A Trace Formula on Stationary Kaluza-Klein Spacetimes
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 -bundle with connection over a Lorentzian manifold one can form the Kaluza-Klein metric on :
where is the projection map. In the special case where is globally hyperbolic, stationary and spatially compact we will see in Section 2 that the Kaluza-Klein spacetime takes the form
with metric given by
where is a Riemannian metric on , a smooth function, and a 1-form on satisfying pointwise. Given an integral coadjoint orbit for we have irreducible unitary representations corresponding to for each 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:
As is shown in Section 3.1, for sufficiently large the operator
is self adjoint on with respect to the energy form with a discrete set of eigenvalues
accumulating at with at worst polynomial growth. For fixed, our goal is to study the following question.
Question:
What are the asymptotics of the frequency spectrum of on ?
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 via the tempered distribution on given by:
and we will study its asymptotics via the periodic generating function
The fact that is tempered and is defined on all of 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:
-
•
is the symplectic manifold of affinely parametrized inextendible future-directed null geodesics on modulo the action of translation in the affine parameter,
-
•
is the symplectic reduction of along the coadjoint orbit as in [11],
-
•
and are respectively the Hamiltonian and Hamiltonian flow corresponding to the reduction of the flow on induced by the flow of on .
The clean intersection hypothesis then states that is a regular value for and the fibered product of the flow map
with the diagonal map 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 is contained in:
where this holonomy is taken with respect to a natural -bundle with connection over defined in 2.35.
Theorem 1.2.
Let . Under the clean intersection hypothesis, is a union of the positive parts of finitely many fibers of , and where . Furthermore, we obtain an asymptotic expansion as :
with each a distribution in , bounded in for fixed, and
where by 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 and there existed only finitely many non-degenerate periodic orbits with each . Then we actually obtain a better asymptotic expansion as :
and is of the form:
Where is the minimum positive value of such that , is the linearized Poincaré first return map of , and 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 for vector bundles. Let’s describe this now.
Let denote the Hermitian vector bundle associated to the representation equipped with the covariant derivative induced from on and the representation. We then have a massive Klein-Gordon operator with mass given by the eigenvalue
of the quadratic Casimir on the representation corresponding to . Here is the highest weight for and is the sum of the positive roots. This massive Klein-Gordon operator acts on sections and its kernel is invariant under the operator given by times the covariant derivative in the direction.
Corollary 1.4.
For sufficiently large operator on equipped with the energy form from Section 3.1 is self adjoint with discrete spectrum equal to the spectrum of on , and the multiplicity of is the product of the multiplicity of and the dimension of the irreducible representation corresponding to . Thus if is defined for in the same way is defined for then under the clean intersection hypothesis we have
In general, one can compute the values of and in terms of the dominant integral element . Indeed, if is the set of positive roots then
and as a consequence of the Weyl character formula we have
In particular we see that is a polynomial of degree and so our leading order asymptotics for as are . This is in agreement with [23] where and for all . When and corresponds to the vector representation then and .
We now provide an outline of the paper. In Section 2 we demonstrate that 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 replaced by . In Section 2.1 we introduce the reduced phase space from [11],[24],[21] in the special case where the symplectic manifold is and in Section 2.2 we study two aspects of periodic orbits on this reduced phase space: the linearized Poincaré first return map and the holonomy map . In Section 3 we use the general setup of [22] to discuss the wave equation on . Since we allow for a potential (as long as it is constant along the fibers of and independent of ) the energy quadratic form on the space of solutions 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 for sufficiently large. In Section 3.1 we apply a result from [14] to obtain that 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 be a connected, globally hyperbolic, oriented, time-oriented Lorentzian manifold. For us, “Lorentzian” will mean that has signature . 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.
We will assume that admits a complete timelike Killing vector field , flowing forwards in time with respect to the time-orientation (and we will make a fixed choice of such a ).
-
2.
We will assume that is spatially compact. That is, for some (hence any) choice of Cauchy hypersurface , is compact.
Lemma 2.1.
[15]
All such spacetimes as described above are diffeomorphic to with metric of the form:
where is smooth, a 1-form on , a vector field on , a Riemannian metric on , and a Cauchy hypersurface. We also have pointwise and is the vector field -dual to . Furthermore, will be a Cauchy hypersurface for each .
Notice that such spacetimes are not necessarily static since need not be normal to . Indeed, is the unit normal. Let’s denote this by
noticing that this is a globally defined vector field on and is the unit normal to each 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
where 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 other than . 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 be a compact Lie group and a principal -bundle.
Definition 2.2.
Recalling that is a positive definite -invariant inner product on , given any connection on we obtain an induced Kaluza-Klein metric:
on the total space of . This is again a Lorentzian metric of signature . Furthermore, we endow with the orientation induced from and , and the time-orientation induced from . We let
We also abuse notation and set
Lemma 2.3.
The horizontal lift of is a complete timelike future-oriented Killing vector field for and:
Proof.
Recall that the connection defines a horizontal bundle splitting and that is an isomorphism. The horizontal lift of is then given by
where is the pullback of to a section of . Furthermore, the flow of is the horizontal lift of the flow of on ([3] section 10.1). This immediately implies is complete since is. Furthermore since is Killing on it follows that is invariant under the flow of and since is horizontal we have
Thus indeed is Killing on . Finally, from [3] section 2.2 we know that is invariant under push-forward along the -action on and therefore for all . ∎
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 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]
is globally hyperbolic and each , , a Cauchy hypersurface.
Proof.
