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

A Trace Formula on Stationary Kaluza-Klein Spacetimes

Anthony McCormick
Abstract

We prove relativistic versions of the ladder asymptotics from [11] on principal bundles over globally hyperbolic, stationary, spatially compact spacetimes equipped with a Kaluza-Klein metric. This involves understanding the distribution of the frequency spectrum for the wave equation on a Kaluza-Klein spacetime when restricted to the isotypic subspace of an irreducible representation of the structure group, in the limit that the weight of the representation approaches infinity in the Weyl chamber. This is a direct generalization of the results from [23] and is closely related to [22], [14]. Furthermore we show how to apply these results to frequency asymptotics for the massive Klein-Gordon equation on vector bundles as one takes the representation defining the vector bundle to infinity.

1 Introduction

Given a principal GSO(k)G\subseteq\operatorname{SO}(k)-bundle PP with connection ω\omega over a Lorentzian manifold (Mn+1,g)(M^{n+1},g) one can form the Kaluza-Klein metric on PP:

gω:=πg+Tr(ω()ω()T)g_{\omega}:=\pi^{*}g+\operatorname{Tr}(\omega(-)\omega(-)^{T})

where π:PM\pi:P\to M is the projection map. In the special case where MM is globally hyperbolic, stationary and spatially compact we will see in Section 2 that the Kaluza-Klein spacetime takes the form

P=t×P0,M=t×Σ0,P0Σ0 a principal bundleP=\mathbb{R}_{t}\times P_{0},\ M=\mathbb{R}_{t}\times\Sigma_{0},\ P_{0}\to\Sigma_{0}\mbox{ a principal bundle}

with metric given by

gω=((Nπ)2|πη|πh2)dt2+dt(πη)+(πη)dt+πh+Tr(ω()ω()T)g_{\omega}=-((N\circ\pi)^{2}-|\pi^{*}\eta|^{2}_{\pi^{*}h})dt^{2}+dt\otimes(\pi^{*}\eta)+(\pi^{*}\eta)\otimes dt+\pi^{*}h+\operatorname{Tr}(\omega(-)\omega(-)^{T})

where hh is a Riemannian metric on Σ0\Sigma_{0}, N:Σ0>0N:\Sigma_{0}\to\mathbb{R}_{>0} a smooth function, and η\eta a 1-form on Σ0\Sigma_{0} satisfying |η|h2<N2|\eta|_{h}^{2}<N^{2} pointwise. Given an integral coadjoint orbit 𝒪\mathcal{O} for GG we have irreducible unitary representations corresponding to m𝒪m\mathcal{O} for each m1m\in\mathbb{Z}_{\geq 1} and as discussed in Section 3 we get isotypic subspaces of the space of solutions to the wave equation with respect to the Kaluza-Klein metric:

mkerω corresponding to the representation m𝒪.\mathcal{H}_{m}\subseteq\ker\Box_{\omega}\ \mbox{ corresponding to the representation }m\mathcal{O}.

As is shown in Section 3.1, for mm sufficiently large the operator

DZ:=1itD_{Z}:=\frac{1}{i}\partial_{t}

is self adjoint on m\mathcal{H}_{m} with respect to the energy form with a discrete set of eigenvalues

λm,λm,+1\cdots\leq\lambda_{m,\ell}\leq\lambda_{m,\ell+1}\leq\cdot

accumulating at ±\pm\infty with at worst polynomial growth. For E>0E>0 fixed, our goal is to study the following question.

Question:
What are the mm\to\infty asymptotics of the frequency spectrum of DZmED_{Z}-mE on m\mathcal{H}_{m}?

This is a relativistic analogue of the question studied in [11], and is a generalization to non-trivial principal bundles with arbitrary compact structure groups of the results in [23].

We make precise the notion of the distribution of the frequency spectrum about mEmE via the tempered distribution μ(E,m,)\mu(E,m,-) on \mathbb{R} given by:

μ(E,m,φ):=φ(λm,mE)\mu(E,m,\varphi):=\sum_{\ell\in\mathbb{Z}}\varphi(\lambda_{m,\ell}-mE)

and we will study its asymptotics via the periodic generating function

Υ(φ)(θ):=m=1μ(E,m,φ)eimθ𝒟(S1)=𝒟(/2π).\Upsilon(\varphi)(\theta):=\sum_{m=1}^{\infty}\mu(E,m,\varphi)e^{im\theta}\in\mathcal{D}^{\prime}(S^{1})=\mathcal{D}^{\prime}(\mathbb{R}/2\pi\mathbb{Z}).

The fact that μ(E,m,)\mu(E,m,-) is tempered and Υ(φ)\Upsilon(\varphi) is defined on all of C(S1)C^{\infty}(S^{1}) are proven in Sections 3.1, 4 respectively.

To state our main results we need to make a dynamical assumption akin to the “clean intersection hypotheses” that appear in [23], [10] and [11]. For this we start by introducing the notation:

  • 𝒩\mathcal{N} is the symplectic manifold of affinely parametrized inextendible future-directed null geodesics on PP modulo the action of translation in the affine parameter,

  • 𝒩𝒪\mathcal{N}_{\mathcal{O}} is the symplectic reduction of 𝒩\mathcal{N} along the coadjoint orbit 𝒪\mathcal{O} as in [11],

  • H~Z:𝒩𝒪\widetilde{H}_{Z}:\mathcal{N}_{\mathcal{O}}\to\mathbb{R} and Φ~tZ\widetilde{\Phi}^{Z}_{t} are respectively the Hamiltonian and Hamiltonian flow corresponding to the reduction of the flow on 𝒩\mathcal{N} induced by the flow of Z=tZ=\partial_{t} on PP.

The clean intersection hypothesis then states that E>0E>0 is a regular value for H~Z\widetilde{H}_{Z} and the fibered product 𝔜E\mathfrak{Y}_{E} of the flow map

×H~Z1(E)H~Z1(E)\mathbb{R}\times\widetilde{H}_{Z}^{-1}(E)\to\widetilde{H}_{Z}^{-1}(E)

with the diagonal map H~Z1(E)H~Z1(E)×H~Z1(E)\widetilde{H}_{Z}^{-1}(E)\to\widetilde{H}_{Z}^{-1}(E)\times\widetilde{H}_{Z}^{-1}(E) is a clean fibered product (it is a smooth manifold with tangent spaces given by the non-necessarily-transverse fibered products of the respective tangent spaces of the factors). We now state our main theorems. These are completely analogous to the main theorems in [11] and are direct relativistic generalizations of these.

Theorem 1.1.

The wave front set of Υ(φ)𝒟(S1)\Upsilon(\varphi)\in\mathcal{D}^{\prime}(S^{1}) is contained in:

𝔇φ,E:={(ω,r)S1×>0\displaystyle\mathfrak{D}_{\varphi,E}:=\Big{\{}(\omega,r)\in S^{1}\times\mathbb{R}_{>0}\ :(T,γ))𝔜E with Tsuppφ^\displaystyle:\ \exists(T,\gamma))\in\mathfrak{Y}_{E}\mbox{ with }T\in\operatorname{supp}\widehat{\varphi}
such that ω=Hol𝒪([0,T]tΦ~tZ(γ))}\displaystyle\mbox{ such that }\omega=\operatorname{Hol}_{\mathcal{O}}([0,T]\ni t\mapsto\widetilde{\Phi}^{Z}_{t}(\gamma))\Big{\}}

where this holonomy is taken with respect to a natural U(1)\operatorname{U}(1)-bundle with connection over 𝒩𝒪\mathcal{N}_{\mathcal{O}} defined in 2.35.

Theorem 1.2.

Let n+1=dim(M)n+1=\dim(M). Under the clean intersection hypothesis, 𝔇φ,E\mathfrak{D}_{\varphi,E} is a union of the positive parts of finitely many fibers of TS1T^{*}S^{1}, and Υ(φ)In+1+14(S1;𝔇φ,E)\Upsilon(\varphi)\in I^{n+\ell-1+\frac{1}{4}}(S^{1};\mathfrak{D}_{\varphi,E}) where 2:=dim𝒪2\ell:=\dim\mathcal{O}. Furthermore, we obtain an asymptotic expansion as mm\to\infty:

μ(E,m,φ)k=0mn+1kak(φ,m)\mu(E,m,\varphi)\sim\sum_{k=0}^{\infty}m^{n+\ell-1-k}a_{k}(\varphi,m)

with each ak(φ,m)a_{k}(\varphi,m) a distribution in φ\varphi, bounded in mm for k,φk,\varphi fixed, and

a0(φ,m)=Cn,dφ(0)Vol(H~Z1(E))a_{0}(\varphi,m)=C_{n,d}\varphi(0)\operatorname{Vol}\left(\widetilde{H}_{Z}^{-1}(E)\right)

where by Vol\operatorname{Vol} we mean to take the invariant measure on the energy hypersurface.

Theorem 1.3.

Suppose that, in addition to the clean intersection hypothesis, we assumed that 0supp(φ)0\notin\operatorname{supp}(\varphi) and there existed only finitely many non-degenerate periodic orbits (T1,γ1),,(Tq,γq)𝔜E(T_{1},\gamma_{1}),\cdots,(T_{q},\gamma_{q})\in\mathfrak{Y}_{E} with each Tj0T_{j}\neq 0. Then we actually obtain a better asymptotic expansion as mm\to\infty:

μ(E,m,φ)k=0mkak(φ,m)\mu(E,m,\varphi)\sim\sum_{k=0}^{\infty}m^{-k}a_{k}(\varphi,m)

and a0(φ,m)a_{0}(\varphi,m) is of the form:

a0(φ,m)=Cn,dj=0qHol𝒪(Tj,γj)mTj#2πφ^(Tj)eiπ𝔪j/4|det(IPj)|1/2.a_{0}(\varphi,m)=C_{n,d}\sum_{j=0}^{q}\operatorname{Hol}_{\mathcal{O}}(T_{j},\gamma_{j})^{m}\frac{T_{j}^{\#}}{2\pi}\widehat{\varphi}(T_{j})\frac{e^{i\pi\mathfrak{m}_{j}/4}}{|\det(I-P_{j})|^{1/2}}.

Where Tj#T_{j}^{\#} is the minimum positive value of TT such that Φ~TZ(γj)=γj\widetilde{\Phi}^{Z}_{T}(\gamma_{j})=\gamma_{j}, PjP_{j} is the linearized Poincaré first return map of γj\gamma_{j}, and 𝔪j\mathfrak{m}_{j} is the Conley-Zehnder index.

These theorems are proven in Section 4 using the tools developed in the previous section. From the results of Section 3.1 we also obtain a corollary concerning the frequency distribution of DZD_{Z} for vector bundles. Let’s describe this now.

Let 𝒱mM\mathcal{V}_{m}\to M denote the Hermitian vector bundle associated to the representation κm\kappa_{m} equipped with the covariant derivative induced from ω\omega on PP and the representation. We then have a massive Klein-Gordon operator m\Box_{m} with mass given by the eigenvalue

mΛ0,mΛ0+ρ\langle m\Lambda_{0},\ m\Lambda_{0}+\rho\rangle

of the quadratic Casimir on the representation corresponding to m𝒪m\mathcal{O}. Here Λ0\Lambda_{0} is the highest weight for 𝒪\mathcal{O} and ρ\rho is the sum of the positive roots. This massive Klein-Gordon operator acts on sections and its kernel is invariant under the operator Dm,ZD_{m,Z} given by 1i\frac{1}{i} times the covariant derivative in the ZZ direction.

Corollary 1.4.

For mm sufficiently large operator Dm,ZD_{m,Z} on kerm\ker\Box_{m} equipped with the energy form from Section 3.1 is self adjoint with discrete spectrum equal to the spectrum of DZD_{Z} on m\mathcal{H}_{m}, and the multiplicity of λσ(DZ)\lambda\in\sigma(D_{Z}) is the product of the multiplicity of λσ(Dm,Z)\lambda\in\sigma(D_{m,Z}) and the dimension dmd_{m} of the irreducible representation corresponding to m𝒪m\mathcal{O}. Thus if μ(E,𝒱m,φ)\mu(E,\mathcal{V}_{m},\varphi) is defined for Dm,ZD_{m,Z} in the same way μ(E,m,φ)\mu(E,m,\varphi) is defined for DZD_{Z} then under the clean intersection hypothesis we have

μ(E,𝒱m,φ)1dmk=0mn+1kbk(φ,m).\mu(E,\mathcal{V}_{m},\varphi)\sim\frac{1}{d_{m}}\sum_{k=0}^{\infty}m^{n+\ell-1-k}b_{k}(\varphi,m).

In general, one can compute the values of \ell and dmd_{m} in terms of the dominant integral element Λ0\Lambda_{0}. Indeed, if R+R^{+} is the set of positive roots then

=12dim𝒪= the number of positive roots not orthogonal to Λ0\ell=\frac{1}{2}\dim\mathcal{O}=\ \mbox{ the number of positive roots not orthogonal to }\Lambda_{0}

and as a consequence of the Weyl character formula we have

dm=αR+α,mΛ0+12ραR+α,12ρ.d_{m}=\frac{\prod_{\alpha\in R^{+}}\langle\alpha,\ m\Lambda_{0}+\frac{1}{2}\rho\rangle}{\prod_{\alpha\in R^{+}}\langle\alpha,\frac{1}{2}\rho\rangle}.

In particular we see that dmd_{m} is a polynomial of degree \ell and so our leading order asymptotics for μ(E,𝒱m,φ)\mu(E,\mathcal{V}_{m},\varphi) as mm\to\infty are mn1m^{n-1}. This is in agreement with [23] where =0\ell=0 and dm=1d_{m}=1 for all mm. When G=SU(2)G=\operatorname{SU}(2) and 𝒪\mathcal{O} corresponds to the vector representation then =1\ell=1 and dm=m+1d_{m}=m+1.

We now provide an outline of the paper. In Section 2 we demonstrate that (M,g)(M,g) being globally hyperbolic, stationary and spatially compact implies that this is also true for the Kaluza-Klein metric. The rest of the section is spent recalling the symplectic geometry from [22] with (M,g)(M,g) replaced by (P,gω)(P,g_{\omega}). In Section 2.1 we introduce the reduced phase space from [11],[24],[21] in the special case where the symplectic manifold is 𝒩\mathcal{N} and in Section 2.2 we study two aspects of periodic orbits on this reduced phase space: the linearized Poincaré first return map PγP_{\gamma} and the holonomy map Hol𝒪\operatorname{Hol}_{\mathcal{O}}. In Section 3 we use the general setup of [22] to discuss the wave equation on (P,gω)(P,g_{\omega}). Since we allow for a potential (as long as it is constant along the fibers of PP and independent of tt) the energy quadratic form on the space of solutions kerω\ker\Box_{\omega} need not be positive definite, but we use standard results from harmonic analysis together with some results on Krein and Pontryagin spaces to show that it is positive definite on the isotypic subspace m\mathcal{H}_{m} for mm sufficiently large. In Section 3.1 we apply a result from [14] to obtain that μ(E,m,)\mu(E,m,-) is tempered and then we provide a proof of Corollary 1.4 given Theorems 1.1,1.2,1.3. Finally, in Section 4 we simply combine the techniques from [22] and [11] to obtain our main theorems.

2 The Classical Dynamics: Wong’s Equations

Let (Mn+1,g)(M^{n+1},g) be a connected, globally hyperbolic, oriented, time-oriented Lorentzian manifold. For us, “Lorentzian” will mean that gg has signature (1,+1,,+1)(-1,+1,\cdots,+1). We refer the reader to [18] for an explanation of the various causality assumptions and related terminology we will use. Throughout, we will assume the following:

  1. 1.

    We will assume that (M,g)(M,g) admits a complete timelike Killing vector field ZZ, flowing forwards in time with respect to the time-orientation (and we will make a fixed choice of such a ZZ).

  2. 2.

    We will assume that (M,g)(M,g) is spatially compact. That is, for some (hence any) choice of Cauchy hypersurface ΣM\Sigma\subseteq M, Σ\Sigma is compact.

Lemma 2.1.

[15]
All such spacetimes (M,g)(M,g) as described above are diffeomorphic to t×Σ\mathbb{R}_{t}\times\Sigma with metric of the form:

g\displaystyle g =(N2|η|h2)dt2+dtη+ηdt+h\displaystyle=-(N^{2}-|\eta|_{h}^{2})dt^{2}+dt\otimes\eta+\eta\otimes dt+h
=N2dt2+hij(dxi+βidt)(dxj+βjdt)\displaystyle=-N^{2}dt^{2}+h_{ij}(dx^{i}+\beta^{i}dt)(dx^{j}+\beta^{j}dt)

where N:Σ>0N:\Sigma\to\mathbb{R}_{>0} is smooth, η\eta a 1-form on Σ\Sigma, β=βii\beta=\beta^{i}\partial_{i} a vector field on Σ\Sigma, hh a Riemannian metric on Σ\Sigma, and Σ={0}×Σ\Sigma=\{0\}\times\Sigma a Cauchy hypersurface. We also have N2>|η|h2N^{2}>|\eta|^{2}_{h} pointwise and β\beta is the vector field hh-dual to η\eta. Furthermore, Σt={t}×Σ\Sigma_{t}=\{t\}\times\Sigma will be a Cauchy hypersurface for each tt\in\mathbb{R}.

Notice that such spacetimes (M,g)(M,g) are not necessarily static since Z=tZ=\partial_{t} need not be normal to Σ\Sigma. Indeed, N1(tβ)N^{-1}(\partial_{t}-\beta) is the unit normal. Let’s denote this by

ν:=N1(tβ)\nu:=N^{-1}(\partial_{t}-\beta)

noticing that this is a globally defined vector field on MM and is the unit normal to each Σt\Sigma_{t} when restricted to that submanifold.

The classical dynamics we wish to study are Wong’s equation [25] on a curved spacetime. In [25], the classical limit of a massive quantum particle in an external classical Yang-Mills field was determined to be given by the equations

mx¨i=Tr(qFAij)x˙j,x˙2=1m\ddot{x}^{i}=\operatorname{Tr}(qF^{ij}_{A})\dot{x}_{j},\ \ \ \dot{x}^{2}=-1

where q𝔲(k)q\in\mathfrak{u}(k) is a conserved quantity describing the internal degrees of freedom of the system (a generalization of charge). This is an analogue of the Lorentz force law, generalized to connections with structure groups GG other than U(1)\operatorname{U}(1). Some references for the study of these equations on curved non-relativistic space are [21], [24]. The basic idea is to express these equations as geodesic equations on a principal bundle over space equipped with a Kaluza-Klein metric. The Lorentzian analogue of this is developed below.

Now, let GSO(k)G\subseteq\operatorname{SO}(k) be a compact Lie group and π:PM\pi:P\to M a principal GG-bundle.

Definition 2.2.

Recalling that (X,Y)Tr(XY)(X,Y)\mapsto-\operatorname{Tr}(XY) is a positive definite Ad\operatorname{Ad}-invariant inner product on 𝔤\mathfrak{g}, given any connection ω\omega on PP we obtain an induced Kaluza-Klein metric:

gω=πgTr(ω()ω())g_{\omega}=\pi^{*}g-\operatorname{Tr}(\omega(-)\omega(-))

on the total space of PP. This is again a Lorentzian metric of signature (1,+1,,+1)(-1,+1,\cdots,+1). Furthermore, we endow (P,gω)(P,g_{\omega}) with the orientation induced from MM and 𝔤𝔰𝔬(k)\mathfrak{g}\subseteq\mathfrak{so}(k), and the time-orientation induced from MM. We let

Zω:= the horizontal lift of Z, and ξ^:= the vertical vector field of ξ𝔤.Z^{\omega}:=\mbox{ the horizontal lift of }Z,\ \mbox{ and }\widehat{\xi}:=\mbox{ the vertical vector field of }\xi\in\mathfrak{g}.

We also abuse notation and set

n^:= the horizontal lift of the unit normal N1(tβ)=N1(Zβ).\widehat{n}:=\mbox{ the horizontal lift of the unit normal }N^{-1}(\partial_{t}-\beta)=N^{-1}(Z-\beta).
Lemma 2.3.

The horizontal lift ZωZ^{\omega} of ZZ is a complete timelike future-oriented Killing vector field for gωg_{\omega} and:

[Zω,ξ^]=0 for all ξ𝔤.[Z^{\omega},\widehat{\xi}]=0\mbox{ for all }\xi\in\mathfrak{g}.
Proof.

Recall that the connection ω:TP𝔤\omega:TP\to\mathfrak{g} defines a horizontal bundle HP:=kerωHP:=\ker\omega splitting TPVPP×𝔤TP\to VP\cong P\times\mathfrak{g} and that π|HP:HPπTM\pi_{*}|_{HP}:HP\to\pi^{*}TM is an isomorphism. The horizontal lift ZωZ^{\omega} of ZZ is then given by

(π|HP)1(πZ)\left(\pi_{*}|_{HP}\right)^{-1}(\pi^{*}Z)

where πZ\pi^{*}Z is the pullback of ZZ to a section of πTM\pi^{*}TM. Furthermore, the flow of ZωZ^{\omega} is the horizontal lift of the flow of ZZ on MM ([3] section 10.1). This immediately implies ZωZ^{\omega} is complete since ZZ is. Furthermore since ZZ is Killing on (M,g)(M,g) it follows that πg\pi^{*}g is invariant under the flow of ZωZ^{\omega} and since ZωZ^{\omega} is horizontal we have

ZωTr(ω()ω()T)=0.\mathcal{L}_{Z^{\omega}}\operatorname{Tr}(\omega(-)\omega(-)^{T})=0.

Thus indeed ZωZ^{\omega} is Killing on (P,gω)(P,g_{\omega}). Finally, from [3] section 2.2 we know that ZωZ^{\omega} is invariant under push-forward along the GG-action on PP and therefore [Zω,ξ^]=0[Z^{\omega},\widehat{\xi}]=0 for all ξ𝔤\xi\in\mathfrak{g}. ∎

The next result is an immediate corollary of [19] but we include the proof from [19] for the reader’s convenience. Note that we are using the intrinsic definition of Cauchy hypersurfaces in terms of inextendible causal curves since this is the definition used when proving well-posedness for the wave equation in [2] and we would like to apply their results directly. However, the proof of well-posedness simplifies when the manifold is spatially compact with a complete timelike Killing vector field, and so our discussion of “inextendibility” below is unnecessary for readers working exclusively in this setting.

For a particularly nice discussion of the geometry of these spacetimes we refer the reader to [1] and [5] where it is also shown that when n=3n=3 these spacetimes do not admit non-trivial (i.e. product with a flat Riemannian spacetime) vacuum solutions to the Einstein equations.

Lemma 2.4.

[19]
(P,gω)(P,g_{\omega}) is globally hyperbolic and each Pt=π1(Σt)P_{t}=\pi^{-1}(\Sigma_{t}), tt\in\mathbb{R}, a Cauchy hypersurface.

Proof.

The map ×P0P\mathbb{R}\times P_{0}\to P induced by flowing along ZωZ^{\omega} is a global diffeomorphism and so we can, without loss of generality, assume P=t×P0P=\mathbb{R}_{t}\times P_{0} as a manifold with Zω=tZ^{\omega}=\partial_{t} and Pt={t}×P0P_{t}=\{t\}\times P_{0}. Using our standard form for the metric gg on MM we can write the Kaluza-Klein metric on PP as:

gω=((Nπ)2|πη|πh2)dt2+dt(πη)+(πη)dt+(πhTr(ω()ω())).{}g_{\omega}=-((N\circ\pi)^{2}-|\pi^{*}\eta|^{2}_{\pi^{*}h})dt^{2}+dt\otimes(\pi^{*}\eta)+(\pi^{*}\eta)\otimes dt+(\pi^{*}h-\operatorname{Tr}(\omega(-)\omega(-))). (1)

In particular, the Riemannian metric h~t\widetilde{h}_{t} on PtP_{t} induced by pulling back gωg_{\omega} is independent of tt:

h~t=h~=πhTr(ω()ω()).\widetilde{h}_{t}=\widetilde{h}=\pi^{*}h-\operatorname{Tr}(\omega(-)\omega(-)).

Choose now an arbitrary inextendible causal curve γ\gamma in PP, parametrized with respect to tt so that γ(t)=(t,γ0(t))\gamma(t)=(t,\gamma_{0}(t)). We can always make this parametrization for causal curves in the spacetimes we are considering thanks to the global time function tt. We now write (a,b){±}(a,b)\subseteq\mathbb{R}\cup\{\pm\infty\} for the domain of γ\gamma with b(,]b\in(-\infty,\infty] and a[,b)a\in[-\infty,b). Since γ\gamma is causal we have at all t(a,b)t\in(a,b) that:

N(π(γ0(t)))2+|πη|πh2(γ(t))+2πη,γ0h~+h~(γ0,γ0)0.-N(\pi(\gamma_{0}(t)))^{2}+|\pi^{*}\eta|^{2}_{\pi^{*}h}(\gamma(t))+2\langle\pi^{*}\eta,\gamma_{0}^{\prime}\rangle_{\widetilde{h}}+\widetilde{h}(\gamma_{0}^{\prime},\gamma_{0}^{\prime})\leq 0.

Thus by Cauchy-Schwarz we have:

|γ0|h~22|πη|h~|γ0|h~N2+|πη|h~20.|\gamma_{0}^{\prime}|^{2}_{\widetilde{h}}-2|\pi^{*}\eta|_{\widetilde{h}}|\gamma_{0}^{\prime}|_{\widetilde{h}}-N^{2}+|\pi^{*}\eta|_{\widetilde{h}}^{2}\leq 0.

Rearranging yields:

(|γ0|h~|πη|h~)2N2.(|\gamma_{0}^{\prime}|_{\widetilde{h}}-|\pi^{*}\eta|_{\widetilde{h}})^{2}\leq N^{2}.

Now, suppose for contradiction that b<b<\infty (the case <a-\infty<a is completely analogous) and set:

C:=sup[|b|1,b+1]×P0(N+|πη|h~).C:=\sup_{[-|b|-1,b+1]\times P_{0}}\left(N+|\pi^{*}\eta|_{\widetilde{h}}\right).

