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

Do the Hodge spectra distinguish orbifolds from manifolds?
Part 1

Katie Gittins Department of Mathematical Sciences, Durham University, Mathematical Sciences & Computer Science Building, Upper Mountjoy Campus, Stockton Road, Durham DH1 3LE, United Kingdom. [email protected] Carolyn Gordon Department of Mathematics, Dartmouth College, Hanover, NH, 03755, USA. [email protected] Magda Khalile Institut für Analysis, Leibniz Universität Hannover, Welfengarten 1, 30167, Hannover, Germany. [email protected] Ingrid Membrillo Solis Mathematical Sciences, University of Southampton, University Road, Southampton SO17 1BJ, United Kingdom. [email protected] Mary Sandoval Department of Mathematics, Trinity College, 300 Summit Street, Hartford, CT, 06106, USA. [email protected]  and  Elizabeth Stanhope Department of Mathematical Sciences, Lewis & Clark College, Portland, OR, 97219, USA. [email protected]
Abstract.

We examine the relationship between the singular set of a compact Riemannian orbifold and the spectrum of the Hodge Laplacian on pp-forms by computing the heat invariants associated to the pp-spectrum. We show that the heat invariants of the 0-spectrum together with those of the 11-spectrum for the corresponding Hodge Laplacians are sufficient to distinguish orbifolds with singularities from manifolds as long as the singular sets have codimension 3.\leq 3. This is enough to distinguish orbifolds from manifolds for dimension 3.\leq 3.

Key words and phrases:
Hodge Laplacian, orbifolds, isospectrality
2020 Mathematics Subject Classification:
Primary: 58J53; Secondary: 53C20 58J50 58J37

Introduction

A (Riemannian) orbifold is a versatile generalization of a (Riemannian) manifold that permits the presence of well-structured singular points. Orbifolds appear in a variety of mathematical areas and have applications in physics, in particular to string theory.

Orbifolds are locally the orbit spaces of effective finite group actions on d\mathbb{R}^{d}. Orbifold singularities correspond to orbits with non-trivial isotropy and thus each singularity has an associated “isotropy type.” The notions of the Laplace-Beltrami operator and Hodge Laplacian extend to Riemannian orbifolds and many results on the spectral geometry of Riemannian manifolds, such as Weyl asymptotics and Cheeger’s inequality, extend to this more general setting.

Because the possible presence of singular points is a defining characteristic of the class of orbifolds, the question “Can one hear the singularities of an orbifold?” is fundamental in the study of the spectral geometry of orbifolds. More precisely, one asks:

  1. 1.

    Do spectral data distinguish orbifolds with singularities from manifolds?

  2. 2.

    Do spectral data detect the topology and geometry of the set of singular points including the isotropy types of singularities?

We will always assume all orbifolds under consideration are connected and compact without boundary.

Many authors have addressed these questions for the spectrum of the Laplace-Beltrami operator. However, the question of whether the spectrum of the Laplace-Beltrami operator always distinguishes Riemannian orbifolds with singularities from Riemannian manifolds, remains open. In this article, we focus primarily on question 1 and consider whether additional spectral data, specifically the spectra of the Hodge Laplacians on pp-forms, suffice to detect the presence of singularities. We will say that two closed Riemannian orbifolds are pp-isospectral if their Hodge Laplacians acting on pp-forms are isospectral. In particular, 0-isospectrality means that the Laplace-Beltrami operators are isospectral.

We construct the fundamental solution of the heat equation and the heat trace for the pp-spectrum on closed Riemannian orbifolds, analogous to that of the 0-spectrum in [DGGW08] discussed below, and then apply the resulting spectral invariants to question 1. Our main result is that the 0-spectrum and 1-spectrum together distinguish orbifolds with a sufficiently large singular set from manifolds. Orbifolds are stratified spaces in which the regular points form an open dense stratum, and each lower-dimensional stratum is a connected component of the set of singularities of a given isotropy type. By the codimension of the singular set, we mean the minimum codimension of the singular strata.

Theorem 1.1.

The 0-spectrum and 11-spectrum together distinguish closed Riemannian orbifolds with singular sets of codimension 3\leq 3 from closed Riemannian manifolds.

Consequently:

Corollary 1.2.

For d3d\leq 3, the 0-spectrum and 11-spectrum together distinguish singular closed dd-dimensional Riemannian orbifolds from closed Riemannian manifolds.

The case d=3d=3 is perhaps of particular interest since 3-dimensional orbifolds play a role in Thurston’s Geometrization Conjecture, later proven by Grigori Perel’man.

In a forthcoming paper (Part 2), joint with Juan Pablo Rossetti, we address both questions 1 and 2 for the individual pp-spectra. For example, we give conditions on the codimension of the singular set which guarantee that the volume of the singular set is determined by spectral data, and in many cases we show by providing counterexamples that the conditions are sharp.

1.1. Partial review of known results for the Laplace-Beltrami operator.

The earliest results in the spectral geometry of Riemannian orbifolds are those of Yuan-Jen Chiang [Chi93] who established the existence of the Laplace spectrum and heat kernel of a Riemannian orbifold, and Carla Farsi [Far01] who extended the Weyl law to the orbifold setting. Results from a variety of authors soon followed, including several examples of isospectral non-isometric orbifolds with a variety of properties. For example, Naveed Shams, the last author, and David Webb [SSW06] produced arbitrarily large finite families of mutually strongly isospectral orbifolds with non-isomorphic maximal isotropy groups of the same order. In particular, this implies that they are pp-isospectral for all pp. Shortly afterward, J. P. Rossetti, Dorothee Schueth, and Martin Weilandt [RSW08] produced examples of strongly isospectral orbifolds with maximal isotropy groups of distinct orders.

Emily Dryden and Alexander Strohmaier [DS09] proved that hyperbolic orientable closed 2-orbifolds are 0-isospectral if and only if they have the same geodesic length spectrum and the same number of cone points of each order. Benjamin Linowitz and Jeffrey Meyer [LM17] obtained interesting results for locally symmetric spaces. A series of papers extended to the orbifold setting the result of Robert Brooks, Peter Perry, and Peter Petersen [BPP92] that a set k(d)\mathcal{I}_{k}(d) of 0-isospectral dd-manifolds, sharing a uniform lower bound kk on sectional curvature, contains manifolds of only finitely many homeomorphism types (diffeomorphism types for d4d\neq 4). Specifically, the last author [Sta05] obtained an upper bound on the order of isotropy groups that can arise in elements of the set k(d)\mathcal{I}_{k}(d) (expanded to contain orbifolds), Emily Proctor and the last author [PS10] showed orbifold diffeomorphism finiteness for k(2)\mathcal{I}_{k}(2), Proctor [Pro12] obtained orbifold homeomorphism finiteness for k(d)\mathcal{I}_{k}(d) assuming only isolated singularities, and John Harvey [Har16] obtained orbifold homeomorphism finiteness for the full set k(d)\mathcal{I}_{k}(d). In addition, Farsi, Proctor and Christopher Seaton [FPS14] introduced and studied a stronger notion of isospectrality of orbifolds.

Recall that the trace of the heat kernel of the Laplace-Beltrami operator on a closed dd-dimensional Riemannian manifold M{M} is a primary source of spectral invariants. Denoting the spectrum by {λj}j=1\{\lambda_{j}\}_{j=1}^{\infty}, the heat trace yields an asymptotic expansion of the following form as t0+t\downarrow 0^{+}:

j=1eλjt(4πt)d/2k=0ak(M)tk.\sum_{j=1}^{\infty}e^{-\lambda_{j}t}\sim(4\pi t)^{-d/2}\sum_{k=0}^{\infty}a_{k}({M})\,t^{k}.

The coefficients aka_{k} are spectral invariants referred to as the heat invariants. The first two, for example, give the volume and the total scalar curvature of the Riemannian manifold.

A orbifold is said to be good if it is a global quotient 𝒪=Γ\M{\mathcal{O}}=\Gamma{\backslash}M of a manifold MM by a (possibly infinite) discrete group Γ\Gamma acting effectively and properly by diffeomorphisms. (We say MM is a cover of 𝒪{\mathcal{O}} even though the group action need not be free.) In [Don76], Harold Donnelly obtained an asymptotic expansion of the form

j=1eλjt(4πt)d/2k=0ck(𝒪)tk/2\sum_{j=1}^{\infty}e^{-\lambda_{j}t}\sim(4\pi t)^{-d/2}\sum_{k=0}^{\infty}c_{k}(\mathcal{{\mathcal{O}}})\,t^{k/2}

for the heat trace of a good Riemannian orbifold by adding contributions from each element γΓ\gamma\in\Gamma. Each ck(𝒪)c_{k}({\mathcal{O}}) is a sum of terms ck(𝒪,γ)c_{k}({\mathcal{O}},\gamma), γΓ\gamma\in\Gamma, which in turn is a sum of integrals over the connected components of the fixed point set of γ\gamma. The integrands are universal expressions in the germs of the metric and the isometry γ\gamma.

By applying Donnelly’s heat trace asymptotics for good orbifolds, Craig Sutton [Sut10] showed that if a closed Riemannian orbifold 𝒪{\mathcal{O}} with singularities and a closed Riemannian manifold MM have 0-isospectral finite Riemannian covers then they cannot be 0-isospectral. In particular if 𝒪{\mathcal{O}} and MM have a common finite Riemannian cover, they cannot be 0-isospectral. Juan Pablo Rossetti and the second author showed that the finiteness of the cover can be removed in the special case that the common cover is a homogeneous Riemannian manifold; see [GR03, Proposition 3.4] and its correction in the errata [GR21].

In [DGGW08], Dryden, the second author, Sarah Greenwald and Webb extended the heat trace asymptotics to the case of arbitrary (not necessarily good) closed Riemannian orbifolds expressing the invariants ck(𝒪)c_{k}({\mathcal{O}}) as sums of contributions from the various primary strata of the orbifold. (See Notation and Remarks 2.3 for the definition of primary.) As in the good case, the small-time heat trace expansion is of the form (4πt)d/2k=0ck(𝒪)tk/2(4\pi t)^{-d/2}\sum_{k=0}^{\infty}c_{k}(\mathcal{{\mathcal{O}}})\,t^{k/2}. Primary strata of even, respectively odd, codimension in 𝒪{\mathcal{O}} contribute to coefficients ck(𝒪)c_{k}({\mathcal{O}}) for kk even, odd respectively. As a consequence:

Theorem 1.3 ([DGGW08, Theorem 5.1]).

A closed Riemannian orbifold that contains a primary singular stratum of odd codimension cannot be 0-isospectral to any closed Riemannian manifold.

Sean Richardson and the last author [RS20] proved that an orbifold 𝒪{\mathcal{O}} is locally orientable if and only if 𝒪{\mathcal{O}} does not contain any primary singular strata of odd codimension. They thus concluded that the 0-spectrum determines local orientability of an orbifold. In addition, [DGGW08, Theorem 5.15] shows that the 0-spectrum distinguishes closed, locally orientable, 2-orbifolds with non-negative Euler characteristic from smooth, oriented, closed surfaces (in fact, a stronger result is proven there, see Remark 4.4).


This paper is organized as follows: Section 2 provides background on Riemannian orbifolds, their singular strata, differential forms and the Hodge Laplacian. In Section 3, we construct the heat kernel, heat trace, and heat invariants for pp-forms on orbifolds. Our construction follows the construction in the manifold case by Matthew Gaffney [Gaf58] and Vijay Kumar Patodi [Pat71], uses results of Donnelly and Patodi [DP77] for good orbifolds, and parallels the adaptations in [DGGW08]. Section 4 contains the proof of Theorem 1.1.

In an appendix, we outline the computation of the heat invariants for good orbifolds in [Don76] and [DP77], indicating aspects of the construction that are applicable to more general settings.

2. Riemannian Orbifold Background

2.1. Definitions and Basic Properties

In this section we recall the definition and basic properties of a Riemannian orbifold. For comprehensive information about orbifolds see the paper [Sco83] by Peter Scott, as well as the texts [Thu97] by William Thurston and [ALR07] by Alejandro Adem, Johann Leida, and Yongbin Ruan. Here we follow a somewhat abridged form of the presentation given in [Gor12] and [DGGW08].

Definition 2.1.

Let XX be a second countable Hausdorff space.

  1. 1.

    For a connected open subset UXU\subseteq X, an orbifold chart (of dimension dd) over UU is a triple (U~,GU,πU)(\widetilde{U},G_{U},\pi_{U}) where U~d\widetilde{U}\subseteq\mathbb{R}^{d} is a connected open subset, GUG_{U} is a finite group acting on U~\widetilde{U} effectively and by diffeomorphisms, and πU:U~X\pi_{U}\colon\widetilde{U}\to X is a map inducing a homeomorphism U~/GU\widetilde{U}/G\xrightarrow{\cong}U. The open set UU is sometimes referred to as the image of the chart.

  2. 2.

    If UVXU\subseteq V\subseteq X, an orbifold chart (U~,GU,πU)(\widetilde{U},G_{U},\pi_{U}) is said to inject into an orbifold chart (V~,GV,πV)(\widetilde{V},G_{V},\pi_{V}) if there exists a smooth embedding i:U~V~i\colon\widetilde{U}\to\widetilde{V} and a monomorphism λ:GUGV\lambda\colon G_{U}\to G_{V} such that πU=πVi\pi_{U}=\pi_{V}\circ i and iγ=λ(γ)ii\circ\gamma=\lambda(\gamma)\circ i for all γGU\gamma\in G_{U}.

  3. 3.

    Orbifold charts (U~,GU,πU)(\widetilde{U},G_{U},\pi_{U}) and (V~,GV,πV)(\widetilde{V},G_{V},\pi_{V}) on open sets UU and VV are said to be compatible if for every xUVx\in U\cap V, there exists a neighborhood WUVW\subset U\cap V of xx that admits an orbifold chart injecting into both (U~,GU,πU)(\widetilde{U},G_{U},\pi_{U}) and (V~,GV,πV)(\widetilde{V},G_{V},\pi_{V}). A dd-dimensional orbifold atlas on XX consists of a collection of compatible dd-dimensional orbifold charts whose images cover XX.

    An orbifold is a second countable Hausdorff topological space together with a maximal dd-dimensional atlas.

  4. 4.

    A Riemannian structure on an orbifold 𝒪{\mathcal{O}} is an assignment of a GUG_{U}-invariant Riemannian metric U~\widetilde{U} to each orbifold chart (U~,GU,πU)(\widetilde{U},G_{U},\pi_{U}), compatible in the sense that each injection of charts as in part 2 is an isometric embedding.

  5. 5.

    If an orbifold is the quotient space of a manifold under a smooth proper action of a discrete group it is called a good orbifold. Otherwise, it is called a bad orbifold.

Definition 2.2.

Let 𝒪\mathcal{O} be an orbifold and let x𝒪x\in\mathcal{O}.

  1. 1.

