Hostname: page-component-7bb8b95d7b-fmk2r Total loading time: 0 Render date: 2024-09-27T02:08:52.997Z Has data issue: false hasContentIssue false

Heat kernel asymptotics for real powers of Laplacians

Published online by Cambridge University Press:  23 January 2023

Cipriana Anghel*
Affiliation:
Institute of Mathematics of the Romanian Academy, Bucharest, Romania
*
Rights & Permissions [Opens in a new window]

Abstract

We describe the small-time heat kernel asymptotics of real powers $\operatorname {\Delta }^r$, $r \in (0,1)$ of a non-negative self-adjoint generalized Laplacian $\operatorname {\Delta }$ acting on the sections of a Hermitian vector bundle $\mathcal {E}$ over a closed oriented manifold M. First, we treat separately the asymptotic on the diagonal of $M \times M$ and in a compact set away from it. Logarithmic terms appear only if n is odd and r is rational with even denominator. We prove the non-triviality of the coefficients appearing in the diagonal asymptotics, and also the non-locality of some of the coefficients. In the special case $r=1/2$, we give a simultaneous formula by proving that the heat kernel of $\operatorname {\Delta }^{1/2}$ is a polyhomogeneous conormal section in $\mathcal {E} \boxtimes \mathcal {E}^* $ on the standard blow-up space $\operatorname {M_{heat}}$ of the diagonal at time $t=0$ inside $[0,\infty )\times M \times M$.

Type
Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of The Canadian Mathematical Society

1 Introduction

Let $\operatorname {\Delta }$ be a self-adjoint generalized Laplacian acting on the sections of a Hermitian vector bundle $\mathcal {E}$ over an oriented, compact Riemannian manifold M of dimension n. Denote by $p_t$ the heat kernel of $\operatorname {\Delta }$ , i.e., the Schwartz kernel of the operator $e^{-t\operatorname {\Delta }}$ . It is known since Minakshisundaram–Pleijel [Reference Minakshisundaram and Pleijel21] that $p_t(x,y)$ has an asymptotic expansion as $t\searrow 0$ near the diagonal

(1.1) $$ \begin{align} p_t(x,y) \stackrel{t \searrow 0}{\sim} t^{-n/2} e^{-\frac{d(x,y)^2}{4t}} \sum_{j=0}^{\infty} t^j \Psi_j (x,y), \end{align} $$

where $d(x,y)$ is the geodesic distance between x and y, and the $\Psi _j$ ’s are recursively defined as solutions of certain ODE’s along geodesics (see, e.g., [Reference Berger, Gauduchon and Mazet4, Reference Berline, Getzler and Vergne5]). This asymptotic expansion applied to $\operatorname {D}^*\operatorname {D}$ , where $\operatorname {D}$ is a twisted Dirac operator, plays a leading role in the heat kernel proofs of the Atiyah–Singer index theorem (see [Reference Berline and Vergne6, Reference Bismut7, Reference Getzler12]).

Bär and Moroianu [Reference Bär and Moroianu2] studied the short-time asymptotic behavior of the heat kernel of $ \operatorname {\Delta }^{1/m}$ , $m \in {\mathbb N}^*$ , for a strictly positive self-adjoint generalized Laplacian $\operatorname {\Delta }$ . They give explicit asymptotic formulæ separately in the case when $t \searrow 0$ along the diagonal $\operatorname {Diag} \subset M \times M$ , and when t goes to $0$ in a compact set away from the diagonal. The asymptotic behavior depends on the parity of the dimension n and of the root m. More precisely, logarithmic terms appear when n is odd and m is even. They use the Legendre duplication formula, and the more general Gauss multiplication formula for the $\Gamma $ function (see, e.g., [Reference Paris and Kaminski22]). Another crucial argument in [Reference Bär and Moroianu2] is to use integration by parts in order to show that the Schwartz kernel $q_{-s}$ of the pseudodifferential operator $\operatorname {\Delta }^{-s}$ , $s \in {\mathbb C}$ , defines a meromorphic function when restricted to the diagonal in $M \times M$ .

1.1 Small-time heat asymptotic for real powers of $\operatorname {\Delta }$

The purpose of this paper is to study the short-time asymptotic of the Schwartz kernel $h_t$ of the operator $e^{-t\operatorname {\Delta }^r}$ , where $r \in (0,1)$ and $\operatorname {\Delta }$ is a non-negative self-adjoint generalized Laplacian, like, for instance, $\operatorname {\Delta }=\operatorname {D}^*\operatorname {D}$ for a Dirac operator $\operatorname {D}$ . We give separate formulæ as t goes to $0$ in $[0, \infty ) \times \operatorname {Diag}$ , and when $t \searrow 0$ in $[0,\infty ) \times K$ , where $K \subset M \times M$ is a compact set disjoint from the diagonal. In Theorem 6.1, we obtain that ${h_t}_{\vert [0,\infty ) \times K} \in t \cdot {\mathcal C}^{\infty } \left ( [0, \infty ) \times K \right )$ is a smooth function vanishing at least to order $1$ at $\{ t=0 \}$ . The asymptotic along the diagonal depends on the parity of n (like in [Reference Bär and Moroianu2]) and on the rationality of r. In Theorem 7.1, the most interesting case occurs when logarithmic terms appear. This happens only if n is odd, $r=\frac {\alpha }{\beta }$ is rational, and the denominator $\beta $ is even. In that case,

(1.2)

Similar expansions are proved in Theorem 7.1 in all the other cases. Furthermore, we prove the non-triviality of the coefficients appearing in the diagonal asymptotics (Theorem 1.1), and also the non-locality of some of them (Theorem 1.3).

In the special case $r=1/2$ , Bär and Moroianu [Reference Bär and Moroianu2] described the small-time asymptotic behavior of $h_t$ on the diagonal and away from it separately. In Theorem 1.4, we give an uniform description of the transition between the on- and off-diagonal behavior by proving that the heat kernel of $\operatorname {\Delta }^{1/2}$ is a polyhomogeneous conormal section in $\mathcal {E} \boxtimes \mathcal {E}^* $ on the standard blow-up space $[[0,\infty )\times M \times M, \{ t=0 \} \times \operatorname {Diag}]$ .

1.2 Comparison to previous results

Fahrenwaldt [Reference Fahrenwaldt11] studied the off-diagonal short-time asymptotics of the heat kernel of $e^{-t f(P)}$ , where $f: [0,\infty ) \longrightarrow [0,\infty )$ is a smooth function with certain properties, and P is a positive self-adjoint generalized Laplacian. The function $f(x)=x^r$ , $r \in (0,1)$ does not satisfy the third condition in [Reference Fahrenwaldt11, Hypothesis 3.3], which seems to be crucial for the arguments and statements in that paper, so the results of [Reference Fahrenwaldt11] do not seem to apply here.

Duistermaat and Guillemin [Reference Duistermaat and Guillemin10] give the asymptotic expansion of the heat kernel of $e^{-tP}$ , where P is a scalar positive elliptic self-adjoint pseudodifferential operator. The order of P in [Reference Duistermaat and Guillemin10] seems to be a positive integer. It is claimed in [Reference Agronovič1] that this asymptotic holds true in the context of fiber bundles. Furthermore, Grubb [Reference Grubb16, Theorem 4.2.2] studied the heat asymptotics for $e^{-tP}$ in the context of fiber bundles when the order of P is positive, not necessary an integer. In Theorem 7.1, we obtain the vanishing of some terms appearing in [Reference Grubb16, Corollary 4.2.7] in our particular case when $P=\operatorname {\Delta }^r$ is a real power of a self-adjoint non-negative generalized Laplacian $\operatorname {\Delta }$ , $r \in (0,1)$ . We also show that the remaining terms do not vanish in general.

Theorem 1.1 For each $r \in (0,1)$ , none of the coefficients in the small-time asymptotic expansion of $h_t$ appearing in Theorem 7.1 vanishes identically for every generalized Laplacian $\operatorname {\Delta }$ .

The logarithmic coefficients $B_l$ and the coefficients $A_j$ for $j \notin {\mathbb Z}$ can be computed in terms of the heat coefficients for $e^{-t\Delta }$ appearing in (1.1). It is well known that the heat coefficients of a generalized Laplacian are locally computable in terms of the curvature of the connection on $\mathcal {E}$ , the Riemannian metric of M and their derivatives (see, e.g., [Reference Berline, Getzler and Vergne5]). This is no longer the case for the coefficients of positive integer powers of t from Theorem 7.1 as we shall see now.

By applying Theorem 7.1 for $r \in (0,1)$ and a set of geometric data, namely a hermitic vector bundle $\mathcal {E}$ over an oriented, compact Riemannian manifold $(M,g)$ , a metric connection $\nabla $ and an endomorphism $F \in \operatorname {End} \mathcal {E}$ , $F^*=F$ , we produce an endomorphism $A_l \left ( M,g,\mathcal {E},h_{\mathcal {E}},\nabla ,F \right ) \in {\mathcal C}^{\infty } \left ( M, \operatorname {End} \mathcal {E} \right )$ for each index l appearing in (1.2).

Definition 1.1

  1. (i) We say that a function A which associates to any set of geometric data $(M,g,\mathcal {E},h_{\mathcal {E}},\nabla , F)$ a section in ${\mathcal C}^{\infty }(M,\operatorname {End} \mathcal {E})$ is locally computable if for any two sets of geometric data $(M,g,\mathcal {E},h_{\mathcal {E}},\nabla , F)$ , $(M',g',\mathcal {E}',h_{\mathcal {E}'}, \nabla ', F')$ which agree on an open set (i.e., there exist an isometry $\alpha : U \longrightarrow U'$ between two open sets $U \subset M$ , $U' \subset M'$ , and a metric isomorphism $\beta : \mathcal {E}_{\vert _U} \longrightarrow \mathcal {E}^{\prime }_{\vert _{U'}}$ which preserves the connection and $\beta _x \circ F_x \circ \beta _{\alpha (x)}^{-1}=F^{\prime }_{\alpha (x)}$ ), we have

    $$\begin{align*}\beta_x \circ A_x \circ \beta_{\alpha(x)}^{-1} = A_{\alpha(x)}, \end{align*}$$

    for any $x \in U$ .

  2. (ii) A scalar function a defined on the set of all geometric data $(M,g,\mathcal {E},h_{\mathcal {E}},\nabla ,F)$ with values in ${\mathbb C}$ is called locally computable if there exists a locally computable function C as in (i) above such that $a=\int _M \operatorname {Tr} C \operatorname {dvol}_g$ for any $\left ( M,g, \mathcal {E}, h_{\mathcal {E}}, \nabla , F \right )$ .

  3. (iii) A function A as in (i) is called cohomologically locally computable if there exists a locally computable function C as in (i) such that for any $\left ( M,g, \mathcal {E}, h_{\mathcal {E}}, \nabla , F \right )$ ,

    $$\begin{align*}\left[ \operatorname{Tr} A \operatorname{dvol}_g \right] = \left[ \operatorname{Tr} C \operatorname{dvol}_g \right] \in H^n_{dR} \left( M \right). \end{align*}$$

Remark 1.2

  1. (i) If a function A is locally computable, then the integral ${a:=\int _{M} \operatorname {Tr} A \operatorname {dvol}_g}$ is locally computable.

  2. (ii) A function A is cohomologically locally computable if and only if ${a:=\int _{M} \operatorname {Tr} A \operatorname {dvol}_g}$ is locally computable.

Theorem 1.3 If r is irrational, the heat coefficients $A_j$ in Theorem 7.1 (and in particular in ( 1.2 )) are not locally computable for integer $j \geq 1$ . If $r=\frac {\alpha }{\beta }$ is rational, then $A_j$ are not locally computable for $j \in {\mathbb N} \setminus \{ l \beta : l \in {\mathbb N} \}$ . All the other coefficients can be written in terms of the heat coefficients of $e^{-t\operatorname {\Delta }}$ , hence they are locally computable.

Consider the asymptotic expansion in [Reference Duistermaat and Guillemin10, Corollary 2.2’] for a scalar admissible operator, i.e., an elliptic, self-adjoint, positive pseudodifferential operator P of positive integer order d:

$$\begin{align*}e^{-tP} \stackrel{t \searrow 0}{\sim} \sum_{l=0}^{\infty} A_l(P) t^{(l-n)/d} + \sum_{k=1}^{\infty} B_k(P) t^k \log t. \end{align*}$$

Gilkey and Grubb [Reference Gilkey and Grubb14, Theorem 1.4] proved that the coefficients $a_l(P)$ for $l \geq 0$ and $b_k(P)$ for $k \geq 1$ from the corresponding small-time heat trace expansion

(1.3) $$ \begin{align} \operatorname{Tr} e^{-tP} \stackrel{t \searrow 0}{\sim} \sum_{l=0}^{\infty} a_l(P) t^{(l-n)/d} + \sum_{k=1}^{\infty} b_k(P) t^k \log t \end{align} $$

are generically non-zero in the above class of admissible operators. In Theorem 1.1, we prove the same type of statement. However, in our case, the order of the operator $\Delta ^r$ is $2r$ ; thus, it is integer only for $r = 1/2$ . Even in this case, the non-vanishing result in Theorem 1.1 is not a consequence of [Reference Gilkey and Grubb14, Theorem 1.4] since, in our case, we do not consider the whole class of admissible operators of fixed integer order d in the sense of Gilkey and Grubb [Reference Gilkey and Grubb14], but the smaller class of square roots of generalized Laplacians.

Furthermore, in [Reference Gilkey and Grubb14, Theorem 1.7], it is proved that the coefficients $a_l(P)$ in (1.3) corresponding to $t^{(l-n)/d}$ , for $(l-n)/d \in {\mathbb N}$ , are not locally computable. Remark that the meaning of “locally computable” in [Reference Gilkey and Grubb14] is different from our Definition 1.1. More precisely, in the definition of Gilkey and Grubb, a locally computable function A has to be a smooth function in the jets of the homogeneous components of the total symbol of the operator. A locally computable coefficient in the sense of Gilkey and Grubb [Reference Gilkey and Grubb14] is clearly locally computable in the sense of Definition 1.1(ii).

For $r=1/2$ , Bär and Moroianu [Reference Bär and Moroianu2] remark that for odd $k=1,3,\ldots $ , the coefficients $A_k$ in (1.2) corresponding to $t^k$ appear to be non-local. In Section 9, we clarify this remark by proving that they are indeed non-local in the sense of Definition 1.1 (i) (Theorem 1.3). In fact, we prove that the $A_k$ ’s are not cohomologically local. By Remark 1.2 (ii), it also follows that the integrals $a_k:=\int _M \operatorname {Tr} A_k \operatorname {dvol}_g$ are not locally computable in the sense of Definition 1.1 (ii). Therefore, the $a_k$ ’s for odd k are also not locally computable in the sense of Gilkey and Grubb [Reference Gilkey and Grubb14].

For $d=1$ , the non-local coefficients in the heat expansion (1.3) in [Reference Gilkey and Grubb14] are $a_{n+1}, a_{n+2},\ldots $ , whereas in our case corresponding to $r=d/2=1/2$ , the non-local coefficients are $a_1,a_3,\ldots $ . Despite some formal resemblances, it appears therefore that the results of the present paper are quite different from those of [Reference Gilkey and Grubb14].