The map induced by flowing along is a global diffeomorphism and so we can, without loss of generality, assume as a manifold with and . Using our standard form for the metric on we can write the Kaluza-Klein metric on as:
(1) |
In particular, the Riemannian metric on induced by pulling back is independent of :
Choose now an arbitrary inextendible causal curve in , parametrized with respect to so that . We can always make this parametrization for causal curves in the spacetimes we are considering thanks to the global time function . We now write for the domain of with and . Since is causal we have at all that:
Thus by Cauchy-Schwarz we have:
Rearranging yields:
Now, suppose for contradiction that (the case is completely analogous) and set:
We have since and are compact, hence and are compact. We then have:
and so the curve must be extendible beyond time , 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 denote our coordinates so that the metric on is . Consider the curve given by and for and for . There exists a time so that is a causal curve for but there exists no on which extends to a causal curve on . So has no causal extensions, but it is not inextendible since it does admit an extension to a curve (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 derived in the above proof, we also obtain the following.
Corollary 2.5.
The horizontal lift of the unit normal to the Cauchy hypersurfaces is the unit normal to the Cauchy hypersurfaces with respect to .
Let’s now discuss the geodesic equations in the Kaluza-Klein spacetime . 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 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
and so
This allows us to split geodesics into three types.
Definition 2.6.
We call a geodesic on lightlike (respectively spacelike and null) if and only if the constant is negative (respectively positive and zero).
As the below lemma explains, despite Wong’s equations describing massive particles in , we will be interested in null geodesics in .
Lemma 2.7.
Let be a geodesic in . Then the value is constant. Thus
being constant implies that the projected curve in has constant. Since (and is zero if and only if ) we see that:
and so timelike or null geodesics in project to timelike or null curves in . In fact, the projection will be timelike unless the geodesic in is null and .
Proof.
The only part of this not proven in the statement is that is constant for a geodesic in . This can be found in [3] theorem 10.1.5. ∎
Lemma 2.8.
In the special case where is flat and , null geodesics in project to solutions to Wong’s equations with a non-positive constant. More generally, null geodesics in project to curves in together with a section of satisfying:
where is the connection on the bundle induced by .
Proof.
Let be a null geodesic in and the projected curve in . We denote and notice that by -equivarance of connection 1-forms on principle bundles this defines a section of . We’ve already seen that is constant so it suffices to prove that is covariantly constant with respect to and that the geodesic equations reduce to Wong’s equations.
The fact that the geodesic equations in reduce to Wong’s equations on is theorem 10.1.6 in [3]. As for , we note that its covariant derivative as a section of is just the horizontal part of its time derivative as a -valued function on a curve in , and this is zero since the entire time derivative vanishes. ∎
Given a null geodesic in , we would like to think of the constant as the “charge”. Unfortunately, unlike the abelian case of the Lorentz force law, different lifts of solutions to Wong’s equations in to geodesics in will have different charges. Indeed, if the two lifts of our curve in are related by the right action of on then the charges of the two lifts will be related by . Identifying the charge with we arrive at the following gauge invariant definition of charge.
Definition 2.9.
Let be a null geodesic in and . The charge of is defined to be the coadjoint orbit:
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 is the Hamiltonian flow on of the Hamiltonian .
Lemma 2.11.
Let and respectively denote the flows on given by the derivatives of the flows of and () on . Then and commute with for every .
Proof.
Since and are derivatives of flows on they are a 1-parameter family of canonical transformations on and therefore, by the Hamiltonian version of Noether’s theorem, it suffices to show that the Hamiltonian is invariant under the flows in order to prove that they commute with . But this is immediate from both and being Killing vector fields for the metric . ∎
One incredibly important subtlety is the following. Since our spacetimes are not necessarily ultrastatic, there is no reason to expect that if then . This is our reason for using the variable instead of . Indeed, we do have (by definition) that for .
Definition 2.12.
We begin by denoting:
Recall that, for us, geodesics are specifically solutions to the equation and hence these are already “affinely parametrized”. By definition, elements of are, in particular, inextendible causal curves and therefore intersect each Cauchy hypersurface exactly once. An invariant way of dealing with the fact that such curves need not satisfy is to define
where acts on by . Notice additionally that the -action on where acts by descends to an -action on the quotient .
The above set with -action is naturally a symplectic manifold with -action and one can view the next lemma as saying the there are -equivariant Cauchy-data symplectomorphisms between 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 :
and
so that .
As in [22] there are natural isomorphisms of bundles over :
which are symplectomorphisms but are not -equivariant! Indeed, the map
is given by .
Lemma 2.14.
Each equivalence class in has a unique representative satisfying . Identifying elements of with these representatives gives us an -equivariant bijection
where is identified with a cotangent vector via . The -action on is given by scalar multiplication in the fibers. Furthermore, the inverse of the above bijection is given by
or, more precisely, maps to the projection of the curve down to .
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.
has a right action on induced by its right action on given by (via the right action on ). The bijection in 2.14 intertwines this right action with the right action on given by dualizing (using ) the action of pushing forward by right multiplication by on .
Proof.
This is immediate from the explicit form of our isomorphism and the fact that acts by isometries and therefore leaves invariant. ∎
Lemma 2.16.
The flows and on induced by 2.14 are Hamiltonian flows with respective Hamiltonians:
where again we have chosen representative geodesics with . Furthermore, the ’s arise (through the exponential map) from the natural right-action of on hence this -action is Hamiltonian.
As the above right -action is Hamiltonian, we can consider its moment-map:
Lemma 2.17.
Under the isomorphism induced by our -invariant inner product on , the moment map is given by
Proof.
