩ(t)H_{\Omega}(t)

We iterate Duhamel’s formula (114) for the inside contribution to study the higher-order asymptotics of HΩ(t)H_{\Omega}(t). We obtain the following expression for IϕI\phi, at order 33:

Iϕ(t,0)=1π0tΩ(1u(τ,))ϕ𝑑σ(tτ)1/2𝑑τ+12π0t0τΩ(1u(τ^,))Nϕ𝑑σ((ττ^)(tτ))1/2𝑑τ^𝑑τ+O(t3/2),I\phi(t,0)=\frac{1}{\sqrt{\pi}}\int_{0}^{t}\int_{\partial\Omega}(1-u(\tau,\cdot))\phi\hskip 0.59998ptd\sigma(t-\tau)^{-1/2}d\tau\\ +\frac{1}{2\pi}\int_{0}^{t}\int_{0}^{\tau}\int_{\partial\Omega}(1-u(\hat{\tau},\cdot))N\phi\hskip 0.59998ptd\sigma((\tau-\hat{\tau})(t-\tau))^{-1/2}d\hat{\tau}\hskip 0.59998ptd\tau+O(t^{3/2}), (129)

where u(t,)=etΔ𝟙Ω()u(t,\cdot)=e^{t\Delta}\mathds{1}_{\Omega}(\cdot) denotes the solution to (36) and NN is the operator defined by

Nϕ=2g(ϕ,δ)+ϕΔδ,ϕCc(Ωr0r0),N\phi=2g(\nabla\phi,\nabla\delta)+\phi\Delta\delta,\qquad\forall\,\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}), (130)

with δ:M\delta\colon M\rightarrow\mathbb{R} the signed distance function from Ω\partial\Omega. The computations for deriving (129) are technical. We refer to Appendix A for further details, and in particular to Lemma A.6. Motivated by (129), we introduce the following functional.

Definition 5.1.

Let MM be a sub-Riemannian manifold, equipped with a smooth measure ω\omega, let ΩM\Omega\subset M be a relatively compact subset with smooth boundary and let vC((0,+)×M)v\in C^{\infty}((0,+\infty)\times M). Define the functional 𝒢v\mathcal{G}_{v}, for any ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}) as:

𝒢v[ϕ](t)=12π0tΩv(τ,)ϕ𝑑σ(tτ)1/2𝑑τ,t0,\mathcal{G}_{v}[\phi](t)=\frac{1}{2\sqrt{\pi}}\int_{0}^{t}\int_{\partial\Omega}v(\tau,\cdot)\phi\hskip 0.59998ptd\sigma(t-\tau)^{-1/2}d\tau,\qquad\forall\,t\geq 0, (131)

where σ\sigma is the sub-Riemannian induced measure on Ω\partial\Omega.

Notice that the functional 𝒢v\mathcal{G}_{v} is linear with respect to the subscript function vv, by linearity of the integral. Moreover, when the function vv is chosen to be the solution to (36), we easily obtain the following corollary of Theorem 3.1, which is just a rewording of (124).

Corollary 5.2.

Let MM be a compact sub-Riemannian manifold, equipped with a smooth measure ω\omega, and let ΩM\Omega\subset M be an open subset whose boundary is smooth and has no characteristic points. Let ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}), then,

𝒢u[ϕ](t)=12πΩϕ(y)𝑑σ(y)t1/2+o(t1/2),as t0.\mathcal{G}_{u}[\phi](t)=\frac{1}{2\sqrt{\pi}}\int_{\partial\Omega}\phi(y)d\sigma(y)t^{1/2}+o(t^{1/2}),\qquad\text{as }t\to 0. (132)

Then, we can rewrite (129) in a compact notation:

Iϕ(t,0)=2𝒢1u[ϕ](t)+1π0t𝒢1u[Nϕ](t)𝑑σ(tτ)1/2𝑑τ+O(t3/2).I\phi(t,0)=2\mathcal{G}_{1-u}[\phi](t)+\frac{1}{\sqrt{\pi}}\int_{0}^{t}\mathcal{G}_{1-u}[N\phi](t)d\sigma(t-\tau)^{-1/2}d\tau+O(t^{3/2}). (133)

However, on the one hand, the application of Corollary 5.2 to (129) does not give any new information on the asymptotics of HΩ(t)H_{\Omega}(t), as the first term produces an error of o(t1/2)o(t^{1/2}). On the other hand, it is clear that an asymptotic series of 𝒢u\mathcal{G}_{u} is enough to deduce the small-time expansion of HΩ(t)H_{\Omega}(t).

5.1 The outside contribution and an asymptotic series for 𝒢u[ϕ]\mathcal{G}_{u}[\phi]

In this section, we deduce an asymptotic series, at order 33, of 𝒢u[ϕ](t)\mathcal{G}_{u}[\phi](t) as t0t\to 0. This is done exploiting the fact that the diffusion of heat is not confined in Ω\Omega, and as a result we can define an outside contribution, namely the quantity obtained from Definition 4.7, applied with base set ΩcM\Omega^{c}\subset M:

Icϕ(t,r)=(Ωc)r(1uc(t,x))ϕ(x)𝑑ω(x),t>0,r0,I^{c}\phi(t,r)=\int_{(\Omega^{c})_{r}}{\left(1-u^{c}(t,x)\right)\phi(x)d\omega(x)},\qquad\forall\,t>0,\quad r\geq 0, (134)

for any ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}), and where σ\sigma denotes the induced measure on the boundary of (Ωc)r(\Omega^{c})_{r} and uc(t,x)=etΔ𝟙Ωc(x)u^{c}(t,x)=e^{t\Delta}\mathds{1}_{\Omega^{c}}(x). We remark that, since Ω\Omega and its complement share the boundary, then (Ωc)r0r0=Ωr0r0(\Omega^{c})_{-r_{0}}^{r_{0}}=\Omega_{-r_{0}}^{r_{0}}. It is convenient to introduce (134), because it turns out that the quantity IϕIcϕI\phi-I^{c}\phi, where IϕI\phi is the inside contribution (119), has an explicit asymptotic series in integer powers of tt.

Proposition 5.3.

Let MM be a compact sub-Riemannian manifold, equipped with a smooth measure ω\omega, and let ΩM\Omega\subset M be an open subset with smooth boundary. Let ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}), then, for any kk\in\mathbb{N},

Iϕ(t,0)Icϕ(t,0)=i=1kai(ϕ)ti+O(tk+1),as t0,I\phi(t,0)-I^{c}\phi(t,0)=\sum_{i=1}^{k}a_{i}(\phi)t^{i}+O(t^{k+1}),\qquad\text{as }t\to 0, (135)

where

ai(ϕ)\displaystyle a_{i}(\phi) =Ωg((Δi1ϕ),δ)𝑑σ,for i1.\displaystyle=\int_{\partial\Omega}g(\nabla(\Delta^{i-1}\phi),\nabla\delta)d\sigma,\qquad\text{for }i\geq 1. (136)
Proof.

Recall that in the definition of the outside contribution (134), the integrand function involves uc(t,x)=etΔ𝟙Ωc(x)u^{c}(t,x)=e^{t\Delta}\mathds{1}_{\Omega^{c}}(x). Since MM is compact, and hence stochastically complete, we have:

1uc(t,x)=etΔ1(x)etΔ𝟙Ωc(x)=u(t,x),t>0,xM,1-u^{c}(t,x)=e^{t\Delta}1(x)-e^{t\Delta}\mathds{1}_{\Omega^{c}}(x)=u(t,x),\qquad\forall t>0,\quad x\in M, (137)

having used the point-wise equality 1𝟙Ωc=𝟙Ω1-\mathds{1}_{\Omega^{c}}=\mathds{1}_{\Omega} in MΩM\setminus\partial\Omega. Therefore, we can write the difference Iϕ(t,0)Icϕ(t,0)I\phi(t,0)-I^{c}\phi(t,0) as follows:

Iϕ(t,0)Icϕ(t,0)\displaystyle I\phi(t,0)-I^{c}\phi(t,0) =Ω(1u(t,))ϕ𝑑ωΩc(1uc(t,))ϕ𝑑ω\displaystyle=\int_{\Omega}\left(1-u(t,\cdot)\right)\phi\hskip 0.59998ptd\omega-\int_{\Omega^{c}}\left(1-u^{c}(t,\cdot)\right)\phi\hskip 0.59998ptd\omega (138)
=Ω(1u(t,))ϕ𝑑ωΩcu(t,)ϕ𝑑ω\displaystyle=\int_{\Omega}\left(1-u(t,\cdot)\right)\phi\hskip 0.59998ptd\omega-\int_{\Omega^{c}}u(t,\cdot)\phi\hskip 0.59998ptd\omega (139)
=Ωϕ(x)𝑑ω(x)Mu(t,x)ϕ(x)𝑑ω(x).\displaystyle=\int_{\Omega}\phi(x)d\omega(x)-\int_{M}u(t,x)\phi(x)d\omega(x). (140)

Since u(t,x)u(t,x) is the solution to (36), the function (140) is smooth as t[0,)t\in[0,\infty). Indeed, the smoothness in the open interval is guaranteed by hypoellipticity of the sub-Laplacian. At t=0t=0, the divergence theorem, together with the fact that ϕ\phi has compact support in MM, implies that

ti(Mu(t,x)ϕ(x)𝑑ω(x))=Mti(u(t,x)ϕ(x))dω(x)=MΔiu(t,x)ϕ(x)𝑑ω(x)=Mu(t,x)Δiϕ(x)𝑑ω(x)t0ΩΔiϕ(x)𝑑ω(x).\begin{split}\partial_{t}^{i}\left(\int_{M}u(t,x)\phi(x)d\omega(x)\right)&=\int_{M}\partial_{t}^{i}\left(u(t,x)\phi(x)\right)d\omega(x)=\int_{M}\Delta^{i}u(t,x)\phi(x)d\omega(x)\\ &=\int_{M}u(t,x)\Delta^{i}\phi(x)d\omega(x)\xrightarrow{t\to 0}\int_{\Omega}\Delta^{i}\phi(x)d\omega(x).\end{split} (141)

The previous limit shows that (140) is smooth at t=0t=0, and also that its asymptotic expansion at order kk, as t0t\to 0, coincides with its kk-th Taylor polynomial at t=0t=0. Finally, we recover (135), applying once again the divergence theorem:

ΩΔiϕ𝑑ω=Ωg((Δi1ϕ),ν)𝑑σ=Ωg((Δi1ϕ),δ)𝑑σ,\int_{\Omega}\Delta^{i}\phi\hskip 0.59998ptd\omega=-\int_{\partial\Omega}g(\nabla(\Delta^{i-1}\phi),\nu)d\sigma=-\int_{\partial\Omega}g(\nabla(\Delta^{i-1}\phi),\nabla\delta)d\sigma, (142)

recalling that ν=δ\nu=\nabla\delta is the inward-pointing normal vector to Ω\Omega at its boundary. ∎

Applying the (iterated) Duhamel’s principle (114) to the difference IϕIcϕI\phi-I^{c}\phi, we are able to obtain relevant information on the functional 𝒢u\mathcal{G}_{u}.

Theorem 5.4.

Let MM be a compact sub-Riemannian manifold, equipped with a smooth measure ω\omega, and let ΩM\Omega\subset M be an open subset whose boundary is smooth and has no characteristic points. Then, for any ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}),

𝒢u[ϕ](t)=12πΩϕ𝑑σt1/2+18ΩϕΔδ𝑑σt+o(t3/2),as t0.\mathcal{G}_{u}[\phi](t)=\frac{1}{2\sqrt{\pi}}\int_{\partial\Omega}\phi\hskip 0.59998ptd\sigma\,t^{1/2}+\frac{1}{8}\int_{\partial\Omega}\phi\Delta\delta\hskip 0.59998ptd\sigma\,t+o(t^{3/2}),\qquad\text{as }t\to 0. (143)
Proof.