1.3 The heat kernel as a conormal section

Recall that a smooth function f on the interior of a manifold with corners is said to be polyhomogeneous conormal if for any boundary hypersurface given by a boundary defining function $\theta $ , f has an expansion with terms of the form $\theta ^k \log ^l \theta $ toward ${\{ \theta =0 \}}$ (only natural powers l are allowed). In [Reference Melrose19], Melrose introduced the heat space $M_H^2$ by performing a parabolic blow-up of the diagonal in $M \times M$ at time $t=0$ . The new space is a manifold with corners with boundary hypersurfaces given by the boundary defining functions $\rho $ and $\omega _0$ . Then the heat kernel $p_t$ has the form $\rho ^{-n} {\mathcal C}^{\infty }(M_H^2)$ , and it vanishes rapidly at $\{ \omega _0=0 \}$ (see [Reference Melrose19, Theorem 7.12]).

In the special case $r=1/2$ , we are able to give a simultaneous formula for the asymptotic behavior of $h_t$ as t goes to zero both on the diagonal and away from it. We can understand better the heat operator $e^{-t \operatorname {\Delta }^{1/2}}$ on a homogeneous (rather than parabolic) blow-up heat space $\operatorname {M_{heat}}$ , the usual blow-up of $\{ 0 \} \times \operatorname {Diag}$ in $[0,\infty ) \times M \,{\times}\, M$ . The new added face is called the front face and we denote it $\operatorname {ff}$ , whereas the lift of the old boundary is the lateral boundary, denoted $\operatorname {lb}$ .

Theorem 1.4 If n is even, then the Schwartz kernel $h_t$ of the operator $e^{-t\operatorname {\Delta }^{1/2}}$ belongs to $ \rho ^{-n}\omega _0 \cdot {\mathcal C}^{\infty } (\operatorname {M_{heat}}) $ , while if n is odd, $h_t \in \rho ^{-n} \omega _0 \cdot {\mathcal C}^{\infty } (\operatorname {M_{heat}}) + \rho \log \rho \cdot \omega _0 \cdot {\mathcal C}^{\infty } (\operatorname {M_{heat}}) $ .

Theorem 1.4 improves the results of [Reference Bär and Moroianu2] twofold. First, it holds true for non-negative generalized Laplacians. Second, while Bär–Moroianu describe the asymptotic behavior of $h_t$ on the diagonal and away from it separately, this theorem also gives a precise, uniform description of the transition between these two regions by showing that $h_t$ is a polyhomogeneous conormal section on $\operatorname {M_{heat}}$ with values in $\mathcal {E} \boxtimes \mathcal {E}^*$ .

Note that throughout the paper, integral kernels act on sections by integration with respect to the fixed Riemannian density from M in the second variable, so $h_t$ does not contain a density factor. We feel that in the present context this exhibits more clearly the asymptotic behavior.

Based on the study of the case $r=1/2$ and on the separate asymptotic expansions of the heat kernel $h_t$ of $\operatorname {\Delta }^r$ , $r \in (0,1)$ as t goes to $0$ given in Theorems 6.1 and 7.1, we can conjecture that the heat kernel $h_t$ is a polyhomogeneous conormal function for all $r \in (0,1)$ on a “transcendental” heat blow-up space $M^r_{heat}$ depending on r. We leave this as a future project.

2 The heat kernel of a generalized Laplacian

Let $\mathcal {E}$ be a Hermitian vector bundle over a compact Riemannian manifold M of dimension n. Consider $\operatorname {\Delta }$ to be a generalized Laplacian, i.e., a second-order differential operator which satisfies

$$\begin{align*}\sigma_2(\operatorname{\Delta})(x,\xi)=|\xi|^2 \cdot \operatorname{id}_{\mathcal{E}}. \end{align*}$$

For example, if $\nabla $ is a connection on $\mathcal {E}$ and $F \in \Gamma (\operatorname {End} \mathcal {E})$ , $F^*=F$ , then $\nabla ^*\nabla +F$ is a symmetric generalized Laplacian on $\mathcal {E}$ .

Suppose that $\operatorname {\Delta }$ is self-adjoint. Since M is compact, the spectrum of $\operatorname {\Delta }$ is discrete and $L^2(M,\mathcal {E})$ splits as an orthogonal Hilbert direct sum

$$\begin{align*}L^2(M,\mathcal{E})=\bigoplus_{\lambda \in \operatorname{Spec} \operatorname{\Delta}}^{\perp} E_{\lambda}, \end{align*}$$

where $E_{\lambda }$ is the eigenspace corresponding to the eigenvalue $\lambda $ of $\operatorname {\Delta }$ . Moreover, ${\dim E_{\lambda } < \infty }$ and by elliptic regularity, the eigensections are smooth (see, e.g., [Reference Bourguignon, Hijazi, Milhorat, Moroianu and Moroianu8]). Let $e^{-t\operatorname {\Delta }}$ be the heat operator defined as

$$\begin{align*}e^{-t\operatorname{\Delta}}\Phi=e^{-t\lambda} \Phi, \end{align*}$$

for any $\Phi \in E_{\lambda }$ , $\lambda \in \operatorname {Spec} \operatorname {\Delta }$ .

Definition 2.1 The heat kernel of a self-adjoint elliptic pseudodifferential operator P acting on the sections of $\mathcal {E}$ is the Schwartz kernel of the operator $e^{-tP}$ .

If we denote by $\lbrace \Phi _j \rbrace $ an orthonormal Hilbert basis of $\operatorname {\Delta }$ -eigensections, then the heat kernel $p_t(x,y)$ satisfies

$$\begin{align*}p_t(x,y)=\sum_{j}e^{-t\lambda_j} \Phi_j(x) \otimes \Phi_j^*(y) \end{align*}$$

in ${\mathcal C}^{\infty } \left ( (0,\infty ) \times M \times M \right )$ .

Recall that the $L^2$ -product of two sections $s_1, s_2 \in \Gamma (\mathcal {E})$ is given by

$$\begin{align*}\langle s_1,s_2 \rangle_{L^2(\mathcal{E})}=\int_M h_{\mathcal{E}}(s_1,s_2) \operatorname{dvol}_g ,\end{align*}$$

where g is the metric on M and $h_{\mathcal {E}}$ is the Hermitian product on $\mathcal {E}$ .

Let $y \in M$ be a fixed point. We work in geodesic normal coordinates defined by the exponential map

$$\begin{align*}\exp_y: T_yM\longrightarrow M. \end{align*}$$

Since M is compact, there exists a global injectivity radius $\epsilon $ . For x close enough to y ( $d(x,y) \leq \epsilon $ ), take $\operatorname {x} \in T_yM$ the unique tangent vector of length smaller than $\epsilon $ such that $x=\exp _y\operatorname {x}$ . Let

$$\begin{align*}\operatorname{j}(\operatorname{x})=\frac{\exp_y^* dx}{d\operatorname{x}}, \end{align*}$$

namely the pull-back of the volume form $dx$ on M through the exponential map $\exp _{y}$ is equal with $\operatorname {j}(\operatorname {x})d\operatorname {x}$ . More precisely,

$$\begin{align*}\operatorname{j}(\operatorname{x})=\vert \det \left( d_{\operatorname{x}}\exp_{x_0} \right) \vert={\det }^{1/2} \left( g_{ij}(\operatorname{x}) \right) .\end{align*}$$

Denote by $\tau _{x}^{y}: \mathcal {E}_x \longrightarrow \mathcal {E}_y$ the parallel transport along the unique minimal geodesic $x_s=\exp _{y} (s\operatorname {x})$ , where $s \in [0,1]$ , which connects the points x and y. The heat kernel $p_t(x,y)$ belongs to the space $ {\mathcal C}^{\infty } \left ( (0,\infty ) \times M \times M, \mathcal {E}_x \otimes \mathcal {E}_y^* \right )$ and $p_t(x,y)$ satisfies the heat equation

$$\begin{align*}\left( \partial_t+{\operatorname{\Delta}}_x \right) p_t(x,y)=0.\end{align*}$$

Furthermore, $ \lim _{t\rightarrow 0} P_t s=s,$ in $\Vert \cdot \Vert _0$ , for any smooth section $s \in \Gamma (M, \mathcal {E})$ , where

$$\begin{align*}(P_t s)(x)=\int_M p_t(x,y)s(y)dg(y),\end{align*}$$

where $dg(y)$ is the Riemannian density of the metric g. The next theorem is due to Minakshisundaram and Pleijel (see, for instance, [Reference Berger, Gauduchon and Mazet4, Reference Minakshisundaram and Pleijel21]).

Theorem 2.1 The heat kernel $p_t$ has the following asymptotic expansion near the diagonal:

$$ \begin{align*} p_t(x,y) \stackrel{t \searrow 0}{\sim} (4 \pi t)^{-n/2} e^{-\frac{d(x,y)^2}{4t}} \sum_{i=0}^{\infty} t^i \Psi_i (x,y), \end{align*} $$

where $\Psi _i: \mathcal {E}_y \longrightarrow \mathcal {E}_x $ are ${\mathcal C}^{\infty }$ sections defined near the diagonal. Moreover, the $\Psi _i$ ’s are given by the following explicit formulæ:

$$ \begin{align*} {}&\Psi_0(x,y)=\operatorname{j}^{-1/2}(\operatorname{x})\tau_{y}^{x}, \\ {}&\tau_{x}^y \Psi_i(x,y)=-\operatorname{j}^{-1/2}(\operatorname{x}) \int_{0}^{1}s^{i-1} \operatorname{j}^{-1/2}(x_s)\tau_{x_s}^{y} \operatorname{\Delta}_x \Psi_{i-1}(x_s,y)ds. \end{align*} $$

The asymptotic sum in Theorem 2.1 can be understood using truncation and bounds of derivatives as in [Reference Berline, Getzler and Vergne5]. We prefer the interpretation given in [Reference Melrose19], where the heat kernel $p_t$ is shown to belong to $\rho ^{-n} {\mathcal C}^{\infty }(M_H^2)$ on the parabolic blow-up space $M_H^2$ and to vanish rapidly at the temporal boundary face $\{ \omega _0 =0 \}$ (see Section 10).

Example 2.2 Let $\mathbb {T}^n=\left ( S^1\right )^n={\mathbb R}^n/(2 \pi {\mathbb Z})^n$ be the n-dimensional torus with the standard product metric $g=d\theta _1^2 \otimes \cdots \otimes d\theta _n^2$ . Consider the trivial bundle $\mathcal {E}= \underline {{\mathbb C}}$ over $\mathbb {T}^n$ with the standard metric $h_{\mathcal {E}}$ , the trivial connection $\nabla =d$ , and the zero endomorphism F. Let $\operatorname {\Delta }_1$ be the Laplacian on $\mathbb {T}^n$ given by the metric g. The eigenvalues of $\operatorname {\Delta }_1$ are $\{k_1^2+\cdots +k_n^2: k_1,\ldots ,k_n \in {\mathbb Z} \}$ . Let $\varphi _l(\xi )=\frac {1}{\sqrt {2\pi }} e^{il\xi }$ be the standard orthonormal basis of eigenfunctions of each $\operatorname {\Delta }_{S^1}$ . Then, for $\theta =(\theta _1,\ldots ,\theta _n) \in \mathbb {T}^n$ , the heat kernel $p_t$ of $\operatorname {\Delta }_1$ is the following:

$$\begin{align*}p_t(\theta,\theta)= \sum_{(k_1,\ldots,k_n) \in {\mathbb Z}^n} e^{-t(k_1^2+\cdots+k_n^2)} \varphi_{k_1}(\theta_1) \overline{\varphi_{k_1}(\theta_1)}\ldots \varphi_{k_n}(\theta_n) \overline{\varphi_{k_n}(\theta_n)}. \end{align*}$$

Since $\varphi _l(\xi )\overline {\varphi _l(\xi )}=\frac {1}{2\pi }$ , for any $\xi \in S^1$ , we get

$$\begin{align*}p_t(\theta,\theta)= \tfrac{1}{(2\pi)^n} \sum_{(k_1,\ldots,k_n) \in {\mathbb Z}^n} e^{-t(k_1^2+\cdots+k_n^2)}. \end{align*}$$

Remark that the Fourier transform of the function $f_t: {\mathbb R}^n \longrightarrow {\mathbb R}$ , $f_t(x)=e^{-t \vert x \vert ^2}$ is given by

$$\begin{align*}\hat{f_t}(\xi)=\tfrac{\pi^{n/2}}{t^{n/2}} e^{-\frac{\vert \xi \vert^2}{4t}}. \end{align*}$$

Using the multidimensional Poisson formula (see, for instance, [Reference Bellman3]), we obtain that

$$ \begin{align*} p_t(\theta,\theta)=\tfrac{1}{(2 \pi)^n} \sum_{k \in {\mathbb Z}^n} f_t(k)=\sum_{k \in {\mathbb Z}^n} \hat{f_t}(2 \pi k)=\tfrac{\pi^{n/2}}{(2 \pi )^n} t^{-n/2} + \tfrac{\pi^{n/2}}{(2 \pi )^n} t^{-n/2} \sum_{k \in {\mathbb Z}^n \setminus \{ 0 \}} e^{-\frac{\pi^2 \vert k \vert^2}{t}}. \end{align*} $$

Since the last sum is of order ${\mathcal O} \left ( e^{-\frac {1}{t}} \right )$ as $t \rightarrow 0$ , it follows that the first coefficient in the asymptotic expansion at small-time t of $p_t$ is $\tfrac {\pi ^{n/2}}{(2 \pi )^n}$ and all the others vanish.

From now on, suppose that $\operatorname {\Delta }$ is non-negative (i.e., $h_{\mathcal {E}} \left ( \operatorname {\Delta } f, f \right ) \geq 0$ , for any $f \in {\mathcal C}^{\infty } (M,\mathcal {E})$ ). For $s \in {\mathbb C}$ , we define the complex powers $\operatorname {\Delta }^{-s} \in \Psi ^{-2s} \left ( M, \mathcal {E} \right )$ of $\operatorname {\Delta }$ as

$$\begin{align*}\operatorname{\Delta}^{-s} \Phi= \left\{ \begin{array}{ll} \lambda^{-s}\Phi, & \mbox{ if } \Phi \in E_{\lambda}, \ \lambda \neq 0, \\ 0, & \mbox{ if } \Phi \in \operatorname{Ker} \operatorname{\Delta}. \end{array} \right. \end{align*}$$

Remark that $(\operatorname {\Delta }^s)_{s \in {\mathbb C}}$ is a holomorphic family of pseudodifferential operators. Let $r \in (0,1)$ . We denote by $h_t$ the heat kernel of $\operatorname {\Delta }^r$ , namely the Schwartz kernel of the operator $e^{-t\operatorname {\Delta }^r}$ . We have seen that

(2.1) $$ \begin{align} p_t(x,x) \stackrel{t \searrow 0}{\sim} t^{-n/2} \sum_{j=0}^{\infty} t^{j}a_{j}(x,x), \end{align} $$

with smooth sections $a_j(x,x) \in \mathcal {E}_x \otimes \mathcal {E}_x^*$ .