    We define the isotropy type of xx as follows: A chart (U~,GU,πU)(\widetilde{U},G_{U},\pi_{U}) about xx defines a smooth action of GUG_{U} on U~d\widetilde{U}\subset\mathbb{R}^{d}. Fix a lift x~U~\widetilde{x}\in\widetilde{U} of xx and let Iso(x~){\operatorname{{Iso}}}(\widetilde{x}) be the isotropy subgroup of GUG_{U} at x~\widetilde{x}. The map γdγx~End(Tx~U~)\gamma\mapsto d\gamma_{\widetilde{x}}\in\operatorname{End}(T_{\widetilde{x}}\widetilde{U}), defines an injective linear representation of Iso(x~){\operatorname{{Iso}}}(\widetilde{x}). Every finite-dimensional linear representation of a compact Lie group is equivalent to an orthogonal representation, unique up to orthogonal equivalence. Thus Iso(x~){\operatorname{{Iso}}}(\widetilde{x}) can be viewed as a subgroup of the orthogonal group O(d)O(d), unique up to conjugacy. The conjugacy class of Iso(x~){\operatorname{{Iso}}}(\widetilde{x}) in O(d)O(d) is independent both of the choice of the lift x~\widetilde{x} of xx in U~\widetilde{U} and of the choice of chart (U~,GU,πU)(\widetilde{U},G_{U},\pi_{U}) and is called the isotropy type of xx.

  2. 2.

    We say that xx is a regular point if it has trivial isotropy type, and a singular point otherwise.

We conclude our discussion of the basic properties of orbifolds by describing their singular stratification.

Notation and Remarks 2.3.

Let 𝒪\mathcal{O} be an orbifold.

  1. 1.

    Define an equivalence relation on 𝒪\mathcal{O} by saying that two points in 𝒪\mathcal{O} are isotropy equivalent if they have the same isotropy type. The connected components of the isotropy equivalence classes of 𝒪\mathcal{O} form a smooth stratification of 𝒪\mathcal{O}. The strata whose corresponding isotropy types are non-trivial are called singular strata. (In the literature, the requirement that a singular stratum be connected is sometimes dropped.)

  2. 2.

    Every orbifold chart (U~,GU,πU)(\widetilde{U},G_{U},\pi_{U}) on 𝒪\mathcal{O} also admits a smooth stratification whose strata are the connected components of the isotropy equivalence classes, where x~\widetilde{x} and y~\widetilde{y} in U~\widetilde{U} are defined to be isotropy equivalent if Iso(x~){\operatorname{{Iso}}}(\widetilde{x}) and Iso(y~){\operatorname{{Iso}}}(\widetilde{y}) are conjugate in GUG_{U}. We will refer to the strata of U~\widetilde{U} that have non-trivial isotropy as pre-singular strata and their union as the pre-singular set of U~\widetilde{U}. If NN is any stratum of 𝒪\mathcal{O} that intersects UU, then πU1(NU)\pi_{U}^{-1}(N\cap U) is a disjoint union of finitely many pre-singular strata of U~\widetilde{U} each of which maps diffeomorphically to NUN\cap U by πU\pi_{U}.

  3. 3.

    Let NN be a singular stratum in 𝒪\mathcal{O} and let Iso(N)<O(d)\operatorname{Iso}(N)<O(d) denote a representative of its isotropy type, unique up to conjugacy. We will refer to Iso(N)\operatorname{Iso}(N) as the isotropy group of the stratum. We denote by Isomax(N)\operatorname{Iso}^{\rm max}(N) the set of all elements γIso(N)\gamma\in\operatorname{Iso}(N) such that the dimension of the 11-eigenspace of γ\gamma is equal to the dimension of NN. We say that a stratum NN is primary if Isomax(N)\operatorname{Iso}^{\rm max}(N) is non-empty. (Item 4 below shows that this definition is equivalent to the one given in [DGGW17].)

  4. 4.

    Similarly, if WW is a pre-singular stratum of an orbifold chart (U~,GU,πU)(\widetilde{U},G_{U},\pi_{U}), we let Iso(W)\operatorname{Iso}(W) denote the common (up to conjugacy) isotropy group of its elements, we denote by Isomax(W)\operatorname{Iso}^{\rm max}(W) the subset of those γIso(W)\gamma\in\operatorname{Iso}(W) for which WW is open in the fixed point set of γ\gamma, and we say that WW is primary if Isomax(W)\operatorname{Iso}^{\rm max}(W) is non-empty. In particular, a pre-singular stratum WW of U~\widetilde{U} is primary if and only if πU(W)\pi_{U}(W) is a (necessarily open) subset of a primary singular stratum of 𝒪{\mathcal{O}}.

An orbifold 𝒪\mathcal{O} may contain some singular strata that are not primary. (See Example 2.16 in [DGGW17].) However, the union of the primary strata is dense in the singular set as the following proposition shows.

Proposition 2.4.

Let 𝒪\mathcal{O} be an orbifold. Then:

  1. 1.

    Every singular stratum of minimal codimension in 𝒪\mathcal{O} is necessarily primary.

  2. 2.

    Any component CC of the singular set that contains a non-primary singular stratum NN must also contain at least three primary singular strata.

Proof.

The first statement is immediate. To prove the second, it suffices to prove the analogous statement for the connected components of the pre-singular set of any orbifold chart (U~,GU,πU)(\widetilde{U},G_{U},\pi_{U}). Let CC be such a connected component that contains a non-primary pre-singular stratum WW. Each γIso(W)\gamma\in\operatorname{Iso}(W) must lie in Isomax(W)\operatorname{Iso}^{\rm max}(W^{\prime}) for some primary stratum WCW^{\prime}\subset C of lower codimension. Iso(W)Iso(W)\operatorname{Iso}(W)\nsubseteq\operatorname{Iso}(W^{\prime}) since codim(W)<codim(W)\operatorname{codim}(W^{\prime})<\operatorname{codim}(W), so the elements of Iso(W)\operatorname{Iso}(W) must be distributed among at least two primary strata WW^{\prime} and W′′W^{\prime\prime} of CC. In particular, we can choose γIso(W)Iso(W)\gamma^{\prime}\in\operatorname{Iso}(W)\cap\operatorname{Iso}(W^{\prime}) and γ′′Iso(W)Iso(W′′)\gamma^{\prime\prime}\in\operatorname{Iso}(W)\cap\operatorname{Iso}(W^{\prime\prime}) such that γ,γ′′Iso(W)Iso(W′′)\gamma^{\prime},\gamma^{\prime\prime}\notin\operatorname{Iso}(W^{\prime})\cap\operatorname{Iso}(W^{\prime\prime}). These conditions imply that γγ′′\gamma^{\prime}\gamma^{\prime\prime} cannot lie in either Iso(W)\operatorname{Iso}(W^{\prime}) or Iso(W′′)\operatorname{Iso}(W^{\prime\prime}) and thus there must be a third primary stratum W′′′W^{\prime\prime\prime} in CC. ∎

Example 2.5.

By considering all the finite subgroups of the orthogonal group O(2)O(2), we obtain a classification of the isotropy types of singular points that can occur in dimension 22:

  1. 1.

    A rotation about the origin in 2\mathbb{R}^{2} through an angle 2πk\frac{2\pi}{k}, for some integer k2k\geq 2, generates a finite cyclic group of order kk. The image of the origin in the quotient of 2\mathbb{R}^{2} under this action is an isolated singular point called a cone point of order kk.

  2. 2.

    A reflection across a line through the origin generates a cyclic group of order 22. The fixed points of the reflection correspond to a 11-dimensional singular stratum in the quotient called a mirror or reflector edge.

  3. 3.

    A dihedral group generated by a pair of reflections across lines forming an angle πk\frac{\pi}{k}, for 2k+2\leq k\in\mathbb{Z}^{+}, yields a 0-dimensional stratum in the quotient called a corner reflector or dihedral point. The corner reflector is not an isolated singular point, rather it is the point where two mirror edges intersect.

Example 2.6.

The direct product 𝒪×𝒪{\mathcal{O}}\times{\mathcal{O}}^{\prime} of two orbifolds is again an orbifold. For example, if 𝒪{\mathcal{O}} is a tear drop (an orbifold with a single cone point x0x_{0} and underlying topological space the 2-sphere) and MM is a closed manifold, then 𝒪×M{\mathcal{O}}\times M is a bad orbifold with a unique singular stratum {x0}×M\{x_{0}\}\times M of codimension 22.

Notation 2.7.

For γO(d)\gamma\in O(d), let rr be the multiplicity of the eigenvalue 1-1 if it occurs, and let {e±iθj}j=1s\{e^{\pm i\theta_{j}}\}_{j=1}^{s}, where θ1,,θs(0,π)\theta_{1},\dots,\theta_{s}\in(0,\pi), be those eigenvalues of γ\gamma with non-zero imaginary part, each repeated according to its multiplicity. The expression E(θ1,θ2,,θs;r)E(\theta_{1},\theta_{2},\dots,\theta_{s};r) will be called the eigenvalue type of γ\gamma. When r=0r=0 we write E(θ1,θ2,,θs;)E(\theta_{1},\theta_{2},\dots,\theta_{s};) and when s=0s=0 we write E(;r)E(;r).

The dimension of the +1+1 eigenspace of γ\gamma is thus d2srd-2s-r; in particular, the parity of rr determines the parity of the codimension of the fixed point set of γ\gamma in d\mathbb{R}^{d}.

Remark 2.8.

We take θjπ\theta_{j}\neq\pi for all 1js1\leq j\leq s for the notational convenience of letting rr count the total number of eigenvalues equal to 1-1. This convention can be dropped without affecting the results of Section 4.

2.2. Differential forms and the Hodge Laplacian

We give a brief description of how to define differential forms on orbifolds. For details of the constructions we refer to [CJ19] or [Wei12]. We take 𝒪\mathcal{O} to be a dd-dimensional orbifold throughout this section.

Definition 2.9.
  1. 1.

    As in the case of manifolds, a pp-form ω\omega on an orbifold 𝒪{\mathcal{O}} is defined to be a section of the pp-th exterior power of the cotangent bundle T(𝒪)T^{*}(\mathcal{O}). The cotangent bundle and its exterior powers are examples of orbibundles; see, for example, [CJ19] or [Gor12] for an expository introduction to orbibundles or the references therein for more in-depth explanations. We will only use the following: If U𝒪U\subset{\mathcal{O}} is an open set on which there exists a coordinate chart (U~,GU,πU)(\widetilde{U},G_{U},\pi_{U}), then any pp-form ω\omega on UU corresponds to a GUG_{U}-invariant pp-form ω~\widetilde{\omega} on U~\widetilde{U}, which we call the lift of ω\omega to U~{\widetilde{U}}. The differential form ω\omega is said to be of class CkC^{k} on UU if ω~\widetilde{\omega} is of class CkC^{k}. The compatibility condition on overlapping charts in Definition 2.1 part 3 leads to a natural compatibility condition on the lifts of pp-forms. One can use partitions of unity to construct globally defined pp-forms from pp-forms defined on an open cover of 𝒪{\mathcal{O}}.

  2. 2.

    For a Riemannian manifold MM, recall that the Hodge Laplacian Δp\Delta^{p} acting on the space of smooth pp-forms Ωp(M)\Omega^{p}(M) is given by

    (1) Δp:=(dδ+δd),\displaystyle\Delta^{p}:=-(d\delta+\delta d),

    where dd is the exterior derivative and δ\delta is the codifferential operator, i.e. the formal adjoint of dd obtained via the Riemannian metric. It is straightforward to check that the Hodge Laplacian commutes with isometries. The notion of Hodge Laplacian extends to Riemannian orbifolds 𝒪{\mathcal{O}} as follows: Let ω\omega be a smooth pp-form on 𝒪{\mathcal{O}}. On each coordinate chart (U~,GU,πU)(\widetilde{U},G_{U},\pi_{U}), the GUG_{U}-invariance of the lift ω~\widetilde{\omega} and the fact that the Hodge Laplacian on Riemannian manifolds commutes with isometries together imply that Δp(ω~)\Delta^{p}(\widetilde{\omega}) is GUG_{U}-invariant. We define Δp(ω)\Delta^{p}(\omega) on UU by

    Δp(ω)~=Δp(ω~).\widetilde{\Delta^{p}(\omega)}=\Delta^{p}(\widetilde{\omega}).

    We can again use the fact that isometries between Riemannian manifolds intertwine the Hodge Laplacians along with the compatibility condition on orbifold charts to conclude that Δp(ω)\Delta^{p}(\omega) is well-defined on 𝒪{\mathcal{O}}.

Remark 2.10.

The results in the present work apply to both orientable and nonorientable orbifolds and manifolds. In what follows the notation dVMdV_{M} indicates the Riemannian volume element in the orientable case, and the Riemannian density in the nonorientable case.

It is known, see e.g. [Buc99, Theorem 4.8.1], that the operator Δp\Delta^{p} viewed as an unbounded operator acting in Lp2(𝒪)L_{p}^{2}(\mathcal{O}), the space of differential pp-forms whose components are square-integrable functions, is essentially self-adjoint and has a purely discrete spectrum. By abuse of notation, in what follows we denote its closure by Δp\Delta^{p}, which is then self-adjoint with compact resolvent. The spectral theorem thus applies to Δp\Delta^{p} and we denote its eigenvalues by 0λ1λ2+0\leq\lambda_{1}\leq\lambda_{2}\leq\dots\to+\infty, with associated smooth eigenforms, denoted by (φi)i(\varphi_{i})_{i}, which form an orthonormal Lp2L_{p}^{2}-basis.

We conclude this section by reviewing the notion of multi pp-forms.

Notation 2.11.
  1. 1.

    Recall that if π:BM\pi:B\to M and π:BM\pi^{\prime}:B^{\prime}\to M^{\prime} are vector bundles over manifolds MM and MM^{\prime}, then the external tensor product BBM×MB\boxtimes B^{\prime}\to M\times M^{\prime} is the vector bundle whose fiber over the point (m,m)(m,m^{\prime}) is π1(m)π1(m)\pi^{-1}(m)\otimes\pi^{-1}(m^{\prime}). Given a CC^{\infty} manifold MM, let MkM^{k} denote the kk-fold Cartesian product M××MM\times\dots\times M. Sections of the kk-fold external tensor product k(pT(M))Mk\boxtimes^{k}(\wedge^{p}T^{*}(M))\to M^{k} are called multi pp-forms on MM of order kk. In particular, a multi pp-form of order 11 is simply a pp-form. Those of order 22 or 33 are called double or triple pp-forms, respectively. Thus for example, if (𝒰,x)(\mathcal{U},x) and (𝒱,y)(\mathcal{V},y) are local coordinate charts on MM and FF is a double pp-form on MM, then F|𝒰×𝒱F_{|\mathcal{U}\times\mathcal{V}} can be expressed as I,JaI,JdxIdyJ,\sum_{I,J}a_{I,J}dx^{I}\otimes dy^{J}, where aI,JC(𝒰×𝒱)a_{I,J}\in C^{\infty}(\mathcal{U}\times\mathcal{V}). (Here II and JJ vary over pp-tuples 1i1<<ipd=dim(M).1\leq i_{1}<\dots<i_{p}\leq d=\dim(M).)