We know that by the definition of a connection on a principal bundle and so the result follows from
since we’re using to identify 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 -action on ), it is in fact quadratic after applying the symplectomorphism .
Lemma 2.18.
Under the symplectomorphism the Hamiltonian becomes:
where is the 1-form on coming from our explicit form for the metric in 1.
Proof.
This follows from the isomorphism being given by and . ∎
Notice in particular that the fact that pointwise implies that is strictly positive. Furthermore, if we had uniformly bounded away from zero on then 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 . We now wish to form the symplectically reduced phase space of solutions with charge . 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 is naturally a symplectic manifold. The symplectic form can be defined as follows. Fix and let denote the stabilizer of under the coadjoint action. Then
induces an isomorphism
which identifies
where is the Lie algebra of . The other tangent spaces of are also identified with by pushforward along the -action. We then have:
We notice that this is well-defined on since
Let denote but equipped with as its symplectic form instead of .
Lemma 2.19.
The extended moment map
is a submersion and acts freely on .
Proof.
The fact that acts freely on simply follows from acting freely on since are principal -bundles. To see that is a submersion, we notice that under the isomorphism we have
and if we use our -invariant inner product to identify then this maps
Forgetting we can already see that is a submersion (and therefore is a submersion). Indeed, it suffices to prove that for every there exists such that . However, is tangent to with so is future-directed, has and as desired. ∎
From the above proof we record as a remark the fact that is precisely the space of pairs where and satisfy
This is precisely the space of solutions with charge , prior to quotienting by gauge transformations.
Definition 2.20.
The reduced phase space is
with symplectic form obtained from the one on .
Lemma 2.21.
The Hamiltonian , extended to to be independent of , is invariant under the -action and therefore descends to a Hamiltonian on with flow .
Proof.
From the definition of we see that what we have to show is that for all and . However:
since is invariant under the 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 and recall from our proof that is a submersion that is also a submersion, hence is automatically a regular value. Furthermore, while the full -action on doesn’t preserve the submanifold , it is preserved by the action of the stabilizer of . The action of on is free since the action of on is free.
Definition 2.22.
The reduced phase space (version II) is the quotient
with the symplectic form induced from that on .
Lemma 2.23.
[11] The map
induces a symplectomorphism intertwining the reductions of the Hamiltonian flow of to and . Here denotes the equivalence class of in the quotient.
2.2 Periodic Orbits
Finally, let’s note that since 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 . There are two aspects of these periodic orbits we will need to consider:
-
1.
the (linearized) Poincaré first return map of a periodic orbit, and
-
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 and restrict ourselves to the contact manifold given by the level surface
This is invariant under the -flow and so we can define the set of periods:
and, for , the set of periodic points:
We say that is the minimum period of if and only if it is the smallest positive time for which . 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 where is its minimum period there exists, in a sufficiently small neighborhood of , a codimension 1 symplectic submanifold
which is transverse to the flow . Furthermore, in a sufficiently small neighborhood of in , the first return time
is well-defined, smooth and satisfies .
Definition 2.25.
With as above, we define the linearized Poincaré first return map to be
This is a linear symplectic map. For any other choice of local symplectic transversal there is a linear symplectic isomorphism
such that
There is actually an alternate, perhaps simpler, description of these maps . 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 with the minimum period of , we define the Floquet operator of to be:
Lemma 2.27.
The subspace
is symplectic, as is the quotient , and is preserved by the Floquet operator. The induced quotient map
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 satisfies the clean intersection hypothesis if and only if is a regular value for and the flow map
admits a clean fibered product over with the diagonal map .
Let’s discuss this hypothesis for a moment. The fibered product is given, as a set, by:
Notice that this contains as a subset and the clean intersection hypothesis implies that is a disjoint union of smooth submanifolds of .
Lemma 2.29.
Under the clean intersection hypothesis, is a clopen subset of and every connected component has
Proof.
Let be any connected component. By the clean intersection hypothesis, for any we must have
Since is linear the only way for the above constraint to be trivial (and not reduce the dimension) is if and if . Indeed, if and we didn’t want the equation to constrain then we would need to constrain to . But now since it follows that implies and so the gradient of the Hamiltonian vanishes at and so is an equilibrium point. However, we assumed that and that was a regular value for , which contradicts vanishing at .
Now, let be the smallest clopen subset containing . We have already shown that for any connected component and so we must have since is a disjoint union of connected components. In particular, since the inclusion
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 is open in , hence open in since is open in . Since is also closed in it follows that it is clopen hence
as desired. ∎
We should remark that there is no reason to expect to be connected even if is connected since we have allowed disconnected structure groups such as .
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 .
Definition 2.30.
Let denote the volume form on induced by the symplectic form and equip with the Riemannian metric induced from the one on and the -invariant inner product on . Using this metric we can define the gradient and the -form on :
Denote:
The is invariant under the Hamiltonian flow and its absolute value defines an invariant measure on the energy hypersurface .
Lemma 2.31.
Under the clean intersection hypothesis, the fibered product comes equipped with a natural volume density. Consider then the case is a union of and finitely many disjoint isolated orbits:
with for all . Then the Poincaré first return map of each is invertible and if is the symplectic volume form on then the induced volume density on is given by:
(2) |
with the Poincaré first return map for one, hence any, choice of symplectic local transversal .
Proof.
Indeed this follows immediately from the expression for the tangent space of derived in the proof of 2.29, noticing that the constraint
in the case of isolated periodic orbits is such that has a 1-dimensional kernel in spanned the vector field corresponding to the reduced flow . Thus by 2.27 the Poincaré first return map is invertible and the induced volume form on is determined by the invariant volume form on and the Poincaré first return map acting on
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 is called integral if and only if the cohomology class of its symplectic form is in the image of .