Let us study the difference of the inside and outside contributions Iϕ(t,0)Icϕ(t,0)I\phi(t,0)-I^{c}\phi(t,0): on the one hand, we have an iterated Duhamel’s principle, cf. Lemma A.7, which we report here:

(IϕIcϕ)(t,0)\displaystyle\left(I\phi-I^{c}\phi\right)(t,0) =2𝒢12u[ϕ](t)+12ΩNϕ𝑑σt\displaystyle=2\mathcal{G}_{1-2u}[\phi](t)+\frac{1}{2}\int_{\partial\Omega}N\phi\hskip 0.59998ptd\sigma\,t (144)
+12π0t0τ𝒢12u[N2ϕ](τ^)((ττ^)(tτ))1/2𝑑τ^𝑑τ\displaystyle\quad+\frac{1}{2\pi}\int_{0}^{t}\int_{0}^{\tau}\mathcal{G}_{1-2u}[N^{2}\phi](\hat{\tau})\left((\tau-\hat{\tau})(t-\tau)\right)^{-1/2}d\hat{\tau}\hskip 0.59998ptd\tau (145)
+14π0tΩ(12u(τ,))(4ΔN2)ϕ𝑑σ(tτ)1/2𝑑τ+O(t2),\displaystyle\quad+\frac{1}{4\sqrt{\pi}}\int_{0}^{t}\int_{\partial\Omega}\left(1-2u(\tau,\cdot)\right)(4\Delta-N^{2})\phi\hskip 0.59998ptd\sigma(t-\tau)^{1/2}d\tau+O(t^{2}), (146)

where we recall that NN is the operator acting on smooth functions compactly supported close to Ω\partial\Omega defined by

Nϕ=2g(ϕ,δ)+ϕΔδ,ϕCc(Ωr0r0).N\phi=2g(\nabla\phi,\nabla\delta)+\phi\Delta\delta,\qquad\forall\,\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}). (147)

Using Corollary 5.2 and the linearity of 𝒢v\mathcal{G}_{v} with respect to vv, we know that

𝒢12u[ϕ](t)=o(t1/2),as t0,ϕCc(Ωr0r0).\mathcal{G}_{1-2u}[\phi](t)=o(t^{1/2}),\qquad\text{as }t\to 0,\quad\forall\,\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}). (148)

Therefore, applying (148) to the function N2ϕCc(Ωr0r0)N^{2}\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}), we obtain

12π0t0τ𝒢12u[N2ϕ](τ^)((ττ^)(tτ))1/2𝑑τ^𝑑τ=o(t3/2),as t0.\frac{1}{2\pi}\int_{0}^{t}\int_{0}^{\tau}\mathcal{G}_{1-2u}[N^{2}\phi](\hat{\tau})\left((\tau-\hat{\tau})(t-\tau)\right)^{-1/2}d\hat{\tau}\hskip 0.59998ptd\tau=o(t^{3/2}),\qquad\text{as }t\to 0. (149)

In addition, an application of Theorem 3.1 and the dominated convergence theorem implies that

0tΩ(12u(τ,))(4ΔN2)ϕ𝑑σ(tτ)1/2𝑑τ=o(t3/2)as t0.\int_{0}^{t}\int_{\partial\Omega}\left(1-2u(\tau,\cdot)\right)(4\Delta-N^{2})\phi\hskip 0.59998ptd\sigma(t-\tau)^{1/2}d\tau=o(t^{3/2})\qquad\text{as }t\to 0. (150)

Thus, using (149) and (150), we can improve (144), obtaining

Iϕ(t,0)Icϕ(t,0)=2𝒢12u[ϕ](t)+12ΩNϕ𝑑σt+o(t3/2).I\phi(t,0)-I^{c}\phi(t,0)=2\mathcal{G}_{1-2u}[\phi](t)+\frac{1}{2}\int_{\partial\Omega}N\phi\hskip 0.59998ptd\sigma\,t+o(t^{3/2}). (151)

On the other hand, the quantity Iϕ(t,0)Icϕ(t,0)I\phi(t,0)-I^{c}\phi(t,0) has a complete asymptotic series by Proposition 5.3, which at order 33 becomes:

Iϕ(t,0)Icϕ(t,0)=Ωg(ϕ,δ)𝑑σt+o(t3/2),as t0.I\phi(t,0)-I^{c}\phi(t,0)=\int_{\partial\Omega}g(\nabla\phi,\nabla\delta)d\sigma\,t+o(t^{3/2}),\qquad\text{as }t\to 0. (152)

Comparing (151) and (152), we deduce that, as t0t\to 0,

2𝒢12u[ϕ](t)\displaystyle 2\mathcal{G}_{1-2u}[\phi](t) =12ΩNϕ𝑑σt+o(t3/2)+Ωg(ϕ,δ)𝑑σt+o(t3/2)\displaystyle=-\frac{1}{2}\int_{\partial\Omega}N\phi\hskip 0.59998ptd\sigma\,t+o(t^{3/2})+\int_{\partial\Omega}g(\nabla\phi,\nabla\delta)d\sigma\,t+o(t^{3/2}) (153)
=12ΩϕΔδ𝑑σt+o(t3/2).\displaystyle=-\frac{1}{2}\int_{\partial\Omega}\phi\Delta\delta\hskip 0.59998ptd\sigma\,t+o(t^{3/2}). (154)

Finally, using the linearity of the functional 𝒢v[ϕ]\mathcal{G}_{v}[\phi] with respect to vv, we conclude the proof. ∎

Remark 5.5.

The asymptotics (143) for the functional 𝒢u[ϕ](t)\mathcal{G}_{u}[\phi](t) is the best result that we are able to achieve. In the expression (144), the problematic term is given by (150), i.e.

0tΩ(12u(τ,))(4ΔN2)ϕ𝑑σ(tτ)1/2𝑑τ,\int_{0}^{t}\int_{\partial\Omega}\left(1-2u(\tau,\cdot)\right)(4\Delta-N^{2})\phi\hskip 0.59998ptd\sigma(t-\tau)^{1/2}d\tau, (155)

which can not be expressed in terms of 𝒢u\mathcal{G}_{u}, hence the only relevant information is given by Theorem 3.1. In conclusion, we can not repeat the strategy of the proof of Theorem 5.4, replacing the series of 𝒢u\mathcal{G}_{u} at order 33 in (144) to deduce the higher-order terms.

5.2 Fourth-order asymptotics

In this section we prove Theorem 1.1. We recall here the statement.

Theorem 5.6.

Let MM be a compact sub-Riemannian manifold, equipped with a smooth measure ω\omega, and let ΩM\Omega\subset M be an open subset whose boundary is smooth and has no characteristic points. Then, as t0t\to 0,

HΩ(t)=ω(Ω)1πσ(Ω)t1/2112πΩ(2g(δ,(Δδ))(Δδ)2)𝑑σt3/2+o(t2).H_{\Omega}(t)=\omega(\Omega)-\frac{1}{\sqrt{\pi}}\sigma(\partial\Omega)t^{1/2}-\frac{1}{12\sqrt{\pi}}\int_{\partial\Omega}\left(2g(\nabla\delta,\nabla(\Delta\delta))-(\Delta\delta)^{2}\right)d\sigma\,t^{3/2}+o(t^{2}). (156)

Before giving the proof of the theorem, let us comment on its strategy. Recall that, on the one hand, for a cutoff function ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}) which is identically 11 close to Ω\partial\Omega, cf. (104), the localization principle (122) holds, namely

ω(Ω)HΩ(t)=Iϕ(t,0)+O(t),as t0.\omega(\Omega)-H_{\Omega}(t)=I\phi(t,0)+O(t^{\infty}),\qquad\text{as }t\to 0. (157)

Moreover, by the iterated Duhamel’s principle for Iϕ(t,0)I\phi(t,0), cf. Lemma A.6, we can deduce expression (133), namely

Iϕ(t,0)=2𝒢1u[ϕ](t)+1π0t𝒢1u[Nϕ](t)𝑑σ(tτ)1/2𝑑τ+O(t3/2).I\phi(t,0)=2\mathcal{G}_{1-u}[\phi](t)+\frac{1}{\sqrt{\pi}}\int_{0}^{t}\mathcal{G}_{1-u}[N\phi](t)d\sigma(t-\tau)^{-1/2}d\tau+O(t^{3/2}). (158)

On the other hand, we have an asymptotic series of the functional 𝒢u\mathcal{G}_{u} at order 33, cf. Theorem 3.1. Therefore, if we naively insert this series in (158), we can obtain, at most, a third-order asymptotic expansion of the relative heat content HΩ(t)H_{\Omega}(t), whereas we are interested in the fourth-order expansion.

Using the outside contribution, we are able to overcome this difficulty. In particular, applying Proposition 5.3, for a function ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}) which is identically 11 close to Ω\partial\Omega, we have the following asymptotic relation:

Iϕ(t,0)=Icϕ(t,0)+O(t),as t0.I\phi(t,0)=I^{c}\phi(t,0)+O(t^{\infty}),\qquad\text{as }t\to 0. (159)

Notice that (159) is a direct consequence of Proposition 5.3 since all the coefficients of the expansion vanish. Therefore, thanks to (159), we can rephrase (157) as follows:

ω(Ω)HΩ(t)=12(Iϕ(t,0)+Icϕ(t,0))+O(t),as t0.\omega(\Omega)-H_{\Omega}(t)=\frac{1}{2}\left(I\phi(t,0)+I^{c}\phi(t,0)\right)+O(t^{\infty}),\qquad\text{as }t\to 0. (160)

The advantage of (160) is that we can now apply the iterated Dirichlet principle for the sum Iϕ+IcϕI\phi+I^{c}\phi, cf. Lemma A.8. Already at order 33, we obtain

(Iϕ+Icϕ)(t,0)=2πΩϕ𝑑σt1/2+1π0t𝒢12u[Nϕ](τ)(tτ)1/2𝑑τ+O(t3/2),\left(I\phi+I^{c}\phi\right)(t,0)=\frac{2}{\sqrt{\pi}}\int_{\partial\Omega}\phi\hskip 0.59998ptd\sigma\,t^{1/2}+\frac{1}{\sqrt{\pi}}\int_{0}^{t}\mathcal{G}_{1-2u}[N\phi](\tau)(t-\tau)^{-1/2}d\tau+O(t^{3/2}), (161)

where NN is the operator defined in (130). As we can see, in (161), the functional 𝒢u\mathcal{G}_{u} occurs for the first time in the second iteration of the Duhamel’s principle, as opposed to the expansion for IϕI\phi, where it appeared already in the first application, cf. (158). Hence we gain an order with respect to the asymptotic series of 𝒢u\mathcal{G}_{u}. More generally, if we were able to develop the kk-th order asymptotics for 𝒢u\mathcal{G}_{u}, this would imply the (k+1)(k+1)-th order expansion for HΩ(t)H_{\Omega}(t).

Proof of Theorem 5.6.