3 The link between the heat kernel and complex powers of the Laplacian

Proposition 1 (Mellin Formula)

With the notations above, for $\Re s>0$ , we have

$$\begin{align*}\operatorname{\Delta}^{-s}=\frac{1}{\Gamma(s)} \int_{0}^{\infty} t^{s-1} \left( e^{-t\operatorname{\Delta}} - \operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}} \right) dt , \end{align*}$$

where $\operatorname {P}_{\operatorname {Ker} \operatorname {\Delta }}$ is the orthogonal projection onto the kernel of $\operatorname {\Delta }$ .

Proof It is straightforward to check that both sides coincide on eigensections $\Phi \in E_{\lambda }$ , $\lambda \in \operatorname {Spec} \operatorname {\Delta }$ . Since $\lbrace \Phi _j \rbrace _{j}$ is a Hilbert basis, the result follows.

We will write $\operatorname {P}_{\operatorname {Ker} \operatorname {\Delta }}(x,y)$ for the Schwartz kernel $\sum _{k} \varphi _k(x) \otimes \varphi _k^*(y)$ , where $\{ \varphi _k \}$ is an orthonormal basis in $\operatorname {Ker} \operatorname {\Delta }$ . Denote by $q_{-s}$ the Schwartz kernel of the operator $\operatorname {\Delta }^{-s}$ . Let us first study the poles and the zeros of $q_{-s}$ away from the diagonal.

Proposition 2 Let K be a compact in $M \times M \setminus \operatorname {Diag}$ . Then, for $(x,y)\in K $ , the function $s \longmapsto {q_{-s}}_{\vert _K} \in {\mathcal C}^{\infty } \left ( K, \mathcal {E} \boxtimes \mathcal {E}^* \right ) $ is entire. Moreover, ${q_{-s}}_{\vert _K}$ vanishes at each negative integer s.

Proof For $\Re s>0$ , let $f_{x,y}(s)= \int _{0}^{\infty } t^{s-1}\left ( p_t(x,y)-\operatorname {P}_{\operatorname {Ker} \operatorname {\Delta }}(x,y) \right ) dt $ . Remark that

$$ \begin{align*} f_{x,y}(s)&=\int_{0}^{\infty} t^{s-1}\left( p_t(x,y)-\operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,y) \right) dt \\&={} \int_{1}^{\infty} t^{s-1} \left( p_t(x,y)-\operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,y) \right) dt \\ &\quad+{} \int_{0}^{1} t^{s-1}p_t(x,y) dt - \operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,y) \cdot \int_{0}^{1} t^{s-1}dt. \end{align*} $$

Since $p_t(x,y)-\operatorname {P}_{\operatorname {Ker} \operatorname {\Delta }}(x,y)$ decays exponentially fast as t goes to $\infty $ , the first integral is absolutely convergent in $ C^k$ norms. The heat kernel $p_t$ vanishes with all of its derivatives as $t \searrow 0$ in the compact K, thus the second integral is also absolutely convergent. The last integral term is well-defined for $\Re s>0$ , and it extends to a meromorphic function on ${\mathbb C}$ with a simple pole in $s=0$ . Therefore, $s \mapsto f_{x,y}(s)$ extends to a meromorphic function on ${\mathbb C}$ . By Proposition 1 and the identity theorem, the equality of meromorphic functions

$$\begin{align*}\Gamma(s)q_{-s}(x,y)= f_{x,y}(s) \end{align*}$$

holds for any $s \in {\mathbb C}$ . In particular, we obtain $q_0(x,y)=- \operatorname {P}_{\operatorname {Ker} \operatorname {\Delta }}(x,y)$ . Furthermore, ${q_{-s}}_{\vert _{K}}$ is an entire function and vanishes in $s=-1,-2,\ldots $ .

Remark 3.1 The fact that ${q_{-s}}_{\vert _K}$ vanishes for negative integers s also follows from the fact that then $\Delta ^{-s}$ is a differential operator.

Now we check the behavior of $q_{-s}$ along the diagonal. It is no longer holomorphic there, and the coefficients $a_j(x,x)$ from (2.1) appear as residues of $q_{-s}(x,x)$ .

Proposition 3 Let $x \in M$ . Then the function $s \mapsto \Gamma (s)q_{-s}(x,x)$ has a meromorphic extension from the set $\{s \in {\mathbb C} : \Re s> \frac {n}{2} \}$ to ${\mathbb C}$ with simple poles in $s \in \lbrace 0 \rbrace \cup \lbrace \frac {n}{2}-j : j \in {\mathbb N} \rbrace $ . The residue of $\Gamma (s)q_{-s}(x,x)$ in $s=\frac {n}{2}-j$ , $j \neq \frac {n}{2}$ , is $a_j(x,x)$ . If n is even, then the residue of $\Gamma (s)q_{-s}(x,x)$ in $s=0$ is $a_{\frac {n}{2}}(x,x)-\operatorname {P}_{\operatorname {Ker} \operatorname {\Delta }}(x,x)$ . If n is odd, the residue in $s=0$ is $-\operatorname {P}_{\operatorname {Ker} \operatorname {\Delta }}(x,x)$ and the meromorphic extension of $q_{-s}(x,x)$ vanishes at ${s \in \{ -1,-2,\ldots \}}$ .

Proof Consider the function $f_{x,x}(s)=\int _{0}^{\infty } t^{s-1}\left ( p_t(x,x)- \operatorname {P}_{\operatorname {Ker} \operatorname {\Delta }}(x,x) \right ) dt$ for $\Re s> \frac {n}{2}$ . We have

$$ \begin{align*} f_{x,x}(s)&=\int_{0}^{\infty} t^{s-1}\left( p_t(x,x)-\operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,x) \right) dt \\&={} \int_{1}^{\infty} t^{s-1} \left( p_t(x,x)-\operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,x) \right) dt \\ &\quad+{} \int_{0}^{1} t^{s-1}p_t(x,x) dt - \operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,x) \cdot \int_{0}^{1} t^{s-1}dt. \end{align*} $$

The first integral is absolutely convergent, as seen in the proof of Proposition 2. The last integral term is meromorphic with a simple pole at $s=0$ with residue $-\operatorname {P}_{\operatorname {Ker} \operatorname {\Delta }}(x,x)$ . Let us analyze the behavior of the second term $A_x(s)=\int _{0}^{1} t^{s-1}p_t(x,x)dt$ .

Using (2.1), we have that for $N \geq 0$ ,

$$\begin{align*}t^{n/2}p_t(x,x)=\sum_{j=0}^N t^j a_j(x,x)+R_{N+1}(t,x), \end{align*}$$

where $R_{N+1}$ is of order ${\mathcal O} (t^{N+1})$ as $t \to 0$ . Furthermore, we obtain

$$ \begin{align*} A_x(s)=\!\int_{0}^{1}\! t^{s-\frac{n}{2}-1} t^{\frac{n}{2}}p_t(x,x)dt=\!&\sum_{j=0}^N \int_{0}^{1}\! t^{s-\frac{n}{2}-1} t^j a_j(x,x)dt + \!\int_{0}^{1} \!t^{s-\frac{n}{2}-1} R_{N+1}(t,x)dt \\ =\!&\sum_{j=0}^N a_j(x,x) \frac{1}{s-\frac{n}{2}+j} + \int_{0}^{1} t^{s-\frac{n}{2}-1} R_{N+1}(t,x)dt. \end{align*} $$

Thus $s \mapsto A_x(s)$ extends to a meromorphic function on ${\mathbb C}$ with simple poles in ${\{ \frac {n}{2}-j : \ j=\overline {0,N+1} \}}$ . Using again Proposition 1 and the identity theorem, we deduce the equality

$$\begin{align*}\Gamma(s)q_{-s}(x,x)=f_{x,x}(s),\end{align*}$$

for any $s \in {\mathbb C}$ . It follows that $\Gamma (s) q_{-s}(x,x)$ is meromorphic on ${\mathbb C}$ with simple poles in $s \in \lbrace 0 \rbrace \cup \lbrace \frac {n}{2}-j : j \in {\mathbb N} \rbrace $ . Moreover, the residue of $\Gamma (s)q_{-s}(x,x)$ in a pole $\frac {n}{2}-j$ is $a_j(x,x)$ , and the conclusion follows.

For $p\in {\mathbb C}$ and $\epsilon>0$ , let $B_{\epsilon }(p)$ be the open disk centered in p of radius $\epsilon $ . We need the following technical result.

Proposition 4 Consider $\alpha < \beta $ , and let $\epsilon>0$ , $l \in {\mathbb N}$ .

  • If K is a compact set disjoint from the diagonal, then the function $ s \longmapsto \Gamma (s){q_{-s}}_{\vert _K} $ is uniformly bounded in $\{ s \in {\mathbb C} : \alpha \leq \Re s \leq \beta \} \setminus B_{\epsilon }(0)$ in the ${\mathcal C}^{l}$ norm on K.

  • The function $ s \longmapsto \Gamma (s){q_{-s}}_{\vert _{\operatorname {Diag}}} $ defined on $\{ s \in {\mathbb C}: \ \alpha \leq \Re s \leq \beta \}\setminus \bigcup _{j \in {\mathbb N} \cup \lbrace \frac {n}{2} \rbrace } B_{\epsilon }(\frac {n}{2}-j) \longrightarrow {\mathcal C}^l \left ( \operatorname {Diag}, \mathcal {E} \otimes \mathcal {E}^* \right )$ is uniformly bounded.

Proof With the same argument as in the proof of Proposition 2, the restriction of the ${\mathcal C}^l$ norm on K of the function $s \mapsto f_{x,y}(s) $ is absolutely convergent in $\{ s \in {\mathbb C} : \alpha \leq \Re s \leq \beta \} \setminus B_{\epsilon }(0)$ , hence it is uniformly bounded.

As in the proof of Proposition 3, the ${\mathcal C}^l$ norm along $\operatorname {Diag}$ of $s \longmapsto f_{x,x}(s)$ converges absolutely in $\{ s \in {\mathbb C}: \ \alpha \leq \Re s \leq \beta \}\setminus \bigcup _{j \in {\mathbb N} \cup \lbrace \frac {n}{2} \rbrace } B_{\epsilon }(\frac {n}{2}-j)$ , thus the conclusion follows.

4 The behavior of quotients of Gamma functions along vertical lines

A fundamental result used in [Reference Bär and Moroianu2] is the Legendre duplication formula

$$\begin{align*}\frac{\Gamma(s)}{\Gamma \left( \frac{s}{2} \right) } = \tfrac{1}{\sqrt{2\pi}} 2^{s- \frac{1}{2}} \Gamma\left( \frac{s+1}{2} \right), \end{align*}$$

together with the rapid decay of the Gamma function in vertical lines $\Re s = \tau $ (see, e.g., [Reference Paris and Kaminski22]). These results are replaced in our case by the following estimate.

Proposition 5 The function $s \longmapsto \frac {\Gamma (s)}{\Gamma (rs)}$ decreases in vertical lines faster than $\vert s \vert ^{-k}$ , for any $k \geq 0$ , uniformly in each strip $\lbrace s \in {\mathbb C} : \alpha \leq \Re (s) \leq \beta \rbrace $ , for any $\alpha ,\beta \in {\mathbb R}$ .

Proof For $z \in {\mathbb C} \setminus {\mathbb R}_{-}$ , recall the Stirling formula (see, for instance, [Reference Whittaker and Watson23])

$$\begin{align*}\log \Gamma(z)=\left( z-\frac{1}{2} \right) \log z -z + \frac{1}{2} \log (2\pi) + \Omega(z), \end{align*}$$

where $\log $ is defined on its principal branch, and $\Omega $ is an analytic function of z. For $|\arg z|<\pi $ and $|z| \to \infty $ , $\Omega $ can be written as

$$\begin{align*}\Omega(z)=\sum_{j=1}^{N-1} \frac{B_{2j}}{2j(2j-1)z^{2j-1}}+R_N(z), \end{align*}$$

where $B_{2j}$ are the Bernoulli numbers $\left ( B_2=\frac {1}{6}, \ B_4=-\frac {1}{30}, \ B_6=\frac {1}{42}, \text {etc.} \right )$ . Moreover, the error term satisfies

$$\begin{align*}|R_N(z)| \leq \frac{|B_{2N}|}{2N(2N-1)} \cdot \frac{\sec^{2N}(\frac{\arg z}{2})}{|z|^{2N-1}}; \end{align*}$$

thus, $R_N(z)$ is of order ${\mathcal O} \left ( |z|^{-2N+1} \right )$ as $|z| \to \infty $ (see, for instance, [Reference Paris and Kaminski22, equation (2.1.6)]). For $s \notin (-\infty ,0)$ , it follows that

$$ \begin{align*} \frac{\Gamma(s)}{\Gamma(rs)}=s^{-s(r-1)}e^{s(r-1)}r^{\frac{1}{2}-rs}e^{\Omega(s)-\Omega(rs)}. \end{align*} $$

Let $s=a+ib$ , $a \in {\mathbb R}$ fixed. As $|b| \to \infty $ , the difference $\vert \Omega (s)-\Omega (rs) \vert \to 0$ ; thus, $\vert e^{\Omega (s)-\Omega (rs)} \vert \to 1$ . Note that $\vert r^{\frac {1}{2}-rs} \vert = \vert r^{\frac {1}{2}-ra} \vert $ and $\vert e^{(r-1)s} \vert = e^{(r-1)a}$ , so these terms are bounded. We show in Lemma 4.1 that for any $k \geq 0$ , $\vert s \vert ^{k} \vert s^s \vert $ goes to $0$ as $\Re s =a$ is fixed and $\vert \operatorname {Im} s \vert $ tends to $\infty $ . It follows that the quotient $\frac {\Gamma (s)}{\Gamma (rs)}$ indeed decreases in vertical lines faster than $\vert s \vert ^{-k}$ , for any $k \geq 0$ , uniformly in vertical strips.

Lemma 4.1 Let $k \geq 0$ . If $a \in {\mathbb R}$ is fixed and $\vert b \vert \to \infty $ , then $\vert (a+ib)^{k+a+ib} \vert $ tends to zero.

Proof Let $s=a+ib \notin (-\infty ,0)$ and set $\log (a+ib)=x+iy$ . Then $x=\log \sqrt {a^2+b^2}$ , $y=\arg s \in (-\pi ,\pi )$ ; hence,

$$\begin{align*}\vert s^{s+k} \vert=\vert e^{(k+a+ib) \log (a+ib)} \vert= e^{(k+a)x-by} = e^{(k+a) \log \sqrt{a^2+b^2}-b \arg s}. \end{align*}$$

Since $b=\tan \arg s \cdot a$ , the exponent is equal to

(4.1) $$ \begin{align} &(k+a)\log \sqrt{a^2+b^2}-b \arg s \\&\quad= (k+a) \log a +\frac{k+a}{2} \log \left( 1+\tan^2 \arg s \right) -a \tan \arg s \cdot \arg s. \nonumber\end{align} $$

If $a>0$ , then $\arg s\nearrow \frac {\pi }{2}$ or $ \arg s \searrow -\frac {\pi }{2}$ , and in both cases $t:=\tan \arg s$ tends to $\infty $ . The exponent (4.1) behaves as the function $t \longmapsto \log (1+t^2)-t$ ; therefore, as $t \to \infty $ , the exponent goes to $-\infty $ and the statement of the claim follows.