We have C<C<\infty since Σ\Sigma and GG are compact, hence P0P_{0} and [|b|1,b+1]×P0[-|b|-1,b+1]\times P_{0} are compact. We then have:

|γ0|h~C on [|b|1,b]|\gamma_{0}^{\prime}|_{\widetilde{h}}\leq C\mbox{ on }[-|b|-1,b]

and so the curve γ\gamma must be extendible beyond time bb, a contradiction. ∎

Here we provide a brief remark on the above proof: recall that in the definition of a Cauchy hypersurface, the assumption is that all curves which are both inextendible and causal intersect the hypersurface exactly once. One does not make this requirement of causal curves which are extendible, but whose extensions are non-causal. To see why, consider the following example on flat 2-dimensional Minkowski space (this example also works in any dimension).

We let t,xt,x denote our coordinates so that the metric on 2\mathbb{R}^{2} is dt2+dx2-dt^{2}+dx^{2}. Consider the curve given by t(s)=st(s)=s and x(s)=0x(s)=0 for s0s\leq 0 and x(s)=e1/s2x(s)=e^{-1/s^{2}} for s>0s>0. There exists a time s0>0s_{0}>0 so that γ(s):=(t(s),x(s))\gamma(s):=(t(s),x(s)) is a causal curve for s(,s0)s\in(-\infty,s_{0}) but there exists no ϵ>0\epsilon>0 on which γ\gamma extends to a causal curve on (,s0+ϵ)(-\infty,s_{0}+\epsilon). So γ\gamma has no causal extensions, but it is not inextendible since it does admit an extension to a curve γ:2\gamma:\mathbb{R}\to\mathbb{R}^{2} (albeit a non-causal one). This clarifies why, in the definition of a Cauchy hypersurface, one only requires all inextendible curves, which are also causal, to intersect the Cauchy hypersurface exactly once.

As an immediate corollary of the specific form of the metric gωg_{\omega} derived in the above proof, we also obtain the following.

Corollary 2.5.

The horizontal lift n^\widehat{n} of the unit normal to the Cauchy hypersurfaces ΣtM\Sigma_{t}\subseteq M is the unit normal to the Cauchy hypersurfaces PtPP_{t}\subseteq P with respect to gωg_{\omega}.

Let’s now discuss the geodesic equations in the Kaluza-Klein spacetime (P,gω)(P,g_{\omega}). Most of the basic facts here can be found in the reference [3] but we include them for the reader’s convenience together with section and/or theorem numbers from [3]. Recall that the Levi-Civita connection, together with the ODE γ˙γ˙=0\nabla_{\dot{\gamma}}\dot{\gamma}=0 defining the geodesic equations are defined on Lorentzian manifolds in the exact same way as they are defined on Riemannian manifolds. Furthermore, we still have

ddsgω(γ˙(s),γ˙(s))=2gω(γ˙(s)γ˙(s),γ˙(s))=0\frac{d}{ds}g_{\omega}(\dot{\gamma}(s),\dot{\gamma}(s))=2g_{\omega}(\nabla_{\dot{\gamma}(s)}\dot{\gamma}(s),\dot{\gamma}(s))=0

and so

gω(γ˙,γ˙) is constant along geoedesics.g_{\omega}(\dot{\gamma},\dot{\gamma})\mbox{ is constant along geoedesics.}

This allows us to split geodesics into three types.

Definition 2.6.

We call a geodesic γ\gamma on (P,gω)(P,g_{\omega}) lightlike (respectively spacelike and null) if and only if the constant gω(γ˙,γ˙)g_{\omega}(\dot{\gamma},\dot{\gamma}) is negative (respectively positive and zero).

As the below lemma explains, despite Wong’s equations describing massive particles in (M,g)(M,g), we will be interested in null geodesics in (P,gω)(P,g_{\omega}).

Lemma 2.7.

Let γ\gamma be a geodesic in (P,gω)(P,g_{\omega}). Then the value ω(γ(t))\omega(\gamma^{\prime}(t)) is constant. Thus

gω(γ˙,γ˙)=g((πγ),(πγ))Tr(ω(γ˙),ω(γ˙))g_{\omega}(\dot{\gamma},\dot{\gamma})=g((\pi\circ\gamma)^{\prime},(\pi\circ\gamma)^{\prime})-\operatorname{Tr}(\omega(\dot{\gamma}),\omega(\dot{\gamma}))

being constant implies that the projected curve πγ\pi\circ\gamma in MM has g((πγ),(πγ))g((\pi\circ\gamma)^{\prime},(\pi\circ\gamma)^{\prime}) constant. Since Tr(ω(γ˙),ω(γ˙))0-\operatorname{Tr}(\omega(\dot{\gamma}),\omega(\dot{\gamma}))\geq 0 (and is zero if and only if ω(γ˙)=0\omega(\dot{\gamma})=0) we see that:

gω(γ˙,γ˙)0 implies g((πγ),(πγ))0g_{\omega}(\dot{\gamma},\dot{\gamma})\leq 0\ \mbox{ implies }\ g((\pi\circ\gamma)^{\prime},(\pi\circ\gamma)^{\prime})\leq 0

and so timelike or null geodesics in (P,gω)(P,g_{\omega}) project to timelike or null curves in (M,g)(M,g). In fact, the projection will be timelike unless the geodesic in PP is null and ω(γ˙)=0\omega(\dot{\gamma})=0.

Proof.

The only part of this not proven in the statement is that ω(γ˙(t))\omega(\dot{\gamma}(t)) is constant for a geodesic γ\gamma in (P,gω)(P,g_{\omega}). This can be found in [3] theorem 10.1.5. ∎

Lemma 2.8.

In the special case where M=t×nM=\mathbb{R}_{t}\times\mathbb{R}^{n} is flat and P=M×GP=M\times G, null geodesics in PP project to solutions to Wong’s equations with x˙2\dot{x}^{2} a non-positive constant. More generally, null geodesics in (P,gω)(P,g_{\omega}) project to curves γ\gamma in (M,g)(M,g) together with a section qq of γAd(P)\gamma^{*}\operatorname{Ad}(P) satisfying:

{γ˙γ˙=γ˙Tr(qFA)(γA)q=0g(γ˙,γ˙)= constant\begin{cases}\nabla_{\dot{\gamma}}\dot{\gamma}&=-\dot{\gamma}\llcorner\operatorname{Tr}(qF_{A})\\ (\gamma^{*}\nabla^{A})q&=0\\ g(\dot{\gamma},\dot{\gamma})&=\mbox{ constant}\end{cases}

where AA is the connection on the bundle Ad(P)\operatorname{Ad}(P) induced by ω\omega.

Proof.

Let γ~\widetilde{\gamma} be a null geodesic in PP and γ\gamma the projected curve in MM. We denote q:=ω(γ~˙(t))q:=\omega(\dot{\widetilde{\gamma}}(t)) and notice that by Ad\operatorname{Ad}-equivarance of connection 1-forms on principle bundles this defines a section of γAd(P)\gamma^{*}\operatorname{Ad}(P). We’ve already seen that g(γ˙,γ˙)g(\dot{\gamma},\dot{\gamma}) is constant so it suffices to prove that qq is covariantly constant with respect to A\nabla^{A} and that the geodesic equations reduce to Wong’s equations.

The fact that the geodesic equations in (P,gω)(P,g_{\omega}) reduce to Wong’s equations on MM is theorem 10.1.6 in [3]. As for qq, we note that its covariant derivative as a section of γAd(P)\gamma^{*}\operatorname{Ad}(P) is just the horizontal part of its time derivative as a 𝔤\mathfrak{g}-valued function on a curve in PP, and this is zero since the entire time derivative vanishes. ∎

Given a null geodesic γ\gamma in (P,gω)(P,g_{\omega}), we would like to think of the constant ω(γ˙)\omega(\dot{\gamma}) as the “charge”. Unfortunately, unlike the abelian case of the Lorentz force law, different lifts of solutions to Wong’s equations in MM to geodesics in (P,gω)(P,g_{\omega}) will have different charges. Indeed, if the two lifts of our curve in MM are related by the right action of gGg\in G on PP then the charges of the two lifts will be related by Adg\operatorname{Ad}_{g}. Identifying the charge qq with Tr(q())𝔤-\operatorname{Tr}(q(-))\in\mathfrak{g}^{*} we arrive at the following gauge invariant definition of charge.

Definition 2.9.

Let γ\gamma be a null geodesic in (P,gω)(P,g_{\omega}) and ξ0:=Tr(ω(γ˙)())𝔤\xi_{0}:=-\operatorname{Tr}(\omega(\dot{\gamma})(-))\in\mathfrak{g}^{*}. The charge of γ\gamma is defined to be the coadjoint orbit:

𝒪:={Adgξ0:gG}𝔤.\mathcal{O}:=\{\operatorname{Ad}_{g}^{*}\xi_{0}\ :\ g\in G\}\subseteq\mathfrak{g}^{*}.

Just as in the flat case, Wong’s equations on a curved spacetime will arise as classical limits of the quantum system. One consequence of this will be charge quantization.

For now, let’s proceed to the Hamiltonian description of the dynamics of these null geodesics. Recall that the relativistic description of the phase space of a system is simply the space of solutions to the equations of motion, and the identification with a cotangent bundle arises from the equations typically being second order ODE and so solutions correspond to initial data. In this way, the following results and definitions can been seen as relativistic versions of the results on the phase space for Wong’s equations from [21],[24].

Definition 2.10.

The null bicharacteristic flow GsG_{s} is the Hamiltonian flow on TP0T^{*}P\setminus 0 of the Hamiltonian ξ12gω1(ξ,ξ)\xi\mapsto\frac{1}{2}g_{\omega}^{-1}(\xi,\xi).

Lemma 2.11.

Let ΦsZ\Phi^{Z}_{s} and Φsξ\Phi^{\xi}_{s} respectively denote the flows on TP0T^{*}P\setminus 0 given by the derivatives of the flows of ZωZ^{\omega} and ξ^\widehat{\xi} (ξ𝔤\xi\in\mathfrak{g}) on PP. Then ΦsZ\Phi_{s}^{Z} and Φsξ\Phi^{\xi}_{s} commute with GsG_{s} for every ξ𝔤\xi\in\mathfrak{g}.

Proof.

Since ΦsZ\Phi^{Z}_{s} and Φsξ\Phi^{\xi}_{s} are derivatives of flows on PP they are a 1-parameter family of canonical transformations on TP0T^{*}P\setminus 0 and therefore, by the Hamiltonian version of Noether’s theorem, it suffices to show that the Hamiltonian ξ12gω1(ξ,ξ)\xi\mapsto\frac{1}{2}g_{\omega}^{-1}(\xi,\xi) is invariant under the flows ΦsZ,Φsξ\Phi^{Z}_{s},\Phi^{\xi}_{s} in order to prove that they commute with GsG_{s}. But this is immediate from both ZωZ^{\omega} and ξ^\widehat{\xi} being Killing vector fields for the metric gωg_{\omega}. ∎

One incredibly important subtlety is the following. Since our spacetimes are not necessarily ultrastatic, there is no reason to expect that if ηTP0\eta\in T^{*}P_{0} then Gs(η)TPsG_{s}(\eta)\in T^{*}P_{s}. This is our reason for using the variable ss instead of tt. Indeed, we do have (by definition) that ΦtZ(η)TPt\Phi^{Z}_{t}(\eta)\in T^{*}P_{t} for ηTP0\eta\in T^{*}P_{0}.

Definition 2.12.

We begin by denoting:

𝒩a:={ all future-directed, inextendible null geodesics in (P,gω)}.\mathcal{N}_{a}:=\{\mbox{ all future-directed, inextendible null geodesics in }(P,g_{\omega})\}.

Recall that, for us, geodesics are specifically solutions to the equation γ˙γ˙=0\nabla_{\dot{\gamma}}\dot{\gamma}=0 and hence these are already “affinely parametrized”. By definition, elements of 𝒩a\mathcal{N}_{a} are, in particular, inextendible causal curves and therefore intersect each Cauchy hypersurface PtP_{t} exactly once. An invariant way of dealing with the fact that such curves γ\gamma need not satisfy γ(0)P0\gamma(0)\in P_{0} is to define

𝒩:=𝒩a/\mathcal{N}:=\mathcal{N}_{a}/\mathbb{R}

where bb\in\mathbb{R} acts on 𝒩a\mathcal{N}_{a} by γ(s)γ(s+b)\gamma(s)\mapsto\gamma(s+b). Notice additionally that the >0\mathbb{R}_{>0}-action on 𝒩a\mathcal{N}_{a} where a>0a\in\mathbb{R}_{>0} acts by γ(s)γ(as)\gamma(s)\mapsto\gamma(as) descends to an >0\mathbb{R}_{>0}-action on the quotient 𝒩\mathcal{N}.

The above set 𝒩\mathcal{N} with >0\mathbb{R}_{>0}-action is naturally a symplectic manifold with >0\mathbb{R}_{>0}-action and one can view the next lemma as saying the there are >0\mathbb{R}_{>0}-equivariant Cauchy-data symplectomorphisms between 𝒩\mathcal{N} and cotangent bundles. Instead, we will simply take the next lemma as a definition of the smooth manifold and symplectic structures. For this, we will need the definition.

Definition 2.13.

We define three sub-cone-bundles of TP0T^{*}P\setminus 0:

T0P:={ζTP0:gω1(ζ,ζ)=0}T^{*}_{0}P:=\{\zeta\in T^{*}P\setminus 0\ :\ g^{-1}_{\omega}(\zeta,\zeta)=0\}

and

T±P:={ζT0P:ζ is future (respectively past) oriented}T^{*}_{\pm}P:=\{\zeta\in T^{*}_{0}P\ :\ \zeta\mbox{ is future (respectively past) oriented}\}

so that T0P=T+PTPT^{*}_{0}P=T_{+}^{*}P\sqcup T_{-}^{*}P.

As in [22] there are natural isomorphisms of bundles over P0P_{0}:

TP00T+P|P0 and TP00TP|P0T^{*}P_{0}\setminus 0\cong T^{*}_{+}P|_{P_{0}}\ \mbox{ and }\ T^{*}P_{0}\setminus 0\cong T^{*}_{-}P|_{P_{0}}

which are symplectomorphisms but are not >0\mathbb{R}_{>0}-equivariant! Indeed, the map

TP00T+P|P0T^{*}P_{0}\setminus 0\xrightarrow{\cong}T^{*}_{+}P|_{P_{0}}

is given by ζζ+|ζ|h2n^\zeta\mapsto\zeta+|\zeta|_{h}^{2}\widehat{n}.

Lemma 2.14.

Each equivalence class in 𝒩\mathcal{N} has a unique representative γ:P\gamma:\mathbb{R}\to P satisfying γ(0)P0\gamma(0)\in P_{0}. Identifying elements of 𝒩\mathcal{N} with these representatives gives us an >0\mathbb{R}_{>0}-equivariant bijection

𝒩\displaystyle\mathcal{N} T+P|P0\displaystyle\xrightarrow{\cong}T^{*}_{+}P|_{P_{0}}
γ\displaystyle\gamma (γ(0),γ˙(0))\displaystyle\mapsto(\gamma(0),\dot{\gamma}(0))

where γ(0)\gamma^{\prime}(0) is identified with a cotangent vector via gωg_{\omega}. The >0\mathbb{R}_{>0}-action on T+P|P0T^{*}_{+}P|_{P_{0}} is given by scalar multiplication in the fibers. Furthermore, the inverse of the above bijection is given by

T+P|P00η(sGs(η))T^{*}_{+}P|_{P_{0}}\setminus 0\ni\eta\mapsto(s\mapsto G_{s}(\eta))

or, more precisely, η\eta maps to the projection of the curve sGs(η)s\mapsto G_{s}(\eta) down to PP.

Proof.

This is immediate from the definition of a future-directed, inextendible null geodesic and the existence and uniqueness of solutions to ODE. ∎

Lemma 2.15.

gGg\in G has a right action on 𝒩\mathcal{N} induced by its right action on 𝒩a\mathcal{N}_{a} given by γ(s)γ(s)g\gamma(s)\mapsto\gamma(s)g (via the right action on PP). The bijection in 2.14 intertwines this right action with the right action on T+P|P0T^{*}_{+}P|_{P_{0}} given by dualizing (using gωg_{\omega}) the action of pushing forward by right multiplication by gg on PP.

Proof.

This is immediate from the explicit form of our isomorphism 𝒩T+P|P0\mathcal{N}\cong T^{*}_{+}P|_{P_{0}} and the fact that GG acts by isometries and therefore leaves T+P|P0T^{*}_{+}P|_{P_{0}} invariant. ∎

Lemma 2.16.

The flows ΦsZ\Phi^{Z}_{s} and Φsξ\Phi^{\xi}_{s} on 𝒩\mathcal{N} induced by 2.14 are Hamiltonian flows with respective Hamiltonians:

HZ(γ)=Zω(γ(0),γ˙(0)) and Hξ(γ)=ξ^(γ(0),γ˙(0))H_{Z}(\gamma)=Z^{\omega}\llcorner(\gamma(0),\dot{\gamma}(0))\ \mbox{ and }\ H_{\xi}(\gamma)=\widehat{\xi}\llcorner(\gamma(0),\dot{\gamma}(0))

where again we have chosen representative geodesics γ\gamma with γ(0)P0\gamma(0)\in P_{0}. Furthermore, the Φsξ\Phi^{\xi}_{s}’s arise (through the exponential map) from the natural right-action of GG on 𝒩\mathcal{N} hence this GG-action is Hamiltonian.

As the above right GG-action is Hamiltonian, we can consider its moment-map:

μ:𝒩\displaystyle\mu:\mathcal{N} 𝔤\displaystyle\to\mathfrak{g}^{*}
μ(γ),ξ\displaystyle\langle\mu(\gamma),\xi\rangle =Hξ(γ).\displaystyle=H_{\xi}(\gamma).
Lemma 2.17.

Under the isomorphism 𝔤𝔤\mathfrak{g}\cong\mathfrak{g}^{*} induced by our Ad\operatorname{Ad}-invariant inner product on 𝔤\mathfrak{g}, the moment map is given by

γω(γ˙).\gamma\mapsto\omega(\dot{\gamma}).
Proof.

We know that ω(ξ^)=ξ\omega(\widehat{\xi})=\xi by the definition of a connection on a principal bundle and so the result follows from

ξ^(γ(0),γ˙(0))=Tr(ω(ξ^)ω(γ˙)T)\widehat{\xi}\llcorner(\gamma(0),\dot{\gamma}(0))=\operatorname{Tr}(\omega(\widehat{\xi})\omega(\dot{\gamma})^{T})

since we’re using gωg_{\omega} to identify γ˙(0)\dot{\gamma}(0) with a covector. ∎

As a final remark before we proceed to symplectic reduction, we demonstrate that, while our Hamiltonian may appear linear (indeed, it is homogeneous of degree 1 with respect to the >0\mathbb{R}_{>0}-action on 𝒩\mathcal{N}), it is in fact quadratic after applying the symplectomorphism TP00𝒩T^{*}P_{0}\setminus 0\cong\mathcal{N}.

Lemma 2.18.

Under the symplectomorphism TP00𝒩T^{*}P_{0}\setminus 0\cong\mathcal{N} the Hamiltonian HZH_{Z} becomes:

HZ:TP00\displaystyle H_{Z}:T^{*}P_{0}\setminus 0 \displaystyle\to\mathbb{R}
ζ\displaystyle\zeta N|ζ|h2+η,ζh\displaystyle\mapsto N|\zeta|_{h}^{2}+\langle\eta,\zeta\rangle_{h}

where η\eta is the 1-form on P0P_{0} coming from our explicit form for the metric gωg_{\omega} in 1.

Proof.

This follows from the isomorphism TP0T+P|P0T^{*}P_{0}\setminus\cong T^{*}_{+}P|_{P_{0}} being given by ζζ+|ζ|h2n^\zeta\mapsto\zeta+|\zeta|_{h}^{2}\widehat{n} and gω(Zω,n^)=N1g_{\omega}(Z^{\omega},\widehat{n})=N^{-1}. ∎

Notice in particular that the fact that N>|η|hN>|\eta|_{h} pointwise implies that HZH_{Z} is strictly positive. Furthermore, if we had N|η|hN-|\eta|_{h} uniformly bounded away from zero on Σ0\Sigma_{0} then HZH_{Z} would both be uniformly bounded away from zero and would have a uniformly positive definite fiberwise Hessian.

2.1 The Reduced Phase Space

Fix a charge, i.e. a coadjoint orbit 𝒪𝔤\mathcal{O}\subseteq\mathfrak{g}^{*}. We now wish to form the symplectically reduced phase space of solutions with charge 𝒪\mathcal{O}. The construction of this in Riemannian signature, and its relationship to Wong’s equations can be found in [9] and it generalizes with almost no modifications to our setting.

Recall that our coadjoint orbit 𝒪\mathcal{O} is naturally a symplectic manifold. The symplectic form ω𝒪\omega_{\mathcal{O}} can be defined as follows. Fix ξ0𝒪\xi_{0}\in\mathcal{O} and let Gξ0G_{\xi_{0}} denote the stabilizer of ξ0\xi_{0} under the coadjoint action. Then

G\displaystyle G 𝒪\displaystyle\to\mathcal{O}
g\displaystyle g Adgξ0\displaystyle\mapsto\operatorname{Ad}_{g}^{*}\xi_{0}

induces an isomorphism

G/Gξ0𝒪G/G_{\xi_{0}}\cong\mathcal{O}

which identifies

Tξ0𝒪𝔤/𝔤ξ0T_{\xi_{0}}\mathcal{O}\cong\mathfrak{g}/\mathfrak{g}_{\xi_{0}}

where 𝔤ξ0\mathfrak{g}_{\xi_{0}} is the Lie algebra of Gξ0G_{\xi_{0}}. The other tangent spaces of 𝒪\mathcal{O} are also identified with 𝔤/𝔤ξ0\mathfrak{g}/\mathfrak{g}_{\xi_{0}} by pushforward along the GG-action. We then have:

ω𝒪(X,Y)=ξ0,[X,Y].\omega_{\mathcal{O}}(X,Y)=\langle\xi_{0},[X,Y]\rangle.

We notice that this is well-defined on 𝔤/𝔤ξ0\mathfrak{g}/\mathfrak{g}_{\xi_{0}} since

𝔤ξ0={X𝔤:ξ0,[X,Y]=0 for all Y𝔤}.\mathfrak{g}_{\xi_{0}}=\{X\in\mathfrak{g}\ :\ \langle\xi_{0},[X,Y]\rangle=0\ \mbox{ for all }Y\in\mathfrak{g}\}.

Let 𝒪¯\overline{\mathcal{O}} denote 𝒪\mathcal{O} but equipped with ω𝒪-\omega_{\mathcal{O}} as its symplectic form instead of 𝒪\mathcal{O}.

Lemma 2.19.

The extended moment map

μ𝒪:𝒩×𝒪¯\displaystyle\mu_{\mathcal{O}}:\mathcal{N}\times\overline{\mathcal{O}} 𝔤\displaystyle\to\mathfrak{g}^{*}
μ𝒪(γ,ξ)\displaystyle\mu_{\mathcal{O}}(\gamma,\xi) :=μ(γ)ξ\displaystyle:=\mu(\gamma)-\xi

is a submersion and GG acts freely on μ𝒪1(0)\mu_{\mathcal{O}}^{-1}(0).

Proof.

The fact that GG acts freely on μ𝒪1(0)\mu_{\mathcal{O}}^{-1}(0) simply follows from GG acting freely on 𝒩T+P|P0\mathcal{N}\cong T^{*}_{+}P|_{P_{0}} since P,P0P,P_{0} are principal GG-bundles. To see that μ𝒪\mu_{\mathcal{O}} is a submersion, we notice that under the isomorphism 𝒩T+P|P0\mathcal{N}\cong T^{*}_{+}P|_{P_{0}} we have

μ𝒪:T+P|P0×𝒪¯\displaystyle\mu_{\mathcal{O}}:T^{*}_{+}P|_{P_{0}}\times\overline{\mathcal{O}} 𝔤\displaystyle\to\mathfrak{g}^{*}
(ζ,ξ)\displaystyle(\zeta,\xi) Tr(ω(ζ)Tω())ξ\displaystyle\mapsto\operatorname{Tr}(\omega(\zeta)^{T}\omega(-))-\xi

and if we use our Ad\operatorname{Ad}-invariant inner product to identify 𝔤𝔤\mathfrak{g}\cong\mathfrak{g}^{*} then this maps

(ζ,ξ)ω(ζ)ξ.(\zeta,\xi)\mapsto\omega(\zeta)-\xi.

Forgetting ξ\xi we can already see that ζω(ζ)\zeta\mapsto\omega(\zeta) is a submersion (and therefore μ𝒪\mu_{\mathcal{O}} is a submersion). Indeed, it suffices to prove that for every ξ𝔤\xi\in\mathfrak{g} there exists ζT+P|P0\zeta\in T^{*}_{+}P|_{P_{0}} such that ω(ζ)=ξ\omega(\zeta)=\xi. However, ξ^\widehat{\xi} is tangent to P0P_{0} with gω(ξ^,ξ^)=Tr(ξξT)g_{\omega}(\widehat{\xi},\widehat{\xi})=\operatorname{Tr}(\xi\xi^{T}) so ζ:=Tr(ξξT)n^+ξ^\zeta:=\operatorname{Tr}(\xi\xi^{T})\widehat{n}+\widehat{\xi} is future-directed, has ω(ζ)=ξ\omega(\zeta)=\xi and gω(ζ,ζ)=0g_{\omega}(\zeta,\zeta)=0 as desired. ∎

From the above proof we record as a remark the fact that μ𝒪1(0)\mu_{\mathcal{O}}^{-1}(0) is precisely the space of pairs (γ,ξ)(\gamma,\xi) where γ𝒩\gamma\in\mathcal{N} and ξ𝒪\xi\in\mathcal{O} satisfy

Tr(ω(γ˙)T())=ξ.\operatorname{Tr}(\omega(\dot{\gamma})^{T}(-))=\xi.