Following the discussion above, it is enough to expand the sum Iϕ+IcϕI\phi+I^{c}\phi, with ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}). For this quantity, Lemma A.8 holds, namely we have the following iterated version of Duhamel’s principle:

(Iϕ+Icϕ)(t,0)\displaystyle\left(I\phi+I^{c}\phi\right)(t,0) =2πΩϕ𝑑σt1/2+1π0t𝒢12u[Nϕ](τ)(tτ)1/2𝑑τ\displaystyle=\frac{2}{\sqrt{\pi}}\int_{\partial\Omega}\phi\hskip 0.59998ptd\sigma\,t^{1/2}+\frac{1}{\sqrt{\pi}}\int_{0}^{t}\mathcal{G}_{1-2u}[N\phi](\tau)(t-\tau)^{-1/2}d\tau (162)
+16πΩ(4Δ+N2)ϕ𝑑σt3/2\displaystyle\quad+\frac{1}{6\sqrt{\pi}}\int_{\partial\Omega}(4\Delta+N^{2})\phi\hskip 0.59998ptd\sigma\,t^{3/2}
+14π3/20t0τ0τ^𝒢12u[N3ϕ](s)((τ^s)(ττ^)(tτ))1/2𝑑s𝑑τ^𝑑τ\displaystyle\quad+\frac{1}{4\pi^{3/2}}\int_{0}^{t}\!\int_{0}^{\tau}\!\int_{0}^{\hat{\tau}}\!\mathcal{G}_{1-2u}[N^{3}\phi](s)\left((\hat{\tau}-s)(\tau-\hat{\tau})(t-\tau)\right)^{-1/2}\!ds\hskip 0.59998ptd\hat{\tau}\hskip 0.59998ptd\tau
+14π0t𝒢12u[(6NΔN32ΔN)ϕ](τ)(tτ)1/2𝑑τ+O(t5/2),\displaystyle\quad+\frac{1}{4\sqrt{\pi}}\int_{0}^{t}\mathcal{G}_{1-2u}[(6N\Delta-N^{3}-2\Delta N)\phi](\tau)(t-\tau)^{1/2}d\tau+O(t^{5/2}),

where NN is defined in (130). Moreover, recall that by Theorem 5.4, the functional 𝒢12u[ϕ]\mathcal{G}_{1-2u}[\phi] has the following expansion for any ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}):

𝒢12u[ϕ](t)=14ΩϕΔδ𝑑σt+o(t3/2),as t0.\mathcal{G}_{1-2u}[\phi](t)=-\frac{1}{4}\int_{\partial\Omega}\phi\Delta\delta\hskip 0.59998ptd\sigma\,t+o(t^{3/2}),\qquad\text{as }t\to 0. (163)

Thus, replacing the term 𝒢12u[Nϕ]\mathcal{G}_{1-2u}[N\phi] in (162) with the expansion (163) for NϕCc(Ωr0r0)N\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}), we obtain the following asymptotic as t0t\to 0:

Iϕ(t,0)+Icϕ(t,0)=2πΩϕ𝑑σt1/213π(ΩNϕΔδ𝑑σ)t3/2+16πΩ(4Δ+N2)ϕ𝑑σt3/2+o(t2),I\phi(t,0)+I^{c}\phi(t,0)=\frac{2}{\sqrt{\pi}}\int_{\partial\Omega}\phi\hskip 0.59998ptd\sigma\,t^{1/2}-\frac{1}{3\sqrt{\pi}}\left(\int_{\partial\Omega}N\phi\Delta\delta\hskip 0.59998ptd\sigma\right)t^{3/2}\\ +\frac{1}{6\sqrt{\pi}}\int_{\partial\Omega}(4\Delta+N^{2})\phi\hskip 0.59998ptd\sigma\,t^{3/2}+o(t^{2}), (164)

for any ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}). In particular, if we choose ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}) such that ϕ1\phi\equiv 1 close to Ω\partial\Omega, then on the one hand, from (164), we obtain, as t0t\to 0,

Iϕ(t,0)+Icϕ(t,0)=2πσ(Ω)t1/2+16πΩ(2g(δ,(Δδ))(Δδ)2)𝑑σt3/2+o(t2).I\phi(t,0)+I^{c}\phi(t,0)=\frac{2}{\sqrt{\pi}}\sigma(\partial\Omega)t^{1/2}\\ +\frac{1}{6\sqrt{\pi}}\int_{\partial\Omega}\left(2g(\nabla\delta,\nabla(\Delta\delta))-(\Delta\delta)^{2}\right)d\sigma\,t^{3/2}+o(t^{2}). (165)

On the other hand, the asymptotic relation (160) holds. This concludes the proof. ∎

Third-order vs fourth-order asymptotics.

We stress that we could have obtained the third-order asymptotic expansion of HΩ(t)H_{\Omega}(t) without introducing the sum of the inside and outside contributions Iϕ+IcϕI\phi+I^{c}\phi, and only using the Duhamel’s principle for IϕI\phi, cf. Lemma A.6, and the asymptotic series for 𝒢u\mathcal{G}_{u}, cf. Theorem 5.4. However, for the improvement to the fourth-order asymptotics, the argument of the sum of contributions seems necessary.

5.3 The weighted relative heat content

Adapting the proof of Theorem 5.6, one can prove a slightly more general result which we state here for completeness.

Proposition 5.7.

Let MM be a compact sub-Riemannian manifold, equipped with a smooth measure ω\omega, and let ΩM\Omega\subset M be an open subset whose boundary is smooth and has no characteristic points. Let χCc(M)\chi\in C_{c}^{\infty}(M) and define the weighted relative heat content

HΩχ(t)=Ωu(t,x)χ(x)𝑑ω(x),t>0.H_{\Omega}^{\chi}(t)=\int_{\Omega}u(t,x)\chi(x)d\omega(x),\qquad\forall\,t>0. (166)

Then, as t0t\to 0,

HΩχ(t)=Ωχ𝑑ω1πΩχ𝑑σt1/212Ωg(χ,δ)𝑑σt(112πΩ(4Δ+N2)χ𝑑σ16πΩ(Nχ)Δδ𝑑σ)t3/212Ωg((Δχ),δ)𝑑σt2+o(t2).\begin{split}H_{\Omega}^{\chi}(t)&=\int_{\Omega}\chi\hskip 0.59998ptd\omega-\frac{1}{\sqrt{\pi}}\int_{\partial\Omega}\chi\hskip 0.59998ptd\sigma\,t^{1/2}-\frac{1}{2}\int_{\partial\Omega}g(\nabla\chi,\nabla\delta)d\sigma\,t\\ &\quad-\left(\frac{1}{12\sqrt{\pi}}\int_{\partial\Omega}(4\Delta+N^{2})\chi\hskip 0.59998ptd\sigma-\frac{1}{6\sqrt{\pi}}\int_{\partial\Omega}(N\chi)\Delta\delta\hskip 0.59998ptd\sigma\right)t^{3/2}\\ &\quad-\frac{1}{2}\int_{\partial\Omega}g(\nabla(\Delta\chi),\nabla\delta)d\sigma\,t^{2}+o(t^{2}).\end{split} (167)
Proof.

Let us consider a cutoff function ϕ\phi as in (104). Then, applying the usual localization argument, cf. (107), we have:

Ωχ(x)𝑑ω(x)HΩχ(t)=I[ϕχ](t,0)+O(t),as t0,\int_{\Omega}\chi(x)d\omega(x)-H_{\Omega}^{\chi}(t)=I[\phi\chi](t,0)+O(t^{\infty}),\qquad\text{as }t\to 0, (168)

where now the function ϕχCc(Ωr0r0)\phi\chi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}) and ϕχ=χ\phi\chi=\chi close to Ω\partial\Omega.

As we did in the proof of Theorem 5.6, we relate HΩχ(t)H_{\Omega}^{\chi}(t) with the sum of contributions. Applying Proposition 5.3, we have the following asymptotic relation at order 44:

I[ϕχ](t,0)Ic[ϕχ](t,0)=Ωg(χ,δ)𝑑σt+Ωg((Δχ),δ)𝑑σt2+o(t2),I[\phi\chi](t,0)-I^{c}[\phi\chi](t,0)=\int_{\partial\Omega}g(\nabla\chi,\nabla\delta)d\sigma\,t+\int_{\partial\Omega}g(\nabla(\Delta\chi),\nabla\delta)d\sigma\,t^{2}+o(t^{2}), (169)

as t0t\to 0, having used the fact that ϕχχ\phi\chi\equiv\chi close to Ω\partial\Omega. Notice that this relation coincides with (159) when χ1\chi\equiv 1 close to Ω\partial\Omega. Thus, we obtain

Ωχ(x)𝑑ω(x)HΩχ(t)=12(I[ϕχ](t,0)+Ic[ϕχ](t,0))+Ωg(χ,δ)𝑑σt+Ωg((Δχ),δ)𝑑σt2+o(t2),as t0.\int_{\Omega}\chi(x)d\omega(x)-H_{\Omega}^{\chi}(t)=\frac{1}{2}\left(I[\phi\chi](t,0)+I^{c}[\phi\chi](t,0)\right)\\ +\int_{\partial\Omega}g(\nabla\chi,\nabla\delta)d\sigma\,t+\int_{\partial\Omega}g(\nabla(\Delta\chi),\nabla\delta)d\sigma\,t^{2}+o(t^{2}),\qquad\text{as }t\to 0. (170)

Finally, applying (164) for I[ϕχ](t,0)+Ic[ϕχ](t,0)I[\phi\chi](t,0)+I^{c}[\phi\chi](t,0), we conclude. ∎

Remark 5.8.

We compare the coefficients of the expansions of HΩ(t)H_{\Omega}(t) and QΩ(t)Q_{\Omega}(t), defined in (11), respectively. On the one hand, by [RR21, Thm. 5.8], the kk-th coefficient of the expansion of QΩ(t)Q_{\Omega}(t) is of the form

ΩDk(χ)𝑑σ,χCc(M),-\int_{\partial\Omega}D_{k}(\chi)d\sigma,\qquad\forall\chi\in C_{c}^{\infty}(M), (171)

where DkD_{k} is a differential operator acting on Cc(M)C_{c}^{\infty}(M) and belonging to span{Δ,N}\mathrm{span}_{\mathbb{R}}\{\Delta,N\} as algebra of operators. On the other hand, Proposition 5.7 shows that this is no longer true for the third coefficient of the expansion of HΩ(t)H_{\Omega}(t), as we need to add the operator multiplication by Δδ\Delta\delta.

6 An alternative approach using the heat kernel asymptotics

As we can see by a first application of Duhamel’s principle, cf. (16), and its iterations, the small-time asymptotics of u(t,)|Ωu(t,\cdot)\rvert_{\partial\Omega}, together with uniform estimates on the remainder with respect to xΩx\in\partial\Omega, would be enough to determine the asymptotic expansion of the relative heat content, at any order.

In Theorem 3.1, we studied the zero-order asymptotics of u(t,)|Ωu(t,\cdot)\rvert_{\partial\Omega}. The technique used for its proof does not work at higher-order, since the exponential remainder term in (84) would be unbounded as t0t\to 0. In this section, we comment how such an higher-order asymptotics of u(t,)|Ωu(t,\cdot)\rvert_{\partial\Omega} can be obtained exploiting the asymptotic formula for the heat kernel proved in [CdVHT21, Thm. A].

Let MM be a compact sub-Riemannian manifold and ΩM\Omega\subset M an open subset with smooth boundary. For xΩx\in\partial\Omega, let us consider ψ=(z1,,zn):UV\psi=(z_{1},\ldots,z_{n})\colon U\rightarrow V a chart of privileged coordinates centered at xx, with UU a relatively compact set. Since the heat kernel is exponentially decaying outside the diagonal, cf. (66),

u(t,x)=Ωpt(x,y)𝑑ω(y)=ΩUpt(x,y)𝑑ω(y)+O(t)=V1pt(0,z)𝑑ω(z)+O(t),\begin{split}u(t,x)&=\int_{\Omega}p_{t}(x,y)d\omega(y)=\int_{\Omega\cap U}p_{t}(x,y)d\omega(y)+O(t^{\infty})\\ &=\int_{V_{1}}p_{t}(0,z)d\omega(z)+O(t^{\infty}),\end{split} (172)

where V1=ψ(UΩ)V_{1}=\psi(U\cap\Omega), and we denote with the same symbols ω\omega and pt(0,z)p_{t}(0,z) the coordinate expression of the measure and heat kernel, respectively. For example, if xΩx\in\partial\Omega is non-characteristic, we may choose ψ\psi as in (63), then V1=V{z1>0}V_{1}=V\cap\{z_{1}>0\}. Recall the asymptotic expansion of the heat kernel of Theorem 2.9, evaluated in (0,z)(0,z): for any mm\in\mathbb{N} and compact set K(0,)×VK\subset(0,\infty)\times V,

|ε|𝒬pε2τ(0,δε(z))=p^τ(0,z)+i=0mεifi(τ,0,z)+o(|ε|m),as ε0,|\varepsilon|^{\mathcal{Q}}p_{\varepsilon^{2}\tau}(0,\delta_{\varepsilon}(z))=\hat{p}_{\tau}(0,z)+\sum_{i=0}^{m}\varepsilon^{i}f_{i}(\tau,0,z)+o(|\varepsilon|^{m}),\qquad\text{as }\varepsilon\to 0, (173)

uniformly as (τ,z)K(\tau,z)\in K, where 𝒬\mathcal{Q}, p^\hat{p} and fif_{i}’s are defined in Section 2. We will omit the dependance on the center of the privileged coordinates xx, it being fixed for the moment. At this point, we would like to integrate (173) to get information of u(t,x)u(t,x) as t0t\to 0. Proceeding formally, let us choose the parameters ε,τ\varepsilon,\tau in (173) such that:

ε2τ=t,ε=tα2α+1,τ=t12α+1,\varepsilon^{2}\tau=t,\qquad\varepsilon=t^{\frac{\alpha}{2\alpha+1}},\qquad\tau=t^{\frac{1}{2\alpha+1}}, (174)

for some α>0\alpha>0 to be fixed. For convenience of notation, set

Vs=δs(V1)s[1,1],V_{s}=\delta_{s}(V_{1})\qquad\forall\,s\in[-1,1], (175)

then, split the integral over V1V_{1} in (172) in two, so that the first one is computed on VεV_{\varepsilon} and the second one is computed on its complement in V1V_{1}, i.e. V1VεV_{1}\setminus V_{\varepsilon}. Notice that, by usual off-diagonal estimates, see [JSC86, Prop. 3] and our choice of the parameter ε\varepsilon as in (174), the following is a remainder term, independently of the value of α\alpha:

V1Vεpt(0,z)𝑑ω(z)=O(eβε2t)=O(t),as t0.\int_{V_{1}\setminus V_{\varepsilon}}p_{t}(0,z)d\omega(z)=O\left(e^{-\beta\frac{\varepsilon^{2}}{t}}\right)=O(t^{\infty}),\qquad\text{as }t\to 0. (176)

Thus, writing the measure in coordinates dω(z)=ω(z)dzd\omega(z)=\omega(z)dz with ω()C(V1)\omega(\cdot)\in C^{\infty}(V_{1}), we have, as t0t\to 0,

u(t,x)\displaystyle u(t,x) =Vεpt(0,z)ω(z)𝑑z+O(t)=V1εQpε2τ(0,δε(z))ω(δε(z))𝑑z+O(t)\displaystyle=\int_{V_{\varepsilon}}p_{t}(0,z)\omega(z)dz+O(t^{\infty})=\int_{V_{1}}\varepsilon^{Q}p_{\varepsilon^{2}\tau}(0,\delta_{\varepsilon}(z))\omega(\delta_{\varepsilon}(z))dz+O(t^{\infty}) (177)
=V1(p^τ(0,z)+i=0m1εifi(τ,0,z)+εmRm(ε,τ,z))ω(δε(z))𝑑z+O(t),\displaystyle=\int_{V_{1}}\left(\hat{p}_{\tau}(0,z)+\sum_{i=0}^{m-1}\varepsilon^{i}f_{i}(\tau,0,z)+\varepsilon^{m}R_{m}(\varepsilon,\tau,z)\right)\omega(\delta_{\varepsilon}(z))dz+O(t^{\infty}),\qquad (178)

where RmR_{m} is a smooth function on [1,1]×(0,)×n[-1,1]\times(0,\infty)\times\mathbb{R}^{n}, such that

supε[1,1]zK|Rm(ε,τ,z)|Cm(τ,K),\sup_{\begin{subarray}{c}\varepsilon\in[-1,1]\\ z\in K\end{subarray}}\left|R_{m}(\varepsilon,\tau,z)\right|\leq C_{m}(\tau,K), (179)

for any compact set KnK\subset\mathbb{R}^{n}, according to (173). Up to restricting the domain of privileged coordinates UU, we can assume that (179) holds on V¯\overline{V}. By our choices (174), we would like the following term

tmα2α+1V1|Rm(tα2α+1,t12α+1,z)|ω(δtα/(2α+1)(z))𝑑zt^{\frac{m\alpha}{2\alpha+1}}\int_{V_{1}}\left|R_{m}\left(t^{\frac{\alpha}{2\alpha+1}},t^{\frac{1}{2\alpha+1}},z\right)\right|\omega(\delta_{t^{\alpha/(2\alpha+1)}}(z))dz (180)

to be an error term of order greater than m12\frac{m-1}{2}, as t0t\to 0. Thus, assume for the moment that KV\forall K\subset V compact and m\forall m\in\mathbb{N}, =(m,K)\exists\ell=\ell(m,K)\in\mathbb{N} and Cm(K)>0C_{m}(K)>0 such that

supε[1,1]zK|Rm(ε,τ,z)|Cm(K)τ,τ(0,1).\sup_{\begin{subarray}{c}\varepsilon\in[-1,1]\\ z\in K\end{subarray}}\left|R_{m}(\varepsilon,\tau,z)\right|\leq\frac{C_{m}(K)}{\tau^{\ell}},\qquad\forall\,\tau\in(0,1). (𝐇\mathbf{H})

Thanks to assumption (𝐇\mathbf{H}), choosing α\alpha large enough, we see that (180) is a o(tm12)o(t^{\frac{m-1}{2}}). Performing the change of variables zδ1/τ(z)z\mapsto\delta_{1/\sqrt{\tau}}(z) in (178), and exploiting the homogeneity properties of p^\hat{p} and fif_{i}, namely (60), we finally obtain the following expression for uu as t0t\to 0:

u(t,x)=Vt1/(2(2α+1))(p^1(0,z)+i=0m1ti/2ai(z))ω(δt(z))𝑑z+o(tm12),u(t,x)=\int_{V_{t^{-1/(2(2\alpha+1))}}}\left(\hat{p}_{1}(0,z)+\sum_{i=0}^{m-1}t^{i/2}a_{i}(z)\right)\omega(\delta_{\sqrt{t}}(z))dz+o(t^{\frac{m-1}{2}}), (181)

having set ai(z)=fi(1,0,z)a_{i}(z)=f_{i}(1,0,z), for all ii\in\mathbb{N}. Therefore, we find an asymptotic expansion of u(t,x)u(t,x) under assumption (𝐇\mathbf{H}), which is crucial to overcome the fact that (173) is formulated on an asymptotic neighborhood of the diagonal, and not uniformly as τ0\tau\to 0. It is likely666Personal communication by Yves Colin de Verdière, Luc Hillairet and Emmanuel Trélat. that (𝐇\mathbf{H}) can be proven in the nilpotent case, and more generally when the ambient manifold is M=nM=\mathbb{R}^{n} and the generating family of the sub-Riemannian structure, {X1,,XN}\{X_{1},\ldots,X_{N}\} satisfies the uniform Hörmander polynomial condition, see [CdVHT21, App. B] for details. Although this strategy could be used to prove the existence of an asymptotic expansion of HΩ(t)H_{\Omega}(t), we refrain to go in this direction since two technical difficulties would arise nonetheless:

  • Uniformity of the expansion of u(t,x)u(t,x) with respect to xΩx\in\partial\Omega. In the non-equiregular case, cf. Section 2.3 for details, the expansion (173) is not uniform as xx varies in compact subsets of MM, hence the same would be true for the expansion (181).

  • Computations of the coefficients. The coefficients appearing in (181) depend on the nilpotent approximation at xΩx\in\partial\Omega and are not clearly related to the invariants of Ω\partial\Omega.

Our strategy avoids almost completely the knowledge of the small-time asymptotics of u(t,)Ωu(t,\cdot)_{\partial\Omega}, it being based on an asymptotic series of the auxiliary functional 𝒢u\mathcal{G}_{u}. Moreover, we stress that our method to prove the asymptotics of HΩ(t)H_{\Omega}(t) up to order 44, cf. Theorem 1.1, holds for any sub-Riemannian manifold, including also the non-equiregular ones.

Remark 6.1.

In order to pass from (181) to the asymptotic expansion of HΩ(t)H_{\Omega}(t), we would use Duhamel’s formula, which holds under the non-characteristic assumption. This means that, even though (173) of course is true even in presence of characteristic points, we can’t say much about the asymptotics of HΩ(t)H_{\Omega}(t) in the general case.

7 The non-compact case

In the non-compact case, we have the following difficulties:

  • The localization principle, cf. Proposition 4.5, may fail.

  • Set u(t,x)=etΔ𝟙Ω(x)u(t,x)=e^{t\Delta}\mathds{1}_{\Omega}(x) and uc(t,x)=etΔ𝟙Ωc(x)u^{c}(t,x)=e^{t\Delta}\mathds{1}_{\Omega^{c}}(x). If the manifold is not stochastically complete, the relation u(t,x)+uc(t,x)=1u(t,x)+u^{c}(t,x)=1 does not hold.

  • The Gaussian bounds for the heat kernel and its time-derivatives, à la Jerison and Sanchez-Calle [JSC86, Thm. 3], may not hold, thus Lemma A.3 may not be true.

Definition 7.1.

Let MM be a sub-Riemannian manifold, equipped with a smooth measure ω\omega. We say that (M,ω)(M,\omega) is (globally) doubling if there exist constants CD>0C_{D}>0 such that:

V(x,2ρ)CDV(x,ρ),ρ>0,xM,V(x,2\rho)\leq C_{D}V(x,\rho),\qquad\forall\,\rho>0,\ x\in M, (182)

where V(x,ρ)=ω(Bρ(x))V(x,\rho)=\omega(B_{\rho}(x)). We say that (M,ω)(M,\omega) satisfies a (global) weak Poincaré inequality, if there exist constants CP>0C_{P}>0 such that,

Bρ(x)|ffx,ρ|2𝑑ωCPρ2B2ρ(x)f2𝑑ω,ρ>0,xM,\int_{B_{\rho}(x)}\left|f-f_{x,\rho}\right|^{2}d\omega\leq C_{P}\rho^{2}\int_{B_{2\rho}(x)}\|\nabla f\|^{2}d\omega,\qquad\rho>0,\ x\in M, (183)

for any smooth function fC(M)f\in C^{\infty}(M). Here fx,ρ=1V(x,ρ)Bρ(x)f𝑑ωf_{x,\rho}=\frac{1}{V(x,\rho)}\int_{B_{\rho}(x)}fd\omega. We refer to these properties as local whenever they hold for any ρ<ρ0\rho<\rho_{0}.

Remark 7.2.

If MM is a sub-Riemannian manifold, equipped with a smooth globally doubling measure ω\omega, then it is stochastically complete, namely

Mpt(x,y)𝑑ω(y)=1,t>0,xM.\int_{M}p_{t}(x,y)d\omega(y)=1,\qquad\forall\,t>0,\ x\in M. (184)

This is a straightforward consequence of the characterization given by [Stu94, Thm. 4] on the volume growth of balls.

Theorem 7.3.

Let MM be a complete sub-Riemannian manifold, equipped with a smooth measure ω\omega. Assume that (M,ω)(M,\omega) is globally doubling and satisfies a global weak Poincaré inequality. Then, there exist constants Ck,ck>0C_{k},c_{k}>0, for any integer k0k\geq 0, depending only on CD,CPC_{D},C_{P}, such that, for any x,yMx,y\in M and t>0t>0,

|tkpt(x,y)|CktkV(x,t)exp(dSR2(x,y)ckt),|\partial_{t}^{k}p_{t}(x,y)|\leq\frac{C_{k}t^{-k}}{V(x,\sqrt{t})}\exp\left(-\frac{d_{\mathrm{SR}}^{2}(x,y)}{c_{k}t}\right), (185)

where we recall V(x,t)=ω(Bt(x))V(x,\sqrt{t})=\omega(B_{\sqrt{t}}(x)).

In addition, there exists constants C,c>0C_{\ell},c_{\ell}>0, depending only on CD,CPC_{D},C_{P}, such that, for any x,yMx,y\in M and t>0t>0,

pt(x,y)CV(x,t)exp(dSR2(x,y)ct).p_{t}(x,y)\geq\frac{C_{\ell}}{V(x,\sqrt{t})}\exp\left(-\frac{d_{\mathrm{SR}}^{2}(x,y)}{c_{\ell}t}\right). (186)
Proof.