Lemma 2.33.
A coadjoint orbit is integral if and only if there exists a character
such that
where is the identity matrix.
So, when our coadjoint orbit is integral we have a -bundle defined by the character:
where is the quotient of by the relation
The right -action on yields a right -action on the total space since the stabilizer is a normal subgroup. Through this, we identify every tangent space of the total space with the tangent space at the equivalence class of .
Lemma 2.34.
We have a natural isomorphism
Furthermore, there is a principal -connection on such that and, under our above isomorphism, it is given by:
This -equivariant bundle with connection over gives us a natural -bundle with connection over the reduced phase space , which we describe now.
Definition 2.35.
The -Bundle With Connection: Construction I
Recalling that we can consider the -equivariant -bundle:
If denotes the Liouville 1-form on and the inclusion then we have a -invariant 1-form on the total space of this bundle given by:
We then set:
with connection 1-form
Definition 2.36.
The -Bundle with Connection: Construction II
Here we instead extend our right -action on so via the character . We then set
with connection 1-form
where now 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 be a periodic orbit of the -flow (i.e. for some and ) and assume that is the minimum period of . We denote:
A key point is that while our construction of the -bundle with connection relied on a choice of character as well as a choice of , the element 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 is locally constant. Furthermore, if we consider the symplectomorphism and suppose we had with and , such that then if denotes the image in the quotient we have:
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 since is a dimensional quantity, but instead sends an observable such as or to infinity (here and 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 . The relativistic version of this is defined below.
Definition 3.1.
Fix a smooth function satisfying to act as a time-independent potential. We denote
acting on . We also define operators:
Lemma 3.2.
We have
Proof.
Indeed, isometries and hence Lie derivatives along Killing vector fields commute with the wave operator so it suffices to shown that for all . since and is the horizontal lift of to . since is constant on the fibers of and the vector fields are vertical. ∎
It should be noted that these results are of interest even when . Nevertheless, we include the potential term in order to allow our results to apply to the conformal wave equation
for a dimensional constant and the scalar curvature of . The origin of this variant of the wave equation comes from considering conformal variations of the Hilbert-Einstein action. Indeed, setting one can compute:
It’s also worth noting that if denotes the (constant) scalar curvature of the fibers of then from [3] Theorem 9.3.7 we have
and so the scalar curvature of does indeed satisfy our assumptions on the potential.
Returning to our operator , the vanishing of our commutators with and tells us that the operators and leave the kernel
of invariant. In fact, we want to complete the kernel
of 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.
Definition 3.3.
For we denote:
and we will often interpret elements of as -valued functions on . For the moment, let’s write:
Lemma 3.4.
Definition 3.5.
We denote
Lemma 3.6.
[2] We have if and only if and there exists a such that . Furthermore, the restriction map
is a vector space isomorphism for all and the locally convex topology on obtained by declaring this to be a homeomorphism is independent of our choice of .
Lemma 3.7.
[2] For each the map
is an isomorphism of locally convex spaces (recall: is the future-directed unit normal of the Cauchy hypersurfaces ). Thus has the topology of a Hilbert space.
Our goal is to study the semiclassical asymptotics of the action of time translation on the space .
Definition 3.8.
We denote by
the isomorphism given by precomposing functions with the time flow along the Killing vector field .
At the moment 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 is not unitary on with respect to any of the Hilbert space structures defined by a fixed Cauchy-data isomorphism since . Instead, we proceed as in [22] and notice that the equation arises from a variational problem and therefore has an associated stress-energy tensor.
Definition 3.9.
Given we define the stress-energy tensor of to be
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 , the stress-energy tensor has divergence with respect to the metric .
Proof.
We compute, using that the metric is divergence free to get:
and since by assumption we’re done. ∎
As such, we can use the stress-energy tensor to define a quadratic form on and extend it to a Hermitian form on via the polarization identity.
Definition 3.11.
For we denote
and extend this quadratic form to a Hermitian one via the polarization identity.
Lemma 3.12.
is invariant under both the action of and .
Proof.
First we recall the proof of -invariance from [22]. Since it suffices to show that the time derivative of vanishes at . We do this for real-valued. Writing for the Hodge- on (not on !) we can compute using Cartan’s formula for the Lie derivative:
where denotes the full divergence on . From the previous lemma we have and since Killing vector fields are divergence-free it follows that
as desired.
Next let’s look at the -action. We write for the function and also continue to use the notation for the induced right action of on covectors . Since acts by isometries we have and therefore
But and are both invariant under the -action so
Finally, the volume form is invariant under the -action since it is an action by isometries hence we can perform the change of variables in the integral to get
as desired. ∎
Unfortunately: since we have allowed possibly negative potentials our quadratic form 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 , etc. on certain closed subspaces of and will be positive definite on these subspaces.
Lemma 3.13.
[22]
is finite
dimensional and consists of functions.
Furthermore, for all we have:
In particular, if is the Hermitian form on induced by then is a Pontryagin space.
Lemma 3.14.
descends to a Krein-self-adjoint operator on whose domain contains the dense -invariant subspace given by the image of in the quotient. Furthermore, the spectrum of on this Krein space is discrete consisting of eigenvalues of finite multiplicity, invariant under and , accumulates at only, and has only finitely many non-real eigenvalues.
Proof.
Lemma 3.15.
[22],[16],[6] Let denote the induced quadratic form on the quotient . Then there exists a maximal negative definite subspace
which is invariant under and . Furthermore, it is finite-dimensional with dimension an invariant of the Krein space and themselves. Finally, is invariant under the action of .