This μ𝒪1(0)\mu_{\mathcal{O}}^{-1}(0) is precisely the space of solutions with charge 𝒪\mathcal{O}, prior to quotienting by gauge transformations.

Definition 2.20.

The reduced phase space is

𝒩𝒪:=μ𝒪1(0)/G\mathcal{N}_{\mathcal{O}}:=\mu_{\mathcal{O}}^{-1}(0)/G

with symplectic form obtained from the one on 𝒩×𝒪¯\mathcal{N}\times\overline{\mathcal{O}}.

Lemma 2.21.

The Hamiltonian HZH_{Z}, extended to 𝒩×𝒪¯\mathcal{N}\times\overline{\mathcal{O}} to be independent of 𝒪¯\overline{\mathcal{O}}, is invariant under the GG-action and therefore descends to a Hamiltonian H~Z\widetilde{H}_{Z} on 𝒩𝒪\mathcal{N}_{\mathcal{O}} with flow Φ~sZ\widetilde{\Phi}^{Z}_{s}.

Proof.

From the definition of HZH_{Z} we see that what we have to show is that gω(Zω,γ˙g)=gω(Zω,γ˙)g_{\omega}(Z^{\omega},\dot{\gamma}\cdot g)=g_{\omega}(Z^{\omega},\dot{\gamma}) for all gGg\in G and γ𝒩\gamma\in\mathcal{N}. However:

gω(Zω,γ˙g)=gω(Zωg1,γ˙)=gω(Zω,γ˙)g_{\omega}(Z^{\omega},\dot{\gamma}\cdot g)=g_{\omega}(Z^{\omega}\cdot g^{-1},\dot{\gamma})=g_{\omega}(Z^{\omega},\dot{\gamma})

since Zω=tZ^{\omega}=\partial_{t} is invariant under the GG action. ∎

The point of the previous construction is its manifestly gauge-invariant nature. Below we give an alternative characterization that might be more familiar to some readers, although we will not use it in our proof.

Fix ξ0𝒪\xi_{0}\in\mathcal{O} and recall from our proof that μ𝒪\mu_{\mathcal{O}} is a submersion that μ\mu is also a submersion, hence ξ0\xi_{0} is automatically a regular value. Furthermore, while the full GG-action on 𝒩\mathcal{N} doesn’t preserve the submanifold μ1(ξ0)\mu^{-1}(\xi_{0}), it is preserved by the action of the stabilizer Gξ0G_{\xi_{0}} of ξ0\xi_{0}. The action of Gξ0G_{\xi_{0}} on μ1(ξ0)\mu^{-1}(\xi_{0}) is free since the action of GG on 𝒩\mathcal{N} is free.

Definition 2.22.

The reduced phase space (version II) is the quotient

μ1(ξ0)/Gξ0\mu^{-1}(\xi_{0})/G_{\xi_{0}}

with the symplectic form induced from that on 𝒩\mathcal{N}.

Lemma 2.23.

[11] The map

μ1(ξ0)\displaystyle\mu^{-1}(\xi_{0}) 𝒩𝒪\displaystyle\to\mathcal{N}_{\mathcal{O}}
γ\displaystyle\gamma [(γ,ξ0)]\displaystyle\mapsto[(\gamma,\xi_{0})]

induces a symplectomorphism μ1(ξ0)/Gξ0𝒩𝒪\mu^{-1}(\xi_{0})/G_{\xi_{0}}\cong\mathcal{N}_{\mathcal{O}} intertwining the reductions of the Hamiltonian flow of HZH_{Z} to μ1(ξ0)/Gξ0\mu^{-1}(\xi_{0})/G_{\xi_{0}} and 𝒩𝒪\mathcal{N}_{\mathcal{O}}. Here [(γ,ξ0)][(\gamma,\xi_{0})] denotes the equivalence class of (γ,ξ0)(\gamma,\xi_{0}) in the quotient.

2.2 Periodic Orbits

Finally, let’s note that since MM is assumed to be spatially compact we expect the quantum system to have discrete spectrum and hence bound states. The leading order singularities in our distributional trace of the propagator will be therefore expressed as a sum over classical bound states: periodic orbits of null geodesics under Φ~sZ\widetilde{\Phi}^{Z}_{s}. There are two aspects of these periodic orbits we will need to consider:

  1. 1.

    the (linearized) Poincaré first return map of a periodic orbit, and

  2. 2.

    the phase change due to a periodic orbit for the Aharonov-Bohm effect.

The first of these points relates to the classical dynamics of periodic orbits, while the second of these is only relevant for the quantum effects we will discuss later.

Following [22], we fix an energy EE\in\mathbb{R} and restrict ourselves to the contact manifold given by the level surface

H~Z1(E)𝒩𝒪.\widetilde{H}_{Z}^{-1}(E)\subseteq\mathcal{N}_{\mathcal{O}}.

This is invariant under the Φ~sZ\widetilde{\Phi}^{Z}_{s}-flow and so we can define the set of periods:

𝒫E:={T{0}:zH~Z1(E) such that Φ~TZ(z)=z}\mathcal{P}_{E}:=\{T\in\mathbb{R}\setminus\{0\}\ :\ \exists z\in\widetilde{H}_{Z}^{-1}(E)\ \mbox{ such that }\ \widetilde{\Phi}^{Z}_{T}(z)=z\}

and, for T𝒫ET\in\mathcal{P}_{E}, the set of periodic points:

𝒫E,T:={zH~Z1(E):Φ~TZ(z)=z.}\mathcal{P}_{E,T}:=\{z\in\widetilde{H}_{Z}^{-1}(E)\ :\ \widetilde{\Phi}_{T}^{Z}(z)=z.\}

We say that T>0T>0 is the minimum period of zz if and only if it is the smallest positive time for which Φ~TZ(z)=z\widetilde{\Phi}_{T}^{Z}(z)=z. The below result is a general fact concerning Hamiltonian dynamics and is a simple consequence of the implicit function theorem.

Lemma 2.24.

([17] Prop 8.5.3)
Given a periodic point z0𝒫E,Tz_{0}\in\mathcal{P}_{E,T} where TT is its minimum period there exists, in a sufficiently small neighborhood of z0z_{0}, a codimension 1 symplectic submanifold

z0SH~Z1(E)z_{0}\in S\subseteq\widetilde{H}_{Z}^{-1}(E)

which is transverse to the flow Φ~sZ\widetilde{\Phi}_{s}^{Z}. Furthermore, in a sufficiently small neighborhood of z0z_{0} in SS, the first return time

𝒯(z):=min{t>0:Φ~tZ(z)S}\mathcal{T}(z):=\min\{t>0\ :\ \widetilde{\Phi}_{t}^{Z}(z)\in S\}

is well-defined, smooth and satisfies 𝒯(z0)=T\mathcal{T}(z_{0})=T.

Definition 2.25.

With z0,S,Tz_{0},S,T as above, we define the linearized Poincaré first return map to be

Pz0,S:=z|z=z0Φ~𝒯(z)Z(z):Tz0STz0S.P_{z_{0},S}:=\frac{\partial}{\partial z}\Big{|}_{z=z_{0}}\widetilde{\Phi}^{Z}_{\mathcal{T}(z)}(z):T_{z_{0}}S\to T_{z_{0}}S.

This is a linear symplectic map. For any other choice of local symplectic transversal SS^{\prime} there is a linear symplectic isomorphism

L:Tz0STz0SL:T_{z_{0}}S^{\prime}\xrightarrow{\cong}T_{z_{0}}S

such that

Pz0,S=L1Pz0,SL.P_{z_{0},S^{\prime}}=L^{-1}\circ P_{z_{0},S}\circ L.

There is actually an alternate, perhaps simpler, description of these maps Pz0,SP_{z_{0},S}. This alternate description is analogous to the more standard definition of the linearized Poincaré first return map for geodesic flow on Riemannian or Lorentzian manifolds, which is usually defined with the aid of Jacobi fields.

Definition 2.26.

Given z0𝒫E,Tz_{0}\in\mathcal{P}_{E,T} with TT the minimum period of z0z_{0}, we define the Floquet operator of z0z_{0} to be:

Vz0(T):=ddz|z=z0Φ~TZ(z):Tz0𝒩𝒪Tz0𝒩𝒪.V_{z_{0}}(T):=\frac{d}{dz}\Big{|}_{z=z_{0}}\widetilde{\Phi}^{Z}_{T}(z):T_{z_{0}}\mathcal{N}_{\mathcal{O}}\to T_{z_{0}}\mathcal{N}_{\mathcal{O}}.
Lemma 2.27.

The subspace

Wz0:=Span{Z~(z0),H~(z0)}W_{z_{0}}:=\operatorname{Span}\{\widetilde{Z}(z_{0}),\ \nabla\widetilde{H}(z_{0})\}

is symplectic, as is the quotient Tz0𝒩𝒪/Wz0T_{z_{0}}\mathcal{N}_{\mathcal{O}}/W_{z_{0}}, and Wz0W_{z_{0}} is preserved by the Floquet operator. The induced quotient map

Vz0(T):Tz0𝒩𝒪/Wz0Tz0𝒩𝒪/Wz0V_{z_{0}}(T):T_{z_{0}}\mathcal{N}_{\mathcal{O}}/W_{z_{0}}\to T_{z_{0}}\mathcal{N}_{\mathcal{O}}/W_{z_{0}}

is conjugate via a linear symplectomorphism to the linearized Poincaré first return map.

Let’s discuss for some time the significance of these operators to us. For this, we will need the following assumption.

Definition 2.28.

We say that EE satisfies the clean intersection hypothesis if and only if EE is a regular value for H~Z\widetilde{H}_{Z} and the flow map

×H~Z1(E)\displaystyle\mathbb{R}\times\widetilde{H}_{Z}^{-1}(E) H~Z1(E)×H~Z1(E)\displaystyle\to\widetilde{H}_{Z}^{-1}(E)\times\widetilde{H}_{Z}^{-1}(E)
(t,γ)\displaystyle(t,\gamma) (γ,Φ~tZ(γ))\displaystyle\mapsto(\gamma,\widetilde{\Phi}^{Z}_{t}(\gamma))

admits a clean fibered product over H~1(E)×H~1(E)\widetilde{H}^{-1}(E)\times\widetilde{H}^{-1}(E) with the diagonal map H~1(E)H~1(E)×H~1(E)\widetilde{H}^{-1}(E)\to\widetilde{H}^{-1}(E)\times\widetilde{H}^{-1}(E).

Let’s discuss this hypothesis for a moment. The fibered product is given, as a set, by:

𝔜E:={(T,γ)×H~Z1(E):Φ~TZ(γ)=γ}.\mathfrak{Y}_{E}:=\{(T,\gamma)\in\mathbb{R}\times\widetilde{H}^{-1}_{Z}(E)\ :\ \widetilde{\Phi}^{Z}_{T}(\gamma)=\gamma\}.

Notice that this contains {0}×H~Z1(E)\{0\}\times\widetilde{H}_{Z}^{-1}(E) as a subset and the clean intersection hypothesis implies that 𝔜E\mathfrak{Y}_{E} is a disjoint union of smooth submanifolds of ×H~Z1(E)\mathbb{R}\times\widetilde{H}^{-1}_{Z}(E).

Lemma 2.29.

Under the clean intersection hypothesis, {0}×H~Z1(E)\{0\}\times\widetilde{H}_{Z}^{-1}(E) is a clopen subset of 𝔜E\mathfrak{Y}_{E} and every connected component Y𝔜EY\subseteq\mathfrak{Y}_{E} has

dim(Y)dimH~Z1(E)=2n+dim𝒪1.\dim(Y)\leq\dim\widetilde{H}_{Z}^{-1}(E)=2n+\dim\mathcal{O}-1.
Proof.

Let Y𝔜EY\subseteq\mathfrak{Y}_{E} be any connected component. By the clean intersection hypothesis, for any (T,γ)Y(T,\gamma)\in Y we must have

T(T,γ)Y={(τ,ζ)TT×TγH~Z1(E):τddt|t=TΦ~tZ(γ)+DΦ~TZ(ζ)=ζ}T_{(T,\gamma)}Y=\left\{(\tau,\zeta)\in T_{T}\mathbb{R}\times T_{\gamma}\widetilde{H}^{-1}_{Z}(E)\ :\ \tau\frac{d}{dt}\Big{|}_{t=T}\widetilde{\Phi}^{Z}_{t}(\gamma)+D\widetilde{\Phi}^{Z}_{T}(\zeta)=\zeta\right\}

Since ζζDΦ~TZ(ζ)\zeta\mapsto\zeta-D\widetilde{\Phi}^{Z}_{T}(\zeta) is linear the only way for the above constraint to be trivial (and not reduce the dimension) is if ddt|t=TΦ~tZ(γ)=0\frac{d}{dt}|_{t=T}\widetilde{\Phi}^{Z}_{t}(\gamma)=0 and if DΦ~TZ=idD\widetilde{\Phi}^{Z}_{T}=\operatorname{id}. Indeed, if ddt|t=TΦ~tZ(γ)0\frac{d}{dt}|_{t=T}\widetilde{\Phi}^{Z}_{t}(\gamma)\neq 0 and we didn’t want the equation to constrain ζ\zeta then we would need to constrain τ\tau to τ=0\tau=0. But now since Φ~TZ(γ)=γ\widetilde{\Phi}^{Z}_{T}(\gamma)=\gamma it follows that ddt|t=TΦ~tZ(γ)=0\frac{d}{dt}|_{t=T}\widetilde{\Phi}^{Z}_{t}(\gamma)=0 implies ddt|t=0Φ~tZ(γ)=0\frac{d}{dt}|_{t=0}\widetilde{\Phi}^{Z}_{t}(\gamma)=0 and so the gradient of the Hamiltonian H~Z\nabla\widetilde{H}_{Z} vanishes at γ\gamma and so γ\gamma is an equilibrium point. However, we assumed that γH~Z1(E)\gamma\in\widetilde{H}^{-1}_{Z}(E) and that EE was a regular value for H~Z\widetilde{H}_{Z}, which contradicts H~Z\nabla\widetilde{H}_{Z} vanishing at γ\gamma.

Now, let YY be the smallest clopen subset containing {0}×H~Z1(E)\{0\}\times\widetilde{H}_{Z}^{-1}(E). We have already shown that dim(Y)dimH~Z1(E)\dim(Y^{\prime})\leq\dim\widetilde{H}_{Z}^{-1}(E) for any connected component YY^{\prime} and so we must have dim(Y)=dimH~Z1(E)\dim(Y)=\dim\widetilde{H}_{Z}^{-1}(E) since YY is a disjoint union of connected components. In particular, since the inclusion

{0}×H~Z1(E)Y\{0\}\times\widetilde{H}_{Z}^{-1}(E)\hookrightarrow Y

is an immersion it is automatically a submersion as well and hence a local diffeomorphism. Local diffeomorphisms are local homeomorphisms and are hence open maps. Thus the image {0}×H~Z1(E)\{0\}\times\widetilde{H}_{Z}^{-1}(E) is open in YY, hence open in 𝔜E\mathfrak{Y}_{E} since YY is open in 𝔜E\mathfrak{Y}_{E}. Since {0}×H~Z1(E)\{0\}\times\widetilde{H}_{Z}^{-1}(E) is also closed in 𝔜E\mathfrak{Y}_{E} it follows that it is clopen hence

{0}×H~Z1(E)=Y\{0\}\times\widetilde{H}_{Z}^{-1}(E)=Y

as desired. ∎

We should remark that there is no reason to expect H~Z1(E)\widetilde{H}_{Z}^{-1}(E) to be connected even if MM is connected since we have allowed disconnected structure groups such as G=O(d)G=\operatorname{O}(d).

In our trace formula, the leading order singularities of the distributional trace will have symbols given by integrals over components of the above clean intersection. The linearized Poincaré map gives us a dynamical description of the volume density on these components. To describe how, let’s first recall the invariant volume density on the energy hypersurface H~Z1(E)\widetilde{H}^{-1}_{Z}(E).

Definition 2.30.

Let Ω\Omega denote the volume form on 𝒩𝒪\mathcal{N}_{\mathcal{O}} induced by the symplectic form and equip 𝒩𝒪\mathcal{N}_{\mathcal{O}} with the Riemannian metric h0h_{0} induced from the one on 𝒩T+P|P0TP00\mathcal{N}\cong T^{*}_{+}P|_{P_{0}}\cong T^{*}P_{0}\setminus 0 and the Ad\operatorname{Ad}-invariant inner product on 𝔤\mathfrak{g}. Using this metric we can define the gradient H\nabla H and the 2(n+)12(n+\ell)-1-form on 𝒩𝒪\mathcal{N}_{\mathcal{O}}:

|H~Z|h02H~ZΩ.|\nabla\widetilde{H}_{Z}|_{h_{0}}^{-2}\nabla\widetilde{H}_{Z}\llcorner\Omega.

Denote:

νE:= the pullback of the above form to H~Z1(E).\nu_{E}:=\mbox{ the pullback of the above form to }\widetilde{H}_{Z}^{-1}(E).

The νE\nu_{E} is invariant under the Hamiltonian flow Φ~tZ\widetilde{\Phi}_{t}^{Z} and its absolute value |νE||\nu_{E}| defines an invariant measure on the energy hypersurface H~Z1(E)\widetilde{H}_{Z}^{-1}(E).

Lemma 2.31.

Under the clean intersection hypothesis, the fibered product 𝔜E\mathfrak{Y}_{E} comes equipped with a natural volume density. Consider then the case 𝔜E\mathfrak{Y}_{E} is a union of {0}×H~Z1(E)\{0\}\times\widetilde{H}_{Z}^{-1}(E) and finitely many disjoint isolated orbits:

Y1:={(T1,Φ~tZ(γ1)):t[0,T1]},,Yq:={(Tq,Φ~tZ(γq)):t[0,Tq]}Y_{1}:=\{(T_{1},\widetilde{\Phi}_{t}^{Z}(\gamma_{1}))\ :\ t\in[0,T_{1}]\},\ ...,\ Y_{q}:=\{(T_{q},\widetilde{\Phi}^{Z}_{t}(\gamma_{q}))\ :\ t\in[0,T_{q}]\}

with Tj0T_{j}\neq 0 for all jj. Then the Poincaré first return map of each γj\gamma_{j} is invertible and if Ωγ\Omega_{\gamma} is the symplectic volume form on Tγ𝒩𝒪T^{*}_{\gamma}\mathcal{N}_{\mathcal{O}} then the induced volume density on TγYjT^{*}_{\gamma}Y_{j} is given by:

|det(IPγ)|1/2|νE||\det(I-P_{\gamma})|^{-1/2}|\nu_{E}| (2)

with PγP_{\gamma} the Poincaré first return map for one, hence any, choice of symplectic local transversal SS.

Proof.

Indeed this follows immediately from the expression for the tangent space of YY derived in the proof of 2.29, noticing that the constraint

τddt|t=TΦ~tZ(γ)+DΦ~TZ(ζ)=ζ\tau\frac{d}{dt}\Big{|}_{t=T}\widetilde{\Phi}_{t}^{Z}(\gamma)+D\widetilde{\Phi}^{Z}_{T}(\zeta)=\zeta

in the case of isolated periodic orbits is such that ζζDΦ~TZ(ζ)\zeta\mapsto\zeta-D\widetilde{\Phi}^{Z}_{T}(\zeta) has a 1-dimensional kernel in TγH~Z1(E)=(H~Z(γ))T_{\gamma}\widetilde{H}^{-1}_{Z}(E)=(\nabla\widetilde{H}_{Z}(\gamma))^{\perp} spanned the vector field Z~\widetilde{Z} corresponding to the reduced flow Φ~tZ\widetilde{\Phi}^{Z}_{t}. Thus by 2.27 the Poincaré first return map is invertible and the induced volume form on TYjTY_{j} is determined by the invariant volume form νE\nu_{E} on H~Z1(E)\widetilde{H}_{Z}^{-1}(E) and the Poincaré first return map acting on

T𝒩𝒪/Span{Z~,H~Z}TH~Z1(E)/Span{Z~}T\mathcal{N}_{\mathcal{O}}/\operatorname{Span}\{\widetilde{Z},\nabla\widetilde{H}_{Z}\}\cong T\widetilde{H}_{Z}^{-1}(E)/\operatorname{Span}\{\widetilde{Z}\}

yielding the formula 2. ∎

Next, let’s discuss the phase associated to a periodic orbit. For this we need the following basic result from representation theory.

Definition 2.32.

The coadjoint orbit 𝒪𝔤\mathcal{O}\subseteq\mathfrak{g}^{*} is called integral if and only if the cohomology class [ω𝒪][\omega_{\mathcal{O}}] of its symplectic form ω𝒪\omega_{\mathcal{O}} is in the image of H2(𝒪;)H2(𝒪;)HdR2(𝒪;)H^{2}(\mathcal{O};\mathbb{Z})\to H^{2}(\mathcal{O};\mathbb{R})\cong H^{2}_{dR}(\mathcal{O};\mathbb{R}).

Lemma 2.33.

A coadjoint orbit 𝒪=Gξ0\mathcal{O}=G\cdot\xi_{0} is integral if and only if there exists a character

χξ0:Gξ0U(1)\chi_{\xi_{0}}:G_{\xi_{0}}\to\operatorname{U}(1)

such that

(dχξ0)I=2πiξ0,:𝔤ξ0i(d\chi_{\xi_{0}})_{I}=2\pi i\langle\xi_{0},-\rangle:\mathfrak{g}_{\xi_{0}}\to i\mathbb{R}

where IGI\in G is the identity matrix.

So, when our coadjoint orbit is integral we have a U(1)\operatorname{U}(1)-bundle defined by the character:

G×χξ0U(1)𝒪G\times_{\chi_{\xi_{0}}}\operatorname{U}(1)\to\mathcal{O}

where G×χξ0U(1)G\times_{\chi_{\xi_{0}}}\operatorname{U}(1) is the quotient of G×U(1)G\times\operatorname{U}(1) by the relation

(g,z)(gh,χξ0(h1)z) for all hGξ0.(g,z)\sim(gh,\chi_{\xi_{0}}(h^{-1})z)\ \mbox{ for all }\ h\in G_{\xi_{0}}.

The right GG-action on GG yields a right GG-action on the total space G×χξ0U(1)G\times_{\chi_{\xi_{0}}}\operatorname{U}(1) since the stabilizer Gξ0G_{\xi_{0}} is a normal subgroup. Through this, we identify every tangent space of the total space with the tangent space at the equivalence class [I,1]G×ξ0U(1)[I,1]\in G\times_{\xi_{0}}\operatorname{U}(1) of (I,1)G×U(1)(I,1)\in G\times\operatorname{U}(1).

Lemma 2.34.

We have a natural isomorphism

T[I,1](G×χξ0U(1)){(Y+cξ0#,i2πc):Y𝔤ξ0,c}.T_{[I,1]}(G\times_{\chi_{\xi_{0}}}\operatorname{U}(1))\cong\left\{\left(Y+c\xi_{0}^{\#},\ \frac{i}{2\pi}c\right)\ :\ Y\in\mathfrak{g}_{\xi_{0}}^{\perp},\ c\in\mathbb{R}\right\}.

Furthermore, there is a principal U(1)\operatorname{U}(1)-connection α\alpha on G×ξ0U(1)G\times_{\xi_{0}}\operatorname{U}(1) such that dα=ω𝒪d\alpha=\omega_{\mathcal{O}} and, under our above isomorphism, it is given by:

α(Y+cξ0#,i2πc)=c=ξ0,Y+cξ0#.\alpha\left(Y+c\xi_{0}^{\#},\ \frac{i}{2\pi}c\right)=c=\langle\xi_{0},Y+c\xi_{0}^{\#}\rangle.

This GG-equivariant bundle with connection over 𝒪\mathcal{O} gives us a natural U(1)\operatorname{U}(1)-bundle with connection over the reduced phase space 𝒩𝒪\mathcal{N}_{\mathcal{O}}, which we describe now.

Definition 2.35.

The U(1)\operatorname{U}(1)-Bundle With Connection: Construction I
Recalling that μ𝒪:𝒩×𝒪¯𝔤\mu_{\mathcal{O}}:\mathcal{N}\times\overline{\mathcal{O}}\to\mathfrak{g}^{*} we can consider the GG-equivariant U(1)\operatorname{U}(1)-bundle:

(𝒩×(G×χξ0U(1)))|μ𝒪1(0)μ𝒪1(0).\left(\mathcal{N}\times(G\times_{\chi_{\xi_{0}}}\operatorname{U}(1))\right)\big{|}_{\mu_{\mathcal{O}}^{-1}(0)}\to\mu_{\mathcal{O}}^{-1}(0).

If α0\alpha^{0} denotes the Liouville 1-form on 𝒩\mathcal{N} and i:μ𝒪1(0)𝒩i:\mu_{\mathcal{O}}^{-1}(0)\hookrightarrow\mathcal{N} the inclusion then we have a GG-invariant 1-form on the total space of this bundle given by:

i(α0α).i^{*}(\alpha^{0}-\alpha).

We then set:

Z𝒪:=(𝒩×(G×χξ0U(1)))|μ𝒪1(0)/Gμ𝒪1(0)/G=𝒩𝒪Z_{\mathcal{O}}:=\left(\mathcal{N}\times(G\times_{\chi_{\xi_{0}}}\operatorname{U}(1))\right)\big{|}_{\mu_{\mathcal{O}}^{-1}(0)}\big{/}G\to\mu_{\mathcal{O}}^{-1}(0)/G=\mathcal{N}_{\mathcal{O}}

with connection 1-form

α𝒪:= the reduction of i(α0α) mod G.\alpha_{\mathcal{O}}:=\mbox{ the reduction of }i^{*}(\alpha^{0}-\alpha)\mbox{ mod }G.
Definition 2.36.

