The Wave Kernel on Asymptotically Complex Hyperbolic Manifolds
Abstract.
We study the behavior of the wave kernel of the Laplacian on asymptotically complex hyperbolic manifolds for finite times. We show that the wave kernel on such manifolds belongs to an appropriate class of Fourier integral operators and analyze its trace. This construction proves that the singularities of its trace are contained in the set of lengths of closed geodesics and we obtain an asymptotic expansion for the trace at time zero.
1. Introduction
There is a long-standing research program investigating the spectral and scattering theory of real asymptotically hyperbolic manifolds, see e.g. [An10, AlMa10, AlBaNa20, ChDeLeSk05, FeGr85, GrWi99, GrZw01, JoSá00, Va17] and references contained therein, for a small sample of the surrounding work. However there is comparatively much less work concerning the analogous setting of asymptotically complex hyperbolic manifolds. These spaces were first introduced by Epstein, Mendoza, and Melrose [EpMeMe91], and more recently have been investigated extensively by [GuSa06, FeHi03, GuSa06, HMM17, PeHiTa08, Ma16, Ma18]. This class of manifolds includes certain quotients of complex hyperbolic space by discrete groups, as well as strictly pseudoconvex domains in Stein manifolds equipped with Kähler metrics of Bergman type.
In this work we extend major results which study the wave kernel of asymptotically real hyperbolic manifolds to this complex setting. Joshi-Sá Barreto [JoSá01] study the wave kernel by exhibiting this operator as an element of a certain algebra of Fourier integral operators which have been adapted to the geometry at infinity of this class of real asymptotically hyperbolic manifolds. In the case of both works, moving from the real to the complex case presents new difficulties to the analysis. On the other hand, the original methods of both Vasy and Joshi-Sá Barreto are robust enough to permit an analysis of this class of manifolds of hyperbolic-type.
Before introducing the structure of complex hyperbolic manifolds we briefly recall the geometry of real asymptotically hyperbolic manifolds. A non-compact Riemannian manifold of real dimension is called asymptotically hyperbolic if it compactifies to a manifold with compact boundary , equipped with a boundary defining function , and such that is a metric which is non-degenerate up to the boundary, and moreover that at . This name is due to the fact that the final hypothesis ensures that along any smooth curve in approaching a point in , all sectional curvatures of approach , see e.g. [MaMe87].
As proven in [JoSá00], these geometric hypotheses are equivalent to the existence of a product-type decomposition at infinity , such that
where is a 1-parameter family of metrics on . In this model, the boundary represents the geometric infinity of , analogous to the role played by the at infinity in . In particular the metric fixes a conformal representative of a metric on .
The spectrum of the Laplacian of such manifolds was first studied by [MaMe87]; they determined that it is comprised of finitely many -eigenvalues and the absolutely continuous spectrum . In particular, they prove that the resolvent
is well-defined as a bounded operator on whenever . Further they prove that has a meromorphic extension to , as an operator , and with only finite order poles (this extension is meromorphic on the whole complex plane assuming the metric is even in the sense of [Gu05]).
We now move to introducing the complex analogue of these spaces, and introduce our results. We say a non-compact Riemannian manifold , of complex dimension , is an asymptotically complex hyperbolic manifold (hereafter ACH manifold) if the following holds. We assume compactifies to a manifold , compact with boundary, equipped with a choice of boundary defining function (hereafter, a bdf). This is a smooth nonnegative function on which such that
We further assume the boundary admits: (1) a contact form defined as satisfying ; (2) an almost complex structure ; such that is a symmetric, positive-definite bilinear form on . Then we say is an ACH manifold if there is a tubular neighborhood of the boundary such that
(1.1) |
In particular, for another choice of boundary defining function, , we observe that , for some . Denoting the conformal class of our contact structure by we can consider the boundary as being endowed with the structure of a conformal pseudohermitian manifold . This is analogous to the natural conformal structure on in the real hyperbolic case.
Before continuing, we require an additional hypothesis, which is that is an even metric; i.e., the dual metric defined on has only even powers of in a Taylor expansion at . This is automatic in the case of , and necessary for the existence of a meromorphic continuation of the resolvent of to all of , (in fact, the failure of this hypothesis implies the existence of at least one essential singularity in the continuation of the resolvent, see [Gu05], [GuSa06]).
In the case that the metric of is even in the above sense, we can replace the smooth structure on this manifold with its even smooth structure, denoted . In this case the smooth structure on has been modified by declaring that only functions which are even in are smooth with respect to . This change of the smooth structure permits us to define a square root of our original defining function, and guarantee that it is an element of . Equivalently, the even smooth structure can be defined by declaring is a smooth manifold with boundary, with bdf . Throughout we shall denote the square root of our bdf .
Now we state our main results on the behavior of solutions to the wave equation for small times. This question can be approached by a study of the fundamental solution to the wave equation, as in the work of Joshi-Sá Barreto [JoSá01] who studied the wave operator
in the setting of real asymptotically hyperbolic manifolds. This operator has Schwartz kernel satisfying
and they prove that resides in an algebra of Fourier integral operators. Having shown this, they use the results of [DuGu75, Hö68, Hör71] to study its (regularized) trace.