Define the sub-Riemannian Hamiltonian as the smooth function H:TMH:T^{*}M\to\mathbb{R},

H(λ)=12i=1Nλ,Xi2,λTM,H(\lambda)=\frac{1}{2}\sum_{i=1}^{N}\langle\lambda,X_{i}\rangle^{2},\qquad\lambda\in T^{*}M, (187)

where {X1,,XN}\{X_{1},\ldots,X_{N}\} is a generating family for the sub-Riemannian structure. Then, following the notations of [Stu96], one can easily verify that

(u,v)=M2H(du,dv)𝑑ω,u,vCc(M),\mathcal{E}(u,v)=\int_{M}2H(du,dv)d\omega,\qquad\forall\,u,v\in C_{c}^{\infty}(M), (188)

where HH is the sub-Riemannian Hamiltonian viewed as a bilinear form on fibers, defines a strongly local Dirichlet form with domain dom()=Cc(M)\mathrm{dom}(\mathcal{E})=C_{c}^{\infty}(M). Notice that the Friedrichs extension of \mathcal{E} is exactly the sub-Laplacian, moreover, the intrinsic metric

dI(x,y)=sup{|u(x)u(y)| s.t. uCc(M),|2H(du,du)|1},x,yM.d_{I}(x,y)=\sup\{|u(x)-u(y)|\text{ s.t. }u\in C_{c}^{\infty}(M),\,|2H(du,du)|\leq 1\},\qquad\forall x,y\in M. (189)

coincides with the usual sub-Riemannian distance, as |2H(du,du)|=u2|2H(du,du)|=\|\nabla u\|^{2}, cf.
[BBS16, Ch. 2, Prop. 12.4]. Thus, \mathcal{E} is also strongly regular and, by our assumptions on (M,ω)(M,\omega), [SC92, Thm. 4.3] holds true, proving (185). For the Gaussian lower bound (186), it is enough to apply [Stu96, Cor. 4.10], cf. also [SC92, Thm. 4.2]. This concludes the proof. ∎

Remark 7.4.

Theorem 7.3 ensures that the time-derivatives of the heat kernel satisfy Gaussian bounds, which are sufficient to prove Lemma A.3 in the non-compact case. This lemma is crucial to obtain the asymptotics expansion of HΩ(t)H_{\Omega}(t) at order strictly greater than 11.

We prove now the non-compact analogue of Proposition 4.5.

Corollary 7.5.

Under the assumptions of Theorem 7.3, let ΩM\Omega\subset M be an open subset with smooth boundary. Then, for any KMK\subset M closed subset of MM such that KΩ=K\cap\partial\Omega=\emptyset, we have:

𝟙Ω(x)u(t,x)=O(t),as t0,uniformly for xK,\mathds{1}_{\Omega}(x)-u(t,x)=O(t^{\infty}),\qquad\text{as }t\to 0,\quad\text{uniformly for }x\in K, (190)

where u(t,x)=etΔ𝟙Ω(x)u(t,x)=e^{t\Delta}\mathds{1}_{\Omega}(x) is the solution to (36).

Proof.

Let us assume that KΩK\subset\Omega such that KΩ=K\cap\partial\Omega=\emptyset. The other part of the statement can be done similarly.

Since MM is stochastically complete, cf. Remark 7.2, for any xKx\in K, we can write:

𝟙Ω(x)u(t,x)=1etΔ𝟙Ω(x)=etΔ1(x)etΔ𝟙Ω(x)=MΩpt(x,y)𝑑ω(y).\mathds{1}_{\Omega}(x)-u(t,x)=1-e^{t\Delta}\mathds{1}_{\Omega}(x)=e^{t\Delta}1(x)-e^{t\Delta}\mathds{1}_{\Omega}(x)=\int_{M\setminus\Omega}p_{t}(x,y)d\omega(y). (191)

Thanks to Theorem 7.3, we can apply (185) for k=0k=0 obtaining

MΩpt(x,y)𝑑ω(y)MΩC0V(x,t)exp(dSR2(x,y)c0t)𝑑ω(y),\int_{M\setminus\Omega}p_{t}(x,y)d\omega(y)\leq\int_{M\setminus\Omega}\frac{C_{0}}{V(x,\sqrt{t})}\exp\left(-\frac{d_{\mathrm{SR}}^{2}(x,y)}{c_{0}t}\right)d\omega(y), (192)

for suitable constants C0,c0>0C_{0},c_{0}>0 not depending on x,yMx,y\in M, t>0t>0. Now, fix L>1L>1: since KΩK\subset\Omega is closed with empty intersection with Ω\partial\Omega, and thus well-separated from Ω\partial\Omega, we deduce there exists a=a(K)>0a=a(K)>0 such that dSR(x,y)>ad_{\mathrm{SR}}(x,y)>a for any xK,yMΩx\in K,\,y\in M\setminus\Omega, and so

MΩpt(x,y)𝑑ω(y)\displaystyle\int_{M\setminus\Omega}p_{t}(x,y)d\omega(y) MΩC0V(x,t)exp(dSR2(x,y)c0t)𝑑ω(y)\displaystyle\leq\int_{M\setminus\Omega}\frac{C_{0}}{V(x,\sqrt{t})}\exp\left(-\frac{d_{\mathrm{SR}}^{2}(x,y)}{c_{0}t}\right)d\omega(y) (193)
exp(C(a,L)c0t)MΩC0V(x,t)exp(dSR2(x,y)2Lc0t)𝑑ω(y),\displaystyle\leq\exp\left(-\frac{C(a,L)}{c_{0}t}\right)\int_{M\setminus\Omega}\frac{C_{0}}{V(x,\sqrt{t})}\exp\left(-\frac{d_{\mathrm{SR}}^{2}(x,y)}{2^{L}c_{0}t}\right)d\omega(y),\qquad (194)

where C(a,L)=a2(2L1)2L>0C(a,L)=\frac{a^{2}(2^{L}-1)}{2^{L}}>0. Thus, if we prove that the integral in (194) is finite, we conclude. Firstly, recall the Gaussian lower bound (186), which holds thanks to Theorem 7.3:

pt(x,y)CV(x,t)exp(dSR2(x,y)ct).p_{t}(x,y)\geq\frac{C_{\ell}}{V(x,\sqrt{t})}\exp\left(-\frac{d_{\mathrm{SR}}^{2}(x,y)}{c_{\ell}t}\right). (195)

for suitable constants constants C,c>0C_{\ell},c_{\ell}>0, not depending on x,yMx,y\in M, t>0t>0. Secondly, by the doubling property of ω\omega, it is well-known that there exists CD,s>0C_{D}^{\prime},s>0 depending only on CDC_{D} such that

V(x,R)CD(Rρ)sV(x,ρ),ρR.V(x,R)\leq C_{D}^{\prime}\left(\frac{R}{\rho}\right)^{s}V(x,\rho),\qquad\forall\,\rho\leq R. (196)

Therefore, choosing L>1L>1 so big that c~2=(2Lc0)/c>1\tilde{c}^{2}=(2^{L}c_{0})/c_{\ell}>1 and applying (196) for ρ=t\rho=\sqrt{t} and R=c~tR=\tilde{c}\sqrt{t}, we have R>ρR>\rho and

V(x,c~t)C~V(x,t),t>0,V\left(x,\tilde{c}\sqrt{t}\right)\leq\tilde{C}V(x,\sqrt{t}),\qquad\forall\,t>0, (197)

having denoted by C~=CDc~s>0\tilde{C}=C_{D}^{\prime}\tilde{c}^{s}>0. Finally, using (197) and the Gaussian lower bound (195), we can estimate the integral in (194) as follows:

MΩ1V(x,t)exp(dSR2(x,y)2Lc0t)𝑑ω(y)MC~V(x,c~t)exp(dSR2(x,y)cc~t)𝑑ω(y)C~CMpt~(x,y)𝑑ω(y)C~C,\int_{M\setminus\Omega}\frac{1}{V(x,\sqrt{t})}\exp\left(-\frac{d_{\mathrm{SR}}^{2}(x,y)}{2^{L}c_{0}t}\right)d\omega(y)\\ \leq\int_{M}\frac{\tilde{C}}{V(x,\tilde{c}\sqrt{t})}\exp\left(-\frac{d_{\mathrm{SR}}^{2}(x,y)}{c_{\ell}\tilde{c}t}\right)d\omega(y)\leq\frac{\tilde{C}}{C_{\ell}}\int_{M}p_{\tilde{t}}(x,y)d\omega(y)\leq\frac{\tilde{C}}{C_{\ell}}, (198)

where t~=c~t\tilde{t}=\tilde{c}t. Since the resulting constant does not depend on xKx\in K, we conclude the proof. ∎

Using Corollary 7.5 and adopting the same strategy of the compact case, one can finally prove the following result.

Theorem 7.6.

Let MM be a complete sub-Riemannian manifold, equipped with a smooth measure ω\omega. Assume that (M,ω)(M,\omega) is globally doubling and satisfies a global weak Poincaré inequality. Let ΩM\Omega\subset M be an open and bounded subset whose boundary is smooth and has no characteristic points. Then, as t0t\to 0,

HΩ(t)=ω(Ω)1πσ(Ω)t1/2112πΩ(2g(δ,(Δδ))(Δδ)2)𝑑σt3/2+o(t2).H_{\Omega}(t)=\omega(\Omega)-\frac{1}{\sqrt{\pi}}\sigma(\partial\Omega)t^{1/2}-\frac{1}{12\sqrt{\pi}}\int_{\partial\Omega}\left(2g(\nabla\delta,\nabla(\Delta\delta))-(\Delta\delta)^{2}\right)d\sigma\,t^{3/2}+o(t^{2}). (199)
Remark 7.7.

Theorem 7.6 holds true also for the weighted relative heat content, cf. Section 5.3. In both cases, we do not know whether its assumptions are sharp in the non-compact case.

7.1 Notable examples

We list here some notable examples of sub-Riemannian manifolds satisfying the assumptions of Theorem 7.3. For these examples Theorem 7.6 is valid.

  • MM is a Lie group with polynomial volume growth, the distribution is generated by a family of left-invariant vector fields satisfying the Hörmander condition and ω\omega is the Haar measure. This family includes also Carnot groups. See for example [Var96, SC92, GS12].

  • M=nM=\mathbb{R}^{n}, equipped with a sub-Riemannian structure induced by a family of vector fields {Y1,,YN}\{Y_{1},\ldots,Y_{N}\} with bounded coefficients together with their derivatives, and satisfying the Hörmander condition. Under these assumptions, the Lebesgue measure is doubling, cf. [NSW85, Thm. 1], and the Poincaré inequality is verified in [Jer86]. We remark that these works provide the local properties of Definition 7.1, with constants depending only on the CkC^{k}-norms of the vector fields YiY_{i}, for i=1,,Ni=1,\ldots,N. Thus, if the CkC_{k}-norms are globally bounded, we obtain the corresponding global properties.

  • MM is a complete Riemannian manifold with metric gg, equipped with the Riemannian measure, and with non-negative Ricci curvature.

We mention that a Riemannian manifold MM with Ricci curvature bounded below by a negative constant satisfies only locally Definition 7.1, i.e. for some ρ0<\rho_{0}<\infty, depending on the Ricci bound. Nevertheless, we can prove Corollary 7.5 in this case, as Li and Yau provides an upper Gaussian bound, see [LY86, Cor. 3.1], and a lower bound as (186) holds, cf.  [BQ99, Cor. 2]. Thus, the first-order asymptotic expansion of HΩ(t)H_{\Omega}(t), cf. Theorem 4.13, is valid in this setting.

Appendix A Iterated Duhamel’s principle for IΩϕ(t,0)I_{\Omega}\phi(t,0)

In this section, we study the iterated Duhamel’s principle for the IΩϕI_{\Omega}\phi, cf. Definition 4.7. The main result is Lemma A.6, which will imply formulas (129), (144) and (162).

The next proposition is a version of the iterated Duhamel’s principle taken from [RR21, Prop. A.1], which we recall here.

Proposition A.1.

Let FC((0,)×[0,+))F\in C^{\infty}((0,\infty)\times[0,+\infty)) be a smooth function compactly supported in the second variable and let L=tr2L=\partial_{t}-\partial_{r}^{2}. Assume that the following conditions hold:

  • (i)

    LkF(0,r)=limt0LkF(t,r)\displaystyle L^{k}F(0,r)=\lim_{t\to 0}L^{k}F(t,r) exists in the sense of distributions777Namely, for any ψC([0,))\psi\in C^{\infty}([0,\infty)), there exists finite limt00f(t,r)ψ(r)𝑑r\lim_{t\to 0}\int_{0}^{\infty}f(t,r)\psi(r)dr. With a slight abuse of notation, we define 0f(0,r)ψ(r)𝑑r=limt00f(t,r)ψ(r)𝑑r\int_{0}^{\infty}f(0,r)\psi(r)dr=\lim_{t\to 0}\int_{0}^{\infty}f(t,r)\psi(r)dr. for any k0k\geq 0;

  • (ii)

    LkF(t,0)L^{k}F(t,0) and rLkF(t,0)\partial_{r}L^{k}F(t,0) converge to a finite limit as t0t\to 0, for any k0k\geq 0.