The U(1)\operatorname{U}(1)-Bundle with Connection: Construction II
Here we instead extend our right Gξ0G_{\xi_{0}}-action on μ1(ξ0)\mu^{-1}(\xi_{0}) so μ1(ξ0)×U(1)\mu^{-1}(\xi_{0})\times\operatorname{U}(1) via the character χξ0\chi_{\xi_{0}}. We then set

Z𝒪:=(μ1(ξ0)×U(1))/Gξ0μ1(ξ0)/Gξ0=𝒩𝒪Z_{\mathcal{O}}:=\left(\mu^{-1}(\xi_{0})\times\operatorname{U}(1)\right)\big{/}G_{\xi_{0}}\to\mu^{-1}(\xi_{0})/G_{\xi_{0}}=\mathcal{N}_{\mathcal{O}}

with connection 1-form

α𝒪:= the reduction of iα0+dθ mod Gξ0\alpha_{\mathcal{O}}:=\mbox{ the reduction of }i^{*}\alpha^{0}+d\theta\mbox{ mod }G_{\xi_{0}}

where now i:μ1(ξ0)𝒩i:\mu^{-1}(\xi_{0})\hookrightarrow\mathcal{N} is the inclusion.

Finally we arrive at the holonomies that describe the quantum phase translation that occurs upon traveling along a classical periodic orbit.

Definition 2.37.

Let γ:[0,T]𝒩𝒪\gamma:[0,T]\to\mathcal{N}_{\mathcal{O}} be a periodic orbit of the Φ~sZ\widetilde{\Phi}^{Z}_{s}-flow (i.e. γ(s)=Φ~sZ(z0)\gamma(s)=\widetilde{\Phi}^{Z}_{s}(z_{0}) for some z0z_{0} and γ(0)=γ(T)\gamma(0)=\gamma(T)) and assume that TT is the minimum period of γ\gamma. We denote:

Hol𝒪(γ):= the holonomy of α𝒪 about the loop γ.\operatorname{Hol}_{\mathcal{O}}(\gamma):=\mbox{ the holonomy of }\alpha_{\mathcal{O}}\mbox{ about the loop }\gamma.

A key point is that while our construction of the U(1)\operatorname{U}(1)-bundle with connection relied on a choice of character as well as a choice of ξ0𝒪\xi_{0}\in\mathcal{O}, the element Hol𝒪(γ)U(1)\operatorname{Hol}_{\mathcal{O}}(\gamma)\in\operatorname{U}(1) is independent of these choices.

The following proposition is from [11] section 4. Their result applies here since it applies in the general context of symplectic reduction along an integral coadjoint orbit.

Proposition 2.38.

The map Hol𝒪:𝔜EU(1)\operatorname{Hol}_{\mathcal{O}}:\mathfrak{Y}_{E}\to\operatorname{U}(1) is locally constant. Furthermore, if we consider the symplectomorphism 𝒩𝒪μ1(ξ0)/Gξ0\mathcal{N}_{\mathcal{O}}\cong\mu^{-1}(\xi_{0})/G_{\xi_{0}} and suppose we had γμ1(ξ0)\gamma\in\mu^{-1}(\xi_{0}) with HZ(γ)=EH_{Z}(\gamma)=E and TT\in\mathbb{R}, gGξ0g\in G_{\xi_{0}} such that ΦTZ(γ)=γg\Phi^{Z}_{T}(\gamma)=\gamma\cdot g then if [γ]𝒩𝒪[\gamma]\in\mathcal{N}_{\mathcal{O}} denotes the image in the quotient we have:

Hol𝒪(T,[γ])=χξ0(g)eiTE.\operatorname{Hol}_{\mathcal{O}}(T,[\gamma])=\chi_{\xi_{0}}(g)e^{iTE}.

3 The Wave Equation on a Kaluza-Klein Spacetime

Similar to the classical phase space, the quantum-mechanical phase space is the space of solutions to the equations of motion. Usually, for quantum particles in a classical gauge field, one solves Schrödinger’s equations for sections of a vector bundle with connection. A choice of such a vector bundle corresponds to a choice of representation (usually irreducible) and hence a choice of fixed “charge”.

When performing semiclassical asymptotics, one doesn’t simply send 0\hbar\to 0 since \hbar is a dimensional quantity, but instead sends an observable such as S^/\widehat{S}/\hbar or J^/\widehat{J}/\hbar to infinity (here S^\widehat{S} and J^\widehat{J} are respectively action and angular momentum). We will work as in [12],[11] and send “charge” to infinity while holding the ratio of charge to energy fixed. Thus we need a quantum phase space which allows for varying representations of our structure group GG. The relativistic version of this is defined below.

Definition 3.1.

Fix a smooth function VC(M,)V\in C^{\infty}(M,\mathbb{R}) satisfying ZV=0\mathcal{L}_{Z}V=0 to act as a time-independent potential. We denote

ω:=dd+Vπ\Box_{\omega}:=d^{*}d+V\circ\pi

acting on C(P,)C^{\infty}(P,\mathbb{C}). We also define operators:

DZ:=1iZω and Dξ:=1iξ^.D_{Z}:=\frac{1}{i}\mathcal{L}_{Z^{\omega}}\ \mbox{ and }\ D_{\xi}:=\frac{1}{i}\mathcal{L}_{\widehat{\xi}}.
Lemma 3.2.

We have

[ω,DZ]=0=[ω,Dξ] for all ξ𝔤.[\Box_{\omega},D_{Z}]=0=[\Box_{\omega},D_{\xi}]\ \mbox{ for all }\xi\in\mathfrak{g}.
Proof.

Indeed, isometries and hence Lie derivatives along Killing vector fields commute with the wave operator ddd^{*}d so it suffices to shown that DZ(Vπ)=0=Dξ(Vπ)D_{Z}(V\circ\pi)=0=D_{\xi}(V\circ\pi) for all ξ𝔤\xi\in\mathfrak{g}. DZ(Vπ)=0D_{Z}(V\circ\pi)=0 since ZV=0\mathcal{L}_{Z}V=0 and ZωZ^{\omega} is the horizontal lift of ZZ to PP. Dξ(Vπ)=0D_{\xi}(V\circ\pi)=0 since VπV\circ\pi is constant on the fibers of PP and the vector fields ξ^\widehat{\xi} are vertical. ∎

It should be noted that these results are of interest even when V=0V=0. Nevertheless, we include the potential term in order to allow our results to apply to the conformal wave equation

ddϕ+CnSgωϕ=0d^{*}d\phi+C_{n}S_{g_{\omega}}\phi=0

for CnC_{n} a dimensional constant and SgωS_{g_{\omega}} the scalar curvature of gωg_{\omega}. The origin of this variant of the wave equation comes from considering conformal variations g~ω:=e2fgω\widetilde{g}_{\omega}:=e^{2f}g_{\omega} of the Hilbert-Einstein action. Indeed, setting ϕ:=e(n2)f/2\phi:=e^{(n-2)f/2} one can compute:

PSg~ω𝑑Vg~ω=Pϕ(ddϕ+CnSgωϕ)𝑑Vgω.\int_{P}S_{\widetilde{g}_{\omega}}dV_{\widetilde{g}_{\omega}}=\int_{P}\phi(d^{*}d\phi+C_{n}S_{g_{\omega}}\phi)dV_{g_{\omega}}.

It’s also worth noting that if SGS_{G} denotes the (constant) scalar curvature of the fibers of PP then from [3] Theorem 9.3.7 we have

Sg~ω=Sgπ+12|Fω|2π+SGS_{\widetilde{g}_{\omega}}=S_{g}\circ\pi+\frac{1}{2}|F_{\omega}|^{2}\circ\pi+S_{G}

and so the scalar curvature of (P,gω)(P,g_{\omega}) does indeed satisfy our assumptions on the potential.

Returning to our operator ω\Box_{\omega}, the vanishing of our commutators with DZD_{Z} and DξD_{\xi} tells us that the operators DZD_{Z} and DξD_{\xi} leave the kernel of ω\Box_{\omega} invariant. In fact, we want to complete the kernel of ω\Box_{\omega} to a Hilbert space of sorts since this is the quantum mechanical phase space. Indeed, the phase space in either classical or quantum mechanics is most naturally viewed as the space of solutions to the equations of motion (from a relativistic point of view). Any choice of Cauchy hypersurface then provides a natural identification of this phase space with a cotangent bundle; a more common non-relativistic description of phase space.

As is done in [22], we adapt several definitions and results from [2].

Definition 3.3.

For T>0T>0 we denote:

FE(P|t|T):=W1,2([T,T],L2(P0))L2([T,T],W1,2(P0))\operatorname{FE}(P_{|t|\leq T}):=W^{1,2}([-T,T],L^{2}(P_{0}))\cap L^{2}([-T,T],W^{1,2}(P_{0}))

and we will often interpret elements of FE(P|t|T)\operatorname{FE}(P_{|t|\leq T}) as \mathbb{C}-valued functions on P|t|T:=π1([T,T]×Σ0)P_{|t|\leq T}:=\pi^{-1}([-T,T]\times\Sigma_{0}). For the moment, let’s write:

kerT(ω):={ϕFE(P|t|T)|ωϕ=0}.\ker_{T}(\Box_{\omega}):=\{\phi\in\operatorname{FE}(P_{|t|\leq T})\ |\ \Box_{\omega}\phi=0\}.
Lemma 3.4.

[2] For 0<T1<T20<T_{1}<T_{2}, the natural restriction map

kerT2(ω)kerT1(ω)\ker_{T_{2}}(\Box_{\omega})\to\ker_{T_{1}}(\Box_{\omega})

is an isomorphism of locally convex spaces.

Definition 3.5.

We denote

kerω:={ϕLloc2(P):ϕ|P|t|TkerT(ω) for all T>0}.\ker\Box_{\omega}:=\{\phi\in L^{2}_{loc}(P)\ :\ \phi|_{P_{|t|\leq T}}\in\ker_{T}(\Box_{\omega})\mbox{ for all }T>0\}.
Lemma 3.6.

[2] We have ϕkerω\phi\in\ker\Box_{\omega} if and only if ϕLloc2(P)\phi\in L^{2}_{loc}(P) and there exists a T>0T>0 such that ϕ|P|t|TkerT(ω)\phi|_{P_{|t|\leq T}}\in\ker_{T}(\Box_{\omega}). Furthermore, the restriction map

kerωkerT(ω)\ker\Box_{\omega}\to\ker_{T}(\Box_{\omega})

is a vector space isomorphism for all T>0T>0 and the locally convex topology on kerω\ker\Box_{\omega} obtained by declaring this to be a homeomorphism is independent of our choice of T>0T>0.

Lemma 3.7.

[2] For each tt\in\mathbb{R} the map

CDt:kerω\displaystyle\operatorname{CD}_{t}:\ker\Box_{\omega} W1,2(Pt)L2(Pt)\displaystyle\to W^{1,2}(P_{t})\oplus L^{2}(P_{t})
ϕ\displaystyle\phi (ϕ|Pt,(n^ϕ)|Pt)\displaystyle\mapsto(\phi|_{P_{t}},\ (\mathcal{L}_{\widehat{n}}\phi)|_{P_{t}})

is an isomorphism of locally convex spaces (recall: ν\nu is the future-directed unit normal of the Cauchy hypersurfaces PtP_{t}). Thus kerω\ker\Box_{\omega} has the topology of a Hilbert space.

Our goal is to study the semiclassical asymptotics of the action of time translation on the space kerω\ker\Box_{\omega}.

Definition 3.8.

We denote by

eitDZ:kerωkerωe^{-itD_{Z}}:\ker\Box_{\omega}\to\ker\Box_{\omega}

the isomorphism given by precomposing functions ϕkerω\phi\in\ker\Box_{\omega} with the time t-t flow PPP\to P along the Killing vector field ZωZ^{\omega}.

At the moment eitDZe^{-itD_{Z}} is merely a notation since we do not have a preferred Hilbert space inner product with which to perform a functional calculus. Let’s now describe how we perform the quantum mechanical analogue of symplectic reduction.

Towards this end, we should notice that eitDZe^{-itD_{Z}} is not unitary on kerω\ker\Box_{\omega} with respect to any of the Hilbert space structures defined by a fixed Cauchy-data isomorphism since Zωn^Z^{\omega}\neq\widehat{n}. Instead, we proceed as in [22] and notice that the equation ωϕ=0\Box_{\omega}\phi=0 arises from a variational problem and therefore has an associated stress-energy tensor.

Definition 3.9.

Given ϕkerωC(P,)\phi\in\ker\Box_{\omega}\cap C^{\infty}(P,\mathbb{C}) we define the stress-energy tensor of ϕ\phi to be

T(ϕ):=dϕdϕ12(|dϕ|gω2+|ϕ|2V)gω.T(\phi):=d\phi\otimes d\phi-\frac{1}{2}\left(|d\phi|_{g_{\omega}}^{2}+|\phi|^{2}V\right)\cdot g_{\omega}.

The proof of the next few results are in [22] but we sketch them here since the computations will be useful to us later.

Lemma 3.10.

[22] For ϕkerωC(P,)\phi\in\ker\Box_{\omega}\cap C^{\infty}(P,\mathbb{C}), the stress-energy tensor T(ϕ)T(\phi) has divergence 12ϕ2dV-\frac{1}{2}\phi^{2}dV with respect to the metric gωg_{\omega}.

Proof.

We compute, using that the metric is divergence free to get:

div(dϕdϕ12(|dϕ|gω2+ϕ2V)gω)\displaystyle\operatorname{div}\left(d\phi\otimes d\phi-\frac{1}{2}\left(|d\phi|_{g_{\omega}}^{2}+\phi^{2}V\right)g_{\omega}\right) =2(ddϕ)dϕ\displaystyle=-2(d^{*}d\phi)d\phi
12(22ϕ,ϕgω+2ϕVϕ+ϕ2V)gω\displaystyle\ \ \ \ \ -\frac{1}{2}\left(2\langle\nabla^{2}\phi,\nabla\phi\rangle_{g_{\omega}}+2\phi V\nabla\phi+\phi^{2}\nabla V\right)\llcorner g_{\omega}
=(ωϕ)dϕ12ϕ2dV\displaystyle=-(\Box_{\omega}\phi)d\phi-\frac{1}{2}\phi^{2}dV

and since ωϕ=0\Box_{\omega}\phi=0 by assumption we’re done. ∎

As such, we can use the stress-energy tensor to define a quadratic form on kerωC(P,)\ker\Box_{\omega}\cap C^{\infty}(P,\mathbb{R}) and extend it to a Hermitian form on kerω\ker\Box_{\omega} via the polarization identity.

Definition 3.11.

For ϕkerωC(P,)\phi\in\ker\Box_{\omega}\cap C^{\infty}(P,\mathbb{R}) we denote

Qω(ϕ):=P0T(ϕ)(Zω,n^)𝑑VP0Q_{\omega}(\phi):=\int_{P_{0}}T(\phi)(Z^{\omega},\widehat{n})dV_{P_{0}}

and extend this quadratic form QωQ_{\omega} to a Hermitian one via the polarization identity.

Lemma 3.12.

QωQ_{\omega} is invariant under both the action of GG and eitDZe^{-itD_{Z}}.

Proof.

First we recall the proof of eitDZe^{-itD_{Z}}-invariance from [22]. Since ei(t1+t2)DZ=eit1DZeit2DZe^{-i(t_{1}+t_{2})D_{Z}}=e^{-it_{1}D_{Z}}e^{-it_{2}D_{Z}} it suffices to show that the time derivative of Q(eitDZϕ)Q(e^{-itD_{Z}}\phi) vanishes at t=0t=0. We do this for ϕ\phi real-valued. Writing \ast for the Hodge-\ast on PP (not on P0P_{0}!) we can compute using Cartan’s formula for the Lie derivative:

ddt|t=0Q(eitDZϕ)\displaystyle\frac{d}{dt}\Big{|}_{t=0}Q(e^{-itD_{Z}}\phi) =ddt|t=0P0T(ϕ)(Zω)\displaystyle=\frac{d}{dt}\Big{|}_{t=0}\int_{P_{0}}\ast T(\phi)(Z^{\omega})
=P0Zd(T(ϕ)(Z)) since the pullback of an exact form is exact\displaystyle=\int_{P_{0}}Z\llcorner d(\ast T(\phi)(Z))\ \mbox{ since the pullback of an exact form is exact}
=P0NdivP(T(ϕ)(Z))𝑑VP0\displaystyle=\int_{P_{0}}N\operatorname{div}_{P}(T(\phi)(Z))dV_{P_{0}}

where divP\operatorname{div}_{P} denotes the full divergence on PP. From the previous lemma we have divP(T(ϕ))(Z)=12ϕ2ZωV=0\operatorname{div}_{P}(T(\phi))(Z)=\frac{1}{2}\phi^{2}\mathcal{L}_{Z^{\omega}}V=0 and since Killing vector fields are divergence-free it follows that

divP(T(ϕ)(Zω))=(divP(T(ϕ)))(Zω)=0\operatorname{div}_{P}(T(\phi)(Z^{\omega}))=(\operatorname{div}_{P}(T(\phi)))(Z^{\omega})=0

as desired.

Next let’s look at the GG-action. We write ϕg\phi\cdot g for the function xϕ(xg1)x\mapsto\phi(xg^{-1}) and also continue to use the notation ζg\zeta\cdot g for the induced right action of GG on covectors ζ\zeta. Since GG acts by isometries we have |d(ϕg)|gω2=|dϕ|gω2g|d(\phi\cdot g)|_{g_{\omega}}^{2}=|d\phi|_{g_{\omega}}^{2}\cdot g and therefore

T(ϕg)(Zω,n^)=T(ϕ)(g1Zω,g1n^)g.T(\phi\cdot g)(Z^{\omega},\widehat{n})=T(\phi)(g^{-1}\cdot Z^{\omega},g^{-1}\cdot\widehat{n})\cdot g.

But Zω=tZ^{\omega}=\partial_{t} and n^=N1(tβ)\widehat{n}=N^{-1}(\partial_{t}-\beta) are both invariant under the GG-action so

T(ϕg)(Zω,n^)=T(ϕ)(Zω,n^)g.T(\phi\cdot g)(Z^{\omega},\widehat{n})=T(\phi)(Z^{\omega},\widehat{n})\cdot g.

Finally, the volume form dVP0dV_{P_{0}} is invariant under the GG-action since it is an action by isometries hence we can perform the change of variables xxg1x\mapsto xg^{-1} in the integral to get

P0T(ϕg)(Zω,n^)𝑑VP0=P0T(ϕ)(Zω,n^)g𝑑VP0=P0T(ϕ)(Zω,n^)𝑑VP0\int_{P_{0}}T(\phi\cdot g)(Z^{\omega},\widehat{n})dV_{P_{0}}=\int_{P_{0}}T(\phi)(Z^{\omega},\widehat{n})\cdot gdV_{P_{0}}=\int_{P_{0}}T(\phi)(Z^{\omega},\widehat{n})dV_{P_{0}}

as desired. ∎

Unfortunately: since we have allowed possibly negative potentials VV our quadratic form QωQ_{\omega} need not be positive definite. Just as in [22], we apply several results on the general theory of Pontryagin and Krein spaces [16],[6]. These are “Hilbert spaces” for which the inner product is permitted to have finite dimensional negative-definite and/or degenerate subspaces. As we will see below in 3.19, we only care about the operators DZ,DξD_{Z},D_{\xi}, etc. on certain closed subspaces m\mathcal{H}_{m} of kerω\ker\Box_{\omega} and QωQ_{\omega} will be positive definite on these subspaces.

Lemma 3.13.

[22]
kerQωkerω\ker Q_{\omega}\subseteq\ker\Box_{\omega} is finite dimensional and consists of CC^{\infty} functions. Furthermore, for all ϕkerQω\phi\in\ker Q_{\omega} we have:

DZϕ=0.D_{Z}\phi=0.

In particular, if Q~ω\widetilde{Q}_{\omega} is the Hermitian form on kerω/kerQω\ker\Box_{\omega}/\ker Q_{\omega} induced by QωQ_{\omega} then (kerω/kerQω)(\ker\Box_{\omega}/\ker Q_{\omega}) is a Pontryagin space.

Lemma 3.14.

DZD_{Z} descends to a Krein-self-adjoint operator on kerω/kerQω\ker\Box_{\omega}/\ker Q_{\omega} whose domain contains the dense GG-invariant subspace given by the image of kerωC(P)\ker\Box_{\omega}\cap C^{\infty}(P) in the quotient. Furthermore, the spectrum of DZD_{Z} on this Krein space is discrete consisting of eigenvalues of finite multiplicity, invariant under λλ¯\lambda\mapsto\overline{\lambda} and λλ\lambda\mapsto-\lambda, accumulates at ±\pm\infty only, and has only finitely many non-real eigenvalues.

Proof.

The only part of this not proven in [22] was the GG-invariance of the subspace kerωC(P)\ker\Box_{\omega}\cap C^{\infty}(P). However, the GG-action preserves C(P)C^{\infty}(P) since it is smooth and therefore preserves kerωC(P)\ker\Box_{\omega}\cap C^{\infty}(P) by 3.2. ∎

Lemma 3.15.

[22],[16],[6] Let Q~ω\widetilde{Q}_{\omega} denote the induced quadratic form on the quotient kerω/kerQω\ker\Box_{\omega}/\ker Q_{\omega}. Then there exists a maximal negative definite subspace

V~(kerω/kerQω,Q~ω)\widetilde{V}^{-}\subseteq(\ker\Box_{\omega}/\ker Q_{\omega},\ \widetilde{Q}_{\omega})

which is invariant under DZD_{Z} and eitDZe^{-itD_{Z}}. Furthermore, it is finite-dimensional with dimension an invariant of the Krein space and DZD_{Z} themselves. Finally, V~\widetilde{V}^{-} is invariant under the action of GG.

Proof.

The only part of this not proven in the above-cited papers is the GG-invariance. Indeed, suppose for contradiction that there was some gGg\in G and vV~v\in\widetilde{V}^{-} with gvV~g\cdot v\notin\widetilde{V}^{-}. Consider the subspace W~:=gV~\widetilde{W}^{-}:=g\cdot\widetilde{V}^{-}. Then, as gvV~g\cdot v\notin\widetilde{V}^{-} we have that the subspace W~+V~\widetilde{W}^{-}+\widetilde{V}^{-} properly contains V~\widetilde{V}^{-}. Furthermore, it is invariant under both DZD_{Z} and eitDZe^{-itD_{Z}} since DZD_{Z} commutes with the GG-action. Finally, Q~ω\widetilde{Q}_{\omega} is negative-definite on W~\widetilde{W}^{-} since it is negative-definite on V~\widetilde{V}^{-} and invariant under the GG-action, hence Q~ω\widetilde{Q}_{\omega} is negative definite on W~+V~\widetilde{W}^{-}+\widetilde{V}^{-}, contradicting maximality. ∎

Since Q~ω\widetilde{Q}_{\omega} is non-degenerate and invariant under both the GG-action and eitDZe^{-itD_{Z}} we obtain the following immediate corollary.

Corollary 3.16.

The subspace

V~+:=(V~)Q~ω\widetilde{V}^{+}:=(\widetilde{V}^{-})^{\perp\widetilde{Q}_{\omega}}

is a Hilbert space with inner product Q~ω\widetilde{Q}_{\omega}, and is equipped with a unitary representation of ×G\mathbb{R}\times G given by the restriction of eitDZe^{-itD_{Z}} and the GG-action from above.

We can now begin the process of showing that QωQ_{\omega} is positive definite on isotypic subspaces for irreducible representations with sufficiently large dominant integral weights.

Lemma 3.17.

Let VV^{-} be the preimage of V~\widetilde{V}^{-} in kerω\ker\Box_{\omega} under the quotient map kerωkerω/kerQω\ker\Box_{\omega}\to\ker\Box_{\omega}/\ker Q_{\omega}. Then VV^{-} is finite dimensional and contains kerQω\ker Q_{\omega}.

Proof.

Indeed the quotient map restricts to a map VV~V^{-}\to\widetilde{V}^{-} with kernel kerQω\ker Q_{\omega}. Choosing a splitting of this linear surjection gives us an isomorphism of vector spaces VV~kerQωV^{-}\cong\widetilde{V}^{-}\oplus\ker Q_{\omega} and since V~kerQω\widetilde{V}^{-}\oplus\ker Q_{\omega} so is VV^{-}. ∎

Definition 3.18.

For 𝒪\mathcal{O} our integral coadjoint orbit and m1m\in\mathbb{Z}_{\geq 1} we let κm\kappa_{m} denote the irreducible representation corresponding to the integral coadjoint orbit m𝒪𝔤m\mathcal{O}\subseteq\mathfrak{g}^{*}.

Proposition 3.19.

There exists an m01m_{0}\in\mathbb{Z}_{\geq 1} depending only on 𝒪\mathcal{O}, DZD_{Z} and the Krein space (kerω/kerQω,Q~ω)(\ker\Box_{\omega}/\ker Q_{\omega},\widetilde{Q}_{\omega}) such that for any mm0m\geq m_{0} and any ϕkerω\phi\in\ker\Box_{\omega} which generates a cyclic GG-representation VϕkerωV_{\phi}\subseteq\ker\Box_{\omega} isomorphic to κm\kappa_{m} we have

VϕV={0}.V_{\phi}\cap V^{-}=\{0\}.

Thus for each mm0m\geq m_{0} we have a closed subspace

m:=Span{ϕkerω:Vϕκm}¯\mathcal{H}_{m}:=\overline{\operatorname{Span}_{\mathbb{C}}\{\phi\in\ker\Box_{\omega}\ :\ V_{\phi}\cong\kappa_{m}\}}

on which QωQ_{\omega} restricts to a positive definite Hilbert space inner product. Furthermore, our representation of ×G\mathbb{R}\times G arising as the product of the GG-action and eitDZe^{-itD_{Z}} leaves m\mathcal{H}_{m} invariant and is unitary.