If $a<0$ , then $\arg s \searrow \frac {\pi }{2}$ or $\arg s \nearrow -\frac {\pi }{2}$ . In the first case when $\arg s \searrow \frac {\pi }{2}$ , it follows that $ t = \tan \arg s \to -\infty $ . The exponent (4.1) behaves as $ \pm \log (1+t^2)+t $ ; hence, the conclusion follows. While if $\arg s \nearrow -\frac {\pi }{2}$ , then $t \to \infty $ , and the exponent (4.1) behaves as $ \pm \log (1+t^2)-t $ ; thus, the exponent tends again to $-\infty $ . Therefore, $\vert s^{k+s} \vert $ goes to zero, which ends the proof.

5 Link between the complex powers of $\operatorname {\Delta }$ and the heat kernel of $\operatorname {\Delta }^r$

Proposition 6 (Inverse Mellin Formula)

For $\Re \tau>0$ , the operators $e^{-t\operatorname {\Delta }^r}$ and $\operatorname {\Delta }^{-s}$ are related by the following formula:

$$\begin{align*}e^{-t \operatorname{\Delta}^r} -\operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}= \frac{1}{2\pi i}\int_{\Re s= \tau} t^{-s} \Gamma(s) \operatorname{\Delta}^{-rs} ds. \end{align*}$$

Proof The equality holds on each eigensection $\Phi _j$ corresponding to an eigenvalue $\lambda _j \in \operatorname {Spec} \operatorname {\Delta }$ . Since $\lbrace \Phi _j \rbrace _{j}$ is a Hilbert basis, the result follows.

Set $\tau> \frac {n}{2r}$ . Then the Schwartz kernel $q_{-rs}$ of $\operatorname {\Delta }^{-rs}$ is continuous and by the inverse Mellin formula, we get an identity which relates the Schwartz kernels $h_t$ and $q_{-rs}$ :

$$ \begin{align*} h_t(x,y)- \operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,y) ={}& \tfrac{1}{2 \pi i} \int_{\Re s=\tau} t^{-s}\Gamma(s) q_{-rs}(x,y) ds \\ ={}&\tfrac{1}{2 \pi i} \int_{\Re s = \tau} t^{-s}\frac{\Gamma(s)}{\Gamma(rs)} \cdot \Gamma(rs)q_{-rs}(x,y) ds. \end{align*} $$

Now let $k>0$ . By changing $\tau $ to $\tau + \epsilon $ (for a small $\epsilon>0$ ) if needed, we can assume that $\tau -k \notin \lbrace \frac {n}{2}-j :j \in {\mathbb N} \rbrace \cup \{ 0 \}$ . Using Propositions 4 and 5, we can apply the residue formula and move the line of integration to the left:

(5.1) $$ \begin{align} \begin{aligned} h_t(x,y) ={}&\tfrac{1}{2 \pi i} \int_{\Re s = \tau -k} t^{-s}\frac{\Gamma(s)}{\Gamma(rs)} \cdot \Gamma(rs)q_{-rs}(x,y) ds \\ {}& +\sum_{s \in -{\mathbb N} \cup \lbrace \frac{n-2j}{2r} : \ j \in {\mathbb N} \rbrace} \operatorname{Res}_{s} \left( t^{-s} \Gamma(s) q_{-rs}(x,y) \right) + \operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,y). \end{aligned} \end{align} $$

Notice that $-{\mathbb N} \cup \lbrace \frac {n-2j}{2r} : \ j \in {\mathbb N} \rbrace $ is the set of all possible poles of $s \mapsto \Gamma (s) q_{-rs}(x,y)$ , but some of them might actually be regular points. We will study the sum (5.1) in detail in Theorems 6.1 and 7.1.

Let K be a compact set in $M \times M \setminus \operatorname {Diag}$ and $l \in {\mathbb N}$ . Remark that the integral term in (5.1) is of order ${\mathcal O} \left ( t^{k-\tau } \right )$ in ${\mathcal C}^l (K, \mathcal {E} \boxtimes \mathcal {E}^*)$ . Indeed,

$$ \begin{align*} \left\Vert \int_{\Re s = \tau-k} t^{-s}\Gamma(s){q_{-rs}}_{\vert_K} ds \right\Vert{}_l \leq t^{-\tau+k} \cdot \int_{ s = \tau-k+i u} \left\Vert \frac{\Gamma(s)}{\Gamma(rs)} \cdot \Gamma(rs) {q_{-rs}}_{\vert_K} \right\Vert{}_l du, \end{align*} $$

and using again Propositions 4 and 5, the claim follows. Furthermore, when k goes to $\infty $ , we get

(5.2) $$ \begin{align} {h_t}_{\vert_K} \stackrel{t \searrow 0}{\sim} \sum_{\alpha=0}^{\infty} t^{\alpha} \cdot \operatorname{Res}_{s=-\alpha} \left( \Gamma(s) {q_{-rs}}_{\vert_K} \right) + t^0 \cdot {\operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}}_{\vert_K}, \end{align} $$

The meaning of the asymptotic sign in (5.2) is that if we set $h_t^N$ to be the right-hand side in (5.2) restricted to $\alpha \leq N$ , then the difference $\vert \partial _t^j \left ( {h_t}_{\vert _K}- h_t^N \right ) \vert $ is of order ${\mathcal O} (t^{N+1-j})$ in ${\mathcal C}^l(K, \mathcal {E} \boxtimes \mathcal {E}^*)$ , for any $N,j \in {\mathbb N}$ .

Remark that using again Propositions 4 and 5, the integral term in (5.1) is of order ${\mathcal O} \left ( t^{k-\tau } \right )$ in ${\mathcal C}^l (\operatorname {Diag}, \mathcal {E} \otimes \mathcal {E}^*)$ . Therefore when k tends to $\infty $ , we obtain

(5.3) $$ \begin{align} {h_t}_{\vert_{\operatorname{Diag}}} \stackrel{t \searrow 0}{\sim} \sum_{\alpha \in \left( -{\mathbb N} \right) \cup \lbrace \frac{n-2j}{2r}: j \in {\mathbb N} \rbrace } t^{-\alpha} \cdot \operatorname{Res}_{s=\alpha} \left( \Gamma(s) {q_{-rs}}_{\vert_{\operatorname{Diag}}} \right) + t^0 \cdot {\operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}}_{\vert_{\operatorname{Diag}}}, \end{align} $$

in the sense of the following:

Definition 5.1 Consider $l \in {\mathbb N}$ and let $A,B \subset {\mathbb R}$ . We say that $ {h_t}_{\vert _{\operatorname {Diag}}} \stackrel {t \searrow 0}{\sim } \sum _{\alpha \in A} t^{\alpha } {c_{\alpha }} + \sum _{\beta \in B} t^{\beta } \log t \cdot c_{\beta } $ if for any $k,N \in {\mathbb N}$ , the difference

$$\begin{align*}\partial_t^j \left( {h_t}_{\vert_{\operatorname{Diag}}}- \sum_{\alpha \leq N} t^{\alpha} {c_{\alpha}} - \sum_{\beta \leq N} t^{\beta} \log t \cdot c_{\beta} \right) \end{align*}$$

is of order ${\mathcal O} (t^{N+1-j} \log t)$ in ${\mathcal C}^l(\operatorname {Diag}, \mathcal {E} \otimes \mathcal {E}^*)$ .

6 The asymptotic expansion of $h_t$ away from the diagonal

Theorem 6.1 The Schwartz kernel $h_t$ of the operator $e^{-t\operatorname {\Delta }^r}$ is ${\mathcal C}^{\infty }$ on $[0,\infty )\times \left ( M \times M \setminus \operatorname {Diag} \right )$ . Furthermore, let $K \subset M \times M \setminus \operatorname {Diag}$ be a compact set. Then the Taylor series of ${h_t}_{\vert _{K}}$ as $t \searrow 0$ is the following:

$$\begin{align*}{h_t}_{\vert_K} \stackrel{t \searrow 0}{\sim} \sum_{j=1}^{\infty} t^{j} {q_{rj}}_{\vert_K} \frac{(-1)^j}{j!}. \end{align*}$$

Moreover, if $r=\frac {\alpha }{\beta }$ is rational with $\alpha ,\beta $ coprime, then the coefficient of $t^j$ vanishes for $j \in \beta {\mathbb N}^*$ .

Proof Let $j \in {\mathbb N}$ . Using Propositions 4 and 5, $(-s)(-s-1)\ldots (-s-j+1) t^{-s-j} \frac {\Gamma (s)}{\Gamma (rs) } \Gamma (rs){q_{-rs}}_{\vert _{K}} $ is $L^1$ integrable on $ \Re s = \tau -k$ in ${\mathcal C}^l (K, \mathcal {E} \boxtimes \mathcal {E}^*)$ , for sufficiently large k and for any $l \in {\mathbb N}$ . It follows that $h_t$ is ${\mathcal C}^{\infty }$ on $(0,\infty )\times \left ( M \times M \setminus \operatorname {Diag} \right )$ . By Proposition 2, the function $s \mapsto q_{-rs}(x,y)$ is entire for any $(x,y) \in K$ . Since $\operatorname {Res}_{s=-j} \Gamma (s)= \frac {(-1)^j}{j!} $ , using (5.2) we get

$$\begin{align*}{h_t}_{\vert_K} \stackrel{t \searrow 0}{\sim} \sum_{j=0}^{\infty} t^{j} {q_{rj}}_{\vert_K} \frac{(-1)^j}{j!} +{\operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}}_{\vert_K}.\end{align*}$$

We obtained in the proof of Proposition 2 that ${q_0}_{\vert _K}=-{\operatorname {P}_{\operatorname {Ker} \operatorname {\Delta }}}_{\vert _K}$ ; thus,

$$\begin{align*}{h_t}_{\vert_K} \stackrel{t \searrow 0}{\sim} \sum_{j=1}^{\infty} t^{j} {q_{rj}}_{\vert_K} \frac{(-1)^j}{j!}, \end{align*}$$

and therefore $h_{t_{\vert _K}}$ is ${\mathcal C}^{\infty }$ also at $t=0$ , and vanishes at order $1$ . Moreover, using again Proposition 2, if $r=\frac {\alpha }{\beta }$ is rational and j is a non-zero multiple of $\beta $ , then $q{_{rj}}_{\vert _K} \equiv 0$ and the conclusion follows.

7 The asymptotic expansion of $h_t$ along the diagonal

To obtain the coefficients in the asymptotic of $h_t$ along the diagonal as $t\searrow 0$ , we need to compute the residues from (5.3). Some of them are related to the heat coefficients $a_j$ ’s of $p_t$ due to Proposition 3. We will distinguish three cases. If n is even, $\Gamma (s)q_{-rs}(x)$ has simple poles in $\lbrace \frac {n}{2r},\frac {n-2}{2r},\ldots ,\frac {2}{2r} \rbrace \cup \lbrace 0,-1,\ldots \rbrace $ and the residues will give rise to real powers of t. If n is odd and either r is irrational or r is rational with odd denominator, $\Gamma (s)q_{-rs}(x)$ has simple poles in $\lbrace 0,-1,\ldots \rbrace \cup \lbrace \frac {n-2j}{2r} : j=0,1,\ldots \rbrace $ . Otherwise, if n is odd and r is rational with even denominator, then there exist some double poles which give rise to logarithmic terms in the asymptotic expansion of $h_t$ .

Theorem 7.1 Let $a_j(x,x)$ be the coefficients in ( 2.1 ) of the heat kernel $p_t$ of the non-negative self-adjoint generalized Laplacian $\operatorname {\Delta }$ . The asymptotic expansion of the Schwartz kernel $h_t$ of the operator $e^{-t\operatorname {\Delta }^r}$ , $r \in (0,1)$ along the diagonal when $t\searrow 0$ is the following:

  1. (1) If n is even, then

    $$\begin{align*}{h_t}_{\vert_{\operatorname{Diag}}} \stackrel{t \searrow 0}{\sim} \sum_{j=0}^{n/2-1} t^{- \frac{n-2j}{2r}} \cdot A_{-\frac{n-2j}{2r}} + a_{n/2} + \sum_{j=1}^{\infty} t^j \cdot A_j. \end{align*}$$

    If $r=\frac {\alpha }{\beta }$ is rational, for $j=l \beta $ , $ l \in {\mathbb N}^*$ , we obtain that $q_{rj}(x,x)=(-1)^j \cdot j! \cdot a_{\frac {n}{2}+l \alpha }(x,x)$ , and the coefficient of $t^{l \beta }$ can be described more precisely as

    $$\begin{align*}A_{l \beta}=a_{\frac{n}{2}+l \alpha}. \end{align*}$$
  2. (2) If n is odd and either $r \in {\mathbb R} \setminus {\mathbb Q}$ or the denominator of r is odd, then

    $$\begin{align*}{h_t}_{\vert_{\operatorname{Diag}}} \stackrel{t \searrow 0}{\sim} \sum_{j=0}^{(n-1)/2} t^{- \frac{n-2j}{2r}} \cdot A_{-\frac{n-2j}{2r}} + \sum_{j=1}^{\infty} t^j \cdot A_j + \sum_{j=1}^{\infty} t^{\frac{2j+1}{2r}} \cdot A_{\frac{2j+1}{2r}}. \end{align*}$$

    Moreover, if $r=\frac {\alpha }{\beta }$ is rational and $\beta $ is odd, then $ A_{l \beta } \equiv 0$ for any $l \in {\mathbb N}^*$ .

  3. (3) If n is odd, $r=\frac {\alpha }{\beta }$ is rational and its denominator $\beta $ is even, then

In all these cases, the coefficients are

$$ \begin{align*} {}&A_{-\frac{n-2j}{2r}}(x)= \frac{\Gamma \left( \frac{n-2j}{2r} \right)} {\Gamma \left( \frac{n-2j}{2} \right)} \cdot \frac{1}{r} \cdot a_j(x,x), && A_{j}(x)=\frac{(-1)^j}{j!} \cdot q_{rj}(x,x), \\ {}&A_{\frac{2j+1}{2r}}(x)=\frac{\Gamma \left( - \frac{2j+1}{2r} \right)} {\Gamma \left( - \frac{2j+1}{2} \right)} \cdot \frac{1}{r} \cdot a_{\frac{n+2j+1}{2}}(x,x), && B_{l\frac{\beta}{2}}(x)=\frac{(-1)^{l\frac{\beta}{2}}}{r \left( l\frac{\beta}{2} \right) ! \Gamma \left( - l\frac{\beta}{2} \cdot r \right)} \cdot a_{\frac{n+l \alpha}{2}}(x,x), \end{align*} $$
$$\begin{align*}A_{l \frac{\beta}{2}}(x)= \frac{(-1)^{l \frac{\beta}{2}}}{(l \frac{\beta}{2})! \Gamma(-rl \frac{\beta}{2})} \cdot \operatorname{FP}_{s=-l \frac{\beta}{2}} \left( \Gamma(rs)q_{-rs}(x,x) \kern-1pt\right) + \operatorname{FP}_{s=-l \frac{\beta}{2}} \!\left(\kern-1pt \frac{\Gamma(s)}{\Gamma(rs)} \kern-1pt\right) \cdot \frac{a_{\frac{n+l\alpha}{2}(x,x)}}{r}. \end{align*}$$

Proof We compute the coefficients from (5.3) by using Proposition 3.