  2. 2.

    Given smooth manifolds MM and NN and smooth maps γ:MM\gamma:M\to M and σ:NN\sigma:N\to N, denote by γ1,\gamma_{1,\cdot} and σ,2\sigma_{\cdot,2} the maps M×NM×NM\times N\to M\times N given by (a,b)(γ(a),b)(a,b)\mapsto(\gamma(a),b) and (a,b)(a,σ(b))(a,b)\mapsto(a,\sigma(b)). If M=NM=N, we also define γ1,2:M2M2\gamma_{1,2}:M^{2}\to M^{2} by (a,b)(γ(a),γ(b))(a,b)\mapsto(\gamma(a),\gamma(b)). We use analogous notation for maps of higher order Cartesian products. Thus, for example, the map γ1,,3:M3M3\gamma_{1,\cdot,3}:M^{3}\to M^{3} is given by (m1,m2,m3)(γ(m1),m2,γ(m3)).(m_{1},m_{2},m_{3})\mapsto(\gamma(m_{1}),m_{2},\gamma(m_{3})).

  3. 3.

    Analogous to the case of pp-forms in Definition 2.9, we can use the notion of orbibundle to extend the definition of multiple pp-forms directly to the orbifold setting, but we will not need the orbibundle formalism in what follows. If UU and VV are open sets in an orbifold 𝒪{\mathcal{O}} on which there exist orbifold coordinate charts (U~,GU,πU)(\widetilde{U},G_{U},\pi_{U}) and (V~,GV,πV)(\widetilde{V},G_{V},\pi_{V}), then the restriction to U×VU\times V of a double pp-form FF on 𝒪{\mathcal{O}} lifts to a section F~\widetilde{F} of pT(U~)pT(V~)\wedge^{p}T^{*}({\widetilde{U}})\boxtimes\wedge^{p}T^{*}(\widetilde{V}) that is invariant under pullback by each of the maps γ1,\gamma_{1,\cdot} and σ,2\sigma_{\cdot,2} for γGU\gamma\in G_{U} and σGV\sigma\in G_{V}. Again, the compatibility condition on overlapping charts in Definition 2.1 part 3 of an orbifold leads to a natural compatibility condition on the lifts of double pp-forms. One can use partitions of unity to construct double pp-forms on 𝒪{\mathcal{O}} from ones on an open cover of 𝒪×𝒪{\mathcal{O}}\times{\mathcal{O}}. Multiple pp-forms of any order are defined similarly.

Observe that if FF and HH are multi pp-forms, say of orders kk and \ell, respectively, then FHF\otimes H is a multi pp-form of order k+k+\ell.

Definition 2.12.

Recall that if V1,,VkV_{1},\dots,V_{k} are inner product spaces and if 1i<jk1\leq i<j\leq k with Vi=VjV_{i}=V_{j}, then the contraction Ci,j:V1VnV1Vi^Vj^VkC_{i,j}:V_{1}\otimes\cdots\otimes V_{n}\to V_{1}\otimes\cdots\otimes\widehat{V_{i}}\otimes\cdots\otimes\widehat{V_{j}}\otimes\cdots\otimes V_{k} is the linear map whose value on decomposable vectors is given by

α1αnαi,αjα1αi^αj^αk\alpha_{1}\otimes\cdots\otimes\alpha_{n}\mapsto\langle\alpha_{i},\alpha_{j}\rangle\alpha_{1}\otimes\cdots\otimes\widehat{\alpha_{i}}\otimes\cdots\otimes\widehat{\alpha_{j}}\otimes\cdots\otimes\alpha_{k}

where ,\langle\,,\,\rangle is the inner product on ViV_{i}. (Here we use hat notation such as Vi^\widehat{V_{i}} or αi^\widehat{\alpha_{i}} to indicate that the corresponding factor is removed from the product.)

In particular, let (M,g)(M,g) be a Riemannian manifold, FF a multi pp-form of order kk on MM, and suppose m=(m1,,mk)Mkm=(m_{1},\dots,m_{k})\in M^{k} satisfies mi=mjm_{i}=m_{j}. Then we can use the inner product on pTmi(M)\wedge^{p}T^{*}_{m_{i}}(M) to define the contraction Ci,j(F(m)).C_{i,j}(F(m)).

To define the contraction of such a multi pp-form in the setting of an orbifold 𝒪{\mathcal{O}} at a point m=(m1,,mk)𝒪km=(m_{1},\dots,m_{k})\in{\mathcal{O}}^{k} with mi=mjm_{i}=m_{j}, let F~\widetilde{F} be a local lift on a neighborhood U1××UkU_{1}\times\dots\times U_{k} as in Notation 2.11 part 3. Choose a lift m~=(m~1,,m~k)\widetilde{m}=(\widetilde{m}_{1},\dots,\widetilde{m}_{k}) in such a way that m~i=m~j\widetilde{m}_{i}=\widetilde{m}_{j}. We can then define Cij(F(m))C_{ij}(F(m)) by the condition that it lifts to Cij(F~(m~))C_{ij}(\widetilde{F}(\widetilde{m})). This definition is independent of the choice of m~\widetilde{m} (subject only to the condition m~i=m~j\widetilde{m}_{i}=\widetilde{m}_{j}), and it is independent of the choice of chart since the compatibility condition on charts respects the Riemannian structure.

3. Heat trace for differential forms on orbifolds

In Subsection 3.1 we will construct the fundamental solution of the heat equation for the Hodge Laplacian on arbitrary closed Riemannian orbifolds by first constructing a parametrix. Our construction follows and generalizes the construction in the manifold case by Gaffney [Gaf58] and Patodi [Pat71]. (The assumption of orientability in the work of Gaffney and Patodi is not needed.) In Subsection 3.2, we translate a result of Donnelly into our context and employ it to develop the small-time asymptotic expansion of the heat trace. Our arguments are similar to those used in [DGGW08] to construct the heat kernel and heat trace asymptotics for the Laplacian acting on functions on a closed Riemannian orbifold.

3.1. Heat kernel for pp-forms on manifolds and orbifolds

Definition 3.1.

We say that Kp:(0,)×𝒪×𝒪Λp(T𝒪)Λp(T𝒪)K^{p}:\mathbb{(}0,\infty)\times{\mathcal{O}}\times{\mathcal{O}}\to\Lambda^{p}(T^{*}{\mathcal{O}})\otimes\Lambda^{p}(T^{*}{\mathcal{O}}) is the heat kernel or fundamental solution of the heat equation for pp-forms if it satisfies

  1. 1.

    Kp(t,,)K^{p}(t,\cdot,\cdot) is a double pp-form for any t(0,)t\in(0,\infty);

  2. 2.

    KpK^{p} is continuous in the three variables, C1C^{1} in the first variable and C2C^{2} in the second variable;

  3. 3.

    (t+Δxp)Kp(t,x,y)=0(\partial_{t}+\Delta^{p}_{x})K^{p}(t,x,y)=0 for each fixed y𝒪y\in{\mathcal{O}}, where Δxp\Delta^{p}_{x} is the Hodge Laplacian with respect to the variable xx;

  4. 4.

    For every continuous pp-form ω\omega on 𝒪{\mathcal{O}} and for all x𝒪x\in{\mathcal{O}}, we have

    limt0+𝒪C2,3Kp(t,x,y)ω(y)𝑑V𝒪(y)=ω(x).\lim_{t\to 0^{+}}\,\int_{\mathcal{O}}\,C_{2,3}K^{p}(t,x,y)\otimes\omega(y)\,dV_{\mathcal{O}}(y)=\omega(x).

(Because the variable tt in the preceding definition is a real parameter, in contrast to a space variable, it is not counted when specifying indices in expressions using Notation 2.11 part 2, or in expressions involving contractions using the notation from Definition 2.12. Also, the variables over which a contraction is occuring will be repeated. For example, in part 4 above, C2,3Kp(t,x,y)ω(y)C_{2,3}K^{p}(t,x,y)\otimes\omega(y) indicates contraction in the second and third entries, not counting tt.)

The same argument as in the manifold case shows that if the heat kernel exists, then it is unique and given by

(2) Kp(t,x,y)=j=1φjφj(x,y)eλjtK^{p}(t,x,y)=\sum_{j=1}^{\infty}\varphi_{j}\otimes\varphi_{j}(x,y)e^{-\lambda_{j}t}

where {φj}j=1\{\varphi_{j}\}_{j=1}^{\infty} is an orthonormal basis of Lp2(𝒪)L_{p}^{2}({\mathcal{O}}) consisting of eigenfunctions and the λj\lambda_{j} are the corresponding eigenvalues as above. In particular, the heat kernel is invariant under any isometry γ\gamma, i.e.,

(3) (γ1,2Kp)(t,x,y)=Kp(t,x,y)(\gamma^{*}_{1,2}K^{p})(t,x,y)=K^{p}(t,x,y)

in the notation of Notation 2.11 part 2.

Definition 3.2.

A parametrix for the heat operator on pp-forms on 𝒪{\mathcal{O}} is a function H:(0,)×𝒪×𝒪Λp(T𝒪)Λp(T𝒪)H:(0,\infty)\times{\mathcal{O}}\times{\mathcal{O}}\to\Lambda^{p}(T^{*}{\mathcal{O}})\otimes\Lambda^{p}(T^{*}{\mathcal{O}}) satisfying:

  1. 1.

    H(t,,)H(t,\cdot,\cdot) is a double pp-form for each t(0,)t\in\mathbb{(}0,\infty);

  2. 2.

    HH is CC^{\infty} on (0,)×𝒪×𝒪(0,\infty)\times{\mathcal{O}}\times{\mathcal{O}};

  3. 3.

    (t+Δxp)H(t,x,y)(\partial_{t}+\Delta^{p}_{x})H(t,x,y) extends continuously to [0,)×𝒪×𝒪[0,\infty)\times{\mathcal{O}}\times{\mathcal{O}};

  4. 4.

    For every continuous pp-form ω\omega on 𝒪{\mathcal{O}} and for all x𝒪x\in{\mathcal{O}}, we have

    limt0+𝒪C2,3H(t,x,y)ω(y)𝑑V𝒪(y)=ω(x).\lim_{t\to 0^{+}}\,\int_{\mathcal{O}}\,C_{2,3}H(t,x,y)\otimes\omega(y)\,dV_{\mathcal{O}}(y)=\omega(x).

We first recall the construction of a parametrix by Patodi, see also Gaffney [Gaf58].

Proposition 3.3 ([Pat71]).

There exist smooth double pp-forms uipu^{p}_{i} for i=0,1,2,i=0,1,2,\dots defined on a neighborhood of the diagonal of M×MM\times M in any Riemannian manifold MM of dimension dd (more precisely, uip(x,y)u^{p}_{i}(x,y) is well-defined whenever yy is in a normal neighborhood about xx) satisfying the following.

  1. 1.

    If {ω1,,ωr}\{\omega_{1},\dots,\omega_{r}\}, where r=(dp)r=\binom{d}{p}, is an orthonormal basis of Λp(TxM)\Lambda^{p}\,(T^{*}_{x}M), then

    u0p(x,x)=j=1rωjωj.u_{0}^{p}(x,x)=\sum_{j=1}^{r}\,\omega_{j}\otimes\omega_{j}.

    (Note that this expression is independent of the choice of orthonormal basis). Equivalently under the identification of Λp(TxM)Λp(TxM)\Lambda^{p}\,(T^{*}_{x}M)\otimes\Lambda^{p}\,(T^{*}_{x}M) with End(Λp(TxM))\operatorname{End}(\Lambda^{p}\,(T^{*}_{x}M)), u0p(x,x)u^{p}_{0}(x,x) is the identity endomorphism of Λp(TxM)\Lambda^{p}(T_{x}^{*}M).

  2. 2.

    For each m=1,2,m=1,2,\dots, we have

    (t+Δxp)H(m)(t,x,y)=(4π)d/2ed2(x,y)/4ttmd2Δxpump(x,y),(\partial_{t}+\Delta^{p}_{x})H^{(m)}(t,x,y)=(4\pi)^{-d/2}e^{-d^{2}(x,y)/4t}t^{m-\frac{d}{2}}\Delta^{p}_{x}u^{p}_{m}(x,y),

    where

    H(m)(t,x,y):=(4πt)d/2ed2(x,y)/4t(u0p(x,y)++tmump(x,y)),H^{(m)}(t,x,y):={(4\pi t)^{-d/2}e^{-d^{2}(x,y)/4t}}{(u^{p}_{0}(x,y)+\dots+t^{m}u^{p}_{m}(x,y))},

    and d2(x,y)d^{2}(x,y) is the square of the Riemannian distance between xx and yy. In particular, if ψ:M×M\psi:M\times M\to\mathbb{R} is supported on a sufficiently small neighborhood of the diagonal and is identically one on a smaller neighborhood of the diagonal, then ψH(m)\psi H^{(m)} is a parametrix for the heat operator on pp-forms when m>d2m>\frac{d}{2}.

  3. 3.

    The uipu_{i}^{p} are the unique double pp-forms satisfying conditions 1 and 2. If γ\gamma is an isometry of an open set in the domain of uipu^{p}_{i}, then γ1,2uip=uip\gamma^{*}_{1,2}u^{p}_{i}=u^{p}_{i}.

Notation and Remarks 3.4.

Fix ϵ>0{\epsilon}>0 so that each x𝒪x\in{\mathcal{O}} is the center of a “convex geodesic ball” WW of radius ϵ\epsilon. By this we mean that W admits a chart (W~,GW,πW)(\widetilde{W},G_{W},\pi_{W}), where W~\widetilde{W} is a convex geodesic ball of radius ϵ\epsilon centered at the necessarily unique point x~{\widetilde{x}} satisfying πW(x~)=x\pi_{W}({\widetilde{x}})=x. (In particular, x~{\widetilde{x}} has isotropy GWG_{W}.) Cover 𝒪{\mathcal{O}} by finitely many such geodesic balls Wα{W_{\alpha}} with charts (W~α,Gα,πα)({\widetilde{W}_{\alpha}},G_{\alpha},\pi_{\alpha}), α=1,,s\alpha=1,\dots,s. (Here we write GαG_{\alpha} for GWαG_{{W_{\alpha}}} and πα\pi_{\alpha} for πWα\pi_{W_{\alpha}}.) Let U~αW~α{\widetilde{U}_{\alpha}}\subset{\widetilde{W}_{\alpha}}, respectively V~αW~α{\widetilde{V}_{\alpha}}\subset{\widetilde{W}_{\alpha}}, be the concentric geodesic balls of radius ϵ4\frac{{\epsilon}}{4}, respectively ϵ2\frac{{\epsilon}}{2}, and let Uα=πα(U~α){U_{\alpha}}=\pi_{\alpha}({\widetilde{U}_{\alpha}}) and Vα=πα(V~α){V_{\alpha}}=\pi_{\alpha}({\widetilde{V}_{\alpha}}). We may assume that the family of balls {Uα}1αs\{{U_{\alpha}}\}_{1\leq{\alpha}\leq s} still covers 𝒪{\mathcal{O}}.