Proof.

Let ϕkerω\phi\ker\Box_{\omega} generate a cyclic GG-representation VϕV_{\phi} isomorphic to κm\kappa_{m}. Suppose that VϕV{0}V_{\phi}\cap V^{-}\neq\{0\} and so there existed a non-zero ψVϕV\psi\in V_{\phi}\cap V^{-}. Since VV^{-} is a GG-invariant subspace we have VψVV_{\psi}\subseteq V^{-} where VψV_{\psi} is the cyclic GG-representation generated by ψ\psi. Furthermore, 0VψVϕ0\neq V_{\psi}\subseteq V_{\phi} and since VϕV_{\phi} is irreducible it follows that Vψ=VϕV_{\psi}=V_{\phi}. So it follows that:

if Vϕκm and VϕV{0} then VϕV.\mbox{if }V_{\phi}\cong\kappa_{m}\ \mbox{ and }\ V_{\phi}\cap V^{-}\neq\{0\}\ \mbox{ then }\ V_{\phi}\subseteq V^{-}.

Since VV^{-} is finite dimensional this can happen for at most finitely many irreducible cyclic invariant subspaces and hence for at most finitely many mm. In fact, since the dimension of VV^{-} is an invariant of DZD_{Z} and the Krein space kerω/kerQω\ker\Box_{\omega}/\ker Q_{\omega} it follows that for m0m_{0} large enough (with dependence as in the statement of the proposition) and all mm0m\geq m_{0} we have:

if ϕkerω with Vϕκm then VϕV={0}.\mbox{if }\phi\in\ker\Box_{\omega}\mbox{ with }V_{\phi}\cong\kappa_{m}\mbox{ then }V_{\phi}\cap V^{-}=\{0\}.

In particular, for mm0m\geq m_{0} and m\mathcal{H}_{m} defined as in the statement of the proposition, QωQ_{\omega} is positive definite on m\mathcal{H}_{m}.

To show that our ×G\mathbb{R}\times G action leaves m\mathcal{H}_{m} invariant and is unitary it suffices to show that it leaves Span{ϕkerω:Vϕκm}\operatorname{Span}_{\mathbb{C}}\{\phi\in\ker\Box_{\omega}\ :\ V_{\phi}\cong\kappa_{m}\} invariant and is unitary here, since it will then extend to m\mathcal{H}_{m} by uniform continuity. Since QωQ_{\omega} is invariant under the full ×G\mathbb{R}\times G-action, unitarity is immediate. All that remains is to check invariance. However, since κm\kappa_{m} is irreducible it follows that for any ϕ\phi with VϕκmV_{\phi}\cong\kappa_{m} and any gGg\in G we have 0VϕgVϕ0\neq V_{\phi\cdot g}\subseteq V_{\phi} hence Vϕg=VϕV_{\phi\cdot g}=V_{\phi} thus we have invariance, as desired. ∎

It is worth noting that, as remarked in [22], if V0V\geq 0 and there exists some xΣ0x\in\Sigma_{0} for which V(x)>0V(x)>0 then QωQ_{\omega} is positive definite. This is especially true for the massive Klein-Gordon equation where VV is a positive constant. In [23] the special case of our results where G=U(1)G=\operatorname{U}(1) and (P,ω)(P,\omega) were trivial was considered. In this case it was shown that when projected down to MM our parameter m1m\in\mathbb{Z}_{\geq 1} above actually corresponds to mass. We will demonstrate an analogue of this later in 3.1.

Another important remark is that not every ϕm\phi\in\mathcal{H}_{m} has VϕκmV_{\phi}\cong\kappa_{m}. This is most easily seen in the Euclidean-signature case where MM is a single point. Then P=GP=G and our Hilbert space is L2(G)L^{2}(G) which, by the Peter-Weyl theorem, contains every irreducible representation of GG as a cyclic subspace. However, as was shown in [8], since GG is compact Hausdorff and second-countable, the entire representation L2(G)L^{2}(G) is itself a cyclic representation.

Combining our previous facts, for mm0m\geq m_{0} we can decompose:

m=L2m,\mathcal{H}_{m}=\bigoplus_{\ell\in\mathbb{Z}}^{L^{2}}\mathcal{H}_{m,\ell}

with m,\mathcal{H}_{m,\ell} the λm,\lambda_{m,\ell}-eigenspace for DZD_{Z} on m\mathcal{H}_{m}, organized so that λm,λm,+1\lambda_{m,\ell}\leq\lambda_{m,\ell+1} for all \ell\in\mathbb{Z}. If λm,=λm,+1\lambda_{m,\ell}=\lambda_{m,\ell+1} then m,=m,+1\mathcal{H}_{m,\ell}=\mathcal{H}_{m,\ell+1} and otherwise these spaces are orthogonal (this is the sense in which the above is indeed an L2L^{2}-direct sum). We can then further decompose:

m,=j=1μ(m,)κm\mathcal{H}_{m,\ell}=\bigoplus_{j=1}^{\mu(m,\ell)}\kappa_{m}

and it is worth noticing that μ(m,)\mu(m,\ell) is indeed always finite since m,\mathcal{H}_{m,\ell} itself is finite dimensional (being an eigenspace for DZD_{Z}).

Since we will be studying asymptotics as mm\to\infty, there’s no harm in replacing κ\kappa with κm0\kappa_{m_{0}} so that we may assume m0=1m_{0}=1. As such, we want to study the time evolution of quantum states in the subspace

:=m1L2mkerω.\mathcal{H}:=\bigoplus^{L^{2}}_{m\geq 1}\mathcal{H}_{m}\subseteq\ker\Box_{\omega}.

However, we still haven’t fully specified a direction in which to take our large quantum numbers limit. Indeed, for fixed mm the eigenvalues λm,\lambda_{m,\ell} very well might accumulate at ±\pm\infty as \ell tends to ±\pm\infty. Thus for each EE\in\mathbb{R} we could consider eigenvalues satisfying

λm,mE\lambda_{m,\ell}\sim mE

and different choices of EE might very well yield different mm\to\infty asymptotics. Classically this is reflected in the fact that symplectic reduction along 𝒪\mathcal{O} generally leads to phase spaces which are not conical. As such, our problem is broken into two steps:

  1. 1.

    For mm fixed, “count” eigenvalues satisfying λm,mE\lambda_{m,\ell}\sim mE.

  2. 2.

    Understand the asymptotics of the above count as mm\to\infty.

The first step is fairly straight-forward. It is highly unlikely for us to have any eigenvalues satisfying λm,=mE\lambda_{m,\ell}=mE exactly and so we instead sum over all \ell\in\mathbb{Z}, weighting eigenvalues near mEmE the most. By stationary phase, this is described for large frequencies by the distribution:

φTr(φ(t)eit(DZmE)|mdt)=φ^(λm,mE)=:μ(E,m,φ).\varphi\mapsto\operatorname{Tr}\left(\int_{-\infty}^{\infty}\varphi(t)e^{-it(D_{Z}-mE)}|_{\mathcal{H}_{m}}dt\right)=\sum_{\ell\in\mathbb{Z}}\widehat{\varphi}(\lambda_{m,\ell}-mE)=:\mu(E,m,\varphi).

We use the letter μ\mu to denote this distribution since it can be viewed as a multiplicity for the representation on \mathcal{H} of ×G\mathbb{R}\times G associated to the coadjoint orbit {E}×𝒪𝔤\{E\}\times\mathcal{O}\subseteq\mathbb{R}\oplus\mathfrak{g}^{*}. The point is that (modulo factors of 2π2\pi), φ^\widehat{\varphi} approaches δ0\delta_{0} as φ1\varphi\to 1 and so in this limit the right hand side approaches the literal multiplicity of mEmE as an eigenvalue on m\mathcal{H}_{m}. However, this is only a moral since the above limit does not converge. Instead we first notice that μ(E,m,)\mu(E,m,-) defines a linear functional on the collection of all φ𝒮()\varphi\in\mathcal{S}(\mathbb{R}) with compactly supported Fourier transform. Our goal now is to apply a result of [14] which generalizes the Weyl law of [22] to vector bundles in order to prove that μ(E,m,)\mu(E,m,-) is actually tempered and hence μ(E,m,φ)\mu(E,m,\varphi) is defined for any φ𝒮()\varphi\in\mathcal{S}(\mathbb{R}).

3.1 Relation to Vector Bundles

We begin by recalling the well-known fact that for any unitary representation VV of GG there is an isomorphism

C(P,V)GΓ(M,P×GV)C^{\infty}(P,V)^{G}\cong\Gamma(M,P\times_{G}V)

between VV-valued GG-equivariant smooth functions on PP and smooth sections of the associated vector bundle P×GVP\times_{G}V over MM. Furthermore, the Hermitian inner product on VV defines a Hermitian fiber metric on P×GVP\times_{G}V. We will need a less well-known, but related construction.

Definition 3.20.

We fix an mm0m\geq m_{0} so that QωQ_{\omega} is positive definite on mkerω\mathcal{H}_{m}\subseteq\ker\Box_{\omega} and denote by κm:GU(Vm)\kappa_{m}:G\to\operatorname{U}(V_{m}) our irreducible representation corresponding to m𝒪m\mathcal{O}. We also let dm:=dimVmd_{m}:=\dim_{\mathbb{C}}V_{m} and fix an orthonormal basis e1,,edm\vec{e}_{1},...,\vec{e}_{d_{m}} for VmV_{m}, writing ,m\langle-,-\rangle_{m} for our Hermitian inner product on VmV_{m}.

Lemma 3.21.

Let ψC(P,Vm)G\vec{\psi}\in C^{\infty}(P,V_{m})^{G} and vVm\vec{v}\in V_{m} both be non-zero. Define a function

ϕ:P\displaystyle\phi:P \displaystyle\to\mathbb{C}
ϕ(p)\displaystyle\phi(p) :=ψ(p),vm.\displaystyle:=\langle\vec{\psi}(p),\vec{v}\rangle_{m}.

Then VϕVmV_{\phi}\cong V_{m} as GG-representations. Furthermore, if ω\Box_{\omega} is extended to act on VmV_{m}-valued smooth functions it follows that ωψ=0\Box_{\omega}\vec{\psi}=0 if and only if ωϕ=0\Box_{\omega}\phi=0.

Proof.

Since vVm\vec{v}\in V_{m} is non-zero and VmV_{m} is irreducible, it is a cyclic vector and so for each j=1,,dmj=1,...,d_{m} there are finitely many group elements gjiGg_{j}^{i}\in G such that igjiv=ej\sum_{i}g_{j}^{i}\vec{v}=\vec{e}_{j}. Thus

iϕ(p(gji)1)=iψ(p),gjivm=ψ(p),ejm.\sum_{i}\phi(p(g_{j}^{i})^{-1})=\sum_{i}\langle\vec{\psi}(p),g_{j}^{i}\vec{v}\rangle_{m}=\langle\psi(p),\vec{e}_{j}\rangle_{m}.

So the functions ψ(),ejm\langle\vec{\psi}(-),\vec{e_{j}}\rangle_{m} are in VϕV_{\phi} for all j=1,,dmj=1,...,d_{m}. Furthermore every function pψ(pg1)=ψ(p),gvmp\mapsto\psi(pg^{-1})=\langle\vec{\psi}(p),g\vec{v}\rangle_{m} is in the span of the functions ψ(),ejm\langle\vec{\psi}(-),\vec{e}_{j}\rangle_{m} hence

Vϕ=Span{ψ(),e1m,,ψ(),edmm}.V_{\phi}=\operatorname{Span}_{\mathbb{C}}\left\{\langle\vec{\psi}(-),\vec{e}_{1}\rangle_{m},...,\langle\vec{\psi}(-),\vec{e}_{d_{m}}\rangle_{m}\right\}.

The set of functions ψ(),ejm\langle\vec{\psi}(-),\vec{e}_{j}\rangle_{m} are linearly independent since if aja^{j}\in\mathbb{C} are such that ψ(p),ajejm=0\langle\vec{\psi}(p),a^{j}\vec{e}_{j}\rangle_{m}=0 for all pPp\in P then since ψ0\vec{\psi}\neq 0 there exists a pPp\in P with 0ψ(p)Vm0\neq\vec{\psi}(p)\in V_{m}. Since VmV_{m} is irreducible there exists elements gkGg_{k}\in G such that kgψ(p)=ajej\sum_{k}g\vec{\psi}(p)=a^{j}\vec{e}_{j} and so

0=kψ(pgk1),ajek=j|aj|20=\sum_{k}\langle\vec{\psi}(pg_{k}^{-1}),a^{j}\vec{e}_{k}\rangle=\sum_{j}|a^{j}|^{2}

hence aj=0a^{j}=0 for all jj as desired. Therefore the map

ejψ(),ejm\vec{e}_{j}\leftrightarrow\langle\vec{\psi}(-),\vec{e}_{j}\rangle_{m}

induces an isomorphism of GG-representations VmVϕV_{m}\cong V_{\phi}.

If ωψ=0\Box_{\omega}\vec{\psi}=0 then by definition (v\vec{v} and ,m\langle-,-\rangle_{m} are constant) ωϕ=0\Box_{\omega}\phi=0. Conversely, GG-invariance of ω\Box_{\omega} implies that if ωϕ=0\Box_{\omega}\phi=0 then ωf=0\Box_{\omega}f=0 for all fVϕf\in V_{\phi} and hence ωψ(),ejm=0\Box_{\omega}\langle\vec{\psi}(-),\vec{e}_{j}\rangle_{m}=0 for all jj. Therefore ωψ=0\Box_{\omega}\vec{\psi}=0 as desired. ∎

Usually one doesn’t look at the full wave operator ω\Box_{\omega} applied to ψC(P,V)G\vec{\psi}\in C^{\infty}(P,V)^{G} but only at the “horizontal” wave operator. To relate these two wave operators, we fix a root system for 𝔤\mathfrak{g} compatible with our Ad\operatorname{Ad}-invariant inner product and let:

ρ:= the sum of all positive roots\rho:=\ \mbox{ the sum of all positive roots}

and

Λ0:= the dominant integral weight for κm0.\Lambda_{0}:=\ \mbox{ the dominant integral weight for }\kappa_{m_{0}}.
Lemma 3.22.

The wave operator ω\Box_{\omega} on C(P)C^{\infty}(P) splits as a sum of vertical and horizontal parts:

ω=HΔG\Box_{\omega}=\Box_{H}-\Delta_{G}

where H\Box_{H} is the horizontal wave operator (plus the potential) and ΔG\Delta_{G} is the Laplacian on the fibers. These operators commute and if ϕm\phi\in\mathcal{H}_{m} has VϕVmV_{\phi}\cong V_{m} then ΔG\Delta_{G} acts on VmV_{m} as multiplication by a constant. Hence ΔG\Delta_{G} acts by multiplication by a constant on all of m\mathcal{H}_{m} and this constant is given by:

ΔG|m=mΛ0,mΛ0+ρ.\Delta_{G}|_{\mathcal{H}_{m}}=\langle m\Lambda_{0},\ m\Lambda_{0}+\rho\rangle.
Proof.

The existence of the splitting and the fact that [H,ΔG]=0[\Box_{H},\Delta_{G}]=0 follows from [11] section 6. Since Δω\Delta_{\omega} and ΔH\Delta_{H} both commute with the GG-action it follows that ΔG\Delta_{G} does as well hence ΔG\Delta_{G} does indeed preserve VϕV_{\phi}. In fact, by the explicit form of ΔG\Delta_{G} we see that its action on VϕV_{\phi} is precisely the action of the quadratic Casimir and hence is given by multiplication by mΛ0,mΛ0+ρ\langle m\Lambda_{0},\ m\Lambda_{0}+\rho\rangle. ∎

In fact, we see that ΔG\Delta_{G} preserves our space

=mm0L2m\mathcal{H}=\bigoplus_{m\geq m_{0}}^{L^{2}}\mathcal{H}_{m}

and on this space m\mathcal{H}_{m} is precisely the mΛ0,mΛ0+ρ\langle m\Lambda_{0},\ m\Lambda_{0}+\rho\rangle-eigenspace of ΔG\Delta_{G}.

Definition 3.23.

We denote by m\Box_{m} the operator

m:=HmΛ0,mΛ0+ρ.\Box_{m}:=\Box_{H}-\langle m\Lambda_{0},\ m\Lambda_{0}+\rho\rangle.
Lemma 3.24.

Denote by

Grm(P):={VkerωC(P):V is G-invariant and VVm}\operatorname{Gr}_{m}(P):=\{V\subseteq\ker\Box_{\omega}\cap C^{\infty}(P)\ :\ V\mbox{ is }G\mbox{-invariant and }V\cong V_{m}\}

the collection of all invariant subspaces of kerω\ker\Box_{\omega} which are isomorphic to VmV_{m} as GG-representations. Then for each VGrm(P)V\in\operatorname{Gr}_{m}(P) we have VmV\subseteq\mathcal{H}_{m}. Furthermore if Φ:VmV\Phi:V_{m}\to V is any isomorphism of GG-representations then

ψ(p):=j=1dmΦ(ej)(p)ej\vec{\psi}(p):=\sum_{j=1}^{d_{m}}\Phi(\vec{e}_{j})(p)\vec{e}_{j}

is a GG-equivariant VmV_{m}-valued function with

mψ=HψmΛ0,mΛ0+ρψ=0.\Box_{m}\vec{\psi}=\Box_{H}\vec{\psi}-\langle m\Lambda_{0},\ m\Lambda_{0}+\rho\rangle\vec{\psi}=0. (3)

Finally, the definition of ψ\vec{\psi} is independent of our choice of orthonormal basis ej\vec{e}_{j}.

Proof.

Since each Φ(ej)\Phi(\vec{e}_{j}) generates a cyclic representation isomorphic to VmV_{m} it automatically follows that VmV\subseteq\mathcal{H}_{m} and ψ\vec{\psi} satisfies 3. So all that remains to be checked is ψ\vec{\psi}’s equivariance and basis-independence. However since Φ\Phi is an isomorphism of GG-representations we can compute:

ψ(pg1)=j=1dmΦ(ej)(pg1)ej=j=1dmΦ(gej)(p)ej.\vec{\psi}(pg^{-1})=\sum_{j=1}^{d_{m}}\Phi(\vec{e}_{j})(pg^{-1})\vec{e}_{j}=\sum_{j=1}^{d_{m}}\Phi(g\vec{e}_{j})(p)\vec{e}_{j}.

But if we write gej=gjieig\vec{e}_{j}=g^{i}_{j}\vec{e}_{i} then we arrive at:

ψ(pg1)=j=1dmgjiΦ(ei)(p)ej=i=1dmΦ(ei)gei\vec{\psi}(pg^{-1})=\sum_{j=1}^{d_{m}}g^{i}_{j}\Phi(\vec{e}_{i})(p)\vec{e}_{j}=\sum_{i=1}^{d_{m}}\Phi(\vec{e}_{i})\ g\vec{e}_{i}

proving equivariance. Similarly, if fjVm\vec{f}_{j}\in V_{m} is another orthonormal basis then there exists a unitary matrix AA satisfying ej=Ajifi\vec{e}_{j}=A_{j}^{i}\vec{f}_{i} hence

ψ(p)=j=1dmAjiAjkΦ(ei)(p)ek=i=1dmΦ(ei)(p)ei\vec{\psi}(p)=\sum_{j=1}^{d_{m}}A_{j}^{i}A_{j}^{k}\Phi(\vec{e}_{i})(p)\vec{e}_{k}=\sum_{i=1}^{d_{m}}\Phi(\vec{e}_{i})(p)\vec{e}_{i}

as desired. ∎

Since VmV_{m} is irreducible, Schur’s lemma tells us that any two isomorphisms VmVV_{m}\cong V of GG-representations differ by a multiplicative non-zero constant complex number. As such, we obtain the following corollary.

Corollary 3.25.

There is a natural isomorphism

Grm(P)\displaystyle\operatorname{Gr}_{m}(P) {ψC(P,Vm)G:mψ=0}/×\displaystyle\to\{\vec{\psi}\in C^{\infty}(P,V_{m})^{G}\ :\ \Box_{m}\vec{\psi}=0\}/\mathbb{C}^{\times}
V\displaystyle V j=1dmΦ(ej)()ejmod ×\displaystyle\mapsto\sum_{j=1}^{d_{m}}\Phi(\vec{e}_{j})(-)\vec{e}_{j}\ \mbox{mod }\mathbb{C}^{\times}

where in the above expression ej\vec{e}_{j} is any choice of orthonormal basis for VmV_{m} and Φ\Phi is any choice of isomorphism of GG-representations VmVV_{m}\cong V.

Proof.

This is simply a combination of 3.21 and 3.24, taking care to remark that the two constructions from these two lemmas are inverse to one-another (taking v=e1\vec{v}=\vec{e}_{1} in 3.21). ∎

Our final step is to compare elements of C(P,Vm)GC^{\infty}(P,V_{m})^{G} with sections of the associated vector bundle.

Definition 3.26.

We define a map Ψ:C(P,Vm)GΓ(M,P×GVm)\Psi:C^{\infty}(P,V_{m})^{G}\to\Gamma(M,P\times_{G}V_{m}) as follows. Given ψC(P,Vm)G\vec{\psi}\in C^{\infty}(P,V_{m})^{G} and xMx\in M we choose an arbitrary pPp\in P in the fiber over xx and define

Ψ(ψ)(x):= the equivalence class of (p,ψ(p)) in the fiber (P×GVm)x.\Psi(\vec{\psi})(x):=\mbox{ the equivalence class of }(p,\vec{\psi}(p))\mbox{ in the fiber }(P\times_{G}V_{m})_{x}.

We recall from [3] Chapter 3, for example, that Ψ\Psi is an isomorphism. Furthermore there is an induced covariant derivative m\nabla^{m} on P×GVmP\times_{G}V_{m} which corresponds under Ψ\Psi to the horizontal exterior derivative on PP with respect to ω\omega, and there is a Hermitian fiber metric ,m\langle-,-\rangle_{m} on P×GVP\times_{G}V corresponding to the constant Hermitian inner product ,m\langle-,-\rangle_{m} on VmV_{m}.

Now, let’s let VGrm(P)V\in\operatorname{Gr}_{m}(P) and choose an isomorphism Φ:VmV\Phi:V_{m}\to V which is unitary where VV is given the QmQ_{m}-inner product. Writing

ψ(p):=j=1dmΦ(ej)(p)ej\vec{\psi}(p):=\sum_{j=1}^{d_{m}}\Phi(\vec{e}_{j})(p)\vec{e}_{j}

it follows that the expression

Qω(ψ):=j=1dmQω(Φ(ej))Q_{\omega}(\vec{\psi}):=\sum_{j=1}^{d_{m}}Q_{\omega}(\Phi(\vec{e}_{j}))

is independent of our choice of orthonormal basis ej\vec{e}_{j} or unitary isomorphism Φ\Phi. We also have the following explicit formula from [22] where we use Greek μ,ν,\mu,\nu,... for indices of coordinates tangent to Σ0M\Sigma_{0}\subseteq M and Roman a,b,a,b,... indices for coordinates tangent to the fibers of P0P_{0}:

Q(Φ(ej))\displaystyle Q(\Phi(\vec{e}_{j})) =P0N1(|tΦ(ej)|2+(N2hμνβμβν)(μΦ(ej))(νΦ(ej)¯)\displaystyle=\int_{P_{0}}N^{-1}\Big{(}|\partial_{t}\Phi(\vec{e}_{j})|^{2}+(N^{2}h^{\mu\nu}-\beta^{\mu}\beta^{\nu})(\partial_{\mu}\Phi(\vec{e}_{j}))(\partial_{\nu}\overline{\Phi(\vec{e}_{j})})
+Tr(ω(dΦ(ej))ω(dΦ(ej)¯)T)+|Φ(ej)|2V)dVP0\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ +\operatorname{Tr}\left(\omega(d\Phi(\vec{e}_{j}))\omega(d\overline{\Phi(\vec{e}_{j})})^{T}\right)+|\Phi(\vec{e}_{j})|^{2}V\Big{)}dV_{P_{0}}

By equivariance it follows that if ξ1,,ξd\xi_{1},...,\xi_{d} is an orthonormal basis for 𝔤\mathfrak{g} then

ω(dΦ(ej))=a(ξ^aΦ(ej))ξ^a=aΦ(ξaej)ξ^a\omega(d\Phi(\vec{e}_{j}))=\sum_{a}(\mathcal{L}_{\widehat{\xi}_{a}}\Phi(\vec{e}_{j}))\widehat{\xi}_{a}=\sum_{a}\Phi(\xi_{a}\cdot\vec{e}_{j})\widehat{\xi}_{a}

and so

Tr(ω(dΦ(ej))ω(dΦ(ej)¯)T)=a|Φ(ξaej)|2=mΛ0,mΛ0+ρ|Φ(ej)|2.\operatorname{Tr}\left(\omega(d\Phi(\vec{e}_{j}))\omega(d\overline{\Phi(\vec{e}_{j})})^{T}\right)=\sum_{a}|\Phi(\xi_{a}\cdot\vec{e}_{j})|^{2}=\langle m\Lambda_{0},\ m\Lambda_{0}+\rho\rangle|\Phi(\vec{e}_{j})|^{2}.