7.1 The case when n is even

For $j \in \{0,1,\ldots ,n/2-1 \}$ , we have

(7.1) $$ \begin{align} \operatorname{Res}_{s=\frac{n-2j}{2r}} \left( t^{-s} \frac{\Gamma(s)}{\Gamma(rs)} \Gamma(rs)q_{-rs}(x,x) \right) =t^{-\frac{n-2j}{2r}} \cdot \frac{\Gamma(\frac{n-2j}{2r})}{\Gamma(\frac{n-2j}{2})} \cdot \frac{a_j(x,x)}{r}. \end{align} $$

The residue in $s=0$ is given by

$$ \begin{align*} \operatorname{Res}_{s=0} \left( t^{-s} \Gamma(s) q_{-rs}(x,x) \right) ={}& \operatorname{Res}_{s=0} \left( t^{-s} \frac{\Gamma(s)}{\Gamma(rs)} \Gamma(rs)q_{-rs}(x,x) \right) \\ ={}& r \cdot \frac{1}{r} \kern-1.2pt\left(\kern-1pt a_{\frac{n}{2}}(x,x) \kern1pt{-}\kern1pt \operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,x) \kern-1pt\right)\kern1.2pt{=}\kern1.2pt a_{\frac{n}{2}}(x,x)\kern1pt{-}\kern1pt\operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,x), \end{align*} $$

thus the coefficient of $t^0$ in the asymptotic expansion (5.3) is $a_{\frac {n}{2}}(x,x)$ .

7.1.1 The case when n is even and r is irrational

Let $j \in {\mathbb N}^*$ . Then

(7.2) $$ \begin{align} \operatorname{Res}_{s=-j} \left( t^{-s} \Gamma(s) q_{-rs}(x,x) \right) = t^j \frac{(-1)^j}{j!} \cdot q_{rj}(x,x). \end{align} $$

Therefore, in this case, the asymptotic expansion of $h_t$ is the following:

$$\begin{align*}h_t(x,x) \stackrel{t \searrow 0}{\sim} \sum_{j=0}^{n/2-1} t^{-\frac{n-2j}{2r}} \frac{\Gamma \left( \frac{n-2j}{2r} \right)}{\Gamma \left( \frac{n-2j}{2} \right)} \frac{a_j(x,x)}{r} + a_{\frac{n}{2}}(x,x) + \sum_{j=1}^{\infty} t^j \frac{(-1)^j}{j!} q_{rj}(x,x). \end{align*}$$

7.1.2 The case when n is even and $r=\frac {\alpha }{\beta }$ is rational with $(\alpha ,\beta )=1$

Some of the coefficients $q_{rj}(x,x)$ from (7.2) can be expressed in terms of the $a_{k}$ ’s from (2.1). Remark that $\frac {\Gamma (s)}{\Gamma (rs)}$ has simple poles in $\{-1,-2,\ldots \} \setminus \{ \frac {-1}{r}, \frac {-2}{r},\ldots \}$ . For $j \in {\mathbb N}^*$ , $s:=-\frac {j}{r} \in \{-1,-2,\ldots \}$ if and only if j is a multiple of $\alpha $ , which is equivalent to $s=\frac {-l\alpha }{r}=-l\beta $ for some $l \in {\mathbb N}^*$ . In this case, we obtain

$$ \begin{align*} \operatorname{Res}_{s=-l\beta} \left( t^{-s} \Gamma(s) q_{-rs}(x,x) \right) ={}& \operatorname{Res}_{s=-l\beta} \left( t^{-s} \frac{\Gamma(s)}{\Gamma(rs)} \Gamma(rs)q_{-rs}(x,x) \right) \\ ={}& t^{l\beta} r \cdot \frac{1}{r} a_{\frac{n}{2}+l\alpha}(x,x)=t^{l\beta} a_{\frac{n}{2}+l \alpha}(x,x). \end{align*} $$

Hence, for rational $r=\frac {\alpha }{\beta }$ , if $j=l\beta $ , $l \in {\mathbb N}^*$ , we conclude that

(7.3) $$ \begin{align} q_{rj}(x,x)=(-1)^j \cdot j! \cdot a_{\frac{n}{2}+l\alpha}(x,x), \end{align} $$

and $h_t(x,x)$ has the following asymptotic expansion as $t \searrow 0$ :

7.2 The case when n is odd

For $j \in \lbrace 0,1,\ldots ,(n-1)/2 \rbrace $ , the coefficient of $t^{-\frac {n-2j}{2r}}$ is computed as in (7.1). Furthermore, in $s=0$ ,

$$ \begin{align*} \operatorname{Res}_{s=0} \left( t^{-s} \Gamma(s) q_{-rs}(x,x) \right) ={}& \operatorname{Res}_{s=0} \left( t^{-s} \frac{\Gamma(s)}{\Gamma(rs)} \cdot \Gamma(rs)q_{-rs}(x,x) \right) \\ ={}& r \cdot \frac{-1}{r} \cdot \operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,x) =-\operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,x); \end{align*} $$

hence, there is no free term in the asymptotic expansion of $h_t$ as t goes to zero.

Now we have to compute the residues of the function $t^{-s} \Gamma (s) q_{-rs}(x,x)$ in ${s \in \{ -1,-2,\ldots \}}$ and $ s \in \{ \frac {-1}{2r}, \frac {-3}{2r},\ldots \}$ .

7.2.1 The case when n is odd and r is irrational

Then these sets are disjoint; thus, all poles of the function $\Gamma (s) q_{-rs}(x)$ are simple. For $j \in {\mathbb N}^*$ , the coefficient of $t^j$ is obtained as in (7.2). Furthermore, for $j \in {\mathbb N}$ , we get

(7.4) $$ \begin{align} \operatorname{Res}_{s=-\frac{2j+1}{2r}} \left( t^{-s} \frac{\Gamma(s)}{\Gamma(rs)} \cdot \Gamma(rs)q_{-rs}(x,x) \right) =t^{\frac{2j+1}{2r}} \cdot \frac{\Gamma(-\frac{2j+1}{2r})}{\Gamma(-\frac{2j+1}{r})} \cdot \frac{a_{\frac{n+2j+1}{2}}(x,x)}{r}. \end{align} $$

Therefore, the small-time asymptotic expansion of $h_t$ is the following:

$$ \begin{align*} h_t(x,x) \stackrel{t \searrow 0}{\sim}{}& \sum_{j=0}^{n/2-1} t^{-\frac{n-2j}{2r}} \cdot \frac{\Gamma \left( \frac{n-2j}{2r} \right)}{\Gamma \left( \frac{n-2j}{2} \right)} \cdot \frac{a_j(x,x)}{r} + \sum_{j=1}^{\infty} t^j \cdot \frac{(-1)^j}{j!} q_{rj}(x,x) \\ +{}& \sum_{j=0}^{\infty} t^{\frac{2j+1}{2r}} \cdot \frac{\Gamma \left( -\frac{2j+1}{2r} \right)}{\Gamma \left( - \frac{2j+1}{2} \right)} \cdot \frac{a_{\frac{n+2j+1}{2}}(x,x)}{r}. \end{align*} $$

7.2.2 The case when n is odd and $r=\frac {\alpha }{\beta }$ is rational

Consider the sets

$$ \begin{align*} A:=\{ -1,-2,\ldots \}, && B:=\{ \tfrac{-1}{2r},\tfrac{-3}{2r},\ldots \}, && C:=\{ \tfrac{-1}{r},\tfrac{-2}{r},\ldots \}. \end{align*} $$

Remark that A is the set of negative poles of $s \longmapsto t^{-s}\Gamma (s)q_{-rs}(x,x)$ , and $A \setminus C$ is the set of poles of the function $s \longmapsto \frac {\Gamma (s)}{\Gamma (rs)}$ . Clearly B and C are disjoint. Moreover, $A \cap C=\{ -l\beta : \ l \in {\mathbb N}^* \}$ . Furthermore, if $\beta $ is odd, then $A\cap B=\emptyset $ , and otherwise if $\beta $ is even, then $A\cap B=\{-l\frac {\beta }{2}: \ l \in 2 {\mathbb N}+1 \}$ . Such an $s=-\frac {2j+1}{2r}=l\frac {\beta }{2} \in A \cap B$ is a double pole for $\Gamma (s)q_{rs}(x)$ .

7.2.3 Suppose that $\beta $ is odd

Then A and B are disjoint. Thus, for $s=-\frac {2j+1}{2r} \in B$ , $j \in {\mathbb N}$ , the residue of $t^{-s}\Gamma (s)q_{rs}(x,x)$ is the one computed in (7.4).

For $s=-j \in A\setminus C$ (which means that $j \in {\mathbb N}^*$ , $\beta \nmid j$ ), the residue of $t^{-s}\Gamma (s)q_{-rs}(x,x)$ in s is the one computed in (7.2).

If $s=-l\beta =-\frac {l\alpha }{r} \in A\cap C$ for some $l \in {\mathbb N}^*$ , then $\Gamma (s)$ has a simple pole in s and by Proposition 3, (the meromorphic extension of) $q_{-rs}(x,x)$ vanishes at $s=-l\beta $ . Hence, the product $t^{-s}\Gamma (s)q_{-rs}(x,x)$ is holomorphic in $s=-l\beta $ and $t^{l\beta }$ , $l \in {\mathbb N}^*$ , does not appear in the asymptotic expansion.

Therefore, if $r=\frac {\alpha }{\beta }$ is rational and $\beta $ is odd, we obtain

7.2.4 Assume now that $\beta $ is even

For $s=-\frac {2j+1}{2r}\in B \setminus A$ ( $j \in {\mathbb N} $ with $\alpha \nmid 2j+1$ ), the residue is computed as in (7.4). For $s=-j \in A\setminus \left ( B\cup C \right )$ (namely $j \in {\mathbb N}^*$ , $\frac {\beta }{2} \nmid j$ ), the residue is computed as in (7.2).

For $s \in C \cap A$ (namely $s=-l\beta $ , $l \in {\mathbb N}^*$ ), the residue is again $0$ . Indeed, $\Gamma (s)$ has a simple pole in $-l\beta $ and by Proposition 3, (the meromorphic extension of) $q_{-rs}(x,x)$ vanishes in $-l\beta $ , thus $t^{l\beta }$ does not appear in the asymptotic expansion of $h_t$ .

Finally, if $s=-\frac {l\alpha }{2r}=-l\frac {\beta }{2} \in A \cap B$ , $l \in 2 {\mathbb N} +1$ , then s is a double pole for $\Gamma (s)q_{-rs}(x,x)$ . We write the Laurent expansions of the functions $t^{-s}$ , $\frac {\Gamma (s)}{\Gamma (rs)}$ , and $\Gamma (rs)q_{-rs}(x,x),$ respectively, in $s=-\frac {l\alpha }{2r}=-l\frac {\beta }{2}=:-k$ :

$$ \begin{align*} {}&t^{-s}=t^k-t^k \log t + {\mathcal O} (s+k)^2, \\ {}&\frac{\Gamma(s)}{\Gamma(rs)}=\frac{(-1)^k}{k! \cdot \Gamma(-kr)} (s+k)^{-1}+\cdots, \\ {}&\Gamma(rs)(q_{-rs}(x,x))=\frac{1}{r}a_{\frac{n+l\alpha}{2}}(x,x) (s+k)^{-1}+\cdots. \end{align*} $$

Thus, we finally obtain that

$$ \begin{align*} \operatorname{Res}_{s=-k}\left( t^{-s} \cdot \frac{\Gamma(s)}{\Gamma(rs)} \cdot \Gamma(rs)q_{-rs}(x,x)\right)={}& t^k \cdot \frac{(-1)^k}{k! \Gamma(-kr)} \cdot \operatorname{FP}_{s=-k} \left( \Gamma(rs)q_{-rs}(x,x) \right) \\ {}&+ t^k \operatorname{FP}_{s=-k} \left(\frac{\Gamma(s)}{\Gamma(rs)} \right) \cdot\frac{a_{\frac{n+l\alpha}{2}(x,x)}}{r} \\ {}&+t^k \log t \cdot \frac{(-1)^k}{k! \Gamma(-kr)} \frac{a_{\frac{n+l\alpha}{2}(x,x)}}{r}. \end{align*} $$

8 Non-triviality of the coefficients

Let us prove Theorem 1.1. Recall the definition of the zeta function of a non-negative self-adjoint generalized Laplacian $\Delta $ :

$$\begin{align*}\zeta_{\Delta}(s):=\sum_{\lambda \in \operatorname{Spec} \Delta \setminus \{ 0 \}} \lambda^{-s}=\int_{M} q_{-s}(x,x)dg(x). \end{align*}$$

This series is absolutely convergent for $\Re s> \frac {n}{2}$ and extends meromorphically to ${\mathbb C} $ with possible simple poles in the set

$$\begin{align*}\left\{ \frac{n}{2}-j : j \in {\mathbb N} \setminus \left\{ \frac{n}{2} \right\} \right\}\end{align*}$$

(see, for instance, [Reference Gilkey13]).

Consider the trivial bundle ${\mathbb C}$ over a compact Riemannian manifold M. As in [Reference Loya, Moroianu and Ponge17], let $\left ( \operatorname {\Delta } + \xi \right )_{\xi> 0}$ be a family of generalized Laplacians indexed by $\xi>0$ , and denote by $ q_{-s}^{\xi } $ the Schwartz kernels of the operators $(\operatorname {\Delta }+\xi )^{-s}$ . Note that for $\Re s> \frac {n}{2}$ ,

(8.1) $$ \begin{align} \int_M q_{-s}^{\xi} (x,x) dx = \operatorname{Tr} \left( \operatorname{\Delta} +\xi \right)^{-s}=\zeta_{\operatorname{\Delta}+\xi}(s) = \sum_{\lambda_j \in \operatorname{Spec} \operatorname{\Delta}} \left( \lambda_j +\xi \right)^{-s}. \end{align} $$

Since for $\Re s> \frac {n}{2}$ the sum is absolutely convergent, we obtain

$$\begin{align*}\frac{d}{d\xi} \zeta_{\operatorname{\Delta}+\xi} (s) =-s \cdot \sum_{\lambda_j \in \operatorname{Spec} \operatorname{\Delta}} \left( \lambda_j +\xi \right)^{-s-1}= -s\cdot \zeta_{ \operatorname{\Delta} +\xi}( s+1). \end{align*}$$

By induction, it follows that for $\Re s> \frac {n}{2}$ ,

(8.2) $$ \begin{align} \frac{d}{d\xi^k} \zeta_{\operatorname{\Delta}+\xi}(s)=(-1)^k s(s+1)\ldots(s+k-1) \cdot \zeta_{\operatorname{\Delta} +\xi}( s+k). \end{align} $$

Using the identity theorem, (8.2) holds true on ${\mathbb C}$ as an equality of meromorphic functions. Consider $s \in {\mathbb R} \setminus (-{\mathbb N})$ and $k \in {\mathbb N}$ large enough such that $s+k> \frac {n}{2}$ . Since $\zeta _{\operatorname {\Delta }+\xi }(s+k)$ is a convergent sum of strictly positive numbers, the right-hand side is non-zero. Thus, for any fixed $s \in {\mathbb R} \setminus (- {\mathbb N}) $ , on any open set $U \subset (0, \infty )$ , the function $\xi \longmapsto \zeta _{\operatorname {\Delta }+\xi }(s)$ is not identically zero on U, and by (8.1), $q_{-s}^{\xi }(x,x)$ cannot be constant zero on M. Hence, for $s=-rj \notin - {\mathbb N}$ , there exist $\xi _0 \in (0, \infty )$ and $x_0 \in M$ such that the coefficient $q_{rj}^{\xi _0}(x_0,x_0)$ of the asymptotic expansion of the Schwartz kernel $h_t$ of $e^{-t (\operatorname {\Delta }+\xi _0)^r}$ is non-zero.