Then, for all mm\in\mathbb{N} and t>0t>0, we have

F(t,0)=k=0m(tkk!0e(t,r,0)LkF(0,r)𝑑r1πk!0trLkF(τ,0)(tτ)k1/2dτ)+1m!0t0e(tτ,r,0)Lm+1F(τ,r)(tτ)m𝑑r𝑑τ,F(t,0)=\sum_{k=0}^{m}\left(\frac{t^{k}}{k!}\int_{0}^{\infty}e(t,r,0)L^{k}F(0,r)dr-\frac{1}{\sqrt{\pi}k!}\int_{0}^{t}\partial_{r}L^{k}F(\tau,0)(t-\tau)^{k-1/2}d\tau\right)\\ +\frac{1}{m!}\int_{0}^{t}\int_{0}^{\infty}e(t-\tau,r,0)L^{m+1}F(\tau,r)(t-\tau)^{m}dr\hskip 0.59998ptd\tau,\qquad\ (200)

where e(t,r,s)e(t,r,s) is the Neumann heat kernel on the half-line, cf. (115).

We want to apply Proposition A.1 to the function IΩϕ(t,0)I_{\Omega}\phi(t,0), thus, we study in detail the operators LkIΩL^{k}I_{\Omega}, for any k1k\geq 1. Define iteratively the family of matrices of operators, acting on smooth functions:

Mkj=(QkjSkjPkjRkj),M_{kj}=\begin{pmatrix}Q_{kj}&S_{kj}\\ P_{kj}&R_{kj}\end{pmatrix}, (201)

as follows. Set

M10=(ΔΔNΩNΩNΩ2+Δ)andM11=(0NΩ00),M_{10}=\begin{pmatrix}\Delta&-\Delta N_{\Omega}\\ N_{\Omega}&-N_{\Omega}^{2}+\Delta\end{pmatrix}\qquad\text{and}\qquad M_{11}=\begin{pmatrix}0&N_{\Omega}\\ 0&0\end{pmatrix}, (202)

and, for all k1k\geq 1 and 0jk0\leq j\leq k, set

Mkj=M10Mk1,j+M11Mk1,j1,M_{kj}=M_{10}M_{k-1,j}+M_{11}M_{k-1,j-1}, (203)

while Mkj=0M_{kj}=0, for all other values of the indices, i.e. k<0k<0, j<0j<0 or k<jk<j. Here NΩN_{\Omega} is the operator defined in (111), namely

NΩϕ=2g(ϕ,ν)+ϕdivω(ν),ϕCc(Ωr0r0),N_{\Omega}\phi=2g\left(\nabla\phi,\nu\right)+\phi\,\mathrm{div}_{\omega}(\nu),\qquad\,\forall\,\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}), (204)

where ν\nu is the inward-pointing normal from Ω\Omega.

Recall the definition of IΩI_{\Omega} and ΛΩ\Lambda_{\Omega}: for any ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}) and for all t>0,r0t>0,r\geq 0,

IΩϕ(t,r)\displaystyle I_{\Omega}\phi(t,r) =Ωr(1u(t,x))ϕ(x)𝑑ω(x),\displaystyle=\int_{\Omega_{r}}{\left(1-u(t,x)\right)\phi(x)d\omega(x)}, (205)
ΛΩϕ(t,r)\displaystyle\Lambda_{\Omega}\phi(t,r) =rIΩϕ(t,r)=Ωr(1u(t,y))ϕ(y)𝑑σ(y),\displaystyle=-\partial_{r}I_{\Omega}\phi(t,r)=-\int_{\partial\Omega_{r}}{\left(1-u(t,y)\right)\phi(y)d\sigma(y)}, (206)

where u(t,)=etΔ𝟙Ω()u(t,\cdot)=e^{t\Delta}\mathds{1}_{\Omega}(\cdot). Iterations of LkIΩϕL^{k}I_{\Omega}\phi satisfy the following lemma.

Lemma A.2.

Let MM be a sub-Riemannian manifold, equipped with a smooth measure ω\omega, and let ΩM\Omega\subset M be an open relatively compact subset whose boundary is smooth and has no characteristic points. Then, as operators on Cc(Ωr0r0)C^{\infty}_{c}(\Omega_{-r_{0}}^{r_{0}}), we have:

  • (i)

    LIΩ=IΩΔ+ΛΩNΩLI_{\Omega}=I_{\Omega}\Delta+\Lambda_{\Omega}N_{\Omega};

  • (ii)

    LΛΩ=ΛΩ(NΩ2+Δ)+tIΩNΩIΩΔNΩL\Lambda_{\Omega}=\Lambda_{\Omega}\left(-N_{\Omega}^{2}+\Delta\right)+\partial_{t}I_{\Omega}N_{\Omega}-I_{\Omega}\Delta N_{\Omega};

  • (iii)

    For any kk\in\mathbb{N},

    LkIΩ=j=0kjtj(ΛΩPkj+IΩQkj)andLkΛΩ=j=0kjtj(ΛΩRkj+IΩSkj).L^{k}I_{\Omega}=\sum_{j=0}^{k}{\frac{\partial^{j}}{\partial t^{j}}(\Lambda_{\Omega}P_{kj}+I_{\Omega}Q_{kj})}\qquad\text{and}\qquad L^{k}\Lambda_{\Omega}=\sum_{j=0}^{k}{\frac{\partial^{j}}{\partial t^{j}}(\Lambda_{\Omega}R_{kj}+I_{\Omega}S_{kj})}. (207)

Here we mean that, for any ϕCc(Ωr0r0)\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}), the operator LkL^{k} acts on the functions IΩϕ(t,r)I_{\Omega}\phi(t,r), ΛΩϕ(t,r)\Lambda_{\Omega}\phi(t,r). Analogously the right-hand side when evaluated in ϕ\phi is a function of (t,r)(t,r).

Proof.

The proof of items (i)(i) and (ii)(ii) follows from Proposition 4.2 and the divergence theorem, cf. [RR21, Lem. A.2]. We show how to recover the iterative law (203).

Consider the vector V=(IΩ,ΛΩ)V=\left(I_{\Omega},\Lambda_{\Omega}\right), then by items (i)(i) and (ii)(ii), we have

LV=(LIΩ,LΛΩ)=VM10+tVM11.LV=\left(LI_{\Omega},L\Lambda_{\Omega}\right)=VM_{10}+\partial_{t}VM_{11}. (208)

Notice that the operator LkL^{k} contains at most kk derivatives with respect to tt, therefore we have

LkV=j=0ktj(VMkj),k0,L^{k}V=\sum_{j=0}^{k}\partial_{t}^{j}\left(VM_{kj}\right),\qquad\forall\,k\geq 0, (209)

On the other hand, we can evaluate LkVL^{k}V, using (208):

LkV\displaystyle L^{k}V =L(Lk1V)=j=0k1Ltj(VMk1,j)=j=0k1tj(LVMk1,j)\displaystyle=L\left(L^{k-1}V\right)=\sum_{j=0}^{k-1}L\partial_{t}^{j}\left(VM_{k-1,j}\right)=\sum_{j=0}^{k-1}\partial_{t}^{j}\left(LVM_{k-1,j}\right) (210)
=j=0k1tjVM10Mk1,j+j=0k1tj+1VM11Mk1,j.\displaystyle=\sum_{j=0}^{k-1}\partial_{t}^{j}VM_{10}M_{k-1,j}+\sum_{j=0}^{k-1}\partial_{t}^{j+1}VM_{11}M_{k-1,j}. (211)

Reorganizing the sum, we find (203), concluding the proof. ∎

We want to apply Proposition A.1 to IΩϕ(t,r)I_{\Omega}\phi(t,r) for k2k\geq 2, in order to obtain higher-order asymptotics. However, Lemma A.2 shows that LkIΩL^{k}I_{\Omega}, for k2k\geq 2, involves time derivatives of u(t,x)u(t,x) which are not well-defined at Ω\partial\Omega as t0t\to 0. Therefore, we consider the following approximation of IΩϕI_{\Omega}\phi and ΛΩϕ\Lambda_{\Omega}\phi, respectively: fix ϵ>0\epsilon>0 and define, for any t>0,r0t>0,r\geq 0,

Iϵϕ(t,r)=\displaystyle I_{\epsilon}\phi(t,r)= =Ωr(1uϵ(t,x))ϕ(x)𝑑ω(x),\displaystyle=\int_{\Omega_{r}}\left(1-u_{\epsilon}(t,x)\right)\phi(x)d\omega(x), (212)
Λϵϕ(t,r)\displaystyle\Lambda_{\epsilon}\phi(t,r) =rIϵϕ(t,r)=Ωr(1uϵ(t,x))ϕ(y)𝑑σ(y),\displaystyle=-\partial_{r}I_{\epsilon}\phi(t,r)=\int_{\partial\Omega_{r}}\left(1-u_{\epsilon}(t,x)\right)\phi(y)d\sigma(y), (213)

where uϵ(t,x)=etΔ𝟙Ωϵ(x)u_{\epsilon}(t,x)=e^{t\Delta}\mathds{1}_{\Omega_{\epsilon}}(x). We recall that, for any aa\in\mathbb{R}, Ωa={xMδ(x)>a}\Omega_{a}=\{x\in M\mid\delta(x)>a\}. Notice that, by the dominated convergence theorem, we have

Iϵϕ(t,0)ϵ0IΩϕ(t,0),uniformly on [0,T],I_{\epsilon}\phi(t,0)\xrightarrow{\epsilon\to 0}I_{\Omega}\phi(t,0),\qquad\text{uniformly on }[0,T], (214)

and, in addition, Lemma A.2 holds unchanged also for IϵI_{\epsilon} and Λϵ\Lambda_{\epsilon}.

Lemma A.3.

Let MM be a compact sub-Riemannian manifold, equipped with a smooth measure ω\omega, and let ΩM\Omega\subset M be an open subset whose boundary is smooth and has no characteristic points. Let ψC([0,))\psi\in C^{\infty}([0,\infty)), ϵ(0,r0)\epsilon\in(0,r_{0}) and define

ψ(1)(r)=0rψ(s)𝑑s,r0.\psi^{(-1)}(r)=\int_{0}^{r}\psi(s)ds,\qquad\forall r\geq 0. (215)

Then, for any ϕCc(Ωr0r0)\phi\in C^{\infty}_{c}(\Omega_{-r_{0}}^{r_{0}}), the following identities hold:

  • (i)

    limt00jtjΛϵϕ(t,r)ψ(r)𝑑r={Ω0ϵϕ(ψδ)𝑑ω if j=0,ΩϵΔj(ϕ(ψδ))𝑑ω if j1;\displaystyle\lim_{t\to 0}\int_{0}^{\infty}{\frac{\partial^{j}}{\partial t^{j}}\Lambda_{\epsilon}\phi(t,r)\psi(r)dr}=\begin{cases}\displaystyle\int_{\Omega_{0}^{\epsilon}}{\phi(\psi\circ\delta)d\omega}&\text{ if }j=0,\\[10.0pt] \displaystyle-\int_{\Omega_{\epsilon}}{\Delta^{j}(\phi(\psi\circ\delta))d\omega}&\text{ if }j\geq 1;\end{cases}

  • (ii)

    limt00jtjIϵϕ(t,r)ψ(r)𝑑r={Ω0ϵϕ(ψ(1)δ)𝑑ω if j=0,ΩϵΔj(ϕ(ψ(1)δ))𝑑ω if j1;\displaystyle\lim_{t\to 0}\int_{0}^{\infty}{\frac{\partial^{j}}{\partial t^{j}}I_{\epsilon}\phi(t,r)\psi(r)dr}=\begin{cases}\displaystyle\int_{\Omega_{0}^{\epsilon}}{\phi\left(\psi^{(-1)}\circ\delta\right)d\omega}&\text{ if }j=0,\\[10.0pt] \displaystyle-\int_{\Omega_{\epsilon}}{\Delta^{j}\left(\phi\left(\psi^{(-1)}\circ\delta\right)\right)d\omega}&\text{ if }j\geq 1;\end{cases}

  • (iii)

    jtjΛϵϕ(0,0)={Ωϕ𝑑σ if j=0,0 if j1;\displaystyle\frac{\partial^{j}}{\partial t^{j}}\Lambda_{\epsilon}\phi(0,0)=\begin{cases}\displaystyle\int_{\partial\Omega}{\phi\hskip 0.59998ptd\sigma}&\text{ if }j=0,\\[10.0pt] \displaystyle 0&\text{ if }j\geq 1;\end{cases}

  • (iv)

    jtjIϵϕ(0,0)={Ω0ϵϕ𝑑ω if j=0,ΩϵΔjϕ𝑑ω if j1;\displaystyle\frac{\partial^{j}}{\partial t^{j}}I_{\epsilon}\phi(0,0)=\begin{cases}\displaystyle\int_{\Omega_{0}^{\epsilon}}\phi\hskip 0.59998ptd\omega&\text{ if }j=0,\\[10.0pt] -\int_{\Omega_{\epsilon}}{\Delta^{j}\phi\hskip 0.59998ptd\omega}&\text{ if }j\geq 1;\end{cases}

where, we recall, Ωϵ={xMδ(x)>ϵ}\Omega_{\epsilon}=\{x\in M\mid\delta(x)>\epsilon\} and Ω0ϵ=ΩΩϵ\Omega_{0}^{\epsilon}=\Omega\setminus\Omega_{\epsilon}.