For each α{\alpha} and each nonnegative integer mm, we define

H~α(m):(0,)×W~α×W~αΛp(W~α)Λp(W~α){\widetilde{H}^{(m)}_{\alpha}}:(0,\infty)\times{\widetilde{W}_{\alpha}}\times{\widetilde{W}_{\alpha}}\rightarrow\Lambda^{p}({\widetilde{W}_{\alpha}})\otimes\Lambda^{p}({\widetilde{W}_{\alpha}})

by

(4) H~α(m)(t,x~,y~):=(4πt)d/2ed(x~,y~)2/4t(u0p(x~,y~)++tmump(x~,y~)),\displaystyle{\widetilde{H}^{(m)}_{\alpha}}(t,{\widetilde{x}},{\widetilde{y}}):={(4\pi t)^{-d/2}e^{-d({\widetilde{x}},{\widetilde{y}})^{2}/4t}}{(u^{p}_{0}({\widetilde{x}},{\widetilde{y}})+\dots+t^{m}u^{p}_{m}({\widetilde{x}},{\widetilde{y}}))},

where the uipu^{p}_{i} are the double pp-forms defined in Proposition 3.3. Since each γGα{\gamma}\in{G_{\alpha}} is an isometry of W~α{\widetilde{W}_{\alpha}}, we have γ1,2uip=uip\gamma^{*}_{1,2}u^{p}_{i}=u^{p}_{i}.

It follows that the double pp-form (depending on the parameter tt)

γGαγ,2H~α(m)\sum_{\gamma\in{G_{\alpha}}}\gamma^{*}_{\cdot,2}{\widetilde{H}^{(m)}_{\alpha}}

is Gα{G_{\alpha}}-invariant in both x~{\widetilde{x}} and y~{\widetilde{y}} and thus descends to a well-defined function on (0,)×Wα×Wα(0,\infty)\times{W_{\alpha}}\times{W_{\alpha}}, which we denote by Hα(m){H^{(m)}_{\alpha}}.

Let ψα:𝒪{\psi_{\alpha}}:{\mathcal{O}}\to\mathbb{R} be a CC^{\infty} cut-off function which is identically one on Vα{V_{\alpha}} and is supported in Wα{W_{\alpha}}. Let {ηα:α=1,,s}\{{\eta_{\alpha}}:{\alpha}=1,\dots,s\} be a partition of unity on 𝒪{\mathcal{O}} with the support of ηα{\eta_{\alpha}} contained in Uα¯\overline{{U_{\alpha}}}. Define L(m){L^{(m)}} on (0,)×𝒪×𝒪(0,\infty)\times{\mathcal{O}}\times{\mathcal{O}} by

(5) L(m)(t,x,y):=α=1sψα(x)ηα(y)Hα(m)(t,x,y).{L^{(m)}}(t,x,y):=\sum_{{\alpha}=1}^{s}\,{\psi_{\alpha}}(x){\eta_{\alpha}}(y){H^{(m)}_{\alpha}}(t,x,y).
Proposition 3.5.

L(m){L^{(m)}} is a parametrix for the heat kernel on 𝒪{\mathcal{O}} when m>d2m>\frac{d}{2}.

Moreover, the extension of (t+Δxp)L(m)(t,x,y){\left(\partial_{t}+\Delta^{p}_{x}\right)}{L^{(m)}}(t,x,y) to [0,)×𝒪×𝒪[0,\infty)\times{\mathcal{O}}\times{\mathcal{O}} is of class CkC^{k} if m>d2+km>\frac{d}{2}+k for any kk\in\mathbb{N}.

Proof.

The proof is similar to that carried out in [DGGW08] (see also references therein) for the case of the Laplacian acting on functions. The first three properties of a parametrix and the final statement of the proposition are straightforward. We include here the proof of the last property of a parametrix (the reproducing property).

Let ω\omega be a continuous pp-form on 𝒪{\mathcal{O}}. Let ψ~α{\widetilde{\psi}_{\alpha}} and η~α{\widetilde{\eta}_{\alpha}} be the lifts of ψα{\psi_{\alpha}} and ηα{\eta_{\alpha}} to W~α{\widetilde{W}_{\alpha}}. We consider the lifts so as to make use of Proposition 3.3 part 2 and we also use the fact that the properties of a parametrix hold locally (see [DGGW08, Remark 3.7]).

Since supp(ηα)Uα¯Wα{\rm supp}({\eta_{\alpha}})\subset\overline{{U_{\alpha}}}\subset{W_{\alpha}}, we have

(6) πα𝒪ψα(x)ηα(y)C2,3Hα(m)(t,x,y)ω(y)𝑑y\displaystyle\pi_{\alpha}^{*}\int_{\mathcal{O}}{\psi_{\alpha}}(x){\eta_{\alpha}}(y)C_{2,3}{H^{(m)}_{\alpha}}(t,x,y)\otimes\omega(y)\,dy
=ψα(x)|Gα|γGαW~αη~α(y~)C2,3γ.,2H~α(m)(t,x~,y~)ω~α(y~)𝑑y~,\displaystyle=\frac{{\psi_{\alpha}}(x)}{|G_{\alpha}|}\sum_{{\gamma}\in{G_{\alpha}}}\int_{\widetilde{W}_{\alpha}}\,{\widetilde{\eta}_{\alpha}}({\widetilde{y}})C_{2,3}{\gamma}_{.,2}^{*}{\widetilde{H}^{(m)}_{\alpha}}(t,{\widetilde{x}},{\widetilde{y}})\otimes\widetilde{\omega}_{\alpha}({\widetilde{y}})\,d{\widetilde{y}},

where ω~α\widetilde{\omega}_{\alpha} is the pull-back of ω|W~α\omega_{|{\widetilde{W}_{\alpha}}} to W~α{\widetilde{W}_{\alpha}} and x~{\widetilde{x}} is an arbitrarily chosen point in the preimage of xx under the map W~αWα{\widetilde{W}_{\alpha}}\to{W_{\alpha}}. We change variables in each of the integrals in the right-hand side of Equation (6), letting u~=γ1(y~){\widetilde{u}}={\gamma}^{-1}({\widetilde{y}}). Since γ{\gamma} is an isometry and since η~α{\widetilde{\eta}_{\alpha}} and ω~α\widetilde{\omega}_{\alpha} are γ{\gamma}-invariant, each integral in the summand is equal to

(7) W~αη~α(u~)C2,3H~α(m)(t,x~,u~)ω~α(u~)𝑑u~.\int_{\widetilde{W}_{\alpha}}\,{\widetilde{\eta}_{\alpha}}({\widetilde{u}})C_{2,3}{\widetilde{H}^{(m)}_{\alpha}}(t,{\widetilde{x}},{\widetilde{u}})\otimes\widetilde{\omega}_{\alpha}({\widetilde{u}})\,d{\widetilde{u}}.

As t0+{t\to 0^{+}}, the integral (7) above converges to η~α(x~)ω~α(x~)=ηα(x)πα(ω(x)){\widetilde{\eta}_{\alpha}}({\widetilde{x}})\widetilde{\omega}_{\alpha}({\widetilde{x}})={\eta_{\alpha}}(x)\pi_{\alpha}^{*}(\omega(x)). Noting that ψα1{\psi_{\alpha}}\equiv 1 on the support of ηα{\eta_{\alpha}}, it follows that both sides of Equation (6) converge to ηα(x)πα(ω(x)){\eta_{\alpha}}(x)\pi_{\alpha}^{*}(\omega(x)) as t0+{t\to 0^{+}}. Since both sides of Equation (6) are identically zero when xx lies outside of supp(ψα)Wα{\rm supp}({\psi_{\alpha}})\subset{W_{\alpha}}, the left-hand side converges to ηα(x)πα(ω(x)){\eta_{\alpha}}(x)\pi_{\alpha}^{*}(\omega(x)) for each fixed x𝒪x\in{\mathcal{O}}. Thus

limt0+𝒪ψα(x)ηα(y)C2,3Hα(m)(t,x,y)ω(y)𝑑y=ηα(x)ω(x),\lim_{t\to 0^{+}}\int_{\mathcal{O}}{\psi_{\alpha}}(x){\eta_{\alpha}}(y)C_{2,3}{H^{(m)}_{\alpha}}(t,x,y)\otimes\omega(y)\,dy={\eta_{\alpha}}(x)\omega(x),

and by Equation (5) we have

limt0+𝒪L(m)(t,x,y)ω(y)𝑑y=αηα(x)ω(x)=ω(x).\lim_{t\to 0^{+}}\int_{\mathcal{O}}\,{L^{(m)}}(t,x,y)\omega(y)\,dy=\sum_{\alpha}\,{\eta_{\alpha}}(x)\omega(x)=\omega(x).\qed
Notation 3.6.
  1. 1.

    For AA and BB continuous double pp-form valued functions on [0,)×𝒪×𝒪[0,\infty)\times{\mathcal{O}}\times{\mathcal{O}}, we define the convolution ABA*B on (0,)×𝒪×𝒪(0,\infty)\times{\mathcal{O}}\times{\mathcal{O}} as

    AB(t,x,y):=0t𝑑θ𝒪C2,4A(θ,x,z)B(tθ,y,z)𝑑V𝒪(z).A*B(t,x,y):=\int_{0}^{t}\,d\theta\int_{\mathcal{O}}\,C_{2,4}A(\theta,x,z)\otimes B(t-\theta,y,z)dV_{\mathcal{O}}(z).
  2. 2.

    Fix m>d2+2m>\frac{d}{2}+2. Define κ0(t,x,y):=(t+Δx)L(m)(t,x,y)\kappa_{0}(t,x,y):={\left(\frac{\partial}{\partial t}+\Delta_{x}\right)}{L^{(m)}}(t,x,y) and, for j=1,2,j=1,2,\dots, set κj(t,x,y):=κ0κj1\kappa_{j}(t,x,y):=\kappa_{0}*\kappa_{j-1}.

An argument analogous to that in the manifold case [Gaf58] shows that the series

(8) Pm(t,x,y):=j=1(1)j+1κj(t,x,y)P_{m}(t,x,y):=\sum_{j=1}^{\infty}\,(-1)^{j+1}\kappa_{j}(t,x,y)

converges uniformly and absolutely on [0,T]×𝒪×𝒪[0,T]\times{\mathcal{O}}\times{\mathcal{O}} for each T>0T>0. Thus PmP_{m} is continuous. Moreover, PmP_{m} is of class C2C^{2} on +×𝒪×𝒪{\mathbb{R}_{+}}\times{\mathcal{O}}\times{\mathcal{O}} and for any T>0T>0, there exists a constant CC such that

(9) Pm(t,x,y)Ct2P_{m}(t,x,y)\leq Ct^{2}

on [0,T]×𝒪×𝒪[0,T]\times{\mathcal{O}}\times{\mathcal{O}}. (Our indexing is different from that in [Gaf58]. Our κj\kappa_{j} is Gaffney’s κj+1\kappa_{j+1}.)

We are now able to state the main theorem of this section, which follows in the same way as that in [Gaf58].

Theorem 3.7.

Let m>d2+2m>\frac{d}{2}+2. Define PmP_{m} as in Equation (8) and let

Kp(t,x,y):=L(m)(t,x,y)+(PmL(m))(t,x,y).K^{p}(t,x,y):={L^{(m)}}(t,x,y)+(P_{m}*{L^{(m)}})(t,x,y).

Then KpK^{p} is a heat kernel. Moreover,

Kp(t,x,y)=L(m)(t,x,y)+O(tmd2+1), as t0+.K^{p}(t,x,y)={L^{(m)}}(t,x,y)+O(t^{m-\frac{d}{2}+1}),\text{ as }t\to 0^{+}.
Notation 3.8.

Let

H~α(t,x~,y~)=γGα(4πt)d/2ed(x~,γ(y~))2/4t(γ.,2u0p(x~,y~)+tγ.,2u1p(x~,y~)+).{\widetilde{H}_{\alpha}}(t,{\widetilde{x}},{\widetilde{y}})=\sum_{{\gamma}\in{G_{\alpha}}}{(4\pi t)^{-d/2}e^{-d({\widetilde{x}},{\gamma}({\widetilde{y}}))^{2}/4t}}{({\gamma}_{.,2}^{*}u^{p}_{0}({\widetilde{x}},{\widetilde{y}})+t{\gamma}_{.,2}^{*}u^{p}_{1}({\widetilde{x}},{\widetilde{y}})+\dots)}.

Observe that H~α{\widetilde{H}_{\alpha}} is Gα{G_{\alpha}}-invariant in both x~{\widetilde{x}} and y~{\widetilde{y}} and thus is the pull-back of a double pp-form valued function, which we denote by Hα{H_{\alpha}}, on (0,)×Wα×Wα(0,\infty)\times{W_{\alpha}}\times{W_{\alpha}}.

As a consequence of Theorem 3.7, we have:

Theorem 3.9.

In the notation of Equation (5) and Notation 3.8, the trace of the heat kernel has an asymptotic expansion as t0+{t\to 0^{+}} given by

trKp(t):=𝒪C1,2Kp(t,x,x)𝑑V𝒪(x)α=1s𝒪ηα(x)C1,2Hα(t,x,x)𝑑V𝒪(x).\operatorname{tr}K^{p}(t):=\int_{\mathcal{O}}\,C_{1,2}K^{p}(t,x,x)\,dV_{\mathcal{O}}(x)\sim\,\sum_{{\alpha}=1}^{s}\,\int_{\mathcal{O}}\,{\eta_{\alpha}}(x)C_{1,2}{H_{\alpha}}(t,x,x)\,dV_{\mathcal{O}}(x).

3.2. Heat invariants for pp-forms on orbifolds

We will use Theorem 3.9 together with a theorem of Donnelly [Don76, Theorem 4.1] (see also [DP77, Theorem 3.1]) to compute the small-time asymptotics of the heat trace for the Laplacian on pp-forms on closed Riemannian orbifolds. Our presentation is similar to that in [DGGW08] where the case p=0p=0 is carried out.

Notation and Remarks 3.10.
  1. 1.