Thus we obtain

Q(Φ(ej))\displaystyle Q(\Phi(\vec{e}_{j})) =P0N1(|tΦ(ej)|2+(N2hμνβμβν)(μΦ(ej))(νΦ(ej)¯)\displaystyle=\int_{P_{0}}N^{-1}\Big{(}|\partial_{t}\Phi(\vec{e}_{j})|^{2}+(N^{2}h^{\mu\nu}-\beta^{\mu}\beta^{\nu})(\partial_{\mu}\Phi(\vec{e}_{j}))(\partial_{\nu}\overline{\Phi(\vec{e}_{j})})
+|Φ(ej)|2(V+mΛ0,mΛ0+ρ))dVP0\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ +|\Phi(\vec{e}_{j})|^{2}\left(V+\langle m\Lambda_{0},\ m\Lambda_{0}+\rho\rangle\right)\Big{)}dV_{P_{0}}

Furthermore, from this explicit expression we see that the sum

j=1dmN1(|tΦ(ej)|2\displaystyle\sum_{j=1}^{d_{m}}N^{-1}\Big{(}|\partial_{t}\Phi(\vec{e}_{j})|^{2} +(N2hμνβμβν)(μΦ(ej))(νΦ(ej)¯)\displaystyle+(N^{2}h^{\mu\nu}-\beta^{\mu}\beta^{\nu})(\partial_{\mu}\Phi(\vec{e}_{j}))(\partial_{\nu}\overline{\Phi(\vec{e}_{j})}) (4)
+|Φ(ej)|2(V+mΛ0,mΛ0+ρ))\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ +|\Phi(\vec{e}_{j})|^{2}(V+\langle m\Lambda_{0},\ m\Lambda_{0}+\rho\rangle)\Big{)}

is invariant under the GG action.

Definition 3.27.

Given a section sΓ(M,P×GVm)s\in\Gamma(M,P\times_{G}V_{m}) we define the bundle stress-energy tensor Tm(s)T_{m}(s) to be the symmetric 2-tensor on MM given by:

Tm(s)ij:=ims,jms12(|ms|2+|s|2(V+mΛ0,mΛ0+ρ))gijT_{m}(s)_{ij}:=\langle\nabla_{i}^{m}s,\nabla_{j}^{m}s\rangle-\frac{1}{2}\left(|\nabla^{m}s|^{2}+|s|^{2}\left(V+\langle m\Lambda_{0},\ m\Lambda_{0}+\rho\rangle\right)\right)g_{ij}

where we recall that m\nabla^{m} is the covariant derivative on P×GVmP\times_{G}V_{m} induced by the connection ω\omega.

Since mΛ0,mΛ0+ρ\langle m\Lambda_{0},\ m\Lambda_{0}+\rho\rangle is a constant and the connection m\nabla^{m} is compatible with the fiber metric it follows exactly as in the scalar case that if we abuse notation and also use m\Box_{m} to denote

m=(m)m+V+mΛ0,mΛ0+ρ\Box_{m}=(\nabla^{m})^{*}\nabla^{m}+V+\langle m\Lambda_{0},\ m\Lambda_{0}+\rho\rangle

acting on sections of P×GVmP\times_{G}V_{m} then

divM(Tm(s))\displaystyle\operatorname{div}_{M}(T_{m}(s)) =ms,ms12|s|2dV\displaystyle=-\langle\Box_{m}s,\nabla^{m}s\rangle-\frac{1}{2}|s|^{2}dV
divM(Tm(s)(Z))\displaystyle\operatorname{div}_{M}(T_{m}(s)(Z)) =(divMTm(s))(Z)=0 if ms=0\displaystyle=(\operatorname{div}_{M}T_{m}(s))(Z)=0\ \mbox{ if }\Box_{m}s=0

where we note that despite the raised and lowered mm’s appearing, we are not summing over them: they merely denote the representation of GG we are considering.

Just as in the scalar case, we can define the space of finite-energy solutions ss to ms=0\Box_{m}s=0 and one has Cauchy-data isomorphisms:

s(s|Σ0,(n^ms)|Σ0)s\mapsto\left(s|_{\Sigma_{0}},\ (\nabla^{m}_{\widehat{n}}s)|_{\Sigma_{0}}\right)

which give kerm\ker\Box_{m} the topology of a Hilbert space. Furthermore, since ZZ is Killing the covariant derivative Zm\nabla^{m}_{Z} commutes with m\Box_{m} and we have the densely defined operator

Dm,Z:=1iZm:kermCkermC.D_{m,Z}:=\frac{1}{i}\nabla^{m}_{Z}:\ker\Box_{m}\cap C^{\infty}\to\ker\Box_{m}\cap C^{\infty}.

Combining all of our results in this section and especially using 4 we arrive at the following result.

Proposition 3.28.

Let VGrm(P)V\in\operatorname{Gr}_{m}(P) and Φ:VmV\Phi:V_{m}\to V a unitary isomorphism so that we can define

ψ(p):=j=1dmΦ(ej)(p)ej.\vec{\psi}(p):=\sum_{j=1}^{d_{m}}\Phi(\vec{e}_{j})(p)\vec{e}_{j}.

Then Ψ(ψ)kerm\Psi(\vec{\psi})\in\ker\Box_{m} and

Qm(Ψ(ψ)):=Σ0Tm(Ψ(ψ))(Z,n^)𝑑VΣ0=Vol(G)Qω(ψ)Q_{m}(\Psi(\vec{\psi})):=\int_{\Sigma_{0}}T_{m}(\Psi(\vec{\psi}))(Z,\widehat{n})dV_{\Sigma_{0}}=\operatorname{Vol}(G)Q_{\omega}(\vec{\psi})

where Vol(G)\operatorname{Vol}(G) is taken with respect to the volume form induced by our Ad\operatorname{Ad}-invariant inner product on 𝔤\mathfrak{g}. Furthermore, since mm0m\geq m_{0} by assumption it follows that QmQ_{m} is positive definite on the finite energy space kerm\ker\Box_{m}.

We are now ready to apply the results of [14]. Really we are using a very special case of these results since we only need them to show that our multiplicity distributions μ(E,m,)\mu(E,m,-) are tempered.

Theorem 3.29.

[14] The operator Dm,ZD_{m,Z} is self-adjoint on (kerm,Qm)(\ker\Box_{m},Q_{m}) with σ(Dm,Z)\sigma(D_{m,Z})\subseteq\mathbb{R} discrete and accumulating at ±\pm\infty with polynomial growth.

Corollary 3.30.

The spectrum of DZD_{Z} on m\mathcal{H}_{m} is real, discrete and accumulates at ±\pm\infty with polynomial growth. Furthermore, the multiplicity of λσ(DZ)\lambda\in\sigma(D_{Z}) is equal to dm=dim(Vm)d_{m}=\dim(V_{m}) times the multiplicity of λσ(Dm,Z)\lambda\in\sigma(D_{m,Z}).

Corollary 3.31.

The distribution μ(E,m,)\mu(E,m,-) given by

μ(E,m,φ):=φ^(λm,mE)\mu(E,m,\varphi):=\sum_{\ell\in\mathbb{Z}}\widehat{\varphi}(\lambda_{m,\ell}-mE)

is a tempered distribution on \mathbb{R}. Here we recall that λm,λm,+1\cdots\leq\lambda_{m,\ell}\leq\lambda_{m,\ell+1}\leq\cdots are the eigenvalues of DZD_{Z} on m\mathcal{H}_{m}.

4 Proofs of Main Theorems

We are now prepared to study the mm\to\infty asymptotics of μ(E,m,φ)\mu(E,m,\varphi). As it turns out, this will depend significantly on whether or not 0suppφ^0\in\operatorname{supp}\widehat{\varphi}. For now we illustrate the method from Section 7 of [10] where φ\varphi is fixed and arbitrary. This method takes advantage of the periodicity and “positive frequency” property of our distributions to express them in terms of linear combinations of the basic homogeneous periodic distributions

m=1mkzmeimθ\sum_{m=1}^{\infty}m^{k}z^{-m}e^{im\theta}

with zS1z\in S^{1} and k0k\in\mathbb{Z}_{\geq 0} determining the location of the singularity and the homogeneity respectively. A key advantage of these techniques from [10] is that it circumvents the need for general Tauberian theorems.

From now on we replace 𝒪\mathcal{O} with m0𝒪m_{0}\mathcal{O} so that we may assume m0=1m_{0}=1.

Definition 4.1.

We define the generating function of the multiplicities μ(E,m,φ)\mu(E,m,\varphi) to be the periodic distribution in the real variable θ\theta:

Υ(φ)(θ):=m=1μ(E,m,φ)eimθ\Upsilon(\varphi)(\theta):=\sum_{m=1}^{\infty}\mu(E,m,\varphi)e^{im\theta}

defined for any function f(θ)f(\theta) which is the Fourier transform of a compactly supported function on \mathbb{R}.

Distributions of the form m=1ameimθ\sum_{m=1}^{\infty}a_{m}e^{im\theta} with ama_{m} real are called Hardy distributions. These are precisely the distributions on the sphere S1S^{1} whose negative Fourier coefficients all vanish and so they have nice descriptions in terms of boundary values of holomorphic functions on the unit disk via the Paley-Weiner theorem. The asymptotics of the Fourier coefficients of such distributions, especially when ama_{m} is a homogeneous function of mm, have been studied in books such as [4] Sections 12 and 13, and applied to spectral asymptotics in [10],[11],[23] for example.

Later in this section we will write Υ(φ)\Upsilon(\varphi) as a composition of Fourier integral operators and through this we will show that it is actually in 𝒟(/2π)\mathcal{D}^{\prime}(\mathbb{R}/2\pi\mathbb{Z}). For now we illustrate how the asymptotics of the Fourier coefficients of a general Lagrangian distribution Υ\Upsilon on S1S^{1} can be related to its principal symbol.

Definition 4.2.

Let Υ𝒟(/2π)\Upsilon\in\mathcal{D}^{\prime}(\mathbb{R}/2\pi\mathbb{Z}). An element s0singsupp(Υ)s_{0}\in\mbox{singsupp}(\Upsilon) is called classical of degree kk if and only if when interpreting Υ\Upsilon as a 2π2\pi\mathbb{Z}-periodic distribution on \mathbb{R} we have:

  1. 1.

    s0s_{0} is an isolated singularity, and

  2. 2.

    for any ρCc()\rho\in C^{\infty}_{c}(\mathbb{R}) with ρ1\rho\equiv 1 on a neighborhood of s0s_{0} and singsupp(Υ)supp(ρ)={s0}\operatorname{singsupp}(\Upsilon)\cap\operatorname{supp}(\rho)=\{s_{0}\} we have asymptotic expansions:

    ρΥ^(ξ)\displaystyle\widehat{\rho\Upsilon}(\xi) eis0ξ=0c+ξk as ξ+ and\displaystyle\sim e^{-is_{0}\xi}\sum_{\ell=0}^{\infty}c_{\ell}^{+}\xi^{k-\ell}\ \mbox{ as }\ \xi\to+\infty\ \mbox{ and}
    ρΥ^(ξ)\displaystyle\widehat{\rho\Upsilon}(\xi) eis0ξ=0cξk as ξ.\displaystyle\sim e^{-is_{0}\xi}\sum_{\ell=0}^{\infty}c_{\ell}^{-}\xi^{k-\ell}\ \mbox{ as }\ \xi\to-\infty.
Lemma 4.3.

Let s0s_{0}\in\mathbb{R} be a classical singularity of Υ\Upsilon of degree kk, and let ρCc()\rho\in C^{\infty}_{c}(\mathbb{R}) have ρ1\rho\equiv 1 on a neighborhood of s0s_{0} and singsupp(Υ(φ))supp(ρ)={s0}\operatorname{singsupp}(\Upsilon(\varphi))\cap\operatorname{supp}(\rho)=\{s_{0}\}. Then

ρΥ(φ)Ik+1/4(,Λ)\rho\Upsilon(\varphi)\in I^{k+1/4}(\mathbb{R},\Lambda)

where Λ={(s0,ξ)T0:ξ0}\Lambda=\{(s_{0},\xi)\in T^{*}\mathbb{R}\setminus 0\ :\ \xi\neq 0\}. If c=0c_{\ell}^{-}=0 for all \ell (in which case the singularity is called positive) then instead Λ={(s0,ξ)T0:ξ>0}\Lambda=\{(s_{0},\xi)\in T^{*}\mathbb{R}\setminus 0\ :\ \xi>0\}.

Proof.

We can write the distribution ρΥ\rho\Upsilon as

ρΥ,ψ=ei(ss0)ξ(eis0ξρΥ^(ξ))ψ(s)𝑑s𝑑ξ\langle\rho\Upsilon,\psi\rangle=\int_{\mathbb{R}}e^{i(s-s_{0})\xi}(e^{is_{0}\xi}\widehat{\rho\Upsilon}(\xi))\psi(s)dsd\xi

and so it suffices to check whether the function eis0ξρΥ^(ξ)e^{is_{0}\xi}\widehat{\rho\Upsilon}(\xi) lives in the correct symbol class. Since ρΥ()\rho\Upsilon\in\mathcal{E}^{\prime}(\mathbb{R}) its Fourier transform is a smooth function and our asymptotics precisely tell us that it lives in the symbol class

Sk(s×ξ) (it is independent of s).S^{k}(\mathbb{R}_{s}\times\mathbb{R}_{\xi})\ \mbox{ (it is independent of }s).

Since dim(s)=1=dim(ξ)\dim(\mathbb{R}_{s})=1=\dim(\mathbb{R}_{\xi}) this is the correct order for a symbol to define an FIO of order

k(121)/4=k+1/4.k-(1-2\cdot 1)/4=k+1/4.

As for the Lagrangian, one simply notices first that the ξ\xi-critical points of the phase are precisely the set of (s,ξ)(s,\xi) with s=s0s=s_{0}, meanwhile the support of eis0ξρΥ^(ξ)e^{is_{0}\xi}\widehat{\rho\Upsilon}(\xi) is everywhere in the non-positive singularity case and is a positive ray in the case of a positive singularity. ∎

Lemma 4.4.

Suppose Υ\Upsilon had only finitely many singularities z1,,zqS1z_{1},...,z_{q}\in S^{1} and that for s1,,sq[0,2π]s_{1},...,s_{q}\in[0,2\pi] with eis1=z1,,eisq=zqe^{-is_{1}}=z_{1},...,e^{-is_{q}}=z_{q} the singularities s1,,sqs_{1},...,s_{q} were all classical with respective degrees k1,,kqk_{1},...,k_{q}. For some ρjCc()\rho_{j}\in C^{\infty}_{c}(\mathbb{R}) smooth cutoffs with ρj1\rho_{j}\equiv 1 on a neighborhood of sjs_{j} and singsupp(Υ)supp(ρj)={sj}\operatorname{singsupp}(\Upsilon)\cap\operatorname{supp}(\rho_{j})=\{s_{j}\}, and for c±,jc^{\pm,j}_{\ell} the coefficients of our asymptotic expansions for ρjΥ^\widehat{\rho_{j}\Upsilon}:

ρjΥ^(ξ)=0c±,jξkj as ξ±\widehat{\rho_{j}\Upsilon}(\xi)\sim\sum_{\ell=0}^{\infty}c^{\pm,j}_{\ell}\xi^{k_{j}-\ell}\ \mbox{ as }\ \xi\to\pm\infty

we have:

12π02πeimsΥ(θ)𝑑s=0j=1qc+,jωjmmkj as m.\frac{1}{2\pi}\int_{0}^{2\pi}e^{-ims}\Upsilon(\theta)ds\sim\sum_{\ell=0}^{\infty}\sum_{j=1}^{q}c^{+,j}_{\ell}\omega_{j}^{-m}m^{k_{j}-\ell}\ \mbox{ as }\ m\to\infty.
Proof.

Choose our cutoffs ρj\rho_{j} to be non-negative with disjoint supports and such that there exists ηCc()\eta\in C^{\infty}_{c}(\mathbb{R}) with 0η10\leq\eta\leq 1 such that

ρ1++ρq+η1 on [0,2π] and singsupp(Υ)supp(η)=.\rho_{1}+\cdots+\rho_{q}+\eta\equiv 1\ \mbox{ on }\ [0,2\pi]\ \mbox{ and }\ \operatorname{singsupp}(\Upsilon)\cap\operatorname{supp}(\eta)=\emptyset.

Then, taking Fourier transforms we have

12π02πeimsΥ(s)𝑑s=12πj=1q02πeimsρj(s)Υ(s)𝑑s+12π02πeimsη(s)Υ(s)𝑑s.\frac{1}{2\pi}\int_{0}^{2\pi}e^{-ims}\Upsilon(s)ds=\frac{1}{2\pi}\sum_{j=1}^{q}\int_{0}^{2\pi}e^{-ims}\rho_{j}(s)\Upsilon(s)ds+\frac{1}{2\pi}\int_{0}^{2\pi}e^{-ims}\eta(s)\Upsilon(s)ds.

Since ηΥCc()\eta\Upsilon\in C^{\infty}_{c}(\mathbb{R}) we have that the last term is going to 0 rapidly as mm\to\infty. For the remaining terms we have

12π02πeimsρj(s)Υ(s)𝑑s=ρjΥ^(m)eisjm=0c+,jmkj as m.\frac{1}{2\pi}\int_{0}^{2\pi}e^{-ims}\rho_{j}(s)\Upsilon(s)ds=\widehat{\rho_{j}\Upsilon}(m)\sim e^{-is_{j}m}\sum_{\ell=0}^{\infty}c^{+,j}_{\ell}m^{k_{j}-\ell}\ \mbox{ as }\ m\to\infty.

Summing these asymptotics together then yields our desired result. ∎

So we see that in order to obtain the leading order asymptotics of μ(E,m,φ)\mu(E,m,\varphi) as mm\to\infty it suffices to demonstrate that the singularities of Υ(φ)\Upsilon(\varphi) are classical and to compute both its order as an FIO, and the leading terms c0+,jc^{+,j}_{0} in the asymptotic expansions of its Fourier transform. Let’s now check how to obtain c0+,jc^{+,j}_{0} from the principal symbol.

Lemma 4.5.

Let s0s_{0} be a classical singularity of degree kk of Υ\Upsilon, let ρCc()\rho\in C^{\infty}_{c}(\mathbb{R}) be a cutoff as in the previous lemma and let a(s,ξ)a(s,\xi) be any principal symbol for ρΥ\rho\Upsilon. i.e.

a(s,ξ)eis0ξρΥ^(ξ)Sk1(s×ξ).a(s,\xi)-e^{is_{0}\xi}\widehat{\rho\Upsilon}(\xi)\in S^{k-1}(\mathbb{R}_{s}\times\mathbb{R}_{\xi}).

Then

c0±=limξ±a(s,ξ)ξk.c^{\pm}_{0}=\lim_{\xi\to\pm\infty}a(s,\xi)\xi^{-k}.
Proof.

Indeed, if a(s,ξ)a(s,\xi) is any principal symbol for ρΥ\rho\Upsilon then, by definition

|a(s,ξ)eis0ξρΥ^(ξ)|(1+|ξ|)k1|a(s,\xi)-e^{is_{0}\xi}\widehat{\rho\Upsilon}(\xi)|\lesssim(1+|\xi|)^{k-1}

and so dividing by ξk\xi^{k} and taking limits yields our desired result. ∎

So, our goal has now been reduced to writing Υ(φ)\Upsilon(\varphi) as a composition of well-understood FIOs and computing the order and principal symbol of the composition in terms of its constituents. Let’s begin by introducing the relevant operators from [11] and [22].

Definition 4.6.

Let EadvE_{adv} and EretE_{ret} respectively denote the advanced and retarded fundamental solutions for ω\Box_{\omega}. Explicitly, for fCc(P)f\in C^{\infty}_{c}(P), u:=Eadvfu:=E_{adv}f is the unique solution to ωu=f\Box_{\omega}u=f whose support is contained in the forward causal set of supp(f)\operatorname{supp}(f). i.e. uu solves ωu=f\Box_{\omega}u=f with vanishing Cauchy data in the past before supp(f)\operatorname{supp}(f). Similarly, EretfE_{ret}f is the unique solution to the Cauchy problem ωu=f\Box_{\omega}u=f with vanishing Cauchy data in the Causal future of ff (in the future after supp(f)\operatorname{supp}(f)).

Lemma 4.7.

[22]
let fC(P0)f\in C^{\infty}(P_{0}) and let |dVP0||dV_{P_{0}}| denote the measure on PP given by integration over P0P_{0} with respect to the induced volume measure on P0P_{0} from the metric. Write E:=EadvEretE:=E_{adv}-E_{ret}. Then

u:=E(f1|dVP0|+n^(f2|dVP0|))u:=E(f_{1}|dV_{P_{0}}|+\partial_{\widehat{n}}(f_{2}|dV_{P_{0}}|))

(where n^(f2|dVP0|)\partial_{\widehat{n}}(f_{2}|dV_{P_{0}}|) is a distributional derivative of a measure) is the unique solution to the Cauchy problem Δωu=0\Delta_{\omega}u=0 with Cauchy data

{u(x,0)=f1(x) on P0,(n^u)(x,0)=f2(x) on P0.\begin{cases}u(x,0)&=f_{1}(x)\ \mbox{ on }P_{0},\\ (\mathcal{L}_{\widehat{n}}u)(x,0)&=f_{2}(x)\ \mbox{ on }P_{0}.\end{cases}

Furthermore EI3/2(P×P;C1)E\in I^{-3/2}(P\times P;C_{1}^{\prime}) with the canonical relation C1C_{1} given by

C1={(ζ1;ζ2)T0P×T0P:s such that ζ2=Gs(ζ1)}.C_{1}=\{(\zeta_{1};\ \zeta_{2})\in T^{*}_{0}P\times T^{*}_{0}P\ :\ \exists s\in\mathbb{R}\mbox{ such that }\zeta_{2}=G_{-s}(\zeta_{1})\}.

Parametrizing the left copy of T0PT^{*}_{0}P in C1C_{1} by T0P|P0×𝒩×sT^{*}_{0}P|_{P_{0}}\times\mathbb{R}\cong\mathcal{N}\times\mathbb{R}_{s^{\prime}} via the geodesic flow and then the ζ2=Gs(ζ1)\zeta_{2}=G_{-s}(\zeta_{1}) by the parameter ss, the principal symbol of EE is given by the half-density

|dC1|1/2:=12|Ω𝒩|1/2|ds|1/2|ds|1/2|d_{C_{1}}|^{1/2}:=-\frac{1}{2}|\Omega_{\mathcal{N}}|^{1/2}\otimes|ds^{\prime}|^{1/2}\otimes|ds|^{1/2}

where Ω𝒩\Omega_{\mathcal{N}} is the Liouville volume form on 𝒩\mathcal{N} induced by the symplectic form.

Before we get to composing FIO’s, let’s recall how this works [13]. Suppose we had smooth manifolds X,Y,ZX,Y,Z of respective dimensions nX,nY,nZn_{X},n_{Y},n_{Z} respectively and C1(TZ0)×(TY0)C_{1}\subseteq(T^{*}Z\setminus 0)\times(T^{*}Y\setminus 0), C2(TY0)×(TX0)0C_{2}\subseteq(T^{*}Y\setminus 0)\times(T^{*}X\setminus 0)\setminus 0 canonical relations. We write CjC_{j}^{\prime} for the result of multiplying the left fiber variables by 1-1 so that the result is a Lagrangian submanifold. Given

A1Id1(Z×Y;C1) and A2Id2(Y×X;C2)A_{1}\in I^{d_{1}}(Z\times Y;C_{1}^{\prime})\ \mbox{ and }\ A_{2}\in I^{d_{2}}(Y\times X;C_{2}^{\prime})

we interpret A1A_{1} and A2A_{2} as operators

A1:Cc(Y)𝒟(Z) and A2:Cc(X)𝒟(Y).A_{1}:C^{\infty}_{c}(Y)\to\mathcal{D}^{\prime}(Z)\ \mbox{ and }\ A_{2}:C^{\infty}_{c}(X)\to\mathcal{D}^{\prime}(Y).

One can then often form the composition

A1A2Id1+d2+e2(Z×X;(C1C2))A_{1}\circ A_{2}\in I^{d_{1}+d_{2}+\frac{e}{2}}(Z\times X;(C_{1}\circ C_{2})^{\prime})

where ee and (C1C2)(C_{1}\circ C_{2})^{\prime} are defined as follows. Since C1C_{1} and C2C_{2} are Lagrangian they have dimensions:

dim(C1)=nX+nY and dim(C2)=nY+nZ.\dim(C_{1})=n_{X}+n_{Y}\ \mbox{ and }\ \dim(C_{2})=n_{Y}+n_{Z}.

The product C1×C2C_{1}\times C_{2} lives in TZ×(TY)×2×TXT^{*}Z\times(T^{*}Y)^{\times 2}\times T^{*}X and has dimension

dim(C1×C2)=nX+2nY+nZ.\dim(C_{1}\times C_{2})=n_{X}+2n_{Y}+n_{Z}.

Meanwhile we also have a diagonal submanifold

D:=(TZ0)×diag(TY0)×(TX0)D:=(T^{*}Z\setminus 0)\times\operatorname{diag}(T^{*}Y\setminus 0)\times(T^{*}X\setminus 0)

of dimension dim(D)=2nZ+2nY+2nX\dim(D)=2n_{Z}+2n_{Y}+2n_{X}. Since the total space TZ×(TY)×2×TXT^{*}Z\times(T^{*}Y)^{\times 2}\times T^{*}X has dimension 2nZ+4nY+2nX2n_{Z}+4n_{Y}+2n_{X} it follows that if DD and C1×C2C_{1}\times C_{2} intersected transversely then the intersection would have dimension

dim(D(C1×C2))=nZ+nX\dim(D\cap(C_{1}\times C_{2}))=n_{Z}+n_{X}

and if πX,πZ\pi_{X},\pi_{Z} are respectively the projection maps from TZ×(TY)×2×TXT^{*}Z\times(T^{*}Y)^{\times 2}\times T^{*}X to TXT^{*}X and TZT^{*}Z then the restriction

πZ×πX|D(C1×C2):D(C1×C2)C1C2:=(πZ×πX)(D(C1×C2))\pi_{Z}\times\pi_{X}|_{D\cap(C_{1}\times C_{2})}:D\cap(C_{1}\times C_{2})\to C_{1}\circ C_{2}:=(\pi_{Z}\times\pi_{X})(D\cap(C_{1}\times C_{2}))

is a local diffeomorphism and C1C2C_{1}\circ C_{2} is a Lagrangian submanifold of (TZ0)×(TX0)(T^{*}Z\setminus 0)\times(T^{*}X\setminus 0). In this case, as long as everything is properly supported, we can take e=0e=0 and we have

A1A2Id1+d2(Z×X;(C1C2)).A_{1}\circ A_{2}\in I^{d_{1}+d_{2}}(Z\times X;(C_{1}\circ C_{2})^{\prime}).