Proof.
The only part of this not proven in the above-cited papers is the -invariance. Indeed, suppose for contradiction that there was some and with . Consider the subspace . Then, as we have that the subspace properly contains . Furthermore, it is invariant under both and since commutes with the -action. Finally, is negative-definite on since it is negative-definite on and invariant under the -action, hence is negative definite on , contradicting maximality. ∎
Since is non-degenerate and invariant under both the -action and we obtain the following immediate corollary.
Corollary 3.16.
The subspace
is a Hilbert space with inner product , and is equipped with a unitary representation of given by the restriction of and the -action from above.
We can now begin the process of showing that is positive definite on isotypic subspaces for irreducible representations with sufficiently large dominant integral weights.
Lemma 3.17.
Let be the preimage of in under the quotient map . Then is finite dimensional and contains .
Proof.
Indeed the quotient map restricts to a map with kernel . Choosing a splitting of this linear surjection gives us an isomorphism of vector spaces and since so is . ∎
Definition 3.18.
For our integral coadjoint orbit and we let denote the irreducible representation corresponding to the integral coadjoint orbit .
Proposition 3.19.
There exists an depending only on , and the Krein space such that for any and any which generates a cyclic -representation isomorphic to we have
Thus for each we have a closed subspace
on which restricts to a positive definite Hilbert space inner product. Furthermore, our representation of arising as the product of the -action and leaves invariant and is unitary.
Proof.
Let generate a cyclic -representation isomorphic to . Suppose that and so there existed a non-zero . Since is a -invariant subspace we have where is the cyclic -representation generated by . Furthermore, and since is irreducible it follows that . So it follows that:
Since is finite dimensional this can happen for at most finitely many irreducible cyclic invariant subspaces and hence for at most finitely many . In fact, since the dimension of is an invariant of and the Krein space it follows that for large enough (with dependence as in the statement of the proposition) and all we have:
In particular, for and defined as in the statement of the proposition, is positive definite on .
To show that our action leaves invariant and is unitary it suffices to show that it leaves invariant and is unitary here, since it will then extend to by uniform continuity. Since is invariant under the full -action, unitarity is immediate. All that remains is to check invariance. However, since is irreducible it follows that for any with and any we have hence thus we have invariance, as desired. ∎
It is worth noting that, as remarked in [22], if and there exists some for which then is positive definite. This is especially true for the massive Klein-Gordon equation where is a positive constant. In [23] the special case of our results where and were trivial was considered. In this case it was shown that when projected down to our parameter above actually corresponds to mass. We will demonstrate an analogue of this later in 3.1.
Another important remark is that not every has . This is most easily seen in the Euclidean-signature case where is a single point. Then and our Hilbert space is which, by the Peter-Weyl theorem, contains every irreducible representation of as a cyclic subspace. However, as was shown in [8], since is compact Hausdorff and second-countable, the entire representation is itself a cyclic representation.
Combining our previous facts, for we can decompose:
with the -eigenspace for on , organized so that for all . If then and otherwise these spaces are orthogonal (this is the sense in which the above is indeed an -direct sum). We can then further decompose:
and it is worth noticing that is indeed always finite
since itself is finite dimensional (being an eigenspace
for ).
Since we will be studying asymptotics as , there’s no harm in replacing with so that we may assume . As such, we want to study the time evolution of quantum states in the subspace
However, we still haven’t fully specified a direction in which to take our large quantum numbers limit. Indeed, for fixed the eigenvalues very well might accumulate at as tends to . Thus for each we could consider eigenvalues satisfying
and different choices of might very well yield different asymptotics. Classically this is reflected in the fact that symplectic reduction along generally leads to phase spaces which are not conical. As such, our problem is broken into two steps:
-
1.
For fixed, “count” eigenvalues satisfying .
-
2.
Understand the asymptotics of the above count as .
The first step is fairly straight-forward. It is highly unlikely for us to have any eigenvalues satisfying exactly and so we instead sum over all , weighting eigenvalues near the most. By stationary phase, this is described for large frequencies by the distribution:
We use the letter to denote this distribution since it can be
viewed as a multiplicity for the representation on of associated to the coadjoint orbit . The point is that (modulo factors of ),
approaches as and so in
this limit the right hand side approaches the literal
multiplicity of as an eigenvalue on . However, this is
only a moral since the above limit does not converge.
Instead we first notice that defines a linear functional on the collection of all 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 is actually tempered and hence is defined for any .
3.1 Relation to Vector Bundles
We begin by recalling the well-known fact that for any unitary representation of there is an isomorphism
between -valued -equivariant smooth functions on and smooth sections of the associated vector bundle over . Furthermore, the Hermitian inner product on defines a Hermitian fiber metric on . We will need a less well-known, but related construction.
Definition 3.20.
We fix an so that is positive definite on and denote by our irreducible representation corresponding to . We also let and fix an orthonormal basis for , writing for our Hermitian inner product on .
Lemma 3.21.
Let and both be non-zero. Define a function
Then as -representations. Furthermore, if is extended to act on -valued smooth functions it follows that if and only if .
Proof.
Since is non-zero and is irreducible, it is a cyclic vector and so for each there are finitely many group elements such that . Thus
So the functions are in for all . Furthermore every function is in the span of the functions hence
The set of functions are linearly independent since if are such that for all then since there exists a with . Since is irreducible there exists elements such that and so
hence for all as desired. Therefore the map
induces an isomorphism of -representations .