    Let 𝒯\mathcal{T} be the set of all triples (M,γ,a)(M,\gamma,a) where MM is a Riemannian manifold, γ\gamma is an isometry of MM and aa is in the fixed point set Fix(γ){\operatorname{{Fix}}}(\gamma) of γ\gamma. A function h:𝒯h:\mathcal{T}\to\mathbb{R} is said to satisfy the locality property if h(M,γ,a)h(M,\gamma,a) depends only on the germs at aa of the Riemannian metric and of γ\gamma. The function hh is said to be universal if whenever σ:M1M2\sigma:M_{1}\to M_{2} is an isometry between two Riemannian manifolds and γi\gamma_{i} is an isometry of MiM_{i} with γ2=σγ1σ1\gamma_{2}=\sigma\circ\gamma_{1}\circ\sigma^{-1}, then h(M1,γ1,a)=h(M2,γ2,σ(a))h(M_{1},\gamma_{1},a)=h(M_{2},\gamma_{2},\sigma(a)) for all aFix(γ1)a\in{\operatorname{{Fix}}}(\gamma_{1}).

  2. 2.

    We will also consider functions on the set 𝒯\mathcal{T}^{\prime} of all (M,η,γ,a)(M,\eta,\gamma,a) where (M,γ,a)(M,\gamma,a) is given as in part 1 and η\eta is a γ\gamma-invariant function on MM. The locality and universality properties are analogously defined. (In the definition of locality, h(M,η,γ,a)h(M,\eta,\gamma,a) may also depend on the germ of η\eta at aa. In the definition of of universality, we have h(M1,η1,γ1,a)=h(M2,η2,γ2,σ(a))h(M_{1},\eta_{1},\gamma_{1},a)=h(M_{2},\eta_{2},\gamma_{2},\sigma(a)) if η2σ=η1\eta_{2}\circ\sigma=\eta_{1} and the other conditions in part 1 hold.)

  3. 3.

    Let MM be a Riemannian manifold of dimension dd and let γ:MM\gamma:M\to M be an isometry. Then each component of the fixed point set Fix(γ){\operatorname{{Fix}}}(\gamma) is a totally geodesic, closed submanifold of MM. (See [Kob58].) If MM is compact, then Fix(γ){\operatorname{{Fix}}}(\gamma) has only finitely many components. In any Riemannian manifold, at most one component of Fix(γ){\operatorname{{Fix}}}(\gamma) can intersect any given geodesically convex ball. For aFix(γ)a\in{\operatorname{{Fix}}}(\gamma), note that the differential dγa:TaMTaMd\gamma_{a}:T_{a}M\to T_{a}M fixes TaQT_{a}Q and restricts to an isomorphism of Ta(Q)T_{a}(Q)^{\perp}, where QQ is the connected component of Fix(γ){\operatorname{{Fix}}}(\gamma) containing aa. In particular, we have det(Idcodim(Q)Aa)0\det({\operatorname{{Id}}}_{\operatorname{codim}(Q)}-A_{a})\neq 0, where AaA_{a} is the restriction of the action of dγad\gamma_{a} to (TaQ)(T_{a}Q)^{\perp} and codim(Q)\operatorname{codim}(Q) denotes the codimension of QQ.

  4. 4.

    For α\alpha in the orthogonal group O(d)O(d), we denote by trp(α)\operatorname{tr}_{p}(\alpha) the trace of the natural action of α\alpha on p(d)\wedge^{p}(\mathbb{R}^{d}). For MM and γ\gamma as in part 1, the Riemannian inner product on TaMT_{a}M allows us to identify dγad\gamma_{a} with an element of O(d)O(d) and trp(dγa)\operatorname{tr}_{p}(d\gamma_{a}) coincides with the trace of γa:p(TaM)p(TaM)\gamma_{a}^{*}:\wedge^{p}(T_{a}^{*}M)\to\wedge^{p}(T_{a}^{*}M).

  5. 5.

    Let MM be a compact Riemannian manifold and γ:MM\gamma:M\to M be an isometry. We denote the set of connected components of the fixed point set of γ\gamma by 𝒞(Fix(γ))\mathcal{C}({\operatorname{{Fix}}}(\gamma)).

Theorem 3.11.


  1. 1.

    [Don76, Theorem 4.1]; [DP77, Theorem 3.1]. In the notation of Notation 3.10, there exist functions bkb_{k}, k=0,1,2,k=0,1,2,\dots, on 𝒯\mathcal{T} satisfying both the locality and universality properties such that if MM is a compact Riemannian manifold and γ:MM\gamma:M\to M is an isometry, then, as t0+t\rightarrow 0^{+},

    MC1,2γ,2Kp(t,x,x)𝑑VM(x)Q𝒞(Fix(γ))(4πt)dim(Q)/2k=0tkQbkp(γ,a)𝑑VQ(a).\int_{M}C_{1,2}\gamma^{*}_{\cdot,2}K^{p}(t,x,x)\,dV_{M}(x)\sim\sum_{Q\in\mathcal{C}({\operatorname{{Fix}}}(\gamma))}(4\pi t)^{-\dim(Q)/2}\sum_{k=0}^{\infty}t^{k}\int_{Q}b^{p}_{k}(\gamma,a)\,dV_{Q}(a).

    (Here we are writing bk(γ,a)b_{k}(\gamma,a) for bk(M,γ,a)b_{k}(M,\gamma,a).) Moreover,

    b0p(γ,a)=trp(dγa)|det(Idcodim(Q)Aa)|.b^{p}_{0}(\gamma,a)=\frac{\operatorname{tr}_{p}(d\gamma_{a})}{\lvert\det({\operatorname{{Id}}}_{\operatorname{codim}(Q)}-A_{a})\rvert}.
  2. 2.

    (Local version.) There exist functions ckc_{k}, k=0,1,2,k=0,1,2,\dots, on 𝒯\mathcal{T}^{\prime} satisfying both the locality and universality properties such that the following condition holds: if MM is an arbitrary Riemannian manifold, γ\gamma is an isometry of MM, and η\eta is a γ\gamma-invariant function on MM whose support is compact and geodesically convex, then as t0+t\to 0^{+}, we have

    M(4πt)d/2ed2(x,γ(x))4tη(x)(j=0tjC1,2γ,2ujp(x,x))𝑑VM(x)Q𝒞(Fix(γ))(4πt)dim(Q)/2k=0tkQckp(η,γ,a)𝑑VQ(a),\int_{M}\,(4\pi t)^{-d/2}e^{-\frac{d^{2}(x,\gamma(x))}{4t}}\eta(x)\left(\sum_{j=0}^{\infty}t^{j}C_{1,2}\gamma^{*}_{\cdot,2}u^{p}_{j}(x,x)\right)\,dV_{M}(x)\\ \sim\sum_{Q\in\mathcal{C}({\operatorname{{Fix}}}(\gamma))}(4\pi t)^{-\dim(Q)/2}\sum_{k=0}^{\infty}t^{k}\int_{Q}c^{p}_{k}(\eta,\gamma,a)\,dV_{Q}(a),

    where the ujpu^{p}_{j}, j=1,2,j=1,2,\dots, are the double pp-forms defined in Proposition 3.3. (We are writing ck(η,γ,a)c_{k}(\eta,\gamma,a) for ck(M,η,γ,a).c_{k}(M,\eta,\gamma,a). Note that there is at most one component QQ of Fix(γ){\operatorname{{Fix}}}(\gamma) that intersects the support of η\eta and thus at most one non-zero term in the sum.)

    Moreover, the dependence on η\eta is linear. In particular, if η1\eta\equiv 1 near aa, then ckp(η,γ,a)=bkp(γ,a)c_{k}^{p}(\eta,\gamma,a)=b_{k}^{p}(\gamma,a).

Remark 3.12.

In the first statement in the previous theorem, we have omitted the hypotheses in [DP77] that MM is orientable and that γ\gamma is orientation-preserving as they are not needed for the proof.

The proof of the second statement is a minor adaptation of Donnelly’s proof of the first statement. In the appendix, we will give an exposition of the proof, making clear that aspects of the statement apply to much more general integrals.

Recall that for closed Riemannian manifolds, the heat trace has a small-time asymptotic expansion

(10) trKp(t):=MC1,2Kp(t,m,m)𝑑VM(m)t0+(4πt)d/2i=0aip(M)ti,\operatorname{tr}K^{p}(t):=\int_{M}C_{1,2}K^{p}(t,m,m)\,dV_{M}(m)\sim_{t\to 0^{+}}(4\pi t)^{-d/2}\sum_{i=0}^{\infty}a^{p}_{i}(M)t^{i},

where

aip(M):=MC1,2uip(m,m)𝑑V(m).a^{p}_{i}(M):=\int_{M}C_{1,2}u^{p}_{i}(m,m)\,dV(m).

The aipa_{i}^{p} are called the heat invariants.

The first two heat invariants for pp-forms on manifolds are given by

(11) a0p(M)\displaystyle a_{0}^{p}({M}) =(dp)vol(M),\displaystyle=\binom{d}{p}{\rm vol}({M}),
(12) a1p(M)\displaystyle a_{1}^{p}({M}) =(16(dp)(d2p1))Mτ(m)𝑑VM(m),\displaystyle=\left(\frac{1}{6}\binom{d}{p}-\binom{d-2}{p-1}\right)\int_{{M}}\tau(m)\,dV_{M}(m),

where τ\tau is the scalar curvature, and we use the convention that (mn)=0\binom{m}{n}=0 whenever n<0n<0 or n>mn>m. (See, for example, [Pat70] and references therein.)

We now address the heat invariants for the Hodge Laplacian on closed Riemannian orbifolds.

Notation 3.13.

Let 𝒪{\mathcal{O}} be a closed Riemannian orbifold of dimension dd.

  1. 1.

    For k=0,1,2,k=0,1,2,\dots, define UkpC(𝒪)U_{k}^{p}\in C^{\infty}({\mathcal{O}}) as follows: For x𝒪x\in{\mathcal{O}}, choose an orbifold chart (U~,GU,πU)(\widetilde{U},G_{U},\pi_{U}) about xx, let x~{\widetilde{x}} be any lift of xx in U~\widetilde{U} and set Ukp(x):=C1,2ukp(x~,x~)U_{k}^{p}(x):=C_{1,2}u_{k}^{p}({\widetilde{x}},{\widetilde{x}}) where ukpu_{k}^{p} is the double pp-form defined in Proposition 3.3. This definition is independent of both the choice of chart and the choice of lift x~{\widetilde{x}} since the ukpu_{k}^{p} are local isometry invariants.

    Define

    (13) akp(𝒪):=𝒪Ukp(x)𝑑V𝒪(x).a_{k}^{p}({\mathcal{O}}):=\int_{\mathcal{O}}\,U_{k}^{p}(x)\,dV_{\mathcal{O}}(x).

    One easily verifies that the resulting expressions for the akp(𝒪)a_{k}^{p}({\mathcal{O}}) are identical to those for manifolds. E.g., a0p(𝒪)=(dp)vol(O)a_{0}^{p}({{\mathcal{O}}})=\binom{d}{p}{\rm vol}({O}) and a1p(𝒪)a_{1}^{p}({\mathcal{O}}) is given by Equation (12) with MM replaced by 𝒪{\mathcal{O}}.

    Set

    I0p(t):=(4πt)d/2k=0akp(𝒪)tk.I_{0}^{p}(t):=(4\pi t)^{-d/2}\sum_{k=0}^{\infty}\,a_{k}^{p}({\mathcal{O}})t^{k}.
  2. 2.

    Let NN be a stratum of the singular set of 𝒪{\mathcal{O}}. For γIsomax(N)\gamma\in\operatorname{Iso}^{\rm max}(N), define bkp(γ,):Nb_{k}^{p}(\gamma,\cdot):N\to\mathbb{R} as follows: For aNa\in N, let (U~,GU,πU)(\widetilde{U},G_{U},\pi_{U}) be an orbifold chart about aa in 𝒪{\mathcal{O}} and let a~\tilde{a} be a lift of aa in U~\widetilde{U}. Note that a~\tilde{a} lies in a pre-singular stratum WW of U~\widetilde{U} (see Notation and Remarks 2.3) and γ\gamma is naturally identified with an element of Isomax(W)\operatorname{Iso}^{\rm max}(W). We define

    bkp(γ,a):=bkp(γ,a~),b_{k}^{p}(\gamma,a):=b_{k}^{p}(\gamma,\tilde{a}),

    where the right-hand side is defined as in Theorem 3.11 part 1. The universality of the functions bkp:𝒯b_{k}^{p}:\mathcal{T}\to\mathbb{R} (as stated in Theorem 3.11 part 1) implies that the left-hand side is independent both of the choice of chart and of the choice of lift a~\tilde{a}. Define bkp(N,):Nb_{k}^{p}(N,\cdot):N\to\mathbb{R} by

    bkp(N,a):=γIsomax(N)bkp(γ,a),b_{k}^{p}(N,a):=\sum_{\gamma\in\operatorname{Iso}^{\rm max}(N)}\,b_{k}^{p}(\gamma,a),

    and set

    (14) bkp(N):=γIsomax(N)Nbkp(γ,a)𝑑VN(a),\displaystyle b_{k}^{p}(N):=\sum_{\gamma\in\operatorname{Iso}^{\rm max}(N)}\,\int_{N}\,b_{k}^{p}(\gamma,a)\,dV_{N}(a),

    where dVNdV_{N} is the volume element on NN for the Riemannian metric on NN induced by that of 𝒪{\mathcal{O}}.

    Aside: this notation differs slightly from that in [DGGW08] where the case p=0p=0 is studied.

    Set

    INp(t):=(4πt)dim(N)/2k=0bkp(N)tk.I_{N}^{p}(t):=(4\pi t)^{-\dim(N)/2}\sum_{k=0}^{\infty}\,b_{k}^{p}(N)t^{k}.

    Observe that INp(t)=0I_{N}^{p}(t)=0 if NN is not a primary stratum.

Remark 3.14.

Since each stratum NN consists of points of a fixed isotropy type, the expression for b0p(γ,a)b_{0}^{p}(\gamma,a) in Theorem 3.11 part 1 is independent of aNa\in N and we denote it by b0p(γ)b_{0}^{p}(\gamma). In the notation of Equation (14), we have

(15) b0p(N)=vol(N)γIsomax(N)b0p(γ).b_{0}^{p}(N)={\rm vol}(N)\sum_{\gamma\in\operatorname{Iso}^{\rm max}(N)}\,b_{0}^{p}(\gamma).
Theorem 3.15.