We call this a transverse composition of FIO’s. Furthermore, in this case if a1,a2a_{1},a_{2} are the principal symbols of A1,A2A_{1},A_{2} then:

the principal symbol of A1A2 is given by the restriction of a1×a2 to (C1C2).\mbox{the principal symbol of }A_{1}\circ A_{2}\mbox{ is given by the restriction of }a_{1}\times a_{2}\mbox{ to }(C_{1}\circ C_{2})^{\prime}.

However, one can still form the composition A1A2A_{1}\circ A_{2} if the intersection of DD and C1×C2C_{1}\times C_{2} in TZ×(TY)×2×TXT^{*}Z\times(T^{*}Y)^{\times 2}\times T^{*}X is merely clean. In this case the intersection is still a smooth manifold, its tangent spaces are given by the intersections of the tangent spaces of DD and C1×C2C_{1}\times C_{2}, C1C2C_{1}\circ C_{2} is still defined in the same way, but now the projection map

πZ×πX|D(C1×C2):D(C1×C2)C1C2\pi_{Z}\times\pi_{X}|_{D\cap(C_{1}\times C_{2})}:D\cap(C_{1}\times C_{2})\to C_{1}\circ C_{2}

is merely required to be a submersion. Since we’re assuming everything is properly supported it follows that the fibers are compact manifolds. We define:

e:= the dimension of the fibers of πZ×πX|D(C1×C2).e:=\mbox{ the dimension of the fibers of }\pi_{Z}\times\pi_{X}|_{D\cap(C_{1}\times C_{2})}.

This is called the excess. Then from Proposition 25.1.5’ in [13] we have

A1A2Id1+d2+e2(Z×X;(C1C2))A_{1}\circ A_{2}\in I^{d_{1}+d_{2}+\frac{e}{2}}(Z\times X;(C_{1}\circ C_{2})^{\prime})

where if a1,a2a_{1},a_{2} are the principal symbols of A1,A2A_{1},A_{2} respectively then the principal symbol of A1A2A_{1}\circ A_{2} at a point z(C1C2)z\in(C_{1}\circ C_{2})^{\prime} is given by

Fza1×a2 where Fz is the fiber over z.\int_{F_{z}}a_{1}\times a_{2}\ \mbox{ where }\ F_{z}\ \mbox{ is the fiber over }z.

We will call this a clean composition of FIOs. It should be noted that the above results can be occasionally tweaked to apply when some of the hypotheses (such as CjC_{j} being a canonical relation) aren’t exactly satisfied as long as one is careful to ensure that the wavefront sets line up correctly in order for the desired products to be defined.

Finally we should say that in our below computations we omit the Maslov index factors until the very end.

We are now ready to apply the above FIO calculus in order to better understand our generating function Υ(φ)\Upsilon(\varphi). Let’s recall our notation from earlier:

  • dd is the dimension of GG,

  • n+1n+1 is the dimension of MM with nn the dimension of Σ0\Sigma_{0},

  • n+1+dn+1+d is the dimension of PP,

  • T0PT^{*}_{0}P is a cone subbundle of TP0T^{*}P\setminus 0 and has dimension 2(n+1+d)12(n+1+d)-1.

  • The restriction T0P|P0T^{*}_{0}P|_{P_{0}} is symplectomorphic to TP00T^{*}P_{0}\setminus 0 (but not in a >0\mathbb{R}_{>0}-equivariant way) and both have dimension 2(n+d)2(n+d).

  • dim𝒪=:2\dim\mathcal{O}=:2\ell so 𝒩𝒪\mathcal{N}_{\mathcal{O}} has dimension 2(n+)2(n+\ell).

The below result is also from [22] and again we state it for the reader’s convenience.

Lemma 4.8.

[22]
Let Et(x,y):=eit(DZ)xE(x,y)E_{t}(x,y):=e^{-it(D_{Z})_{x}}E(x,y). Then

Et(x,y)I7/4(P×P×;C2)E_{t}(x,y)\in I^{-7/4}(P\times P\times\mathbb{R};C_{2}^{\prime})

where C2C_{2} is the canonical relation:

C2\displaystyle C_{2} :={(ζ1;ζ2;t,τ)(T0P)×2×(T0)\displaystyle:=\{(\zeta_{1};\ \zeta_{2};\ t,\tau)\in(T^{*}_{0}P)^{\times 2}\times(T^{*}\mathbb{R}\setminus 0)
:τ+Zω,ζ1=0,s such that ζ2=(GsΦtZ)(ζ1)}.\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ :\ \tau+\langle Z^{\omega},\zeta_{1}\rangle=0,\ \exists s\mbox{ such that }\zeta_{2}=(G_{-s}\circ\Phi_{t}^{Z})(\zeta_{1})\}.

Parametrizing C2C×tC_{2}\cong C\times\mathbb{R}_{t} as the flowout of CC under the ZZ-flow the principal symbol of Et(x,y)E_{t}(x,y) is given by:

i2(2π)3/4|dC1|1/2|dt|1/2 on C±\mp\frac{i}{2}(2\pi)^{3/4}|d_{C_{1}}|^{1/2}\otimes|dt|^{1/2}\ \mbox{ on }\ C_{\pm}

where C±C_{\pm} is the subset of C1C_{1} where both covectors are in T±PT^{*}_{\pm}P.

In the next lemma we begin combining results from [22] and [11].

Lemma 4.9.

The right GG-action gives us an action map C(P)C(P×G)C^{\infty}(P)\to C^{\infty}(P\times G) which is an FIO

FId/4(P×P×G;Γ0)F\in I^{-d/4}(P\times P\times G;\Gamma_{0}^{\prime})

with Γ0\Gamma_{0}^{\prime} the moment Lagrangian, whose canonical relation is:

Γ0:={(ζ;ζg;g,η)(TP0)×2×(TG0):μ(ζ)=η}.\Gamma_{0}:=\{(\zeta;\ \zeta\cdot g;\ g,\eta)\in(T^{*}P\setminus 0)^{\times 2}\times(T^{*}G\setminus 0)\ :\ \mu(\zeta)=\eta\}.

The composition EtFE_{t}\circ F, denoted by Et(x,yg)E_{t}(x,yg), arises from a transverse intersection of canonical relations and is therefore an FIO:

Et(x,yg)I(d+7)/4(P×P×G×;Γ)E_{t}(x,yg)\in I^{-(d+7)/4}(P\times P\times G\times\mathbb{R};\Gamma^{\prime})

with canonical relation

Γ\displaystyle\Gamma :={(ζ1;ζ2;g,η;t,τ)(T0P)×2×(TG0)×(T0)\displaystyle:=\{(\zeta_{1};\ \zeta_{2};\ g,\eta;\ t,\tau)\in(T^{*}_{0}P)^{\times 2}\times(T^{*}G\setminus 0)\times(T^{*}\mathbb{R}\setminus 0)
:τ+Zω,ζ1=0,μ(ζ2g1)=η,s such that ζ2=(GsΦtZ)(ζ1)g}\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ :\tau+\langle Z^{\omega},\zeta_{1}\rangle=0,\ \mu(\zeta_{2}g^{-1})=\eta,\ \exists s\mbox{ such that }\ \zeta_{2}=(G_{-s}\circ\Phi_{t}^{Z})(\zeta_{1})g\}

Parametrizing Γ\Gamma by C2×G×C1×t1×G×t2C_{2}\times G\times\mathbb{R}\cong C_{1}\times\mathbb{R}_{t_{1}}\times G\times\mathbb{R}_{t_{2}} the principal symbol of Et(x,yg)E_{t}(x,yg) is given by:

i2(2π)(d+3)/4|dC1|1/2|dt1|1/2|dg|1/2|dt2|1/2 on C±\mp\frac{i}{2}(2\pi)^{(d+3)/4}|d_{C_{1}}|^{1/2}\otimes|dt_{1}|^{1/2}\otimes|dg|^{1/2}\otimes|dt_{2}|^{1/2}\ \mbox{ on }\ C_{\pm}

where |dg||dg| is the volume measure on GG induced by our Ad\operatorname{Ad}-invariant inner product on the tangent spaces.

Proof.

The expression for the moment Lagrangian and the fact that Γ0Id/4(P×P×G;Γ0)\Gamma_{0}\in I^{-d/4}(P\times P\times G;\Gamma_{0}^{\prime}) is proven in [11]. By construction we have

Γ=C2Γ0\Gamma=C_{2}\circ\Gamma_{0}

and the composition is clean so the orders of the FIOs simply add up. ∎

In [22], the distributional trace of eitDZe^{-itD_{Z}} was expressed in terms of EtE_{t} and so Et(x,yg)E_{t}(x,yg) will play a similar role for our equivariant trace.

Lemma 4.10.

Write n^x,n^y\widehat{n}_{x},\widehat{n}_{y} for the Lie derivatives along the unit normal n^\widehat{n} in the variables xx and yy respectively. Then

:=n^xEt(x,yg)n^yEt(x,yg)I(d+3)/4(P×P×G×;Γ)\mathcal{F}:=\widehat{n}_{x}E_{t}(x,yg)-\widehat{n}_{y}E_{t}(x,yg)\in I^{-(d+3)/4}(P\times P\times G\times\mathbb{R};\Gamma)

with Γ\Gamma given in 4.9. Under the same parametrization of Γ\Gamma as in 4.9, the principal symbol of \mathcal{F} is given by

±12(2π)(d+3)/4n^,|dC1|1/2|dt1|1/2|dg|1/2|dt2|1/2 on C±\pm\frac{1}{2}(2\pi)^{(d+3)/4}\langle\widehat{n},-\rangle|d_{C_{1}}|^{1/2}\otimes|dt_{1}|^{1/2}\otimes|dg|^{1/2}\otimes|dt_{2}|^{1/2}\ \mbox{ on }C_{\pm}

where n^,\langle\widehat{n},-\rangle is the function on Γ\Gamma given by pairing the first cotangent vector with n^\widehat{n}.

Proof.

Differentiation is a differential operator, hence pseudodifferential operator, and so its Lagrangian is just the diagonal. Therefore differentiating an FIO does not affect the Lagrangian and merely increases the order by 1. ∎

For the next couple of lemmas we hold off on computing the principal symbols since it will be easier to directly compute the principal symbol of the wave trace after all of these compositions.

Lemma 4.11.

Let diag:PP×P\operatorname{diag}:P\to P\times P denote the diagonal map so that pulling back along diag\operatorname{diag} is an FIO

diagI(n+1+d)/4(P×P×P;C3)\operatorname{diag}^{*}\in I^{(n+1+d)/4}(P\times P\times P;C_{3}^{\prime})

with Lagrangian C3C_{3}^{\prime} where

C3={(p,ζ2ζ1;p,ζ1;p,ζ2)(TP0)×3}.C_{3}=\{(p,\zeta_{2}-\zeta_{1};\ p,\zeta_{1};\ p,\zeta_{2})\in(T^{*}P\setminus 0)^{\times 3}\}.

Then the composition diag\operatorname{diag}^{*}\mathcal{F} arises from a transverse intersection and is therefore an FIO

diagI(n2)/4(P×G×;Γ1)\operatorname{diag}^{*}\mathcal{F}\in I^{(n-2)/4}(P\times G\times\mathbb{R};\Gamma_{1})

where

Γ1\displaystyle\Gamma_{1} :={(ζ2ζ1;g,η;t,τ)TP×(TG0)×(T0)\displaystyle:=\{(\zeta_{2}-\zeta_{1};\ g,\eta;\ t,\tau)\in T^{*}P\times(T^{*}G\setminus 0)\times(T^{*}\mathbb{R}\setminus 0)
:τ+Zω,ζ1=0,μ(ζ2g1)=η,s such that ζ2=(GsΦtZ)(ζ1)g\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ :\ \tau+\langle Z^{\omega},\zeta_{1}\rangle=0,\ \mu(\zeta_{2}g^{-1})=\eta,\ \exists s\mbox{ such that }\zeta_{2}=(G_{-s}\circ\Phi^{Z}_{t})(\zeta_{1})g
 and ζ1,ζ2T0P live over the same point in P}\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \mbox{ and }\zeta_{1},\zeta_{2}\in T^{*}_{0}P\mbox{ live over the same point in }P\}
Proof.

The expression for Γ1\Gamma_{1} above is precisely the definition of C3ΓC_{3}\circ\Gamma so let’s check that this is indeed a transverse composition. Notice that in the definition of Γ1\Gamma_{1}, one ζ1\zeta_{1} and tt are chosen, ss and gg are uniquely determined by the requirement that (GsΦtZ)(ζ1)g(G_{-s}\circ\Phi^{Z}_{t})(\zeta_{1})g must live over the same point in PP as ζ1\zeta_{1}. Furthermore, the constraint that there must exist s,gs,g so that (GsΦtZ)(ζ1)g(G_{-s}\circ\Phi^{Z}_{t})(\zeta_{1})g lives over the same point in PP adds nn independent constraints on ζ1\zeta_{1} since they must also live over the same point in Σ0\Sigma_{0}. This is unless t=0t=0.

So, given 0t,ζ10\neq t,\zeta_{1} satisfying our nn independent constraints: s,gs,g and therefore ζ2\zeta_{2} and η\eta are completely determined. τ\tau is directly determined by ζ1\zeta_{1}. Hence we see that there are exactly

dim(T0P)+1n=2(n+1+d)1+1n=(n+1+d)+d+1\dim(T^{*}_{0}P)+1-n=2(n+1+d)-1+1-n=(n+1+d)+d+1

independent directions in both the composition C3ΓC_{3}\circ\Gamma and in the fiber over the point corresponding to 0t,ζ10\neq t,\zeta_{1}.

In the case t=0t=0 we necessarily have s=0s=0 and g=1Gg=1\in G, however the t=0t=0 local is a proper submanifold of Γ1\Gamma_{1} and the tangent space to Γ1\Gamma_{1} at t=0t=0 has d+1d+1 tangent directions arising from how s,gs,g vary as we move off the t=0t=0 local. Within the t=0t=0 local we then have ζ2=ζ1\zeta_{2}=\zeta_{1} and τ,η\tau,\eta are determined by ζ1=ζ2\zeta_{1}=\zeta_{2}. While, in this case, we do have dim(T0P)=2(n+1+d)1\dim(T^{*}_{0}P)=2(n+1+d)-1 choices for ζ1\zeta_{1}, it is the quantity ζ2ζ1\zeta_{2}-\zeta_{1} that appears in Γ1\Gamma_{1} and so the fiber coordinates of the first component of Γ1\Gamma_{1} always vanish. Thus in both Γ1\Gamma_{1} and in the fiber over the point corresponding to (0,ζ1)(0,\zeta_{1}) we have

dim(P)+d+1=(n+1+d)+d+1\dim(P)+d+1=(n+1+d)+d+1

tangent directions. Hence indeed we have a transverse intersection and the order of diag\operatorname{diag}^{*}\mathcal{F} is the sum of the orders of diag\operatorname{diag}^{*} and \mathcal{F}. ∎

Lemma 4.12.

Let ι:P0P\iota:P_{0}\hookrightarrow P denote the inclusion. Pulling back along ι\iota is an FIO

ιI1/4(P0×P;C4)\iota^{*}\in I^{1/4}(P_{0}\times P;C_{4}^{\prime})

with Lagrangian C4C_{4}^{\prime} defined by

C4={(x,ζ1;x,ζ2)TP0×TP|P0:ζ2|TP0=ζ1}.C_{4}=\{(x,\zeta_{1};\ x,\zeta_{2})\in T^{*}P_{0}\times T^{*}P|_{P_{0}}\ :\ \zeta_{2}|_{TP_{0}}=\zeta_{1}\}.

As in [22], the canonical relation of diag\operatorname{diag}^{*}\mathcal{F} is disjoint from the conormal bundle NP0N^{*}P_{0} and C4Γ1C_{4}\circ\Gamma_{1} arises from a tranverse intersection so the composition ιdiag\iota^{*}\operatorname{diag}^{*}\mathcal{F} can be formed as if it were a transverse composition of FIOs and

ιdiagI(n1)/4(P0×G×;Γ2)\iota^{*}\operatorname{diag}^{*}\mathcal{F}\in I^{(n-1)/4}(P_{0}\times G\times\mathbb{R};\Gamma_{2})

where

Γ2\displaystyle\Gamma_{2} :={((ζ2ζ1)|TP0;g,η;t,τ)TP0×(TG0)×(T0)\displaystyle:=\{((\zeta_{2}-\zeta_{1})|_{TP_{0}};\ g,\eta;\ t,\tau)\in T^{*}P_{0}\times(T^{*}G\setminus 0)\times(T^{*}\mathbb{R}\setminus 0)
:ζ1,ζ2T0P|P0 lie over the same point in P0,τ+Zω,ζ1=0,\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ :\zeta_{1},\zeta_{2}\in T^{*}_{0}P|_{P_{0}}\mbox{ lie over the same point in }P_{0},\ \tau+\langle Z^{\omega},\zeta_{1}\rangle=0,
μ(ζ2g1)=η,s such that ζ2=(GsΦtZ)(ζ1)g}.\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \mu(\zeta_{2}g^{-1})=\eta,\ \exists s\mbox{ such that }\zeta_{2}=(G_{-s}\circ\Phi_{t}^{Z})(\zeta_{1})g\}.
Proof.

The proof that the composition idiagi^{*}\operatorname{diag}^{*}\mathcal{F} can be formed is exactly the same as in Lemma 8.3 and the discussion preceding it in [22], and our Γ2\Gamma_{2} is precisely defined to be C4Γ1C_{4}\circ\Gamma_{1}. Transversality again follows from noticing that the intersection is clean and then dimension-counting, however we should remark that in order to get exactly n+2d+1n+2d+1 degrees of freedom one uses the fact that the restriction of covectors in T0P|P0T^{*}_{0}P|_{P_{0}} to TP0TP_{0} yields the isomorphism T0P|P0TP0T^{*}_{0}P|_{P_{0}}\cong T^{*}P\setminus 0 and hence the only way the fiber variable of the first component is zero is if ζ1=ζ2\zeta_{1}=\zeta_{2}. ∎

So, we’ve arrived at the following object:

ιdiag=(n^xEt(x,yg)n^yEt(x,yg))|x=yP0I(n1)/4(P0×G×;Γ2).\iota^{*}\operatorname{diag}^{*}\mathcal{F}=\left(\widehat{n}_{x}E_{t}(x,yg)-\widehat{n}_{y}E_{t}(x,yg)\right)|_{x=y\in P_{0}}\in I^{(n-1)/4}(P_{0}\times G\times\mathbb{R};\Gamma_{2}).

The importance of this object arises from the following slight generalization of Theorem 4.1 from [22].

Proposition 4.13.

Let Π:C(P0×G×)C(G×)\Pi_{*}:C^{\infty}(P_{0}\times G\times\mathbb{R})\to C^{\infty}(G\times\mathbb{R}) be the operator given by integration over P0P_{0}. Then since the base has dimension d+1d+1 and the fibers have dimension n+dn+d we have

ΠId+12n+d4(G××P0×G×;C5)\Pi_{*}\in I^{\frac{d+1}{2}-\frac{n+d}{4}}(G\times\mathbb{R}\times P_{0}\times G\times\mathbb{R};C_{5}^{\prime})

where

C5\displaystyle C_{5} ={(g,η;t,τ;x,0;g,η;t,τ)\displaystyle=\{(g,\eta;\ t,\tau;\ x,0;\ g,\eta;\ t,\tau)
(TG0)×(T0)×TP0×(TG0)×(T0):xP0}.\displaystyle\ \ \ \ \ \ \ \ \ \ \in(T^{*}G\setminus 0)\times(T^{*}\mathbb{R}\setminus 0)\times T^{*}P_{0}\times(T^{*}G\setminus 0)\times(T^{*}\mathbb{R}\setminus 0)\ :\ x\in P_{0}\}.

Furthermore, since m0=1m_{0}=1 and QωQ_{\omega} is positive definite on =m1L2m\mathcal{H}=\bigoplus_{m\geq 1}^{L^{2}}\mathcal{H}_{m}, if we set

𝒱:=Qω then kerω=𝒱\mathcal{V}:=\mathcal{H}^{\perp Q_{\omega}}\ \mbox{ then }\ \ker\Box_{\omega}=\mathcal{V}\oplus\mathcal{H}

then we have

𝒦(g,t):=Πιdiag\displaystyle\mathcal{K}(g,t):=\Pi_{*}\iota^{*}\operatorname{diag}^{*}\mathcal{F} =P0(n^xEt(x,yg)n^yEt(x,yg))|x=ydVP0(x)\displaystyle=\int_{P_{0}}\left(\widehat{n}_{x}E_{t}(x,yg)-\widehat{n}_{y}E_{t}(x,yg)\right)\big{|}_{x=y}dV_{P_{0}}(x) (5)
=Tr𝒱(eitDZF)+m=1μ(m,)Tr(κm(g))eitλm,.\displaystyle=\operatorname{Tr}_{\mathcal{V}}(e^{-itD_{Z}}\circ F)+\sum_{m=1}^{\infty}\sum_{\ell\in\mathbb{Z}}\mu(m,\ell)\operatorname{Tr}(\kappa_{m}(g))e^{-it\lambda_{m,\ell}}. (6)
Proof.

The basic facts concerning push-forward distributions such as Π\Pi_{*} can be found in section 7.1 of [23] and we omit the proofs here as they are well known.

Let’s now derive the above explicit expression 6 for 𝒦(g,t)\mathcal{K}(g,t). Indeed, by the computation in Theorem 4.1 of [22], 𝒦(g,t)\mathcal{K}(g,t) is the equivariant trace of the operator eitDZFe^{-itD_{Z}}\circ F on kerω\ker\Box_{\omega}. Recalling that μ(m,)\mu(m,\ell) is simply the multiplicity of κm\kappa_{m} in the λm,\lambda_{m,\ell}-eigenspace and that FF acts by κm\kappa_{m} on this eigenspace by definition of m\mathcal{H}_{m} we obtain our above expression 6 for 𝒦(g,t)\mathcal{K}(g,t), as desired. ∎

We will now build a distribution on P0×G××S1P_{0}\times G\times\mathbb{R}\times S^{1} which we will then compose with ιdiag\iota^{*}\operatorname{diag}^{*}\mathcal{F} to produce Υ(φ)\Upsilon(\varphi). A key motivating fact in the below definition is the orthogonality of the functions gTr(κm(g))g\mapsto\operatorname{Tr}(\kappa_{m}(g)) for different mm’s. This is a well-known fact from abstract harmonic analysis (see Section 5.3 of [7], for example), however one should take care not to confuse the two distinct notions of “character” of a representation.

Lemma 4.14.

[11, 10]
The operator 𝒪:C(G)𝒟(S1)\mathcal{L}_{\mathcal{O}}:C^{\infty}(G)\to\mathcal{D}^{\prime}(S^{1}) with Schwartz kernel given by the distribution

(eiθ,g):=m=1Tr(κm(g))eimθ\mathcal{L}(e^{i\theta},g):=\sum_{m=1}^{\infty}\operatorname{Tr}(\kappa_{m}(g))e^{im\theta}

is in

𝒪I(1d)/4(S1×G;Λ𝒪)\mathcal{L}_{\mathcal{O}}\in I^{(1-d)/4}(S^{1}\times G;\Lambda_{\mathcal{O}})

with Lagrangian

Λ𝒪={(z,r;g,rξ)(TS10)×(TG0)|ξ𝒪,gGξ,z=χξ(g)}\Lambda_{\mathcal{O}}=\{(z,r;g,r\xi)\in(T^{*}S^{1}\setminus 0)\times(T^{*}G\setminus 0)\ |\ \xi\in\mathcal{O},\ g\in G_{\xi},\ z=\chi_{\xi}(g)\}

where χξ:GξU(1)\chi_{\xi}:G_{\xi}\to U(1) is the character associated to ξ𝒪\xi\in\mathcal{O}.

Our final to-do before we have, at least morally, obtained a description of Υ(φ)\Upsilon(\varphi) as a composition of FIOs is to localize about the ray λm,mE\lambda_{m,\ell}\sim mE via φ\varphi. Towards this end, we define an operator

Tφ,E:Cc(S1×)𝒟(S1)T_{\varphi,E}:C^{\infty}_{c}(S^{1}\times\mathbb{R})\to\mathcal{D}^{\prime}(S^{1})

by declaring its Schwartz kernel to be given by the oscillatory integral

Tφ,E(θ;θ,t):=(2π)2φ^(t)𝑑seis(θθtE).T_{\varphi,E}(\theta^{\prime};\theta,t):=(2\pi)^{-2}\widehat{\varphi}(t)\int_{-\infty}^{\infty}ds\ e^{is(\theta^{\prime}-\theta-tE)}.
Lemma 4.15.

[11]
Tφ,EI1/4(S1×S1×;ΛE)T_{\varphi,E}\in I^{-1/4}(S^{1}\times S^{1}\times\mathbb{R};\Lambda_{E}) where

ΛE:={(zeitE,r;z,r;t,rE)(TS10)×(TS10)×(T0)|r,zS1}.\Lambda_{E}:=\{(ze^{itE},r;\ z,r;\ t,rE)\in(T^{*}S^{1}\setminus 0)\times(T^{*}S^{1}\setminus 0)\times(T^{*}\mathbb{R}\setminus 0)\ |\ r\in\mathbb{R},\ z\in S^{1}\}.
Lemma 4.16.