Remark A.4.

The only difference with respect to [RR21, Lem. A.4] is item (iii)(iii), which now holds only as t0t\to 0 and not for all positive times.

Proof of Lemma A.3.

We claim that, for any j1j\geq 1,

limt0Ωϕ(x)Δjuϵ(t,x)𝑑ω(x)=ΩϵΔjϕ(x)𝑑ω(x).\lim_{t\to 0}\int_{\Omega}\phi(x)\Delta^{j}u_{\epsilon}(t,x)d\omega(x)=\int_{\Omega_{\epsilon}}\Delta^{j}\phi(x)d\omega(x). (216)

Let us prove it by induction: for j=1j=1, applying the divergence theorem, we have

ΩϕΔuϵ𝑑ω=Ωϕg(uϵ,δ)𝑑σ+Ωuϵg(ϕ,δ)𝑑σ+ΩuϵΔϕ𝑑ω.\int_{\Omega}\phi\Delta u_{\epsilon}\hskip 0.59998ptd\omega=-\int_{\partial\Omega}\phi g\left(\nabla u_{\epsilon},\nabla\delta\right)d\sigma+\int_{\partial\Omega}u_{\epsilon}g\left(\nabla\phi,\nabla\delta\right)d\sigma+\int_{\Omega}u_{\epsilon}\Delta\phi\hskip 0.59998ptd\omega. (217)

Let us discuss the first term in (217): by divergence theorem applied with respect to the set Ωc\Omega^{c}, we have

Ωϕg(uϵ,δ)𝑑σ=ΩcϕΔuϵ𝑑ω+Ωuϵg(ϕ,δ)𝑑σΩcuϵΔϕ𝑑ω,\int_{\partial\Omega}\phi g\left(\nabla u_{\epsilon},\nabla\delta\right)d\sigma=\int_{\Omega^{c}}\phi\Delta u_{\epsilon}\hskip 0.59998ptd\omega+\int_{\partial\Omega}u_{\epsilon}g\left(\nabla\phi,\nabla\delta\right)d\sigma-\int_{\Omega^{c}}u_{\epsilon}\Delta\phi\hskip 0.59998ptd\omega, (218)

then, using [JSC86, Thm. 3] and noticing that dSR(x,y)ϵd_{\mathrm{SR}}(x,y)\geq\epsilon, for any xΩϵx\in\Omega_{\epsilon} and yΩcy\in\Omega^{c}, we conclude that in the limit as t0t\to 0, (218) converges to 0. This proves (216), for j=1j=1. For j>1j>1, proceeding by induction, we conclude. Finally, using the co-area formula (94), we complete the proof of the statement as in the usual argument of [Sav98, Lem. 5.6]. ∎

Remark A.5.

In the non-compact case, under the assumption of Theorem 7.3, the above lemma holds. In particular, on the one hand, the divergence theorem holds since ϕ\phi has compact support. On the other hand, notice that the time derivative estimates (185) are enough to ensure that (218) converges to 0 as t0t\to 0, regardless of the compactness of the set of integration. The same is true for j>1j>1, where higher-order time derivatives appear.

The next step is to apply the iterated Duhamel’s principle (200) to IϵI_{\epsilon}, which now satisfies its assumptions, then, pass to the limit as ϵ0\epsilon\to 0. The computations are long but straightforward: we report here the result at order t5/2t^{5/2}.

Lemma A.6.

Under the same assumptions of Lemma A.3, let ϕCc(Ωr0r0)\phi\in C^{\infty}_{c}(\Omega_{-r_{0}}^{r_{0}}). Then, as t0t\to 0, we have:

IΩϕ(t,0)\displaystyle I_{\Omega}\phi(t,0) =2𝒢1u[ϕ](t)+1π0t𝒢1u[NΩϕ](τ)(tτ)1/2𝑑τ\displaystyle=2\mathcal{G}_{1-u}[\phi](t)+\frac{1}{\sqrt{\pi}}\int_{0}^{t}\mathcal{G}_{1-u}[N_{\Omega}\phi](\tau)(t-\tau)^{-1/2}d\tau (219)
+12π0t0τ𝒢1u[NΩ2ϕ](τ^)((ττ^)(tτ))1/2𝑑τ^𝑑τ\displaystyle\quad+\frac{1}{2\pi}\int_{0}^{t}\int_{0}^{\tau}\mathcal{G}_{1-u}[N_{\Omega}^{2}\phi](\hat{\tau})\left((\tau-\hat{\tau})(t-\tau)\right)^{-1/2}d\hat{\tau}\hskip 0.59998ptd\tau (220)
+14π3/20t0τ0τ^𝒢1u[NΩ3ϕ](s)((τ^s)(ττ^)(tτ))1/2𝑑s𝑑τ^𝑑τ\displaystyle\quad+\frac{1}{4\pi^{3/2}}\int_{0}^{t}\int_{0}^{\tau}\int_{0}^{\hat{\tau}}\mathcal{G}_{1-u}[N_{\Omega}^{3}\phi](s)\left((\hat{\tau}-s)(\tau-\hat{\tau})(t-\tau)\right)^{-1/2}ds\hskip 0.59998ptd\hat{\tau}\hskip 0.59998ptd\tau (221)
+14π0tΩ(1u(τ,))(4ΔNΩ2)ϕ𝑑σ(tτ)1/2𝑑τ\displaystyle\quad+\frac{1}{4\sqrt{\pi}}\int_{0}^{t}\int_{\partial\Omega}\left(1-u(\tau,\cdot)\right)(4\Delta-N_{\Omega}^{2})\phi\hskip 0.59998ptd\sigma(t-\tau)^{1/2}d\tau (222)
+14π0t𝒢1u[(6NΩΔNΩ32ΔNΩ)ϕ](τ)(tτ)1/2𝑑τ+O(t5/2),\displaystyle\quad+\frac{1}{4\sqrt{\pi}}\int_{0}^{t}\mathcal{G}_{1-u}[(6N_{\Omega}\Delta-N_{\Omega}^{3}-2\Delta N_{\Omega})\phi](\tau)(t-\tau)^{1/2}d\tau+O(t^{5/2}), (223)

where u(t,)=etΔ𝟙Ωu(t,\cdot)=e^{t\Delta}\mathds{1}_{\Omega} and 𝒢u[ϕ]\mathcal{G}_{u}[\phi] is the functional defined in (131). We recall that NΩN_{\Omega} is the operator defined in (204).

The expression (129) is a direct consequence of A.6. Moreover, we can apply it, when the base set is chosen to be Ωc\Omega^{c}. Then, evaluating the difference between IΩϕ(t,0)I_{\Omega}\phi(t,0) and IΩcϕ(t,0)I_{\Omega^{c}}\phi(t,0) we obtain the asymptotic equality (144), which is proved in the next lemma. We use the shorthands II, IcI^{c} for IΩI_{\Omega} and IΩcI_{\Omega^{c}} respectively.

Lemma A.7.

Under the same assumptions of Lemma A.3, let ϕCc(Ωr0r0)\phi\in C^{\infty}_{c}(\Omega_{-r_{0}}^{r_{0}}). Then, as t0t\to 0, we have:

(IϕIcϕ)(t,0)\displaystyle\left(I\phi-I^{c}\phi\right)(t,0) =2𝒢12u[ϕ](t)+12ΩNϕ𝑑σt\displaystyle=2\mathcal{G}_{1-2u}[\phi](t)+\frac{1}{2}\int_{\partial\Omega}N\phi\hskip 0.59998ptd\sigma\,t (224)
+12π0t0τ𝒢12u[N2ϕ](τ^)((ττ^)(tτ))1/2𝑑τ^𝑑τ\displaystyle\quad+\frac{1}{2\pi}\int_{0}^{t}\int_{0}^{\tau}\mathcal{G}_{1-2u}[N^{2}\phi](\hat{\tau})\left((\tau-\hat{\tau})(t-\tau)\right)^{-1/2}d\hat{\tau}\hskip 0.59998ptd\tau (225)
+14π0tΩ(12u(τ,))(4ΔN2)ϕ𝑑σ(tτ)1/2𝑑τ+O(t2),\displaystyle\quad+\frac{1}{4\sqrt{\pi}}\int_{0}^{t}\int_{\partial\Omega}\left(1-2u(\tau,\cdot)\right)(4\Delta-N^{2})\phi\hskip 0.59998ptd\sigma(t-\tau)^{1/2}d\tau+O(t^{2}), (226)

where NN is the operator given by

Nϕ=2g(ϕ,δ)+ϕΔδ,ϕCc(Ωr0r0),N\phi=2g(\nabla\phi,\nabla\delta)+\phi\Delta\delta,\qquad\forall\,\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}), (227)

with δ:M\delta\colon M\rightarrow\mathbb{R} the signed distance function from Ω\partial\Omega.

Proof.