Now suppose that $rj \in {\mathbb N}$ . Then $r=\frac {\alpha }{\beta }$ is rational and j is a multiple of $\beta $ , $j:=l\beta $ . If n is odd, we already proved in Theorem 7.1 that $t^{l\beta }$ does not appear in the asymptotic expansion of $h_t$ as $t \searrow 0$ . Furthermore, if n is even, by (7.3), $q_{rj}(x,x)$ is a non-zero multiple of the coefficient $a_{\frac {n}{2}+l\alpha }(x,x)$ in the asymptotic expansion (2.1) of the heat kernel $p_t$ . It is well known that the heat coefficients in (2.1) are non-trivial (see, for instance, [Reference Gilkey13]). It follows that all coefficients obtained in Theorem 7.1 indeed appear in the asymptotic expansion, proving Theorem 1.1.

9 Non-locality of the coefficients $A_j(x)$ in the asymptotic expansions

Let us prove Theorem 1.3. We give an example of an n-dimensional manifold and a Laplacian for which the coefficients $A_{j}(x)=\frac {(-1)^j}{j!} q_{rj(x,x)}$ , $j \in {\mathbb N}^*$ , $rj \notin {\mathbb N}$ appearing in Theorem 7.1 are not locally computable in the sense of Definition 1.1 (i). Let ${\mathbb {T}^n={\mathbb R}^n / \left ( 2\pi {\mathbb Z} \right )^n}$ be the n-dimensional torus from Example 2.2. Let $\operatorname {\Delta }_g$ be the Laplacian on $\mathbb {T}^n$ given by the metric $g=d\theta _1^2 +\cdots + d\theta _n^2$ .

Remark that the eigenvalues of $\operatorname {\Delta }_g$ are $\{ k_1^2 +\cdots +k_n^2 : k_1,\ldots ,k_n \in {\mathbb Z} \}$ . Let $ {\varphi _l(t)= \frac {1}{\sqrt {2\pi }} e^{il t} }$ be the standard orthonormal basis of eigenfunctions of each $\operatorname {\Delta }_{S^1}$ . Then, for $\Re s> \frac {n}{2}$ and $\theta =(\theta _1,\ldots ,\theta _n) \in \mathbb {T}^n$ , the Schwartz kernel of $\operatorname {\Delta }_g^{-s}$ is given by

$$\begin{align*}q_{-s}^{\operatorname{\Delta}_g} \left( \theta , \theta \right) = \sum_{ (k_1,\ldots,k_n) \in {\mathbb Z}^n \setminus \{ 0 \} } \left( k_1^2+\cdots+k_n^2 \right)^{-s} \varphi_{k_1}(\theta_1) \overline{\varphi_{k_1}(\theta_1)}\ldots \varphi_{k_n}(\theta_n) \overline{\varphi_{k_n}(\theta_n)}. \end{align*}$$

Consider the n-dimensional zeta function

$$\begin{align*}\zeta_n(s):= \sum_{(k_1,\ldots,k_n) \in {\mathbb Z}^n \setminus \{ 0 \}} \left( k_1^2 +\cdots+k_n^2 \right)^{-s}= \sum_{k \in {\mathbb N}^*} k^{-s} R_n(k), \end{align*}$$

where $R_n(k)$ is the number of representations of k as a sum of n squares. Since $\varphi _{l}(t) \overline {\varphi _{l}(t)} =\frac {1}{2\pi }$ for any $t \in S^1$ , it follows that

(9.1) $$ \begin{align} q_{-s}^{\operatorname{\Delta}_g} \left( \theta,\theta \right) = \frac{1}{(2 \pi)^n} \zeta_n(s), \end{align} $$

for any $\Re s> \frac {n}{2}$ , and clearly $q_{-s}^{\operatorname {\Delta }_g}$ is independent of $\theta $ .

Now let us change the metric locally on each component $S^1$ . Let U be an open interval in $S^1$ , and $\psi :S^1 \longrightarrow [0,\infty )$ a smooth function with $\operatorname {supp} \psi \subset U$ . Consider the new metric $\left ( 1+\psi (\theta ) \right ) d\theta ^2$ on each $S^1$ . Then there exist $p>0$ and an isometry $\Phi : \left ( S^1, \left ( 1+\psi (\theta ) \right ) d\theta ^2 \right ) \longrightarrow \left ( S^1, p^2 d\theta ^2 \right )$ . Remark that the Laplacian on $S^1$ given by the metric $p^2 d \theta ^2$ corresponds under this isometry to $p^{-2}$ times the Laplacian for the metric $d\theta ^2$ . Let

$$ \begin{align*} \tilde{g}= \sum_{j=1}^n \left( 1+ \psi(\theta_j) \right) d\theta_j^2 && g_p=\sum_{j=1}^n p^2 d\theta_j^2=p^2g. \end{align*} $$

Then clearly $\Phi \times \cdots \times \Phi : (\mathbb {T}^n, \tilde {g}) \longmapsto (\mathbb {T}^n, g_p) $ is an isometry, and let $\tilde {\Delta }$ , $\Delta _{p }$ be the corresponding Laplacians on $\mathbb {T}^n$ . Denote by $q_{-s}^{\tilde {\Delta }}$ and $q_{-s}^{\Delta _{p }}$ the Schwartz kernels of the complex powers $ \tilde {\Delta } ^{-s}$ and $\Delta _{p }^{-s}$ . We have for $\Re s>\frac {n}{2}$ ,

(9.2) $$ \begin{align} q_{-s}^{\operatorname{\Delta}_{p }} \left( \theta, \theta \right) = \frac{1}{(2 \pi p)^n} \sum_{k=(k_1,\ldots,k_n) \in {\mathbb Z}^n \setminus \{ 0 \}} \left( p^{-2}k_1^2 +\cdots+p^{-2}k_n^2 \right)^{-s} = \frac{p^{2s}}{(2 \pi p)^n} \zeta_n(s). \end{align} $$

Remark that

$$\begin{align*}q_{-s}^{\operatorname{\Delta}_{p}} \left( \theta,\theta \right) =q_{-s}^{\tilde{\operatorname{\Delta}}} \left( \Phi(\theta),\Phi(\theta) \right), \end{align*}$$

and both of them are independent of $\theta $ . By (9.2), for $\Re s>\frac {n}{2}$ , we obtain

(9.3) $$ \begin{align} q_{-s}^{\tilde{\operatorname{\Delta}}} \left( \theta,\theta \right) = \frac{p^{2s-n}}{(2 \pi )^n} \zeta_n(s). \end{align} $$

Now we prove that $\zeta _n(s)$ has a meromorphic extension on ${\mathbb C}$ with so-called trivial zeros at $s=-1,-2,\ldots $ . By Proposition 1, for $\Re s> \frac {n}{2}$ , we have

$$\begin{align*}\zeta_n(s) \Gamma(s)=\int_{0}^{\infty} t^{s-1} \sum_{k=(k_1,\ldots,k_n) \in {\mathbb Z}^n \setminus \{ 0 \}} e^{-t(k_1^2+\cdots+k_n^2)} dt = \int_0^{\infty} t^{s-1} F(t) dt, \end{align*}$$

where $F(t):= \sum _{k=(k_1,\ldots ,k_n) \in {\mathbb Z}^n \setminus \{ 0 \}} e^{-t(k_1^2+\cdots +k_n^2)}$ . Using the multidimensional Poisson formula (see, for instance, [Reference Bellman3]), it follows that

$$\begin{align*}1+F(t)= \sum_{k \in {\mathbb Z}^n} f_t(k)=\sum_{k \in {\mathbb Z}^n} \hat{f_t}(2 \pi k)=\pi^{n/2} t^{-n/2} \left( 1 + F \left( \frac{\pi^2}{t} \right) \right), \end{align*}$$

and therefore

$$\begin{align*}F(t)=-1+\pi^{n/2}t^{-n/2} + \pi^{n/2} t^{-n/2} F \left( \frac{\pi^2}{t} \right). \end{align*}$$

Since $F(t)$ goes to $0$ rapidly as $t \to \infty $ , the function $A(s)= \int _{1}^{\infty } t^{s-1} F(\pi t)dt $ is entire. Remark that

$$ \begin{align*} \zeta_n(s) \Gamma(s)={}& \int_{0}^{\pi} t^{-s} F(t)dt + \int_{\pi}^{\infty} t^{s-1}F(t) dt \\ ={}& \pi^s \left( -\frac{1}{s} + \frac{1}{s-\frac{n}{2}} + A\left( \frac{n}{2}-s \right) + A(s)\right), \end{align*} $$

so

(9.4) $$ \begin{align} \pi^{-s} \zeta_n(s) \Gamma(s)= -\frac{1}{s} + \frac{1}{s-\frac{n}{2}} + A\left( \frac{n}{2}-s \right) + A(s). \end{align} $$

Therefore, $\zeta _n$ extends meromorphically to ${\mathbb C}$ with a simple pole in $s=\frac {n}{2}$ and zeros at $s=-1,-2,\ldots $ . Furthermore, since the RHS is invariant through the involution $s \mapsto \frac {n}{2}-s$ , it follows that $\zeta _n(s)$ does not have any other zeros for $s \in (-\infty ,0)$ . We obtain the well-known functional equation of the Epstein zeta function

$$\begin{align*}\pi^{-s} \zeta_n(s) \Gamma(s)= \pi^{s-n/2} \zeta_n \left( \frac{n}{2}-s \right) \Gamma \left( \frac{n}{2}-s \right) \end{align*}$$

(see, for instance, [Reference Chandrasekharan and Narasimhan9, equation (63)]). Remark that for $r \in (0,1)$ and $j \in {\mathbb N}^*$ with $rj \notin {\mathbb N}$ , $\zeta _n (-rj )$ is not zero.

Using the identity theorem, it follows that (9.1) and (9.3) hold true as an equality of meromorphic functions on ${\mathbb C}$ , and furthermore, we get

$$\begin{align*}q_{rj}^{\operatorname{\Delta}_g}(\theta, \theta) \neq q_{rj}^{\tilde{\operatorname{\Delta}}}(\theta,\theta), \end{align*}$$

for $rj \notin {\mathbb N}$ . Since we modified the metric locally in $U^n \subset \mathbb {T}^n$ and the corresponding kernel $q_{rj}^{\tilde {\operatorname {\Delta }}}$ changed its behavior globally, it follows that it is not locally computable in the sense of Definition 1.1 (i).

Furthermore, let us see that the heat coefficients $A_j(x)=\frac {(-1)^j}{j!} q_{rj}(x,x)$ for $j={\mathbb N}^*$ , $rj \notin {\mathbb N}$ are not cohomologically local in the sense of Definition 1.1 (iii). We argue by contradiction. Let j be fixed. Suppose that there exists a function C, locally computable in the sense of Definition 1.1 (i), such that

(9.5) $$ \begin{align} \int_{\mathbb{T}^n} q_{rj}^{\operatorname{\Delta}_g} \operatorname{dvol}_g = \int_{\mathbb{T}^n} C(g) \operatorname{dvol}_g, && \int_{\mathbb{T}^n} q_{rj}^{\tilde{\operatorname{\Delta}}} \operatorname{dvol}_{\tilde{g}} = \int_{\mathbb{T}^n} C(\tilde{g}) \operatorname{dvol}_{\tilde{g}}. \end{align} $$

Using (9.1) and (9.3), it follows that

$$ \begin{align*} (2 \pi)^n \zeta_n(-rj) = \int_{\mathbb{T}^n} C(g) \operatorname{dvol}_g , && &{} (2 \pi p)^n p^{-2rj} \zeta_{n}(-rj)= \int_{\mathbb{T}^n} C(\tilde{g}) \operatorname{dvol}_{\tilde{g}}. \end{align*} $$

Remark that in the case of the trivial bundle with the trivial connection over a locally homogeneous Riemannian manifold $(M,h)$ (i.e., such that every two points have isometric neighborhoods), the function $C(M,h) \in {\mathcal C}^{\infty }(M)$ is constant on M. This follows directly from Definition 1.1 (i). Therefore, $C(g)$ , $C(\tilde {g}),$ and $C(g_p)$ are constant functions.

Since $(\mathbb {T}^n , \tilde {g})$ is (globally) isometric to $(\mathbb {T}^n , g_p)$ , it follows that $C(\tilde {g})=C(g_p)$ . Furthermore, since $(\mathbb {T}^n , g_p)$ is locally isometric to $(\mathbb {T}^n , g)$ and $C(g_p)$ , $C(g)$ are constant functions, it also follows that they are equal: $C(g_p)=C(g)$ . Hence we conclude that $C(\tilde {g})=C(g_p)=C(g)=:C$ , for some $C \in {\mathbb C}$ , and thus we have

(9.6) $$ \begin{align} \int_{\mathbb{T}^n} C \operatorname{dvol}_{\tilde{g}} = \int_{\mathbb{T}^n} C \operatorname{dvol}_{g_p}. \end{align} $$

Since $g_p=p^2 g$ , we obtain that

(9.7) $$ \begin{align} \int_{\mathbb{T}^n} C \operatorname{dvol}_{g_p} = p^n \int_{\mathbb{T}^n} C \operatorname{dvol}_g, \end{align} $$

and then using (9.5)–(9.7), we get

$$\begin{align*}(2 \pi p)^n p^{-2rj} \zeta_n(-rj) = p^n \cdot (2 \pi)^n \zeta_n(-rj). \end{align*}$$

But, we proved above that $\zeta _n (-rj)$ does not vanish for $rj \notin {\mathbb N}$ . We obtain a contradiction because $p^{-2rj} \neq 1$ for $r \in (0,1)$ , $j=1,2,\ldots $ .