We can form the composition

Tφ,E(𝒪id)Id/4(S1×G×;Θφ,E)T_{\varphi,E}\circ(\mathcal{L}_{\mathcal{O}}\otimes\operatorname{id}_{\mathbb{R}})\in I^{-d/4}(S^{1}\times G\times\mathbb{R};\Theta^{\prime}_{\varphi,E})

where

Θφ,E={(χξ(g)eitE,r;g,rξ;t,Er)TS1×TG×T:ξ𝒪,gGξ}.\Theta^{\prime}_{\varphi,E}=\{(\chi_{\xi}(g)e^{itE},r;g,r\xi;t,Er)\in T^{*}S^{1}\times T^{*}G\times T^{*}\mathbb{R}\ :\ \xi\in\mathcal{O},\ g\in G_{\xi}\}.

Furthermore:

(Tφ,E𝒪id)𝒦=Υ(φ).(T_{\varphi,E}\circ\mathcal{L}_{\mathcal{O}}\otimes\operatorname{id}_{\mathbb{R}})\mathcal{K}=\Upsilon(\varphi).
Proof.

The fact that this composition Tφ,E(𝒪id)T_{\varphi,E}\circ(\mathcal{L}_{\mathcal{O}}\otimes\operatorname{id}_{\mathbb{R}}) can be formed, has the above order, and the above canonical relation Θφ,E\Theta^{\prime}_{\varphi,E} is proven in [11]. So, we just need to demonstrate that we do indeed obtain Υ(φ)\Upsilon(\varphi) when applying it to 𝒦\mathcal{K}. Recalling the formula 6 for 𝒦(g,t)\mathcal{K}(g,t) we note that by [22] the trace over 𝒱\mathcal{V} still decomposes as a sum over (possibly generalized) eigenvalues of DZD_{Z} counted with multiplicity, only now not all are real and some may be zero modes. Furthermore, the GG-dependence in the trace over 𝒱\mathcal{V} is still in the form of the Tr(κ(g))\operatorname{Tr}(\kappa(g)) for κ\kappa the representation generated by that specific (generalized) eigenvector. Indeed, while the Hilbert space inner products from the Cauchy data isomorphism are DZD_{Z}-invariant they are still GG-invariant and so 𝒱\mathcal{V} is completely decomposable since it is a unitary GG-representation. Since characters are orthogonal with respect to the Haar measure on GG, we obtain:

(m0𝒪id)𝒦=m=1eitλm,eimθ(\mathcal{L}_{m_{0}\mathcal{O}}\otimes\operatorname{id}_{\mathbb{R}})\mathcal{K}=\sum_{m=1}^{\infty}\sum_{\ell\in\mathbb{Z}}e^{-it\lambda_{m,\ell}}e^{im\theta}

as a distribution on S1×S^{1}\times\mathbb{R}. Finally, applying Tφ,ET_{\varphi,E} we immediately obtain:

(Tφ,E(m0𝒪id))𝒦=m=1φ(λm,mE)eimθ(T_{\varphi,E}\circ(\mathcal{L}_{m_{0}\mathcal{O}}\operatorname{id}_{\mathbb{R}}))\mathcal{K}=\sum_{m=1}^{\infty}\sum_{\ell\in\mathbb{Z}}\varphi(\lambda_{m,\ell}-mE)e^{im\theta}

as desired. ∎

Our next step is to understand the composition (Tφ,E(𝒪id))𝒦(T_{\varphi,E}\circ(\mathcal{L}_{\mathcal{O}}\otimes\operatorname{id}_{\mathbb{R}}))\mathcal{K} as an actual Lagrangian distribution. As it turns out, it is more clear if one first computes (Tφ,E(𝒪id))Π(T_{\varphi,E}\circ(\mathcal{L}_{\mathcal{O}}\otimes\operatorname{id}_{\mathbb{R}}))\circ\Pi_{*}.

Lemma 4.17.

The composition

(Tφ,E(𝒪id))ΠI12n4(S1×P0×G×;C6)(T_{\varphi,E}\circ(\mathcal{L}_{\mathcal{O}}\otimes\operatorname{id}_{\mathbb{R}}))\circ\Pi_{*}\in I^{\frac{1}{2}-\frac{n}{4}}(S^{1}\times P_{0}\times G\times\mathbb{R};C_{6}^{\prime})

is a transverse composition of FIOs with Lagrangian determined by

C6\displaystyle C_{6} ={(χη(g)eitE,r;x,0;g,rη;t,rE)(TS10)×TP0×(TG0)×(T0)\displaystyle=\{(\chi_{\eta}(g)e^{itE},r;\ x,0;\ g,r\eta;\ t,rE)\in(T^{*}S^{1}\setminus 0)\times T^{*}P_{0}\times(T^{*}G\setminus 0)\times(T^{*}\mathbb{R}\setminus 0)
:η𝒪,gGη}.\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ :\ \eta\in\mathcal{O},\ g\in G_{\eta}\}.
Proof.

This is immediate since this composition does not affect the TP0T^{*}P_{0}-variables and in the (TG0)×(T0)(T^{*}G\setminus 0)\times(T^{*}\mathbb{R}\setminus 0)-variables the Lagrangian for Π\Pi_{*} is just the diagonal. ∎

Theorem 4.18.

The clean intersection hypothesis implies that the composition of Tφ,E(𝒪id)ΠT_{\varphi,E}\circ(\mathcal{L}_{\mathcal{O}}\otimes\operatorname{id}_{\mathbb{R}})\circ\Pi_{*} and ιdiag\iota^{*}\operatorname{diag}^{*}\mathcal{F} is a clean composition of FIOs with excess

e=2(n+)2 and therefore order (12n4)+n14+e2=n+1+14e=2(n+\ell)-2\ \mbox{ and therefore order }\ \left(\frac{1}{2}-\frac{n}{4}\right)+\frac{n-1}{4}+\frac{e}{2}=n+\ell-1+\frac{1}{4}

Thus

Υ(φ)In+1+14(S1;E)\Upsilon(\varphi)\in I^{n+\ell-1+\frac{1}{4}}(S^{1};\mathfrak{C}_{E}^{\prime})

where

E\displaystyle\mathfrak{C}_{E}^{\prime} ={(χη(g)eitE,r)TS10:η𝒪,gGη,ζT0P|P0 such that\displaystyle=\{(\chi_{\eta}(g)e^{itE},r)\in T^{*}S^{1}\setminus 0\ :\ \eta\in\mathcal{O},\ g\in G_{\eta},\ \exists\zeta\in T^{*}_{0}P|_{P_{0}}\mbox{ such that}
s with ζ=(GsΦtZ)(ζ)g,Zω,ζ=rE,μ(ζg1)=rη}\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \exists s\mbox{ with }\zeta=(G_{-s}\circ\Phi^{Z}_{t})(\zeta)g,\ \langle Z^{\omega},\zeta\rangle=-rE,\ \mu(\zeta g^{-1})=r\eta\}
Proof.

The main goal here is to compute the fiber over a point (ω,r)TODO(\omega,r)\in TODO. By homogeneity we can assume r=1r=1 and so the fiber is given by:

F(ω,1):={(x,0;\displaystyle F_{(\omega,1)}:=\{(x,0;\ g,η;t,E)TP0×(TG0)×(T0):xP0,ζ(T0P)x\displaystyle g,\eta;\ t,E)\in T^{*}P_{0}\times(T^{*}G\setminus 0)\times(T^{*}\mathbb{R}\setminus 0)\ :\ x\in P_{0},\ \exists\zeta\in(T^{*}_{0}P)_{x}
such thats with ζ=(GsΦtZ)(ζ)g,Zω,ζ=E,μ(ζg1)=η,\displaystyle\mbox{ such that}\exists s\mbox{ with }\zeta=(G_{-s}\circ\Phi_{t}^{Z})(\zeta)g,\ \langle Z^{\omega},\zeta\rangle=-E,\ \mu(\zeta g^{-1})=\eta,
χη(g)eitE=ω, and η𝒪,gGη}\displaystyle\chi_{\eta}(g)e^{itE}=\omega,\mbox{ and }\eta\in\mathcal{O},\ g\in G_{\eta}\}

Since we chose E>0E>0 our constraint Zω,ζ=E\langle Z^{\omega},\zeta\rangle=-E implies ζT+P|P0T0P|P0\zeta\in T^{*}_{+}P|_{P_{0}}\subseteq T^{*}_{0}P|_{P_{0}} and therefore ζ\zeta corresponds to a unique null geodesic γ𝒩\gamma\in\mathcal{N} with γ(0)=x\gamma(0)=x, HZ(γ)=EH_{Z}(\gamma)=E, μ(γg1)=η\mu(\gamma g^{-1})=\eta and γ=ΦtZ(γ)g\gamma=\Phi_{t}^{Z}(\gamma)g. Therefore (γg1,η)μ𝒪1(0)(\gamma g^{-1},\eta)\in\mu_{\mathcal{O}}^{-1}(0) and the image of this in the quotient is a periodic orbit in 𝒩𝒪\mathcal{N}_{\mathcal{O}} with period tt and energy H~Z=E\widetilde{H}_{Z}=E.

Now, let’s write π:×μ𝒪1(0)×𝒩𝒪\pi:\mathbb{R}\times\mu_{\mathcal{O}}^{-1}(0)\to\mathbb{R}\times\mathcal{N}_{\mathcal{O}} for the projection map and recall that 𝔜E×𝒩𝒪\mathfrak{Y}_{E}\subseteq\mathbb{R}\times\mathcal{N}_{\mathcal{O}} is the set of periodic orbits for the reduced flow together with their periods. If we denote

𝔛:={(t,γ,η,g):(t,γ,η)π1(𝔜E) and gGη}\mathfrak{X}:=\{(t,\gamma,\eta,g)\ :\ (t,\gamma,\eta)\in\pi^{-1}(\mathfrak{Y}_{E})\ \mbox{ and }\ g\in G_{\eta}\}

then dim𝔛=dimπ1(𝔜E)+dimGdim𝒪=2d+2n1\dim\mathfrak{X}=\dim\pi^{-1}(\mathfrak{Y}_{E})+\dim G-\dim\mathcal{O}=2d+2n-1 since 𝔜E\mathfrak{Y}_{E} has dimension 2(n+)12(n+\ell)-1 where 2=dim𝒪2\ell=\dim\mathcal{O}. Note: the clean intersection hypothesis implies that 𝔜E\mathfrak{Y}_{E} is a disjoint union of smooth manifolds with the clopen subset {0}×H~Z1(E)\{0\}\times\widetilde{H}_{Z}^{-1}(E) having dimension =2(n+)1=2(n+\ell)-1 and the other components having dimension at most 2(n+)12(n+\ell)-1. Furthermore the map

𝔛\displaystyle\mathfrak{X} F(ω,1)\displaystyle\to F_{(\omega,1)}
(t,γ,η,g)\displaystyle(t,\gamma,\eta,g) (γ(0)g,g,η,t)\displaystyle\mapsto(\gamma(0)g,g,\eta,t)

is a submersion. This, together with the fact that the holonomy map Hol:𝔜EU(1)\operatorname{Hol}:\mathfrak{Y}_{E}\to\operatorname{U}(1) is locally constant, implies that we have a clean composition of FIOs. Since the only part of the derivative γ(0)\gamma^{\prime}(0) of γ\gamma captured in the image of our submersion 𝔛F(ω,1)\mathfrak{X}\to F_{(\omega,1)} is η=μ(γ)\eta=\mu(\gamma) it follows that the kernel of the above submersion at each point contains a 2d22d-2\ell-dimension subspace of tangent vectors orthogonal to the tangent space Tη𝒪T_{\eta}\mathcal{O}. The only other degeneracy comes the 1-dimensional space of vectors tangent to the curve γ\gamma itself and so we arrive at:

dimF(ω,1)=2(n+d)12(d)1=2(n+)2\dim F_{(\omega,1)}=2(n+d)-1-2(d-\ell)-1=2(n+\ell)-2

as desired. ∎

All that remains now is the calculation of the principal symbol of Υ(φ)\Upsilon(\varphi). It’s worth noticing, however, that from the expression for Tφ,ET_{\varphi,E} we see that the actually wave front set WF(Υ(φ))\operatorname{WF}^{\prime}(\Upsilon(\varphi)) will often be a proper subset of E\mathfrak{C}_{E}^{\prime} depending on suppφ^\operatorname{supp}\widehat{\varphi}. This is due to the varying dimensions of the components of 𝔜E\mathfrak{Y}_{E} and the support of the principal symbol of Υ(φ)\Upsilon(\varphi) being constrained by suppφ^\operatorname{supp}\widehat{\varphi}. We compute this principal symbol now.

Theorem 4.19.

We have

WF(Υ(φ))\displaystyle\operatorname{WF}^{\prime}(\Upsilon(\varphi)) {(ω,r)S1×>0:(T,γ)𝔜E\displaystyle\subseteq\{(\omega,r)\in S^{1}\times\mathbb{R}_{>0}\ :\ \exists(T,\gamma)\in\mathfrak{Y}_{E}
with Tsuppφ^ such that Hol𝒪([0,T]tΦ~tZ(γ))=ω}\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \mbox{with }T\in\operatorname{supp}\widehat{\varphi}\mbox{ such that }\operatorname{Hol}_{\mathcal{O}}([0,T]\ni t\mapsto\widetilde{\Phi}_{t}^{Z}(\gamma))=\omega\}

and, under the clean intersection hypothesis, the singularities of Υ(φ)\Upsilon(\varphi) are all classical. Assuming φ^(0)0\widehat{\varphi}(0)\neq 0 the principal symbol at (ω,r)WF(Υ(φ))(\omega,r)\in\operatorname{WF}^{\prime}(\Upsilon(\varphi)) is given by:

Cn,dωmφ^(0)Vol(H~Z1(E))|r|n+1|dωdr|1/2.C_{n,d}\omega^{m}\widehat{\varphi}(0)\operatorname{Vol}\left(\widetilde{H}_{Z}^{-1}(E)\right)|r|^{n+\ell-1}|d\omega\wedge dr|^{1/2}.

Here we omit the Maslov factor since in this case it can be invariantly taken to be constant on WF(Υ(φ))\operatorname{WF}^{\prime}(\Upsilon(\varphi)).

Proof.

The result concerning the wave front set will follow immediately from the calculation of the principal symbol since the constraint Tsuppφ^T\in\operatorname{supp}\widehat{\varphi} comes from the support of the principal symbol.

We can compute a principal symbol for Tφ,E(𝒪id)T_{\varphi,E}\circ(\mathcal{L}_{\mathcal{O}}\otimes\operatorname{id}_{\mathbb{R}}) by composing their explicitly given Schwartz kernels. The Schwartz kernel for the composition is then a distribution on S1×G×S^{1}\times G\times\mathbb{R} with Schwartz kernel

(θ,g,t)m=1(2π)2φ^(t)eim(θtE)Tr(κm(g)).(\theta^{\prime},g,t)\mapsto\sum_{m=1}^{\infty}(2\pi)^{-2}\widehat{\varphi}(t)e^{im(\theta^{\prime}-tE)}\operatorname{Tr}(\kappa_{m}(g)).

Recalling that the principal symbol of \mathcal{F} is given by

±12(2π)(d+3)/4n^,|dC1|1/2|dt1|1/2|dg|1/2|dt2|1/2\pm\frac{1}{2}(2\pi)^{(d+3)/4}\langle\widehat{n},-\rangle|d_{C_{1}}|^{1/2}\otimes|dt_{1}|^{1/2}\otimes|dg|^{1/2}\otimes|dt_{2}|^{1/2}

we have that the principal symbol of Υ(φ)\Upsilon(\varphi) at a point (ω,1)E(\omega,1)\in\mathfrak{C}_{E}^{\prime} is given by the integral over the fiber

F(ω,1) quotient of holonomy ω clopen subset of 𝔜E by the action of the flow Φ~ZF_{(\omega,1)}\cong\mbox{ quotient of holonomy }\omega\mbox{ clopen subset of }\mathfrak{Y}_{E}\mbox{ by the action of the flow }\widetilde{\Phi}^{Z}

of the product of the symbol of \mathcal{F} and the symbol of

Tφ,E(𝒪id)ΠιdiagT_{\varphi,E}\circ(\mathcal{L}_{\mathcal{O}}\otimes\operatorname{id}_{\mathbb{R}})\circ\Pi_{*}\circ\iota^{*}\circ\operatorname{diag}^{*}

restricted to the fiber. The symbol over a more general point (ω,r)E(\omega,r)\in\mathfrak{C}_{E}^{\prime} is then obtained by homogeneity in rr. Since 0suppφ^0\in\operatorname{supp}\widehat{\varphi} and principal symbols are defined modulo symbols of lower order it suffices to compute this integral over the quotient of the clopen subset

{0}×H~Z1(E)𝔜E.\{0\}\times\widetilde{H}_{Z}^{-1}(E)\subseteq\mathfrak{Y}_{E}.

In this fibered product of symbols, the pairing of the gg in the symbol for Tφ,E(𝒪id)T_{\varphi,E}\circ(\mathcal{L}_{\mathcal{O}}\otimes\operatorname{id}_{\mathbb{R}}) and the gg in the symbol for \mathcal{F} amounts to replacing variables in the fibers of T0P|P0𝒩T^{*}_{0}P|_{P_{0}}\cong\mathcal{N} with fiber variables in 𝒩𝒪\mathcal{N}_{\mathcal{O}} (here the “fibers” are diffeomorphic to 𝒪\mathcal{O}). Since our fibered product is just over the quotient of {0}×H~Z\{0\}\times\widetilde{H}_{Z} by the flow, the pairing of the tt-variables in the symbol for Tφ,E(𝒪)T_{\varphi,E}\circ(\mathcal{L}_{\mathcal{O}}\otimes\it_{\mathbb{R}}) and the symbol for \mathcal{F} simply amounts to setting t=0t=0 in both symbols and multiplying by φ^(0)\widehat{\varphi}(0). The effect of restricting to P0P_{0} along the diagonal P0P×PP_{0}\hookrightarrow P\times P on the fibered product of symbols (aside from replacing the volume half-density on Γ\Gamma with the one on H~Z1(E)\widetilde{H}_{Z}^{-1}(E)) is to divide by the function n^,\langle\widehat{n},-\rangle and multiply by a dimensional constant, hence removing the function n^,\langle\widehat{n},-\rangle from our symbol expression. Therefore, denoting by Cn,dC_{n,d} a dimensional constant and writing ω=eiθ\omega=e^{i\theta^{\prime}}, we obtain the symbol over the point (ω,1)(\omega,1) as:

Cn,dωmH~Z1(E)\displaystyle C_{n,d}\omega^{m}\int_{\widetilde{H}^{-1}_{Z}(E)} φ^(0)|dωdr|1/21|H~Z|2H~ZdV𝒩𝒪\displaystyle\widehat{\varphi}(0)|d\omega\wedge dr|^{1/2}\frac{1}{|\nabla\widetilde{H}_{Z}|^{2}}\nabla\widetilde{H}_{Z}\llcorner dV_{\mathcal{N}_{\mathcal{O}}}
=Cn,dωmφ^(0)Vol(H~Z1(E))|dωdr|1/2.\displaystyle=C_{n,d}\omega^{m}\widehat{\varphi}(0)\operatorname{Vol}\left(\widetilde{H}^{-1}_{Z}(E)\right)|d\omega\wedge dr|^{1/2}.

We can now recover the principal symbol over (ω,r)(\omega,r) by scaling. Since the fibers are diffeomorphic to H~Z1(E)/\widetilde{H}_{Z}^{-1}(E)/\mathbb{R} where the \mathbb{R}-action is by the Hamiltonian flow of H~Z\widetilde{H}_{Z} they have dimension 2(n+)22(n+\ell)-2 and so the principal symbol over (ω,r)(\omega,r) is given by:

Cn,dωmφ^(0)Vol(H~Z1(E))|r|(n+)1|dωdr|1/2C_{n,d}\omega^{m}\widehat{\varphi}(0)\operatorname{Vol}\left(\widetilde{H}_{Z}^{-1}(E)\right)|r|^{(n+\ell)-1}|d\omega\wedge dr|^{1/2}

where we note that n+1n+\ell-1 is half the dimension of our fiber. ∎

Theorem 4.20.

Under the assumptions of 2.31 where the time 0\neq 0 part of the set 𝔜E\mathfrak{Y}_{E} consists of finitely many isolated periodic orbits (T1,γ1),.,(Tq,γq)(T_{1},\gamma_{1}),....,(T_{q},\gamma_{q}), and assuming 0suppφ^0\notin\operatorname{supp}\widehat{\varphi} we actually have Υ(φ)I1/4\Upsilon(\varphi)\in I^{1/4} with principal symbol at each (Hol𝒪(Tj,γj),r)(\operatorname{Hol}_{\mathcal{O}}(T_{j},\gamma_{j}),r), j=1,,qj=1,...,q, given by:

Cn,dHol𝒪(Tj,γj)mTj#2πφ^(Tj)|det(IPj)|1/2eiπ𝔪j/4|dωdr|1/2C_{n,d}\operatorname{Hol}_{\mathcal{O}}(T_{j},\gamma_{j})^{m}\frac{T_{j}^{\#}}{2\pi}\widehat{\varphi}(T_{j})|\det(I-P_{j})|^{-1/2}e^{i\pi\mathfrak{m}_{j}/4}|d\omega\wedge dr|^{1/2}

where Tj#T_{j}^{\#} is the primitive period of γj\gamma_{j}, PjP_{j} is the linearized Poincaré first return map of γj\gamma_{j} with respect to any local symplectic transversal and we have included the Maslov factor eiπ𝔪j/4e^{i\pi\mathfrak{m}_{j}/4} where 𝔪j\mathfrak{m}_{j} is the Conley-Zehnder index of γj\gamma_{j} as in [22].

Proof.

The proof is exactly the same as the previous one only instead of integrating over {0}×H~Z1(E)\{0\}\times\widetilde{H}^{-1}_{Z}(E) with respect to its invariant measure we integrate over the respective periodic orbit γj\gamma_{j} with respect to the density from 2.31. ∎

These last three theorems respectively conclude the proofs of Theorems 1.1,1.2 and 1.3.

References

  • [1] Michael T. Anderson. On stationary vacuum solutions to the einstein equations. Ann. Henri Poincaré, 1(5):977–994, 2000.
  • [2] Christian Bär, Nicolas Ginoux, and Frank Pfäffle. Wave equations on Lorentzian manifolds and quantization, volume 3. European Mathematical Society, 2007.
  • [3] David Bleecker. Gauge theory and variational principles. Courier Corporation, 2005.
  • [4] L. Boutet de Monvel and V. Guillemin. The spectral theory of Toeplitz operators, volume 99 of Annals of Mathematics Studies. Princeton University Press, Princeton, NJ; University of Tokyo Press, Tokyo, 1981.
  • [5] Julien Cortier and Vincent Minerbe. On complete stationary vacuum initial data. J. Geom. Phys., 99:20–27, 2016.
  • [6] Michael A Dritschel and James Rovnyak. Operators on indefinite inner product spaces. Lectures on operator theory and its applications (Waterloo, ON, 1994), 3:141–232, 1996.
  • [7] Gerald B Folland. A course in abstract harmonic analysis, volume 29. CRC press, 2016.
  • [8] Frederick Greenleaf and Martin Moskowitz. Cyclic vectors for representations of locally compact groups. Mathematische Annalen, 190(4):265–288, 1971.
  • [9] Victor Guillemin and Shlomo Sternberg. Symplectic techniques in physics. Cambridge University Press, Cambridge, second edition, 1990.
  • [10] Victor Guillemin and Alejandro Uribe. Circular symmetry and the trace formula. Inventiones mathematicae, 96(2):385–423, 1989.
  • [11] Victor Guillemin and Alejandro Uribe. Reduction and the trace formula. Journal of Differential Geometry, 32(2):315–347, 1990.
  • [12] H Hogreve, J Potthoff, and R Schrader. Classical limits for quantum particles in external yang-mills potentials. Communications in mathematical physics, 91(4):573–598, 1983.
  • [13] Lars Hörmander. The analysis of linear partial differential operators IV: Fourier integral operators, volume 17. Springer, 2009.
  • [14] Onirban Islam. A gutzwiller trace formula for dirac operators on a stationary spacetime. arXiv preprint arXiv:2109.09219, 2021.
  • [15] Miguel Angel Javaloyes and Miguel Sánchez. A note on the existence of standard splittings for conformally stationary spacetimes. Classical and Quantum Gravity, 25(16):168001, 2008.
  • [16] Heinz Langer. Spectral functions of definitizable operators in Kreĭn spaces. In Functional analysis (Dubrovnik, 1981), volume 948 of Lecture Notes in Math., pages 1–46. Springer, Berlin-New York, 1982.
  • [17] Kenneth R. Meyer and Daniel C. Offin. Introduction to Hamiltonian dynamical systems and the N-body problem, volume 90 of Applied Mathematical Sciences. Springer, Cham, third edition, 2017.
  • [18] Barret O‘Neil. Semi-riemannian geometry, 1983.
  • [19] M. Sánchez. Some remarks on causality theory and variational methods in Lorenzian manifolds. Conf. Semin. Mat. Univ. Bari, (265):ii+12, 1997.
  • [20] Robert Schrader and Michael E Taylor. Small \hbar asymptotics for quantum partition functions associated to particles in external yang-mills potentials. Communications in mathematical physics, 92(4):555–594, 1984.
  • [21] Shlomo Sternberg. Minimal coupling and the symplectic mechanics of a classical particle in the presence of a yang-mills field. Proceedings of the National Academy of Sciences, 74(12):5253–5254, 1977.
  • [22] Alexander Strohmaier and Steve Zelditch. A Gutzwiller trace formula for stationary space-times. Adv. Math., 376:Paper No. 107434, 53, 2021.
  • [23] Alexander Strohmaier and Steve Zelditch. Semi-classical mass asymptotics on stationary spacetimes. Indag. Math. (N.S.), 32(1):323–363, 2021.
  • [24] Alan Weinstein. A universal phase space for particles in yang-mills fields. Letters in Mathematical Physics, 2(5):417–420, 1978.
  • [25] SK Wong. Field and particle equations for the classical yang-mills field and particles with isotopic spin. Il Nuovo Cimento A (1965-1970), 65(4):689–694, 1970.