Let 𝒪{\mathcal{O}} be a closed dd-dimensional Riemannian orbifold, let p{1,,d}p\in\{1,\dots,d\} and let 0λ1λ2+0\leq\lambda_{1}\leq\lambda_{2}\leq\dots\to+\infty be the spectrum of the Hodge Laplacian acting on smooth pp-forms on 𝒪{\mathcal{O}}. The heat trace has an asymptotic expansion as t0+t\to 0^{+} given by

(16) trKp(t)t0+I0p(t)+NS(𝒪)INp(t)|Iso(N)|,\operatorname{tr}K^{p}(t)\sim_{t\to 0^{+}}\,I_{0}^{p}(t)+\sum_{N\in S({\mathcal{O}})}\,\frac{I_{N}^{p}(t)}{|\operatorname{Iso}(N)|},

where S(𝒪)S({\mathcal{O}}) is the set of all singular 𝒪{\mathcal{O}}-strata and where |Iso(N)||\operatorname{Iso}(N)| is the order of the isotropy group of NN. This asymptotic expansion is of the form

(17) (4πt)d/2j=0cjp(𝒪)tj2\displaystyle(4\pi t)^{-d/2}\sum_{j=0}^{\infty}\,c^{p}_{j}(\mathcal{O})\,t^{\frac{j}{2}}

with cjp(𝒪)c^{p}_{j}(\mathcal{O})\in\mathbb{R}.

Observe that if there are no singular strata, then Equation (16) agrees with Equation (10).

Proof.

By Theorem 3.9, we can write

(18) trKp(t)t0+α=1s𝒪ηα(x)C1,2Hα(t,x,x)𝑑x.\operatorname{tr}K^{p}(t)\sim_{{t\to 0^{+}}}\,\sum_{{\alpha}=1}^{s}\,\int_{\mathcal{O}}\,{\eta_{\alpha}}(x)C_{1,2}{H_{\alpha}}(t,x,x)\,dx.

Recalling the notation introduced in Equation (4), Equation (18), we have

(19) trKp(t)t0+α=1s1|Gα|γGαL(t,α,γ)\operatorname{tr}K^{p}(t)\sim_{{t\to 0^{+}}}\sum_{{\alpha}=1}^{s}\frac{1}{|G_{\alpha}|}\sum_{{\gamma}\in{G_{\alpha}}}L(t,\alpha,\gamma)

where

L(t,α,γ):=U~α(4πt)d/2ed2(x~,γ(x~))/4tη~α(x~)C1,2(γ.,2u0p(x~,x~)+tγ.,2u1p(x~,x~)+)𝑑VU~α(x~).L(t,\alpha,\gamma):=\int_{{\widetilde{U}_{\alpha}}}(4\pi t)^{-d/2}e^{-d^{2}({\widetilde{x}},{\gamma}({\widetilde{x}}))/4t}{\widetilde{\eta}_{\alpha}}({\widetilde{x}})\,C_{1,2}\Big{(}{\gamma}_{.,2}^{*}u^{p}_{0}({\widetilde{x}},{\widetilde{x}})+t{\gamma}_{.,2}^{*}u^{p}_{1}({\widetilde{x}},{\widetilde{x}})+\cdots\Big{)}dV_{{\widetilde{U}_{\alpha}}}({\widetilde{x}}).

We will group together various terms in this double sum. First consider the identity element 1α1_{\alpha} of each GαG_{\alpha}. By Notation 3.13 part 1, we have

α=1s1|Gα|L(t,α,1α)=(4πt)d/2𝒪(U0p(x)+tU1p(x)+)𝑑V𝒪(x)=I0p(t).\sum_{{\alpha}=1}^{s}\frac{1}{|G_{\alpha}|}L(t,{\alpha},1_{\alpha})=(4\pi t)^{-d/2}\int_{\mathcal{O}}(U_{0}^{p}(x)+tU_{1}^{p}(x)+\cdots)\,dV_{\mathcal{O}}(x)=I_{0}^{p}(t).

Next let

(20) I(t):=α=1s1|Gα|1αγGαL(t,α,γ).I^{\prime}(t):=\sum_{{\alpha}=1}^{s}\frac{1}{|G_{\alpha}|}\sum_{1_{\alpha}\neq{\gamma}\in{G_{\alpha}}}L(t,\alpha,\gamma).

It remains to show that

(21) I(t)=NS(𝒪)INp(t)|Iso(N)|.I^{\prime}(t)=\sum_{N\in S({\mathcal{O}})}\frac{I_{N}^{p}(t)}{|\operatorname{Iso}(N)|}.

We will apply Theorem 3.11 to each term in I(t)I^{\prime}(t). For each α{\alpha} and each nontrivial element γGα\gamma\in G_{\alpha}, Theorem 3.11 (with U~α{\widetilde{U}_{\alpha}} playing the role of MM in Theorem 3.11 part 2) expresses L(t,α,γ)L(t,{\alpha},\gamma) as a sum of integrals over the various components of the fixed point set of γ\gamma in U~α{\widetilde{U}_{\alpha}}. Each such component is a union of pre-singular strata in U~α{\widetilde{U}_{\alpha}}. Since the union of those pre-singular strata WW in U~α{\widetilde{U}_{\alpha}} for which γIsomax(W)\gamma\in\operatorname{Iso}^{\rm max}(W) is an open subset of Fix(γ){\operatorname{{Fix}}}(\gamma) of full measure, we can instead add up the integrals over such pre-singular strata. Letting S(U~α)S({\widetilde{U}_{\alpha}}) denote the pre-singular strata of U~α{\widetilde{U}_{\alpha}} and applying Theorem 3.11, we have

(22) I(t)=α=1sWS(U~α)γIsomax(W)1|Gα|(4πt)dim(W)/2k=0tkWckp(η~α,γ,a~)𝑑VW(a~).I^{\prime}(t)=\sum_{{\alpha}=1}^{s}\sum_{W\in S({\widetilde{U}_{\alpha}})}\,\sum_{{\gamma}\in\operatorname{Iso}^{\rm max}(W)}\,\frac{1}{|G_{\alpha}|}(4\pi t)^{-\dim(W)/2}\sum_{k=0}^{\infty}\,t^{k}\int_{W}\,c_{k}^{p}({\widetilde{\eta}_{\alpha}},\gamma,\tilde{a})\,dV_{W}(\tilde{a}).

Let NS(𝒪)N\in S({\mathcal{O}}). For aNa\in N, set

ckp(ηα,γ,a):=ckp(η~α,γ,a~)c_{k}^{p}({\eta_{\alpha}},\gamma,a):=c_{k}^{p}({\widetilde{\eta}_{\alpha}},\gamma,\tilde{a})

where a~\tilde{a} is any lift of aa in U~α{\widetilde{U}_{\alpha}}. This is well-defined since any two lifts differ by an element of GαG_{\alpha} and η~α{\widetilde{\eta}_{\alpha}} is GαG_{\alpha}-invariant. Moreover, Theorem 3.11 part 2 implies that

(23) α=1sckp(ηα,γ,a)=bkp(γ,a).\sum_{{\alpha}=1}^{s}c_{k}^{p}({\eta_{\alpha}},\gamma,a)=b_{k}^{p}(\gamma,a).

For each α\alpha and each WS(U~α)W\in S({\widetilde{U}_{\alpha}}), there exists NS(𝒪)N\in S({\mathcal{O}}) such that πα\pi_{\alpha} maps WW isometrically onto NUαN\cap U_{\alpha}. Let

(24) S(U~α;N)={WS(U~α):πα(W)=NUα}S({\widetilde{U}_{\alpha}};N)=\{W\in S({\widetilde{U}_{\alpha}}):\pi_{\alpha}(W)=N\cap U_{\alpha}\}

and observe that

(25) |S(U~α;N)|=|Gα||Iso(N)|.|S({\widetilde{U}_{\alpha}};N)|=\frac{|G_{\alpha}|}{|{\operatorname{{Iso}}}(N)|}.

For WS(U~α;N)W\in S({\widetilde{U}_{\alpha}};N), we identify Isomax(N)\operatorname{Iso}^{\rm max}(N) with Isomax(W)\operatorname{Iso}^{\rm max}(W). Observe that

(26) Wckp(η~α,γ,a~)𝑑VW(a~)=NUαckp(ηα,γ,a)𝑑VN(a).\int_{W}\,c_{k}^{p}({\widetilde{\eta}_{\alpha}},\gamma,\tilde{a})\,dV_{W}(\tilde{a})=\int_{N\cap U_{\alpha}}\,c_{k}^{p}({\eta_{\alpha}},\gamma,a)\,dV_{N}(a).

Since dim(W)=dim(N)\dim(W)=\dim(N), Equations (22), (23), (24), (25) and (26) yield

I(t)\displaystyle I^{\prime}(t) =NS(𝒪)α=1s|Gα||Iso(N)|γIsomax(N)1|Gα|(4πt)dim(N)/2k=0tkNUαckp(ηα,γ,a)𝑑VN(a)\displaystyle=\sum_{N\in S({\mathcal{O}})}\sum_{{\alpha}=1}^{s}\frac{|G_{\alpha}|}{|{\operatorname{{Iso}}}(N)|}\sum_{{\gamma}\in\operatorname{Iso}^{\rm max}(N)}\,\frac{1}{|G_{\alpha}|}(4\pi t)^{-\dim(N)/2}\sum_{k=0}^{\infty}\,t^{k}\int_{N\cap U_{\alpha}}\,c_{k}^{p}({\eta_{\alpha}},\gamma,a)dV_{N}(a)
=NS(𝒪)1|Iso(N)|γIsomax(N)(4πt)dim(N)/2k=0tkNbkp(γ,a)𝑑VN(a)=NS(𝒪)INp(t)|Iso(N)|.\displaystyle=\sum_{N\in S({\mathcal{O}})}\frac{1}{|{\operatorname{{Iso}}}(N)|}\sum_{{\gamma}\in\operatorname{Iso}^{\rm max}(N)}\,(4\pi t)^{-\dim(N)/2}\sum_{k=0}^{\infty}\,t^{k}\int_{N}\,b_{k}^{p}(\gamma,a)\,dV_{N}(a)=\sum_{N\in S({\mathcal{O}})}\,\frac{I_{N}^{p}(t)}{|\operatorname{Iso}(N)|}.

The theorem follows. ∎

Remark 3.16.

Let 𝒪{\mathcal{O}} be a closed Riemannian orbifold all of whose singular strata NN have co-dimension one; i.e., the singular strata are mirrors. Every such stratum is necessarily totally geodesic in 𝒪{\mathcal{O}}. (Indeed, about any point aa in NN, there exists a coordinate chart (U~,GU,πU)({\widetilde{U}},G_{U},\pi_{U}) where GUG_{U} is generated by a reflection τ\tau, and the Riemannian metric on U=π(U~)U=\pi({\widetilde{U}}) arises from a τ\tau-invariant metric on U~{\widetilde{U}}. The fixed point set of any involutive isometry is necessarily totally geodesic.) The underlying topological space 𝒪¯\underline{{\mathcal{O}}} has the structure of a smooth Riemannian manifold each of whose boundary components is totally geodesic.

The orbifold 𝒪{\mathcal{O}} is good in this case, as we may double 𝒪{\mathcal{O}} over its reflectors to obtain a smooth Riemannian manifold MM admitting a reflection symmetry τ\tau so that 𝒪=τ\M{\mathcal{O}}=\langle\tau\rangle\backslash M. Let P:M𝒪¯P:M\to\underline{{\mathcal{O}}} be the projection. Identify the fixed point set of τ\tau with NN. Let ω\omega be a smooth pp-form on the orbifold 𝒪{\mathcal{O}}. Then ω~:=Pω\widetilde{\omega}:=P^{*}\omega is τ\tau invariant. Let ν\nu be a unit normal vector field along NN in MM and let j:NMj:N\to M be the inclusion. The τ\tau-invariance of ω~\widetilde{\omega} is equivalent to the conditions jινω~=jινdω~=0j^{*}\iota_{\nu}\widetilde{\omega}=j^{*}\iota_{\nu}d\widetilde{\omega}=0. Using the same notation ν\nu for the unit normal PνP_{*}\nu along NN in 𝒪{\mathcal{O}} and now letting j:N𝒪j:N\to{\mathcal{O}} be the inclusion, this gives

jινω=jινdω=0,j^{*}\iota_{\nu}\omega=j^{*}\iota_{\nu}d\omega=0,

and these are precisely the absolute boundary conditions for the manifold 𝒪¯\underline{{\mathcal{O}}}. The spectrum of the Hodge Laplacian for pp-forms on 𝒪{\mathcal{O}} thus coincides with the spectrum of the Hodge Laplacian for pp-forms on 𝒪¯\underline{{\mathcal{O}}} with absolute boundary conditions. It is then straightforward to prove that the spectral invariants bkp(N)b^{p}_{k}(N) computed here for the orbifold agree with the familiar contributions to the heat trace asymptotics arising from the boundary of 𝒪¯\underline{{\mathcal{O}}} as obtained in [Par04, (2) Theorem 3.2] (see also [BG90, Theorem 1.2]).

4. Proof of Theorem 1.1

In this section we use the asymptotic results derived in the previous section to prove our main theorem. For various types of singular strata NN in a dd-dimensional closed orbifold 𝒪\mathcal{O}, we first compute the invariant b0p(γ)b_{0}^{p}(\gamma) for γIsomax(N)\gamma\in\operatorname{Iso}^{\rm max}(N).

Proposition 4.1.

Let NN be a singular stratum of codimension kk in the dd-dimensional closed orbifold 𝒪{\mathcal{O}}. Suppose γIsomax(N)\gamma\in\operatorname{Iso}^{\rm max}(N) has eigenvalue type E(θ1,θ2,,θs;r)E(\theta_{1},\theta_{2},\dots,\theta_{s};r), using Notation 2.7.

Then

(27) b01(γ)=(dkr+j=1s 2cos(θj))(2kj=1scsc2(θj/2)).b_{0}^{1}(\gamma)=\left(d-k-r+\sum_{j=1}^{s}\,2\cos(\theta_{j})\right)\left(2^{-k}\prod_{j=1}^{s}\,\csc^{2}(\theta_{j}/2)\right).

Here we use the convention that when s=0s=0, we have j=1scsc2(θj/2)=1\prod_{j=1}^{s}\,\csc^{2}(\theta_{j}/2)=1.

Proof.

We apply Theorem 3.11. The expression in the first set of large parentheses in Equation (27) is the trace of γ\gamma. The expression in the second set of large parentheses equals 1|det(Idcodim(N)A)|\displaystyle\frac{1}{|\det({\operatorname{{Id}}}_{\operatorname{codim}(N)}-A)|}, where (taking note of Notation 3.13 and Notation and Remarks 3.10), we are writing AA to denote the expression AaA_{a} in Theorem 3.11 part 1. ∎

In preparation for the proof of Theorem 1.1, we consider primary singular strata of codimension two with cyclic isotropy groups.

Proposition 4.2.

Let NN be a stratum of codimension two with cyclic isotropy group of order m2m\geq 2. Then,

(28) b01(N)=((d2)m2112+m26m+56)vol(N).b_{0}^{1}(N)=\left((d-2)\frac{m^{2}-1}{12}+\frac{m^{2}-6m+5}{6}\right){\rm vol}(N).
Proof.