If then by definition ( and are constant) . Conversely, -invariance of implies that if then for all and hence for all . Therefore as desired. ∎
Usually one doesn’t look at the full wave operator applied to but only at the “horizontal” wave operator. To relate these two wave operators, we fix a root system for compatible with our -invariant inner product and let:
and
Lemma 3.22.
The wave operator on splits as a sum of vertical and horizontal parts:
where is the horizontal wave operator (plus the potential) and is the Laplacian on the fibers. These operators commute and if has then acts on as multiplication by a constant. Hence acts by multiplication by a constant on all of and this constant is given by:
Proof.
The existence of the splitting and the fact that follows from [11] section 6. Since and both commute with the -action it follows that does as well hence does indeed preserve . In fact, by the explicit form of we see that its action on is precisely the action of the quadratic Casimir and hence is given by multiplication by . ∎
In fact, we see that preserves our space
and on this space is precisely the -eigenspace of .
Definition 3.23.
We denote by the operator
Lemma 3.24.
Denote by
the collection of all invariant subspaces of which are isomorphic to as -representations. Then for each we have . Furthermore if is any isomorphism of -representations then
is a -equivariant -valued function with
(3) |
Finally, the definition of is independent of our choice of orthonormal basis .
Proof.
Since each generates a cyclic representation isomorphic to it automatically follows that and satisfies 3. So all that remains to be checked is ’s equivariance and basis-independence. However since is an isomorphism of -representations we can compute:
But if we write then we arrive at:
proving equivariance. Similarly, if is another orthonormal basis then there exists a unitary matrix satisfying hence
as desired. ∎
Since is irreducible, Schur’s lemma tells us that any two isomorphisms of -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
where in the above expression is any choice of orthonormal basis for and is any choice of isomorphism of -representations .
Proof.
Our final step is to compare elements of with sections of the associated vector bundle.
Definition 3.26.
We define a map as follows. Given and we choose an arbitrary in the fiber over and define
We recall from [3] Chapter 3, for example, that is an isomorphism. Furthermore there is an induced covariant derivative on which corresponds under to the horizontal exterior derivative on with respect to , and there is a Hermitian fiber metric on corresponding to the constant Hermitian inner product on .
Now, let’s let and choose an isomorphism which is unitary where is given the -inner product. Writing
it follows that the expression
is independent of our choice of orthonormal basis or unitary isomorphism . We also have the following explicit formula from [22] where we use Greek for indices of coordinates tangent to and Roman indices for coordinates tangent to the fibers of :
By equivariance it follows that if is an orthonormal basis for then
and so
Thus we obtain
Furthermore, from this explicit expression we see that the sum
(4) | ||||
is invariant under the action.
Definition 3.27.
Given a section we define the bundle stress-energy tensor to be the symmetric 2-tensor on given by:
where we recall that is the covariant derivative on induced by the connection .
Since is a constant and the connection is compatible with the fiber metric it follows exactly as in the scalar case that if we abuse notation and also use to denote
acting on sections of then
where we note that despite the raised and lowered ’s appearing, we are not summing over them: they merely denote the representation of we are considering.
Just as in the scalar case, we can define the space of finite-energy solutions to and one has Cauchy-data isomorphisms:
which give the topology of a Hilbert space. Furthermore, since is Killing the covariant derivative commutes with and we have the densely defined operator
Combining all of our results in this section and especially using 4 we arrive at the following result.
Proposition 3.28.
Let and a unitary isomorphism so that we can define
Then and
where is taken with respect to the volume form induced by our -invariant inner product on . Furthermore, since by assumption it follows that is positive definite on the finite energy space .
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 are tempered.
Theorem 3.29.
[14] The operator is self-adjoint on with discrete and accumulating at with polynomial growth.
Corollary 3.30.
The spectrum of on is real, discrete and accumulates at with polynomial growth. Furthermore, the multiplicity of is equal to times the multiplicity of .
Corollary 3.31.
The distribution given by
is a tempered distribution on . Here we recall that are the eigenvalues of on .
4 Proofs of Main Theorems
We are now prepared to study the asymptotics of . As it turns out, this will depend significantly on whether or not . For now we illustrate the method from Section 7 of [10] where 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
with and 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 with so that we may assume .
Definition 4.1.
We define the generating function of the multiplicities to be the periodic distribution in the real variable :
defined for any function which is the Fourier transform of a compactly supported function on .
Distributions of the form with real are called Hardy distributions. These are precisely the distributions on the sphere 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 is a homogeneous function of , 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 as a composition of Fourier integral operators and through this we will show that it is actually in . For now we illustrate how the asymptotics of the Fourier coefficients of a general Lagrangian distribution on can be related to its principal symbol.
Definition 4.2.
Let . An element is called classical of degree if and only if when interpreting as a -periodic distribution on we have:
-
1.
is an isolated singularity, and
-
2.
for any with on a neighborhood of and we have asymptotic expansions:
Lemma 4.3.
Let be a classical singularity of of degree , and let have on a neighborhood of and . Then
where . If for all (in which case the singularity is called positive) then instead .
Proof.
We can write the distribution as
and so it suffices to check whether the function lives in the correct symbol class. Since its Fourier transform is a smooth function and our asymptotics precisely tell us that it lives in the symbol class
Since this is the correct order for a symbol to define an FIO of order
As for the Lagrangian, one simply notices first that the -critical points of the phase are precisely the set of with , meanwhile the support of is everywhere in the non-positive singularity case and is a positive ray in the case of a positive singularity. ∎
Lemma 4.4.
Suppose had only finitely many singularities and that for with the singularities were all classical with respective degrees . For some smooth cutoffs with on a neighborhood of and , and for the coefficients of our asymptotic expansions for :
we have:
Proof.