10 Interpretation of $h_t$ on the heat space for $r=1/2$

In Theorems 6.1 and 7.1, we studied the asymptotic behavior of the heat kernel $h_t$ of ${ \operatorname {\Delta }^r}$ , $r \in (0,1)$ for small-time t in two distinct cases: when we approach $t=0$ along the diagonal in $M \times M$ , and when we approach a compact set away from the diagonal. We now give a simultaneous asymptotic expansion formula for both cases when $r=\frac {1}{2}$ . Furthermore, in order to understand the asymptotic behavior as t goes to zero in any direction (not just the case when t goes to $0$ in the vertical one),we will pull-back the formula on a certain linear heat space $\operatorname {M_{heat}}$ .

In [Reference Melrose19], Melrose used his blow-up techniques to give a conceptual interpretation for the asymptotic of the heat kernel $p_t$ . Recall that the heat space $M_H^2$ is obtained by performing a parabolic blow-up of $ \{ t=0 \} \times \operatorname {Diag} $ in $[0,\infty ) \times M \times M$ . The heat space $M_H^2$ is a manifold with corners with boundary hypersurfaces given by the boundary defining functions $\rho $ and $\omega _0$ . The heat kernel $p_t$ belongs to $\rho ^{-n} {\mathcal C}^{\infty }(M_H^2)$ , and vanishes rapidly at the boundary hypersurface $\{ \omega _0=0 \}$ (see [Reference Melrose19, Theorem 7.12]).

In order to study the Schwartz kernel $h_t$ of $e^{-t \operatorname {\Delta }^{1/2}}$ , we introduce the linear heat space $\operatorname {M_{heat}}$ , which is just the standard blow-up of $\{ 0 \} \times \operatorname {Diag}$ in $[0,\infty ) \times M \times M$ (see [Reference Melrose and Mazzeo20] for details regarding the blow-up of a submanifold). Let $\operatorname {ff}$ be the front face, i.e., the newly added face, and denote by $\operatorname {lb}$ the lateral boundary which is the lift of the old boundary $\{ 0 \} \times M \times M$ . The blow down map is given locally by

$$ \begin{align*} \beta_H: \operatorname{M_{heat}} \longrightarrow [0,\infty) \times M \times M && \beta_H(\rho, \omega, x')=(\rho \omega_0, \rho \omega'+x', x'), \end{align*} $$

where

$$\begin{align*}\omega \in {\mathbb S}^n_H= \{ \omega= (\omega_0, \omega') \in {\mathbb R}^{n+1} : \ \omega_0 \geq 0, \ \omega_0^2+|\omega'|^2 =1 \}. \end{align*}$$

Proof of Theorem 1.4

We want to show that $h_t \in \rho ^{-n} \omega _0 \cdot {\mathcal C}^{\infty } (\operatorname {M_{heat}}) + \rho \log \rho \cdot \omega _0 \cdot {\mathcal C}^{\infty } (\operatorname {M_{heat}}) $ , and in fact, the second (logarithmic) term does not occur when n is even. First, we deduce the unified formula for $h_t$ as $t \searrow 0$ both on the diagonal and away from it. By Mellin formula 1 and inverse Mellin formula 6, for $\tau>n$ , we get

$$ \begin{align*} h_t(x,y)-\operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,y)={}& \tfrac{1}{2\pi i} \int_{\Re s=\tau} t^{-s} \frac{\Gamma(s)}{\Gamma \left( \frac{s}{2} \right)} \Gamma \left(\frac{s}{2} \right)q_{-s/2}(x,y)ds \\ ={}& \tfrac{1}{2\pi i} \int_{\Re s = \tau} t^{-s} \frac{\Gamma(s)}{\Gamma \left( \frac{s}{2} \right)} \int_{0}^{\infty} T^{\tfrac{s}{2}-1} \left( p_T(x,y)-\operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,y) \right) dT ds. \end{align*} $$

We use the Legendre duplication formula as in [Reference Bär and Moroianu2] (see, for instance, [Reference Paris and Kaminski22]):

$$\begin{align*}\frac{\Gamma(s)}{\Gamma\left( \frac{s}{2} \right) } = \frac{1}{\sqrt{2\pi}} 2^{s-\tfrac{1}{2}} \Gamma \left( \frac{s+1}{2} \right), \end{align*}$$

obtaining that $h_t(x,y)-\operatorname {P}_{\operatorname {Ker} \operatorname {\Delta }}(x,y)$ is equal to

$$ \begin{align*} \tfrac{1}{\sqrt{4 \pi}} \tfrac{1}{2\pi i} \int_{\Re s = \tau} \int_{0}^{\infty} \left( \frac{2 \sqrt{T}}{t} \right)^s \Gamma \left( \frac{s+1}{2} \right) \left( p_T(x,y)-\operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,y) \right) dT ds. \end{align*} $$

Set $X:=\tfrac {2\sqrt {T}}{t}$ . Using Propositions 4, 5, and Fubini, we first compute the integral in s. Changing the variable $S=\frac {s+1}{2}$ and applying the residue theorem, we get

$$ \begin{align*} \tfrac{1}{2\pi i} \int_{\Re s = \tau} X^s \Gamma \left( \frac{s+1}{2} \right) ds={}& \tfrac{2}{2\pi i} \int_{\Re S = \frac{\tau+1}{2}} X^{2S-1} \Gamma(S) dS = 2 \sum_{k=0}^{\infty} \frac{(-1)^k}{k!} X^{-2k-1} \\ ={}& 2 X^{-1}e^{-X^{-2}}=\frac{t}{\sqrt{T}}e^{-\frac{t^2}{4T}}. \end{align*} $$

Thus, we obtain

(10.1) $$ \begin{align} \begin{aligned} h_t(x,y)-\operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,y)= \tfrac{t}{2\sqrt{\pi}} \int_{0}^{\infty} T^{-3/2} e^{-\frac{t^2}{4T}} \left( p_T(x,y)-\operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,y) \right) dT. \end{aligned} \end{align} $$

Since $p_T(x,y)-\operatorname {P}_{\operatorname {Ker} \operatorname {\Delta }}(x,y)$ decays exponentially as T goes to infinity, it follows that the integral from $1$ to $\infty $ in the right-hand side of equation (10.1) is of the form $t \cdot {\mathcal C}^{\infty }_{t,x,y} \left ( [0,\infty ) \times M^2 \right )$ . Furthermore, by the change of variable $u=\tfrac {t}{2\sqrt {T}}$ , we have

$$ \begin{align*} -\tfrac{t}{2\sqrt{\pi}}\int_{0}^{1} T^{-3/2}e^{-\frac{t^2}{4T}} dT \cdot \operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,y) ={}&-\tfrac{2}{\sqrt{\pi}}\int_{t/2}^{\infty} e^{-u^2} du \cdot \operatorname{P}_{\operatorname{Ker} \operatorname{\Delta}}(x,y). \end{align*} $$

Since $ \int _{t/2}^{\infty } e^{-u^2}du$ tends to $\frac {\sqrt {\pi }}{2}$ as $t \searrow 0$ , the term $-\tfrac {t}{2\sqrt {\pi }}\int _{0}^{1} T^{-3/2}e^{-\frac {t^2}{4T}} dT \operatorname {P}_{\operatorname {Ker} \operatorname {\Delta }}(x,y)$ will cancel in the limit as $t \to 0$ with $- \operatorname {P}_{\operatorname {Ker} \operatorname {\Delta }}(x,y)$ from the left-hand side of (10.1).

Let us study the remaining integral term $\tfrac {t}{2\sqrt {\pi }} \int _{0}^{1} T^{-3/2}e^{-\frac {t^2}{4T}} p_T(x,y)dT$ . By Theorem 2.1,

$$ \begin{align*} p_T(x,y)=T^{-n/2} e^{-\frac{d(x,y)^2}{4T}} \sum_{j=0}^N T^j a_j(x,y) + R_{N+1}(T,x,y), \end{align*} $$

where the remainder $R_{N+1}(T,x,y)$ is of order ${\mathcal O} (T^{N+1})$ ; therefore,

$$ \begin{align*} \tfrac{t}{2\sqrt{\pi}} \int_{0}^{1} T^{-3/2}e^{-\frac{t^2}{4T}} p_T(x,y)dT={}& \tfrac{t}{2\sqrt{\pi}} \int_{0}^{1} T^{-3/2}e^{-\frac{t^2}{4T}} R_{N+1}(T,x,y) dT \\ +{}& \tfrac{t}{2\sqrt{\pi}} \!\int_{0}^{1} \!T^{-3/2}e^{-\frac{t^2}{4T}} T^{-n/2}e^{-\frac{d(x,y)^2}{4T}} \sum_{j=0}^{N} T^j a_{j}(x,y) dT. \end{align*} $$

Since $R_{N+1}(T,x,y)$ is of order ${\mathcal O} (T^{N+1})$ , the first integral is again of type $ t \cdot {\mathcal C}^{\infty }_{t,x,y} $ . By changing the variable $u=\tfrac {t^2+d(x,y)^2}{4T}$ in the second integral, we get

$$ \begin{align*} {}&\tfrac{t}{2\sqrt{\pi}} \sum_{j=0}^N a_{j}(x,y) \int_{0}^{1} T^{-\frac{n+3}{2} +j}e^{-\frac{t^2+d(x,y)^2}{4T}} dT \\ ={}& \tfrac{t}{2 \sqrt{\pi}} \sum_{j=0}^N a_j(x,y) \left( \frac{t^2+d(x,y)^2}{4} \right)^{-\frac{n+1}{2}+j} \int_{\frac{t^2+d(x,y)^2}{4}}^{\infty} u^{\frac{n+1}{2}-j-1} e^{-u}du \\ ={}& \tfrac{t}{2 \sqrt{\pi}} \sum_{j=0}^N a_j(x,y) \Gamma \left( \frac{n+1}{2}-j, \frac{t^2+d(x,y)^2}{4} \right) \left( \frac{t^2+d(x,y)^2}{4} \right)^{-\frac{n+1}{2}+j}, \end{align*} $$

where $\Gamma (z,\xi ):=\int _{\xi }^{\infty } u^{z-1}e^{-u}du$ is the upper incomplete Gamma function. We conclude that $h_t(x,y)$ is equal to

(10.2) $$ \begin{align} t \cdot {\mathcal C}^{\infty}_{t,x,y} + \tfrac{t}{2 \sqrt{\pi}} \sum_{j=0}^N a_j(x,y) \Gamma \left( \frac{n+1}{2}-j, \frac{t^2+d(x,y)^2}{4} \right) \left( \frac{t^2+d(x,y)^2}{4} \right)^{-\frac{n+1}{2}+j}. \end{align} $$

10.1 The case when n is even

If $z>0$ , then one can easily check that $\Gamma (z,\xi ) \in \xi ^z {\mathcal C}^{\infty }_{\xi }[0,\epsilon ) + \Gamma (z)$ , for some $\epsilon>0$ . Furthermore, for $z \in (-\infty ,0] \setminus \lbrace 0,-1,-2,\ldots \rbrace $ ,

$$ \begin{align*} \Gamma(z,\xi)={}&-\frac{1}{z}\xi^ze^{-\xi}+\frac{1}{z} \Gamma(z+1,\xi) \\ ={}&\xi^z e^{-\xi} \sum_{k=0}^{a-1} \frac{-1}{z(z+1)\ldots(z+k)} \xi^k + \frac{1}{z(z+1)\ldots(z+a)} \Gamma(z+a,\xi) \\ ={}& \xi^z {\mathcal C}^{\infty}_{\xi}[0, \epsilon) + \frac{1}{z(z+1)\ldots(z+a-1)} \Gamma(z+a,\xi), \end{align*} $$

where a is a positive integer such that $z+a>0$ . Thus, for a non-integer $z<0$ , we have

$$ \begin{align*} \Gamma(z,\xi)= \xi^z {\mathcal C}^{\infty}_{\xi} [0, \epsilon) + \frac{1}{z(z+1)\ldots(z+a-1)} \Gamma(z+a). \end{align*} $$

We want to interpret equation (10.2) on the heat space $\operatorname {M_{heat}}$ ; thus, we pull back (10.2) through $\beta _H$ :

$$ \begin{align*} \beta_H^*h={}& \rho \omega_0 \beta_H^* {\mathcal C}^{\infty}_{t,x,y} + \tfrac{1}{2 \sqrt \pi} \rho \omega_0 \sum_{j=0}^{N} \left( \tfrac{\rho^2}{4} \right)^{-\frac{n+1}{2}+j} \beta_H^*a_j(x,y) \Gamma\left( \frac{n+1}{2}-j , \frac{\rho^2}{4} \right) \\ ={}& \rho \omega_0 \beta_H^* {\mathcal C}^{\infty}_{t,x,y} + \tfrac{1}{2 \sqrt{\pi}} \rho^{-n}\omega_0 \sum_{j=0}^{n/2} \rho^{2j} 2^{n+1-2j} \beta_H^*a_j(x,y) \Gamma \left( \frac{n+1}{2}-j \right) \\ {}&+ \tfrac{1}{2 \sqrt{\pi}} \rho \omega_0 \sum_{j=0}^{n/2} \beta_H^* a_j(x,y) {\mathcal C}^{\infty}_{\rho^2}[0,\epsilon) + \tfrac{1}{2 \sqrt{\pi}} \rho\omega_0 \sum_{j=n/2 +1}^{N} \beta_H^* a_j(x,y) {\mathcal C}^{\infty}_{\rho^2}[0,\epsilon) \\ {}&+\tfrac{1}{2 \sqrt{\pi}} \rho^{-n}\omega_0 \sum_{j=n/2+1}^{N} \rho^{2j} 2^{n+1-2j} \beta_H^*a_j(x,y) \frac{2^{-n/2+j}}{\left( n+1-2j \right) \left( n+3-2j \right)\ldots(-1)} \Gamma \left( \frac{1}{2} \right). \end{align*} $$

Since $\Gamma \left ( \frac {n+1}{2} -j \right ) = \frac {\sqrt {\pi } (n-2j-1)!!}{2^{n/2-j}} $ for $j \in \{0,1,\ldots ,n/2 \}$ , it follows that

(10.3) $$ \begin{align} \begin{aligned} \beta_H^* h={}& \rho \omega_0 \beta_H^* {\mathcal C}^{\infty}_{t,x,y} + \omega_0 \rho {\mathcal C}^{\infty}_{\rho^2}[0,\epsilon) + \rho^{-n} \omega_0 \sum_{j=0}^{n/2} \rho^{2j} 2^{n/2-j} (n-2j-1)!! \beta_H^* a_j(x,y) \\ {}&+ \rho^{-n}\omega_0 \sum_{j=n/2+1}^{N} \rho^{2j} \frac{ (-1)^{j-n/2} 2^{n/2-j} }{(2j-n-1)!!} \beta_H^*a_j(x,y). \end{aligned} \end{align} $$

The case $\rho \neq 0$ and $\omega _0 \to 0$ corresponds to $x \neq y$ and $t \searrow 0$ before the pull-back. We obtain that $\beta _H^*h$ is in ${\mathcal C}^{\infty }(\operatorname {M_{heat}})$ and it vanishes at first order on $\operatorname {lb}$ , which is compatible with Theorem 6.1.