Firstly, we apply Lemma A.6 to IϕI\phi: we obtain exactly the expression (219), with the operator NΩ=NN_{\Omega}=N. Secondly, for the outside contribution, recall that we have the following equality of smooth functions:

1uc(t,x)=1etΔ𝟙Ωc(x)=etΔ𝟙Ω(x)=u(t,x),t>0,xM.1-u^{c}(t,x)=1-e^{t\Delta}\mathds{1}_{\Omega^{c}}(x)=e^{t\Delta}\mathds{1}_{\Omega}(x)=u(t,x),\qquad\forall\,t>0,\ x\in M. (228)

Therefore, when we apply Lemma A.6 to IcϕI^{c}\phi, we replace 1uc(t,)=1etΔ𝟙Ωc1-u^{c}(t,\cdot)=1-e^{t\Delta}\mathds{1}_{\Omega^{c}} with the function u(t,)=etΔ𝟙Ω()u(t,\cdot)=e^{t\Delta}\mathds{1}_{\Omega}(\cdot). Moreover, the operator NΩcN_{\Omega^{c}} defined in (204), for the set Ωc\Omega^{c}, is equal to N-N, since the inward-pointing normal to Ωc\Omega^{c} is δ-\nabla\delta. Therefore, writing the difference of the two contributions, and noticing that Ω\Omega and its complement share the boundary, we have:

(IϕIcϕ)(t,0)\displaystyle\left(I\phi-I^{c}\phi\right)(t,0) =2𝒢12u[ϕ](t)+1π0t𝒢1[Nϕ](τ)(tτ)1/2𝑑τ\displaystyle=2\mathcal{G}_{1-2u}[\phi](t)+\frac{1}{\sqrt{\pi}}\int_{0}^{t}\mathcal{G}_{1}[N\phi](\tau)(t-\tau)^{-1/2}d\tau (229)
+12π0t0τ𝒢12u[N2ϕ](τ^)((ττ^)(tτ))1/2𝑑τ^𝑑τ\displaystyle\quad+\frac{1}{2\pi}\int_{0}^{t}\!\int_{0}^{\tau}\mathcal{G}_{1-2u}[N^{2}\phi](\hat{\tau})\left((\tau-\hat{\tau})(t-\tau)\right)^{-1/2}d\hat{\tau}\hskip 0.59998ptd\tau (230)
+14π3/20t0τ0τ^𝒢1[N3ϕ](s)((τ^s)(ττ^)(tτ))1/2𝑑s𝑑τ^𝑑τ\displaystyle\quad+\frac{1}{4\pi^{3/2}}\int_{0}^{t}\!\int_{0}^{\tau}\!\int_{0}^{\hat{\tau}}\mathcal{G}_{1}[N^{3}\phi](s)\left((\hat{\tau}-s)(\tau-\hat{\tau})(t-\tau)\right)^{-1/2}ds\hskip 0.59998ptd\hat{\tau}\hskip 0.59998ptd\tau\qquad (231)
+14π0tΩ(12u(τ,))(4ΔN2)ϕ𝑑σ(tτ)1/2𝑑τ\displaystyle\quad+\frac{1}{4\sqrt{\pi}}\int_{0}^{t}\!\int_{\partial\Omega}\left(1-2u(\tau,\cdot)\right)(4\Delta-N^{2})\phi\hskip 0.59998ptd\sigma(t-\tau)^{1/2}d\tau (232)
+14π0t𝒢1[(6NΔN32ΔN)ϕ](τ)(tτ)1/2𝑑τ+O(t5/2).\displaystyle\quad+\frac{1}{4\sqrt{\pi}}\int_{0}^{t}\mathcal{G}_{1}[(6N\Delta-N^{3}-2\Delta N)\phi](\tau)(t-\tau)^{1/2}d\tau+O(t^{5/2}). (233)

To conclude, it is enough to notice that the functional 𝒢1\mathcal{G}_{1} can be explicitly computed:

𝒢1[ϕ](t)=1πΩϕ𝑑σt1/2,ϕCc(Ωr0r0).\mathcal{G}_{1}[\phi](t)=\frac{1}{\sqrt{\pi}}\int_{\partial\Omega}\phi\hskip 0.59998ptd\sigma\,t^{1/2},\qquad\forall\,\phi\in C_{c}^{\infty}(\Omega_{-r_{0}}^{r_{0}}). (234)

Thus, the terms in (231) and (233) are remainder of order O(t2)O(t^{2}). ∎

Applying Lemma A.6 to the sum of IΩϕ(t,0)I_{\Omega}\phi(t,0) and IΩcϕ(t,0)I_{\Omega^{c}}\phi(t,0) instead, we obtain (162). The proof of this result is not provided here, as it is similar to the proof of Lemma A.7.

Lemma A.8.

Under the same assumptions of Lemma A.3, let ϕCc(Ωr0r0)\phi\in C^{\infty}_{c}(\Omega_{-r_{0}}^{r_{0}}). Then, as t0t\to 0, we have:

(Iϕ+Icϕ)(t,0)\displaystyle\left(I\phi+I^{c}\phi\right)(t,0) =2πΩϕ𝑑σt1/2+1π0t𝒢12u[Nϕ](τ)(tτ)1/2𝑑τ\displaystyle=\frac{2}{\sqrt{\pi}}\int_{\partial\Omega}\phi\hskip 0.59998ptd\sigma\,t^{1/2}+\frac{1}{\sqrt{\pi}}\int_{0}^{t}\mathcal{G}_{1-2u}[N\phi](\tau)(t-\tau)^{-1/2}d\tau (235)
+16πΩ(4Δ+N2)ϕ𝑑σt3/2\displaystyle\quad+\frac{1}{6\sqrt{\pi}}\int_{\partial\Omega}(4\Delta+N^{2})\phi\hskip 0.59998ptd\sigma\,t^{3/2}
+14π3/20t0τ0τ^𝒢12u[N3ϕ](s)((τ^s)(ττ^)(tτ))1/2𝑑s𝑑τ^𝑑τ\displaystyle\quad+\frac{1}{4\pi^{3/2}}\int_{0}^{t}\int_{0}^{\tau}\int_{0}^{\hat{\tau}}\mathcal{G}_{1-2u}[N^{3}\phi](s)\left((\hat{\tau}-s)(\tau-\hat{\tau})(t-\tau)\right)^{-1/2}ds\hskip 0.59998ptd\hat{\tau}\hskip 0.59998ptd\tau
+14π0t𝒢12u[(6NΔN32ΔN)ϕ](τ)(tτ)1/2𝑑τ+O(t5/2).\displaystyle\quad+\frac{1}{4\sqrt{\pi}}\int_{0}^{t}\mathcal{G}_{1-2u}[(6N\Delta-N^{3}-2\Delta N)\phi](\tau)(t-\tau)^{1/2}d\tau+O(t^{5/2}).

References

  • [ABB20] 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.
  • [AMM13] L. Angiuli, U. Massari, and M. Miranda, Jr. Geometric properties of the heat content. Manuscripta Math., 140(3-4):497–529, 2013.
  • [Bal03] Z. M. Balogh. Size of characteristic sets and functions with prescribed gradient. J. Reine Angew. Math., 564:63–83, 2003.
  • [BBS16] D. Barilari, U. Boscain, and M. Sigalotti, editors. Geometry, analysis and dynamics on sub-Riemannian manifolds. Vol. 1. EMS Series of Lectures in Mathematics. European Mathematical Society (EMS), Zürich, 2016. Lecture notes from the IHP Trimester held at the Institut Henri Poincaré, Paris and from the CIRM Summer School “Sub-Riemannian Manifolds: From Geodesics to Hypoelliptic Diffusion” held in Luminy, Fall 2014.
  • [Bel96] A. Bellaïche. The tangent space in sub-Riemannian geometry. In Sub-Riemannian geometry, volume 144 of Progr. Math., pages 1–78. Birkhäuser, Basel, 1996.
  • [BMP12] M. Bramanti, M. Miranda, Jr., and D. Pallara. Two characterization of BV functions on Carnot groups via the heat semigroup. Int. Math. Res. Not. IMRN, 2012(17):3845–3876, 2012.
  • [BQ99] D. Bakry and Z. M. Qian. Harnack inequalities on a manifold with positive or negative Ricci curvature. Rev. Mat. Iberoamericana, 15(1):143–179, 1999.
  • [CCSGM13] L. Capogna, G. Citti, and C. Senni Guidotti Magnani. Sub-Riemannian heat kernels and mean curvature flow of graphs. J. Funct. Anal., 264(8):1899–1928, 2013.
  • [CdVHT18] Y. Colin de Verdière, L. Hillairet, and E. Trélat. Spectral asymptotics for sub-Riemannian Laplacians, I: Quantum ergodicity and quantum limits in the 3-dimensional contact case. Duke Math. J., 167(1):109–174, 2018.
  • [CdVHT21] Y. Colin 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.
  • [DG54] E. De Giorgi. Su una teoria generale della misura (r1)(r-1)-dimensionale in uno spazio ad rr dimensioni. Ann. Mat. Pura Appl. (4), 36:191–213, 1954.
  • [DGN12] D. Danielli, N. Garofalo, and D. M. Nhieu. Integrability of the sub-Riemannian mean curvature of surfaces in the Heisenberg group. Proc. Amer. Math. Soc., 140(3):811–821, 2012.
  • [FPR20] V. Franceschi, D. Prandi, and L. Rizzi. On the essential self-adjointness of singular sub-Laplacians. Potential Anal., 53(1):89–112, 2020.
  • [Gri09] A. Grigor’yan. Heat kernel and analysis on manifolds, volume 47 of AMS/IP Studies in Advanced Mathematics. American Mathematical Society, Providence, RI; International Press, Boston, MA, 2009.
  • [Gro96] M. Gromov. Carnot-Carathéodory spaces seen from within. In Sub-Riemannian geometry, volume 144 of Progr. Math., pages 79–323. Birkhäuser, Basel, 1996.
  • [GS12] I. Gallagher and Y. Sire. Besov algebras on Lie groups of polynomial growth. Studia Math., 212(2):119–139, 2012.
  • [GT20] N. Garofalo and G. Tralli. A bourgain-brezis-mironescu-dávila theorem in carnot groups of step two. Communications in Analysis and Geometry (in press), 2020.
  • [Jea14] F. Jean. Control of nonholonomic systems: from sub-Riemannian geometry to motion planning. SpringerBriefs in Mathematics. Springer, Cham, 2014.
  • [Jer86] D. Jerison. The Poincaré inequality for vector fields satisfying Hörmander’s condition. Duke Math. J., 53(2):503–523, 1986.
  • [JSC86] D. S. Jerison and A. Sánchez-Calle. Estimates for the heat kernel for a sum of squares of vector fields. Indiana Univ. Math. J., 35(4):835–854, 1986.
  • [Led94] M. Ledoux. Semigroup proofs of the isoperimetric inequality in Euclidean and Gauss space. Bull. Sci. Math., 118(6):485–510, 1994.
  • [LY86] P. Li and S.-T. Yau. On the parabolic kernel of the Schrödinger operator. Acta Math., 156(3-4):153–201, 1986.
  • [Mit85] J. Mitchell. On Carnot-Carathéodory metrics. J. Differential Geom., 21(1):35–45, 1985.
  • [Mon02] R. Montgomery. A tour of subriemannian geometries, their geodesics and applications, volume 91 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2002.
  • [MPPP07] M. Miranda, Jr., D. Pallara, F. Paronetto, and M. Preunkert. Short-time heat flow and functions of bounded variation in 𝐑N\mathbf{R}^{N}. Ann. Fac. Sci. Toulouse Math. (6), 16(1):125–145, 2007.
  • [NSW85] A. Nagel, E. M. Stein, and S. Wainger. Balls and metrics defined by vector fields. I. Basic properties. Acta Math., 155(1-2):103–147, 1985.
  • [Rif14] L. Rifford. Sub-Riemannian geometry and optimal transport. SpringerBriefs in Mathematics. Springer, Cham, 2014.
  • [Ros21] T. Rossi. Heat content asymptotics in sub-Riemannian geometry. Phd Thesis. Université Grenoble Alpes & International School for Advanced Studies, 2021.
  • [Ros23] T. Rossi. Integrability of the sub-Riemannian mean curvature at degenerate characteristic points in the Heisenberg group. Adv. Calc. Var., 16(1):99–110, 2023.
  • [RR21] L. Rizzi and T. Rossi. Heat content asymptotics for sub-Riemannian manifolds. J. Math. Pures Appl. (9), 148:267–307, 2021.
  • [Sav98] A. Savo. Uniform estimates and the whole asymptotic series of the heat content on manifolds. Geom. Dedicata, 73(2):181–214, 1998.
  • [SC92] L. Saloff-Coste. A note on Poincaré, Sobolev, and Harnack inequalities. Internat. Math. Res. Notices, 1992(2):27–38, 1992.
  • [Str86] R. S. Strichartz. Sub-Riemannian geometry. J. Differential Geom., 24(2):221–263, 1986.
  • [Stu94] K.-T. Sturm. Analysis on local Dirichlet spaces. I. Recurrence, conservativeness and LpL^{p}-Liouville properties. J. Reine Angew. Math., 456:173–196, 1994.
  • [Stu96] K. T. Sturm. Analysis on local Dirichlet spaces. III. The parabolic Harnack inequality. J. Math. Pures Appl. (9), 75(3):273–297, 1996.
  • [Var96] N. T. Varopoulos. Analysis on Lie groups. Rev. Mat. Iberoamericana, 12(3):791–917, 1996.
  • [vdB13] M. van den Berg. Heat flow and perimeter in m\mathbb{R}^{m}. Potential Anal., 39(4):369–387, 2013.
  • [vdBG15] M. van den Berg and P. Gilkey. Heat flow out of a compact manifold. J. Geom. Anal., 25(3):1576–1601, 2015.