Choose our cutoffs to be non-negative with disjoint supports and such that there exists with such that
Then, taking Fourier transforms we have
Since we have that the last term is going to rapidly as . For the remaining terms we have
Summing these asymptotics together then yields our desired result. ∎
So we see that in order to obtain the leading order asymptotics of as it suffices to demonstrate that the singularities of are classical and to compute both its order as an FIO, and the leading terms in the asymptotic expansions of its Fourier transform. Let’s now check how to obtain from the principal symbol.
Lemma 4.5.
Let be a classical singularity of degree of , let be a cutoff as in the previous lemma and let be any principal symbol for . i.e.
Then
Proof.
Indeed, if is any principal symbol for then, by definition
and so dividing by and taking limits yields our desired result. ∎
So, our goal has now been reduced to writing 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 and respectively denote the advanced and retarded fundamental solutions for . Explicitly, for , is the unique solution to whose support is contained in the forward causal set of . i.e. solves with vanishing Cauchy data in the past before . Similarly, is the unique solution to the Cauchy problem with vanishing Cauchy data in the Causal future of (in the future after ).
Lemma 4.7.
[22]
let and let denote the measure on given by integration over with respect to the induced volume measure on from the metric. Write . Then
(where is a distributional derivative of a measure) is the unique solution to the Cauchy problem with Cauchy data
Furthermore with the canonical relation given by
Parametrizing the left copy of in by via the geodesic flow and then the by the parameter , the principal symbol of is given by the half-density
where is the Liouville volume form on induced by the symplectic form.
Before we get to composing FIO’s, let’s recall how this works [13]. Suppose we had smooth manifolds of respective dimensions respectively and , canonical relations. We write for the result of multiplying the left fiber variables by so that the result is a Lagrangian submanifold. Given
we interpret and as operators
One can then often form the composition
where and are defined as follows. Since and are Lagrangian they have dimensions:
The product lives in and has dimension
Meanwhile we also have a diagonal submanifold
of dimension . Since the total space has dimension it follows that if and intersected transversely then the intersection would have dimension
and if are respectively the projection maps from to and then the restriction
is a local diffeomorphism and is a Lagrangian submanifold of . In this case, as long as everything is properly supported, we can take and we have
We call this a transverse composition of FIO’s. Furthermore, in this case if are the principal symbols of then:
However, one can still form the composition if the intersection of and in 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 and , is still defined in the same way, but now the projection map
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:
This is called the excess. Then from Proposition 25.1.5’ in [13] we have
where if are the principal symbols of respectively then the principal symbol of at a point is given by
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 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 . Let’s recall our notation from earlier:
-
•
is the dimension of ,
-
•
is the dimension of with the dimension of ,
-
•
is the dimension of ,
-
•
is a cone subbundle of and has dimension .
-
•
The restriction is symplectomorphic to (but not in a -equivariant way) and both have dimension .
-
•
so has dimension .
The below result is also from [22] and again we state it for the reader’s convenience.
Lemma 4.8.
[22]
Let . Then
where is the canonical relation:
Parametrizing as the flowout of under the -flow the principal symbol of is given by:
where is the subset of where both covectors are in .
Lemma 4.9.
The right -action gives us an action map which is an FIO
with the moment Lagrangian, whose canonical relation is:
The composition , denoted by , arises from a transverse intersection of canonical relations and is therefore an FIO:
with canonical relation
Parametrizing by the principal symbol of is given by:
where is the volume measure on induced by our -invariant inner product on the tangent spaces.
Proof.
The expression for the moment Lagrangian and the fact that is proven in [11]. By construction we have
and the composition is clean so the orders of the FIOs simply add up. ∎
In [22], the distributional trace of was expressed in terms of and so will play a similar role for our equivariant trace.
Lemma 4.10.
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 denote the diagonal map so that pulling back along is an FIO
with Lagrangian where
Then the composition arises from a transverse intersection and is therefore an FIO
where
Proof.
The expression for above is precisely the definition of so let’s check that this is indeed a transverse composition. Notice that in the definition of , one and are chosen, and are uniquely determined by the requirement that must live over the same point in as . Furthermore, the constraint that there must exist so that lives over the same point in adds independent constraints on since they must also live over the same point in . This is unless .
So, given satisfying our independent constraints: and therefore and are completely determined. is directly determined by . Hence we see that there are exactly
independent directions in both the composition and in the fiber over the point corresponding to .
In the case we necessarily have and , however the local is a proper submanifold of and the tangent space to at has tangent directions arising from how vary as we move off the local. Within the local we then have and are determined by . While, in this case, we do have choices for , it is the quantity that appears in and so the fiber coordinates of the first component of always vanish. Thus in both and in the fiber over the point corresponding to we have
tangent directions. Hence indeed we have a transverse intersection and the order of is the sum of the orders of and . ∎
Lemma 4.12.
Let denote the inclusion. Pulling back along is an FIO
with Lagrangian defined by
As in [22], the canonical relation of is disjoint from the conormal bundle and arises from a tranverse intersection so the composition can be formed as if it were a transverse composition of FIOs and
where
Proof.
The proof that the composition can be formed is exactly the same as in Lemma 8.3 and the discussion preceding it in [22], and our is precisely defined to be . Transversality again follows from noticing that the intersection is clean and then dimension-counting, however we should remark that in order to get exactly degrees of freedom one uses the fact that the restriction of covectors in to yields the isomorphism and hence the only way the fiber variable of the first component is zero is if . ∎
So, we’ve arrived at the following object:
The importance of this object arises from the following slight generalization of Theorem 4.1 from [22].