If m=2m=2, then the generator γ\gamma of Iso(N){\operatorname{{Iso}}}(N) must have eigenvalue type E(;2)E(;2). If m3m\geq 3, Iso(N){{\operatorname{{Iso}}}(N)} is generated by an element γ\gamma with eigenvalue type E(2π/m;)E(2\pi/m;). All elements of Iso(N){{\operatorname{{Iso}}}(N)} other than the identity lie in Isomax(N)\operatorname{Iso}^{\rm max}(N). The proposition then follows from Notation 3.13 part 2, Proposition 4.1, and the following two formulas:

j=1m1csc2(πj/m)=m213andj=1m1cos(2πj/m)csc2(πj/m)=m26m+53,\sum_{j=1}^{m-1}\csc^{2}(\pi j/m)=\frac{m^{2}-1}{3}\ \ \ \ \text{and}\ \ \ \ \sum_{j=1}^{m-1}\cos(2\pi j/m)\csc^{2}(\pi j/m)=\frac{m^{2}-6m+5}{3},

assuming here that 2m2\leq m\in\mathbb{Z}. These formulas are proved using the calculus of residues. See [DGGW08, Lemma 5.4] for the first formula. For the second, by [CS12, Equation (2.3)], we have

S3(m,1,1)\displaystyle S_{3}(m,1,1) =j=1m1cos(2πj/m)csc2(πj/m)\displaystyle=\sum_{j=1}^{m-1}\cos(2\pi j/m)\csc^{2}(\pi j/m)
=2α=01(22α)B2α(1/m)B22α(2)(1)m2α\displaystyle=2\sum_{\alpha=0}^{1}\binom{2}{2\alpha}B_{2\alpha}(1/m)B_{2-2\alpha}^{(2)}(1)m^{2\alpha}
=2B0(1/m)B2(2)(1)+2B2(1/m)B0(2)(1)m2=m26m+53,\displaystyle=2B_{0}(1/m)B_{2}^{(2)}(1)+2B_{2}(1/m)B_{0}^{(2)}(1)m^{2}=\frac{m^{2}-6m+5}{3},

where Bn(x)B_{n}(x) are the ordinary Bernoulli polynomials and Bn(m)(x)B_{n}^{(m)}(x) are the higher-order Bernoulli polynomials in xx (see [CS12] and references therein).

Example 4.3.

Although not required for our main result, one can generalize Proposition 4.2 to compute b01(N)b_{0}^{1}(N) for strata for which Iso(N){{\operatorname{{Iso}}}(N)} is cyclic and

  • when m>2m>2, the generator has eigenvalue type E(2π/m,,2π/m;)E(2\pi/m,\dots,2\pi/m;) with 2π/m2\pi/m repeated ss times, or

  • when m=2m=2, the generator has eigenvalue type E(;r)E(;r) with rr even.

Consider the case that NN has codimension 44 in 𝒪{\mathcal{O}}. By [BY02, Equation (5.2) and Corollary 5.4], we have

(29) j=1m1cos(2πj/m)csc4(πj/m)=m420m2+1945.\sum_{j=1}^{m-1}\cos(2\pi j/m)\csc^{4}(\pi j/m)=\frac{m^{4}-20m^{2}+19}{45}.

From [BY02, Equation (5.2) and Corollary 5.2], we have that

(30) j=1m1d4sin4(πj/m)=(d4)(m4+10m211)45.\sum_{j=1}^{m-1}\frac{d-4}{\sin^{4}(\pi j/m)}=\frac{(d-4)(m^{4}+10m^{2}-11)}{45}.

Thus by Proposition 4.1,

b01(N)=vol(N)4(m420m2+19)+(d4)(m4+10m211)720.b_{0}^{1}(N)={\rm vol}(N)\frac{4(m^{4}-20m^{2}+19)+(d-4)(m^{4}+10m^{2}-11)}{720}.

Reference [BY02] also contains formulas analogous to Equations (29) and (30) with the power 44 replaced by the power 2s2s for arbitrary positive ss, thus yielding expressions for b01(N)b_{0}^{1}(N) for rotation strata of constant rotation angle and higher codimension.

Proof of Theorem 1.1.

If the singular set of 𝒪{\mathcal{O}} has odd codimension, then Theorem 1.3 and the fact that strata of minumum codimension are primary imply that the 0-spectrum alone suffices to distinguish 𝒪{\mathcal{O}} from any Riemannian manifold. Thus we may assume that the singular set has codimension two. Let S2(𝒪)S_{2}({\mathcal{O}}) denote the set of all strata of codimension two.

Let MM be a closed Riemannian manifold of dimension dd. In the notation of Equation (17), we have

(31) c2p(M)=a1p(M)whilec2p(𝒪)=a1p(𝒪)+NS2(𝒪)4π|Iso(N)|b0p(N).c_{2}^{p}(M)=a_{1}^{p}(M)\,\,\,\mbox{while}\,\,\,c_{2}^{p}({\mathcal{O}})=a_{1}^{p}({\mathcal{O}})+\sum_{N\in S_{2}({\mathcal{O}})}\,\frac{4\pi}{|\operatorname{Iso}(N)|}b_{0}^{p}(N).

For =M*=M or 𝒪{\mathcal{O}}, we have

(32) a11()=d66τ𝑑V=(d6)a10().a_{1}^{1}(*)=\frac{d-6}{6}\int_{*}\tau\,dV_{*}=(d-6)a_{1}^{0}(*).

We will show that

(33) b01(N)>(d6)b00(N)b_{0}^{1}(N)>(d-6)b_{0}^{0}(N)

for every NS2(𝒪)N\in S_{2}({\mathcal{O}}). Theorem 1.1 will then follow from Theorem 3.15 and Equations (31), (32) and (33).

For NS2(𝒪)N\in S_{2}({\mathcal{O}}), we have Iso(N)A×Id2\operatorname{Iso}(N)\simeq A\times I_{d-2} where AO(2)A\subset O(2) and Id2I_{d-2} is the (d2)×(d2)(d-2)\times(d-2) identity matrix. Since 𝒪{\mathcal{O}} contains no strata of codimension one, AA cannot contain any reflections and thus NN has cyclic isotropy group of order mm contained in SO(2)×Id2SO(2)\times I_{d-2}. The computation of b00(N)b_{0}^{0}(N) is analogous to the case of cone points in dimension 2, carried out in [DGGW08, Proposition 5.5], yielding

b00(N)=(m21)12vol(N).b_{0}^{0}(N)=\frac{(m^{2}-1)}{12}{\rm vol}(N).

Using Proposition 4.2, we have

b01(N)=((d2)m2112+m26m+56)vol(N).b_{0}^{1}(N)=\left((d-2)\frac{m^{2}-1}{12}+\frac{m^{2}-6m+5}{6}\right){\rm vol}(N).

Thus, noting that m2m\geq 2, we have

b01(N)(d6)b00(N)=3m26m+36>0.b_{0}^{1}(N)-(d-6)b_{0}^{0}(N)=\frac{3m^{2}-6m+3}{6}>0.

This proves Equation (33) and the theorem follows. ∎

Remark 4.4.

It is shown in [DGGW08, Theorem 5.15] that the 0-spectrum alone suffices to distinguish singular, closed, locally orientable Riemannian orbisurfaces with nonnegative Euler characteristic from smooth, oriented, closed Riemannian surfaces. In fact, it is shown there that the degree zero term in the small-time heat trace asymptotics for functions gives rise to a complete topological invariant for orbisurfaces satisfying these constraints. In Theorem 1.1, by also appealing to the 11-spectrum, we do not require local orientability or nonnegative Euler characteristic to distinguish Riemannian orbisurfaces from closed Riemannian surfaces.

5. Appendix

We outline the proof of Theorem 3.11. We are closely following the proof of [Don76, Theorem 4.1] but expressing it in a more general context.

Let MM be an arbitrary Riemannian manifold and γ\gamma an isometry of MM. Let d(x,y)d(x,y) denote the induced distance function between xx and yy in MM. Outside of a small tubular neighborhood of Fix(γ){\operatorname{{Fix}}}(\gamma), we have that (4πt)d/2ed2(x,γ(x))4t0(4\pi t)^{-d/2}e^{-\frac{d^{2}(x,\gamma(x))}{4t}}\to 0 uniformly as t0+t\to 0^{+}. Thus for any compactly supported smooth function ff on MM, we have

M(4πt)d/2ed2(x,γ(x))4tf(x)𝑑VM(x)\displaystyle\int_{M}\,(4\pi t)^{-d/2}e^{-\frac{d^{2}(x,\gamma(x))}{4t}}f(x)\,dV_{M}(x)
=Q𝒞(Fix(γ))UQ(4πt)d/2ed2(x,γ(x))4tf(x)𝑑VM(x)+O(ec/t),\displaystyle\qquad=\sum_{Q\in\mathcal{C}({\operatorname{{Fix}}}(\gamma))}\int_{U_{Q}}\,(4\pi t)^{-d/2}e^{-\frac{d^{2}(x,\gamma(x))}{4t}}f(x)\,dV_{M}(x)+O(e^{-c/t}),

where UQU_{Q} is a small tubular neighborhood of QQ, say of radius rr, and c>0c>0 is a constant, where, as in Notation and Remarks 3.10 part 5, 𝒞(Fix(γ))\mathcal{C}({\operatorname{{Fix}}}(\gamma)) denotes the set of connected components of Fix(γ).{\operatorname{{Fix}}}(\gamma).

We will later specialize to the choices of ff that arise in Theorem 3.11, but for now we work in the general setting.

As in [Don76, Theorem 4.1], we analyze each term in the sum above individually and define

IQ(f):=UQ(4πt)d/2ed2(x,γ(x))4tf(x)𝑑VM(x).I_{Q}(f):=\int_{U_{Q}}\,(4\pi t)^{-d/2}e^{-\frac{d^{2}(x,\gamma(x))}{4t}}f(x)\,dV_{M}(x).

Let WQTMW_{Q}\subset TM be the normal disk bundle of QQ of radius rr and let φ:WQQ\varphi:W_{Q}\to Q be the bundle projection. We use the diffeomorphism WQUQW_{Q}\to U_{Q} given by xexpφ(x)x{\rm{x}}\mapsto\exp_{\varphi({\rm{x}})}{\rm{x}}, where exp\exp is the Riemannian exponential map of MM, to express IQ(f)I_{Q}(f) as

IQ(f)=Qφ1(a)(4πt)d/2ed2(x,γ(x))4tf(x)ψ(x)𝑑x𝑑VQ(a).I_{Q}(f)=\int_{Q}\int_{\varphi^{-1}(a)}(4\pi t)^{-d/2}e^{-\frac{d^{2}(x,\gamma(x))}{4t}}f(x)\,\psi({\rm{x}})\,d{\rm{x}}\,dV_{Q}(a).

Here dxd{\rm{x}} is the Euclidean volume element defined by the Riemannian structure on (TaQ)(T_{a}Q)^{\perp}, and for xφ1(a){\rm{x}}\in\varphi^{-1}(a), we are writing x=expa(x)x=\exp_{a}({\rm{x}}). The function ψ\psi arising in this change of variables depends only on the curvature and its covariant derivatives and, as shown in [Don76, Equation (2.6)], satisfies

(34) ψ(x)=112Riαjαxixj16Rikjkxixj+O(x3),\psi({\rm{x}})=1-\frac{1}{2}R_{i\alpha j\alpha}{\rm{x}}^{i}{\rm{x}}^{j}-\frac{1}{6}R_{ikjk}{\rm{x}}^{i}{\rm{x}}^{j}+O({\rm{x}}^{3}),

where the xi{\rm{x}}^{i} are the coordinates of x{\rm{x}} with respect to an orthonormal basis of (TaQ)(T_{a}Q)^{\perp} and where the components of the curvature tensor are evaluated at the point aa.

Letting x¯=xdγa(x)\bar{{\rm{x}}}={\rm{x}}-d\gamma_{a}({\rm{x}}), Donnelly shows that d2(x,γ(x))=(x¯)2+O(x¯)3d^{2}(x,\gamma(x))=(\bar{{\rm{x}}})^{2}+O(\bar{{\rm{x}}})^{3} and then applies a version of the Morse Lemma, see [Don76, Lemma A.1], to find new coordinates y{\rm y} so that

d2(x,γ(x))=i=1syi2=y2,d^{2}(x,\gamma(x))=\sum_{i=1}^{s}{\rm y}^{2}_{i}=\|{\rm y}\|^{2},

Making the two changes of variables xx¯y{\rm{x}}\to\bar{{\rm{x}}}\to{\rm y} on (TaQ)(T_{a}Q)^{\perp}, we have

dx=|det(B(a))|dx¯=|det(B(a))||J(x¯,y)|dy,d{\rm{x}}=|\det(B(a))|\,d\bar{{\rm{x}}}=|\det(B(a))||J(\bar{{\rm{x}}},{\rm y})|d{\rm y},

where

(35) B(a)=(Idcodim(Q)Aa)1.B(a)=({\operatorname{{Id}}}_{\operatorname{codim}(Q)}-A_{a})^{-1}.

Here AaA_{a} is the restriction of dγad\gamma_{a} to (TaQ)(T_{a}Q)^{\perp} and |J(x¯,y)||J(\bar{{\rm{x}}},{\rm y})| denotes the absolute value of the Jacobian determinant for the second change of variables. We are viewing x=exp(x)x=\exp({\rm{x}}), x{\rm{x}} and x¯\bar{{\rm{x}}} as functions of the new variable y{\rm y}. Donnelly shows that |J(x¯,y)||J(\bar{{\rm{x}}},{\rm y})| has a Taylor expansion in y{\rm y} whose coefficients depend only on the values at aa of BB and of the curvature tensor RR of MM and its covariant derivatives:

(36) |J(x¯,y)|=1+16(RikihBksBht+RikshBkiBht+RikthBksBhi)ysyt+O(y3).|J(\bar{{\rm{x}}},{\rm y})|=1+\frac{1}{6}(R_{ikih}B_{ks}B_{ht}+R_{iksh}B_{ki}B_{ht}+R_{ikth}B_{ks}B_{hi}){\rm y}^{s}{\rm y}^{t}+O({\rm y}^{3}).

Letting Ha(f):(TaQ)H_{a}(f):(T_{a}Q)^{\perp}\to\mathbb{R} be given by

Ha(f)(y):=f(x)|det(B(a))||J(x¯,y)|ψ(x)H_{a}(f)({\rm y}):=f(x)|\det(B(a))|\,|J(\overline{{\rm{x}}},{\rm y})|\,\psi({\rm{x}})

we have

(37) IQ(f)=Qφ1(a)(4πt)d/2ey2/4tHa(f)(y)𝑑y𝑑VQ(a).I_{Q}(f)=\int_{Q}\,\int_{\varphi^{-1}(a)}\,(4\pi t)^{-d/2}e^{-\|{\rm y}\|^{2}/4t}\,H_{a}(f)({\rm y})\,d{\rm y}\,dV_{Q}(a).