If $\rho \to 0$ and $\omega _0=1$ , which corresponds to $x=y$ and $t \searrow 0$ , then $\beta _H^*h= \rho ^{-n}\omega _0 \sum _{j=0}^{N} \rho ^{2j} A_{j}(x)$ , where we denoted by $A_j(x)$ the coefficients appearing in (10.3). Again, this result is compatible with Theorem 7.1, and moreover, the coefficients are precisely the ones from [Reference Bär and Moroianu2, Theorem 3.1].

Remark that formula (10.3) is stronger than Theorems 6.1 and 7.1. If both $\rho $ and $\omega _0$ tend to $0$ (with different speeds), it describes the behavior of $h_t$ as t goes to zero from any positive direction (not only the vertical one).

10.2 The case when n is odd

Remark that for small $\xi $ , we have

$$ \begin{align*} \Gamma(0,\xi)={}&\int_{\xi}^{\infty} t^{-1}e^{-t} dt = \int_{\xi}^{1} \frac{e^{-t}-1}{t} dt +\int_{\xi}^{1} t^{-1} dt + \int_{1}^{\infty}t^{-1}e^{-t}dt \\ ={}&- \log \xi + {\mathcal C}^{\infty}_{\xi}[0,\epsilon). \end{align*} $$

Furthermore, if p is a negative integer, inductively we obtain

$$ \begin{align*} \Gamma(-p,\xi)={}& \frac{e^{-\xi}\xi^{-p}}{p!} \sum_{k=0}^{p-1} (-1)^k (p-k-1)! \xi^k + \frac{(-1)^p}{p!} \Gamma(0,\xi) \\ ={}& \xi^{-p} {\mathcal C}^{\infty}_{\xi}[0,\epsilon)-\frac{(-1)^p}{p!} \log \xi +{\mathcal C}^{\infty}_{\xi}[0,\epsilon). \end{align*} $$

We pull-back equation (10.2) on the heat space $\operatorname {M_{heat}}$ :

$$ \begin{align*} \beta_H^*h={}& \rho \omega_0 \beta_H^* {\mathcal C}^{\infty}_{t,x,y} + \tfrac{1}{2 \sqrt \pi} \rho \omega_0 \sum_{j=0}^{N} \left( \tfrac{\rho^2}{4} \right)^{-\frac{n+1}{2}+j} \beta_H^*a_j(x,y) \Gamma\left( \frac{n+1}{2}-j , \frac{\rho^2}{4} \right) \\ ={}& \rho \omega_0 \beta_H^*a_j(x,y) + \tfrac{1}{2\sqrt{\pi}} \rho \omega_0 \sum_{l=0}^{(n-1)/2} \beta_H^* a_j(x,y) {\mathcal C}^{\infty}_{\rho^2}[0,\epsilon) \\ {}&+\frac{1}{\sqrt{\pi}} \rho^{-n}\omega_0 \sum_{j=0}^{(n-1)/2} \rho^{2j} \beta_H^*a_j(x,y) 2^{n-2j} \Gamma \left( \frac{n+1}{2}-j \right) \\ {}&+ \frac{2}{\sqrt{\pi}} \rho^{-n} \omega_0 \sum_{j=(n+1)/2}^{N} \rho^{2j} \log \rho \beta_H^{*} a_j(x,y) 2^{n-2j} \frac{(-1)^{j-\frac{n+1}{2}+1}}{\left( j-\frac{n+1}{2} \right)!} \\ {}&+ \frac{2}{\sqrt{\pi}} \rho^{-n} \omega_0 \sum_{j=(n+1)/2}^{N} \rho^{2j} \beta_H^{*} a_j(x,y) 2^{n-2j} \frac{(-1)^{j-\frac{n+1}{2}}}{\left( j-\frac{n+1}{2} \right)!} \log 2 \\ {}&+\tfrac{1}{2\sqrt{\pi}} \rho \omega_0 \sum_{j=(n+1)/2}^{N} \beta_H^*a_j(x.y) {\mathcal C}^{\infty}_{\rho^2}[0,\epsilon) \\ {}&+ \tfrac{1}{ \sqrt{\pi}} \rho^{-n}\omega_0 \sum_{j=(n+1)/2}^{N} \rho^{2j} \beta_H^*a_j(x,y) 2^{n-2j} \frac{(-1)^{j-\frac{n+1}{2}}}{\left( j-\frac{n+1}{2} \right)!} {\mathcal C}^{\infty}_{\rho^2}[0,\epsilon). \end{align*} $$

Therefore, we obtain

(10.4) $$ \begin{align} \begin{aligned} \beta_H^* h={}& \rho \omega_0 \beta_H^* {\mathcal C}^{\infty}_{t,x,y} + \omega_0 \rho {\mathcal C}^{\infty}_{\rho^2}[0,\epsilon) + \omega_0 \rho^{-n} {\mathcal C}^{\infty}_{\rho^2}[0,\epsilon) \\ {}&+ \frac{1}{\sqrt{\pi}} \rho^{-n} \omega_0 \sum_{j=0}^{(n-1)/2} \rho^{2j} \beta_H^*a_j(x,y) 2^{n-2j} \left( \frac{n+1}{2}-j \right) ! \\ {}&+ \frac{2}{\sqrt{\pi}} \rho^{-n} \omega_0 \sum_{j=(n+1)/2}^{N} \rho^{2j} \log \rho \beta_H^{*} a_j(x,y) 2^{n-2j} \frac{(-1)^{j-\frac{n+1}{2}+1}}{\left( j-\frac{n+1}{2} \right)!} \\ {}&+ \frac{2}{\sqrt{\pi}} \rho^{-n} \omega_0 \sum_{j=(n+1)/2}^{N} \rho^{2j} \beta_H^{*} a_j(x,y) 2^{n-2j} \frac{(-1)^{j-\frac{n+1}{2}}}{\left( j-\frac{n+1}{2} \right)!} \log 2. \end{aligned} \end{align} $$

If $\rho \neq 0$ and $\omega _0 \to 0$ (corresponding to $x \neq y$ and $t \searrow 0$ before the pull-back on $\operatorname {M_{heat}}$ ), we obtain that $\beta _H^* h \in {\mathcal C}^{\infty }(M_{heat})$ and it vanishes at order $1$ at $\operatorname {lb}$ , which is compatible with the result of Theorem 6.1.

In the case $\rho \to 0$ and $\omega _0=1$ which corresponds to $x=y$ and $t \searrow 0$ , we obtain $\beta _H^* h= \rho ^{-n}{\mathcal C}^{\infty }_{\rho ^2} + \rho ^{-n}\sum _{j=0}^N \rho ^{2j} A_j(x) + \rho ^{-n} \sum _{j=(n+1)/2}^N \rho ^{2j} \log \rho B_j(x)$ , where we denoted by $A_j$ and $B_j$ the coefficients appearing in (10.4). This result is compatible with Theorem 7.1 and again, we find some of the coefficients appearing in [Reference Bär and Moroianu2, Theorem 3.1].

11 The heat kernel as a polyhomogeneous conormal section

Let us recall the notions of index family and polyhomogeneous conormal functions on a manifold with corners with two boundary hypersurfaces. (For an accessible introduction, see [Reference Grieser, Gil, Grieser and Lesch15], and for full details of the theory, see [Reference Melrose18].) A discrete subset $F \in {\mathbb C} \times {\mathbb N} $ is called an index set if the following conditions are satisfied:

  1. 1) For any $N \in {\mathbb R}$ , the set $F \cap \{ (z,p): \Re z < N \} $ is finite.

  2. 2) If $p> p_0$ and $(z,p) \in F$ , then $(z, p_0) \in F$ .

If X is a manifold with corners with two boundary hypersurfaces $B_1$ and $B_2$ given by the boundary defining functions x and y, a smooth function f on is said to be polyhomogeneous conormal with index sets E and F, respectively, if in a small neighborhood $[0,\epsilon ) \times B_1$ , f has the asymptotic expansion

$$\begin{align*}f(x,y) \stackrel{x \searrow 0}{\sim} \sum_{(z,p) \in F} a_{z,p}(y)\cdot x^z \log^p x, \end{align*}$$

where $a_{z,p}$ are smooth coefficients on $B_2$ , and for each $a_{z,p}$ there exists a sequence of real numbers $b_{w,q}$ , such that

$$\begin{align*}a_{z,p}(y) \stackrel{y \searrow 0}{\sim} \sum_{(w,q) \in E}b_{w,q} \cdot y^w \log^q y.\end{align*}$$

One can prove that f is a polyhomogeneous conormal function on X with index sets $F_p= \{ (k,0) : k \in {\mathbb Z}, k \geq -p \}$ and $F_0=\{ (n,0) : n \in {\mathbb N} \}$ if and only if $f \in y^{-p} {\mathcal C}^{\infty }(X)$ . Furthermore, f is a polyhomogeneous conormal function on X with index sets $F'=\{ (n,1) : n \in {\mathbb N}^* \}$ and $F_0$ if and only if $f \in {\mathcal C}^{\infty }(X)+ \log y \cdot {\mathcal C}^{\infty }(X)$ . Therefore, we can restate Theorem 1.4 as follows:

Theorem 11.1 For $r=\frac {1}{2}$ , the heat kernel $h_t$ of the operator $e^{-t \operatorname {\Delta }^{1/2}}$ is a polyhomogeneous conormal section on the linear heat space $\operatorname {M_{heat}}$ with values in $\mathcal {E} \boxtimes \mathcal {E}^*$ . The index set for the lateral boundary is

$$\begin{align*}F_{\operatorname{lb}} =\{ (k,0): k \in {\mathbb N}^* \}. \end{align*}$$

If n is even, the index set of the front face is

$$\begin{align*}F_{\operatorname{ff}}=\{(-n+k,0): k \in {\mathbb N} \}, \end{align*}$$

whereas for n odd, the index set toward $\operatorname {ff}$ is given by

$$\begin{align*}F_{\operatorname{ff}}=\{ (-n+k,0): k \in {\mathbb N} \} \cup \{ (k,1) : k \in {\mathbb N}^* \}. \end{align*}$$

It seems reasonable to expect that the Schwartz kernel $h_t$ of the operator $e^{-t\operatorname {\Delta }^r}$ for $r \in (0,1)$ can be lifted to a polyhomogeneous conormal section in a certain “transcendental” heat space $M^r_{Heat}$ depending on r with values in $\mathcal {E} \boxtimes \mathcal {E}^*$ . However, already in the case $r=1/3,$ our method leads to complicated computations involving Bessel modified functions. We therefore leave this investigation open for a future project.

Acknowledgment

I am grateful to my advisor Sergiu Moroianu for many enlightening discussions and for a careful reading of the paper. I would like to thank the anonymous referee for helpful suggestions and remarks leading to the improvement of the presentation.

Footnotes

This work was partially supported from the project PN-III-P4-ID-PCE-2020-0794 funded by UEFSCDI.

References

Agronovič, M. S., Some asymptotic formulas for elliptic pseudodifferential operators . Funktsional. Anal. i Prilozhen. 21(1987), 6365.Google Scholar
Bär, C. and Moroianu, S., Heat kernel asymptotics for roots of generalized Laplacians . Int. J. Math. 14(2003), 397412.CrossRefGoogle Scholar
Bellman, R., A brief introduction to theta functions, Athena Series: Selected Topics in Mathematics, Holt, Rinehart and Winston, New York, 1961.Google Scholar
Berger, M., Gauduchon, P., and Mazet, E., Le spectre d’une variété riemannienne, Lecture Notes in Mathematics, 194, Springer, Berlin and New York, 1971.CrossRefGoogle Scholar
Berline, N., Getzler, E., and Vergne, M., Heat kernels and Dirac operators, Springer, Berlin, 2004.Google Scholar
Berline, N. and Vergne, M., A computation of the equivariant index of the Dirac operator . Bull. Soc. Math. France 113(1985), 305345.CrossRefGoogle Scholar
Bismut, J. M., The Atiyah–Singer theorems: a probabilistic approach . J. Funct. Anal. 57(1984), 329348.CrossRefGoogle Scholar
Bourguignon, J., Hijazi, O., Milhorat, J., Moroianu, A., and Moroianu, S., A spinorial approach to Riemannian and conformal geometry, European Mathematical Society, Zurich, 2015.CrossRefGoogle Scholar
Chandrasekharan, K. and Narasimhan, R., Hecke’s functional equation and arithmetical identities . Ann. Math. 74(1961), 123.CrossRefGoogle Scholar
Duistermaat, J. J. and Guillemin, V. W., The spectrum of positive elliptic operators and periodic bicharacteristics . Invent. Math. 29(1975), 3979.CrossRefGoogle Scholar
Fahrenwaldt, M. A., Off-diagonal heat kernel asymptotics of pseudodifferential operators on closed manifolds and subordinate Brownian motion . Integr. Equ. Oper. Theory 87(2017), 327347.CrossRefGoogle Scholar
Getzler, E., Pseudodifferential operators on supermanifolds and the index theorem . Commun. Math. Phys. 92(1983), 163178.CrossRefGoogle Scholar
Gilkey, P. B., Invariance theory, the heat equation, and the Atiyah–Singer index theorem. 2nd ed., Studies in Advanced Mathematics, CRC Press, Boca Raton, FL, 1995.Google Scholar
Gilkey, P. B. and Grubb, G., Logarithmic terms in asymptotic expansions of heat operator traces . Comm. Partial Differential Equations 23(1998), nos. 5–6, 777792.CrossRefGoogle Scholar
Grieser, D., Basics of the b-calculus . In: Gil, J. B., Grieser, D., and Lesch, M. (eds.), Approaches to singular analysis, Advances in Partial Differential Equations, Birkhäuser, Basel, 2001, pp. 3084.CrossRefGoogle Scholar
Grubb, G., Functional calculus of pseudo-differential boundary problems, Progress in Mathematics, 65, Birkhäuser, Boston, MA, 1986.CrossRefGoogle Scholar
Loya, P., Moroianu, S., and Ponge, R., On the singularities of the zeta and eta functions of an elliptic operator . Int. J. Math. 23(2012), no. 6, 1250020.CrossRefGoogle Scholar
Melrose, R. B., Calculus of conormal distributions on manifolds with corners . Int. Math. Res. Not. 3(1992), 5161.CrossRefGoogle Scholar
Melrose, R. B., The Atiyah–Patodi–Singer index theorem, Research Notes in Mathematics, 4, A K Peters, Ltd., Wellesley, MA, 1993.CrossRefGoogle Scholar
Melrose, R. B. and Mazzeo, R. R., Analytic surgery and the eta invariant . Geom. Funct. Anal. 5(1995), no. $1$ , 1475.Google Scholar
Minakshisundaram, S. and Pleijel, A., Some properties of the eigenfunctions of the Laplace operator on Riemannian manifolds . Can. J. Math. 1(1949), 242256.CrossRefGoogle Scholar
Paris, R. B. and Kaminski, D., Asymptotics and Mellin–Barnes integrals, Cambridge University Press, Cambridge, 2001.CrossRefGoogle Scholar
Whittaker, E. T. and Watson, G. N., A course of modern analysis, Cambridge University Press, Cambridge, 1965.Google Scholar