Proposition 4.13.
Let be the operator given by integration over . Then since the base has dimension and the fibers have dimension we have
where
Furthermore, since and is positive definite on , if we set
then we have
(5) | ||||
(6) |
Proof.
The basic facts concerning push-forward distributions such as 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 . Indeed, by the computation in Theorem 4.1 of [22], is the equivariant trace of the operator on . Recalling that is simply the multiplicity of in the -eigenspace and that acts by on this eigenspace by definition of we obtain our above expression 6 for , as desired. ∎
We will now build a distribution on which we will then compose with to produce . A key motivating fact in the below definition is the orthogonality of the functions for different ’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.
Our final to-do before we have, at least morally, obtained a description of as a composition of FIOs is to localize about the ray via . Towards this end, we define an operator
by declaring its Schwartz kernel to be given by the oscillatory integral
Lemma 4.15.
[11]
where
Lemma 4.16.
We can form the composition
where
Furthermore:
Proof.
The fact that this composition can be formed, has the above order, and the above canonical relation is proven in [11]. So, we just need to demonstrate that we do indeed obtain when applying it to . Recalling the formula 6 for we note that by [22] the trace over still decomposes as a sum over (possibly generalized) eigenvalues of counted with multiplicity, only now not all are real and some may be zero modes. Furthermore, the -dependence in the trace over is still in the form of the for the representation generated by that specific (generalized) eigenvector. Indeed, while the Hilbert space inner products from the Cauchy data isomorphism are -invariant they are still -invariant and so is completely decomposable since it is a unitary -representation. Since characters are orthogonal with respect to the Haar measure on , we obtain:
as a distribution on . Finally, applying we immediately obtain:
as desired. ∎
Our next step is to understand the composition as an actual Lagrangian distribution. As it turns out, it is more clear if one first computes .
Lemma 4.17.
The composition
is a transverse composition of FIOs with Lagrangian determined by
Proof.
This is immediate since this composition does not affect the -variables and in the -variables the Lagrangian for is just the diagonal. ∎
Theorem 4.18.
The clean intersection hypothesis implies that the composition of and is a clean composition of FIOs with excess
Thus
where
Proof.
The main goal here is to compute the fiber over a point . By homogeneity we can assume and so the fiber is given by:
Since we chose our constraint implies and therefore corresponds to a unique null geodesic with , , and . Therefore and the image of this in the quotient is a periodic orbit in with period and energy .
Now, let’s write for the projection map and recall that is the set of periodic orbits for the reduced flow together with their periods. If we denote
then since has dimension where . Note: the clean intersection hypothesis implies that is a disjoint union of smooth manifolds with the clopen subset having dimension and the other components having dimension at most . Furthermore the map
is a submersion. This, together with the fact that the holonomy map is locally constant, implies that we have a clean composition of FIOs. Since the only part of the derivative of captured in the image of our submersion is it follows that the kernel of the above submersion at each point contains a -dimension subspace of tangent vectors orthogonal to the tangent space . The only other degeneracy comes the 1-dimensional space of vectors tangent to the curve itself and so we arrive at:
as desired. ∎
All that remains now is the calculation of the principal symbol of . It’s worth noticing, however, that from the expression for we see that the actually wave front set will often be a proper subset of depending on . This is due to the varying dimensions of the components of and the support of the principal symbol of being constrained by . We compute this principal symbol now.
Theorem 4.19.
We have
and, under the clean intersection hypothesis, the singularities of are all classical. Assuming the principal symbol at is given by:
Here we omit the Maslov factor since in this case it can be invariantly taken to be constant on .
Proof.
The result concerning the wave front set will follow immediately from the calculation of the principal symbol since the constraint comes from the support of the principal symbol.
We can compute a principal symbol for by composing their explicitly given Schwartz kernels. The Schwartz kernel for the composition is then a distribution on with Schwartz kernel
Recalling that the principal symbol of is given by
we have that the principal symbol of at a point is given by the integral over the fiber
of the product of the symbol of and the symbol of
restricted to the fiber. The symbol over a more general point is then obtained by homogeneity in . Since and principal symbols are defined modulo symbols of lower order it suffices to compute this integral over the quotient of the clopen subset
In this fibered product of symbols, the pairing of the in the symbol for and the in the symbol for amounts to replacing variables in the fibers of with fiber variables in (here the “fibers” are diffeomorphic to ). Since our fibered product is just over the quotient of by the flow, the pairing of the -variables in the symbol for and the symbol for simply amounts to setting in both symbols and multiplying by . The effect of restricting to along the diagonal on the fibered product of symbols (aside from replacing the volume half-density on with the one on ) is to divide by the function and multiply by a dimensional constant, hence removing the function from our symbol expression. Therefore, denoting by a dimensional constant and writing , we obtain the symbol over the point as:
We can now recover the principal symbol over by scaling. Since the fibers are diffeomorphic to where the -action is by the Hamiltonian flow of they have dimension and so the principal symbol over is given by:
where we note that is half the dimension of our fiber. ∎
Theorem 4.20.
Under the assumptions of 2.31 where the time part of the set consists of finitely many isolated periodic orbits , and assuming we actually have with principal symbol at each , , given by:
where is the primitive period of , is the linearized Poincaré first return map of with respect to any local symplectic transversal and we have included the Maslov factor where is the Conley-Zehnder index of as in [22].
Proof.
The proof is exactly the same as the previous one only instead of integrating over with respect to its invariant measure we integrate over the respective periodic orbit with respect to the density from 2.31. ∎
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 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.