We expand Ha(f)H_{a}(f) into its Maclaurin series. Because of symmetry, only the terms of even degree contribute to the integral IQ(f)I_{Q}(f). Thus denoting by [Ha(f)]j(y)[H_{a}(f)]_{j}({\rm y}) the homogeneous component of degree 2j2j in the Maclaurin series of Ha(f)H_{a}(f) and then making the change of variable z=yt{\rm z}=\frac{{\rm y}}{\sqrt{t}}, we have

IQ(f)=(4πt)dim(Q)/2Qcodim(Q)(4π)(codim(Q))/2ez2/4k=0Ltk[Ha(f)]k(z)dzdVQ(a)+O(tL+1).I_{Q}(f)=(4\pi t)^{-\dim(Q)/2}\int_{Q}\,\int_{\mathbb{R}^{\operatorname{codim}(Q)}}\,(4\pi)^{-(\operatorname{codim}(Q))/2}e^{-\|{\rm z}\|^{2}/4}\sum_{k=0}^{L}\,t^{k}[H_{a}(f)]_{k}({\rm z})\,d{\rm z}\,dV_{Q}(a)\\ +O(t^{L+1}).

(Here we are using the fact that φ1(a)\varphi^{-1}(a) is a disk of radius rt\frac{r}{\sqrt{t}} in z{\rm z}, which can be replaced by codim(Q)\mathbb{R}^{\operatorname{codim}(Q)} without changing the asymptotics.)

Proceeding exactly as in the proof of [Don76, Theorem 4.1], we complete the integration by doing an iterated integral over the z{\rm z} variables, noting that odd powers of any of the coordinates zi{\rm z}_{i} lead to integrals of the form zi2k+1e(zi2)/4𝑑zi\int_{\mathbb{R}}\,{\rm z}_{i}^{2k+1}e^{-({\rm z}_{i}^{2})/4}\,d{\rm z}_{i} which are equal to zero, while the contributions from the terms of the form zi2me(zi2)/4𝑑zi\int_{\mathbb{R}}{\rm z}_{i}^{2m}e^{-({\rm z}_{i}^{2})/4}\,d{\rm z}_{i} can be computed by using the classical formula

x2mex2𝑑x=135(2m1)2m+1π.\int_{\mathbb{R}}x^{2m}e^{-x^{2}}\,dx=\frac{1\cdot 3\cdot 5\cdots(2m-1)}{2^{m+1}}\sqrt{\pi}.

We are left with

IQ(f)=(4πt)dim(Q)/2k=0Ltkk!Qyk(Ha(f))(0)𝑑VQ(a)+O(tL+1),I_{Q}(f)=(4\pi t)^{-\dim(Q)/2}\sum_{k=0}^{L}\frac{t^{k}}{k!}\int_{Q}\,\Box^{k}_{{\rm y}}(H_{a}(f))(0)\,dV_{Q}(a)+O(t^{L+1}),

where, following the notation of Donnelly in [Don76, Theorem 4.1], we denote

y:=i=1s2yi2.\Box_{\rm y}:=\sum_{i=1}^{s}\frac{\partial^{2}}{\partial{\rm y}_{i}^{2}}.

Thus we conclude:

Proposition 5.1.

For any compactly supported smooth function ff on MM, we have

M(4πt)d/2ed2(x,γ(x))4tf(x)𝑑VM(x)=Q𝒞(Fix(γ))(4πt)dim(Q)/2k=0Ltkk!Qyk(Ha(f))(0)𝑑VQ(a)+O(tL+1)\int_{M}\,(4\pi t)^{-d/2}e^{-\frac{d^{2}(x,\gamma(x))}{4t}}f(x)\,dV_{M}(x)\\ =\sum_{Q\in\mathcal{C}({\operatorname{{Fix}}}(\gamma))}(4\pi t)^{-\dim(Q)/2}\sum_{k=0}^{L}\frac{t^{k}}{k!}\int_{Q}\,\Box^{k}_{{\rm y}}\left(H_{a}(f)\right)(0)\,dV_{Q}(a)+O(t^{L+1})

where Ha(f)=f(x)|J(x¯,y)|ψ(x)|det(Idcodim(Q)Aa)|H_{a}(f)=\frac{f(x)\,|J(\overline{{\rm{x}}},{\rm y})|\,\psi({\rm{x}})}{|\det({\operatorname{{Id}}}_{\operatorname{codim}(Q)}-A_{a})|} with ψ(x)\psi(x) and |J(x¯,y)||J(\overline{{\rm{x}}},{\rm y})| given as in Equations (34) and (36) and with AaA_{a} being the restriction of dγad\gamma_{a} to (TaQ)(T_{a}Q)^{\perp}.

To complete the proof of Theorem 3.11 part 1, write

(38) fjp(x)=C1,2γ,2ujp(x,x).f_{j}^{p}(x)=C_{1,2}\gamma^{*}_{\cdot,2}u^{p}_{j}(x,x).

Then

MC1,2γ,2Kp(t,x,x)𝑑VM(x)=j=0LtjM(4πt)d/2ed2(x,γ(x))4tfjp(x)𝑑VM(x)+O(tL+1)=Q𝒞(Fix(γ))(4πt)dim(Q)/2k=0LtkQbkp(γ,a)𝑑VQ(a)+O(tL+1),\int_{M}C_{1,2}\gamma^{*}_{\cdot,2}K^{p}(t,x,x)dV_{M}(x)=\sum_{j=0}^{L}\,t^{j}\int_{M}\,(4\pi t)^{-d/2}e^{-\frac{d^{2}(x,\gamma(x))}{4t}}f_{j}^{p}(x)dV_{M}(x)+O(t^{L+1})\\ =\sum_{Q\in\mathcal{C}({\operatorname{{Fix}}}(\gamma))}\,(4\pi t)^{-\dim(Q)/2}\sum_{k=0}^{L}t^{k}\,\int_{Q}b^{p}_{k}(\gamma,a)dV_{Q}(a)+O(t^{L+1}),

where

bkp(γ,a)=j=0k1j!yj(Ha(fkjp))(0)b^{p}_{k}(\gamma,a)=\sum_{j=0}^{k}\,\frac{1}{j!}\Box^{j}_{{\rm y}}\left(H_{a}(f^{p}_{k-j})\right)(0)

with fkjpf^{p}_{k-j} defined as in Equation (38).

In particular, when k=0k=0, we have

(39) b0p(γ,a)=Ha(f0p)(0)=C1,2γ,2u0p(a,a)|det(Idcodim(Q)Aa)|=trp(dγa)|det(Idcodim(Q)Aa)|,b^{p}_{0}(\gamma,a)=H_{a}(f_{0}^{p})(0)=\frac{C_{1,2}\gamma^{*}_{\cdot,2}u^{p}_{0}(a,a)}{\lvert\det({\operatorname{{Id}}}_{\operatorname{codim}(Q)}-A_{a})\rvert}=\frac{\operatorname{tr}_{p}(d\gamma_{a})}{\lvert\det({\operatorname{{Id}}}_{\operatorname{codim}(Q)}-A_{a})\rvert},

where trp(dγa)\operatorname{tr}_{p}(d\gamma_{a}) is defined as in Notation and Remarks 3.10 part 4. (The second equality follows from the fact that u0p(a,a)u_{0}^{p}(a,a) is the identity transformation of Λp(TaM)\Lambda^{p}(T^{*}_{a}M) as noted in Proposition 3.3 part 1.)

Theorem 3.11 part 2 also follows with

ckp(η,γ,a)=j=0k1j!yj(Ha(ηfkjp))(0).c^{p}_{k}(\eta,\gamma,a)=\sum_{j=0}^{k}\,\frac{1}{j!}\Box^{j}_{{\rm y}}\left(H_{a}(\eta f^{p}_{k-j})\right)(0).

Funding: This work was supported by the Banff International Research Station and Casa Matemática Oaxaca [19w5115]; the Association for Women in Mathematics National Science Foundation ADVANCE grant [1500481]; and the Swiss Mathematical Society [SMS Travel Grant to KG]. IMS was supported by the Leverhulme Trust grant [RPG-2019-055].

Acknowledgments: We thank Carla Farsi for a helpful discussion of orbifold orientability. We also thank the Banff International Research Station and Casa Matemática Oaxaca for initiating our collaboration by hosting the 2019 Women in Geometry Workshop. In addition, we thank the Association for Women in Mathematics and the organizers of the Women In Geometry Workshop for supporting this collaboration. IMS thanks Prof. Jacek Brodzki for providing her with excellent conditions to work.

References

  • [ALR07] Alejandro Adem, Johann Leida, and Yongbin Ruan. Orbifolds and stringy topology, volume 171 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 2007.
  • [BG90] Thomas P. Branson and Peter B. Gilkey. The asymptotics of the Laplacian on a manifold with boundary. Comm. Partial Differential Equations, 15(2):245–272, 1990.
  • [BPP92] Robert Brooks, Peter Perry, and Peter Petersen, V. Compactness and finiteness theorems for isospectral manifolds. J. Reine Angew. Math., 426:67–89, 1992.
  • [Buc99] Bogdan Bucicovschi. Contributions to the complex powers and zeta function of elliptic pseudodifferential operators. ProQuest LLC, Ann Arbor, MI, 1999. Thesis (Ph.D.)–The Ohio State University.
  • [BY02] Bruce C. Berndt and Boon Pin Yeap. Explicit evaluations and reciprocity theorems for finite trigonometric sums. Adv. in Appl. Math., 29(3):358–385, 2002.
  • [Chi93] Yuan-Jen Chiang. Spectral geometry of VV-manifolds and its application to harmonic maps. In Differential geometry: partial differential equations on manifolds (Los Angeles, CA, 1990), volume 54 of Proc. Sympos. Pure Math., pages 93–99. Amer. Math. Soc., Providence, RI, 1993.
  • [CJ19] Francisco C. Caramello Jr. Introduction to orbifolds, 2019.
  • [CS12] Djurdje Cvijović and Hari Mohan Srivastava. Closed-form summations of Dowker’s and related trigonometric sums. J. Phys. A, 45(37):374015, 10, 2012.
  • [DGGW08] Emily B. Dryden, Carolyn S. Gordon, Sarah J. Greenwald, and David L. Webb. Asymptotic expansion of the heat kernel for orbifolds. Michigan Math. J., 56(1):205–238, 2008.
  • [DGGW17] Emily B. Dryden, Carolyn S. Gordon, Sarah J. Greenwald, and David L. Webb. Erratum to “Asymptotic expansion of the heat kernel for orbifolds” [MR2433665]. Michigan Math. J., 66(1):221–222, 2017.
  • [Don76] Harold Donnelly. Spectrum and the fixed point sets of isometries. I. Math. Ann., 224(2):161–170, 1976.
  • [DP77] Harold Donnelly and V. K. Patodi. Spectrum and the fixed point sets of isometries. II. Topology, 16(1):1–11, 1977.
  • [DS09] Emily B. Dryden and Alexander Strohmaier. Huber’s theorem for hyperbolic orbisurfaces. Canad. Math. Bull., 52(1):66–71, 2009.
  • [Far01] Carla Farsi. Orbifold spectral theory. Rocky Mountain J. Math., 31(1):215–235, 2001.
  • [FPS14] Carla Farsi, Emily Proctor, and Christopher Seaton. Γ\Gamma-extensions of the spectrum of an orbifold. Trans. Amer. Math. Soc., 366(7):3881–3905, 2014.
  • [Gaf58] Matthew P. Gaffney. Asymptotic distributions associated with the Laplacian for forms. Comm. Pure Appl. Math., 11:535–545, 1958.
  • [Gor12] Carolyn S. Gordon. Orbifolds and their spectra. In Spectral geometry, volume 84 of Proc. Sympos. Pure Math., pages 49–71. Amer. Math. Soc., Providence, RI, 2012.
  • [GR03] Carolyn S. Gordon and Juan Pablo Rossetti. Boundary volume and length spectra of Riemannian manifolds: what the middle degree Hodge spectrum doesn’t reveal. Ann. Inst. Fourier (Grenoble), 53(7):2297–2314, 2003.
  • [GR21] Carolyn S. Gordon and Juan Pablo Rossetti. Correction to “Boundary volume and length spectra of Riemannian manifolds: what the middle degree Hodge spectrum doesn’t reveal”. Annales de l’Institut Fourier, to appear, 2021. https://arxiv.org/pdf/math/0111016.pdf.
  • [Har16] John Harvey. Equivariant Alexandrov geometry and orbifold finiteness. J. Geom. Anal., 26(3):1925–1945, 2016.
  • [Kob58] Shoshichi Kobayashi. Fixed points of isometries. Nagoya Math. J., 13:63–68, 1958.
  • [LM17] Benjamin Linowitz and Jeffrey S. Meyer. On the isospectral orbifold-manifold problem for nonpositively curved locally symmetric spaces. Geom. Dedicata, 188:165–169, 2017.
  • [Par04] JeongHyeong Park. The spectral geometry of Einstein manifolds with boundary. J. Korean Math. Soc., 41(5):875–882, 2004.
  • [Pat70] V. K. Patodi. Curvature and the fundamental solution of the heat operator. J. Indian Math. Soc., 34(3-4):269–285 (1971), 1970.
  • [Pat71] V. K. Patodi. Curvature and the eigenforms of the Laplace operator. J. Differential Geometry, 5:233–249, 1971.
  • [Pro12] Emily Proctor. Orbifold homeomorphism finiteness based on geometric constraints. Ann. Global Anal. Geom., 41(1):47–59, 2012.
  • [PS10] Emily Proctor and Elizabeth Stanhope. Spectral and geometric bounds on 2-orbifold diffeomorphism type. Differential Geom. Appl., 28(1):12–18, 2010.
  • [RS20] Sean Richardson and Elizabeth Stanhope. You can hear the local orientability of an orbifold. Differential Geom. Appl., 68:101577, 7, 2020.
  • [RSW08] Juan Pablo Rossetti, Dorothee Schueth, and Martin Weilandt. Isospectral orbifolds with different maximal isotropy orders. Ann. Global Anal. Geom., 34(4):351–366, 2008.
  • [Sco83] Peter Scott. The geometries of 33-manifolds. Bull. London Math. Soc., 15(5):401–487, 1983.
  • [SSW06] Naveed Shams, Elizabeth Stanhope, and David L. Webb. One cannot hear orbifold isotropy type. Arch. Math. (Basel), 87(4):375–384, 2006.
  • [Sta05] Elizabeth Stanhope. Spectral bounds on orbifold isotropy. Ann. Global Anal. Geom., 27(4):355–375, 2005.
  • [Sut10] Craig J. Sutton. Equivariant isospectrality and Sunada’s method. Arch. Math. (Basel), 95(1):75–85, 2010.
  • [Thu97] William P. Thurston. Three-dimensional geometry and topology. Vol. 1, volume 35 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1997. Edited by Silvio Levy.
  • [Wei12] Martin Weilandt. Isospectral metrics on weighted projective spaces. New York J. Math., 18:421–449, 2012.