This construction of the wave group as a Fourier integral operator was motivated by the analysis of the resolvent of a real asymptotically hyperbolic manifold initiated in [MaMe87]. Mazzeo-Melrose obtained their results by exhibiting the resolvent as an element of the “large” calculus of zero pseudodifferential operators ; i.e., those pseudodifferential operators with Schwartz kernels constructed as distributions on the blown-up space , obtained by blowing up the intersection of the the corner with the diagonal in . The new boundary hypersurface resulting from this blow-up is called the front face. (For an extended treatment on such blow-ups see [MaMe87, §3], [Mel96], and [Gr01])
Along such lines [JoSá01] construct a class of zero Fourier integral operators as those operators whose Schwartz kernels, when lifted to , have support away from the left and right boundary faces (i.e. the lifts of and respectively). This greatly simplifies the construction of this class of operators, as typically the corners formed by the intersections of the left face (resp. right) with the front face would need to be incorporated into the definition of the operators; requiring the support of the Schwartz kernels avoid such corners allows their contributions to be neglected. In particular, due to the finite speed of propagation for the wave equation, a distribution which is initially supported only on the front face (such as ) will remain supported in the interior of the front face for all finite time. Thus [JoSá01] can construct a small time parametrix for the wave group while remaining entirely in this restricted calculus of zero Fourier integral operators.
Following this strategy we begin with the notion of the -stretched product, , which is the analogous blow-up of the double space defining the class of -pseudodifferential operators used in the study of the resolvent initiated by [EpMeMe91]. With the appropriate definition of -Fourier integral operators, we can quickly conclude:
Theorem 1.1.
Let be the length functional on , (i.e. the dual metric). For each , the graph of the time- flow-out of the diagonal in by the Hamilton vector field is a canonical relation, denoted . Furthermore, the wave group is a -FIO with respect to this canonical relation.
Once we know the wave group is a -Fourier integral operator, it is straightforward to use the results of [DuGu75, Hör71] to analyze the trace of . One subtlety is that the trace needs to replaced with a regularized trace, defined using a Hadamard regularization procedure using our choice of bdf . Defining the cut-off wave trace,
we obtain
Proposition 1.2.
There exists such that for all , the singular support of is contained in the set of periods of closed geodesics of .
With this result in hand, after choosing a smooth cutoff supported away from the lengths of all non-zero periods of closed geodesics, and using the results of [Hö68] we obtain a Duistermaat-Guillemin type result for the cutoff wave trace.
Theorem 1.3.
There exists such that the renormalized trace satisfies,
as and is rapidly decaying as . The leading term, , is called the renormalized volume, and can be computed as
(1.2) |
where are the unique real numbers such that this limit exists.
Finally, we remark on the appearance of the renormalized volume in Theorem (1.3). In the real hyperbolic setting it is known that the renormalized volume is, in certain dimensions, independent on the choice of representative of the conformal infinity. Namely, for a real asymptotically hyperbolic manifold, one can similarly define the renormalized volume as the finite part of the in the expansion of as , given a choice of bdf . For odd, the real hyperbolic renormalized volume is independent of , the choice of conformal representative. On the other hand, for even, we suddenly have the dependence of the renormalized volume on this choice of representative of . This is result is the so-called holographic anomaly (see [HeSk98]) and motivates much of the interest of asymptotically hyperbolic manifolds in mathematical physics, for their connection with the anti deSitter/conformal field theory (AdS/CFT) correspondence.
More concretely, the volume expansion of of an Einstein asymptotically hyperbolic metric, for even, is given by
and [GrZw01] first made the connection of to Branson’s -curvature [Br95],
for a dimensional constant. In the ACH setting, the renormalized volume was first studied at this level of generality by Matsumoto in [Ma16]. Our construction of the renormalized wave trace thus provides an alternate proof of Matsumoto’s result, via formula (1.2). For a general ACH metric, [Ma16] generalizes this result for an analogue of Bransons -curvature. From his result we obtain as a corollary that the constant in our Theorem 1.3, is given by,
This quantity is a global CR invariant of the boundary, thus leading to a pseudoconformal analogue of the holographic anomaly. Given these results there is strong connection between the renormalized volume of an ACH manifold and its spectrum. On the mathematical physics side there seems to be relatively scarce work on this complex analogue of the AdS/CFT correspondence.
Funding
This work was supported by the National Science Foundation grant numbers DGE-1746047, DMS-1440140, and DMS-1711325, growing out of conversations while the author was in residence at the Mathematical Sciences Research Institute in Berkeley, California during the Fall 2019 semester.
Acknowledgements.
The author is very grateful to acknowledge the kind discussions and suggestions of Andras Vasy, Antônio Sá Barreto, and Pierre Albin.
2. The geometry of asymptotically complex hyperbolic manifolds
Because the construction of our adapted FIO-calculus entails a finer understanding of the geometry of an asymptotically complex hyperbolic manifold, we briefly recall the geometry of the Bergman-type metric our manifold is endowed with.
Let be a non-compact manifold with closed boundary. We assume the boundary admits a contact form and an almost complex structure (i.e., an endomorphism satisfying ) such that is symmetric positive definite on . We consider a metric of the following form: there is a boundary defining function ,
such that in a collar neighborhood it takes the form
(2.1) |
where is a smooth section of . Here, denotes the anisotropic dilation map
with splitting induced by the choice of contact structure , (i.e., ).
We observe that for any other choice of defining function we have
thus it is more natural to associate to a conformal class of 1-forms . The boundary manifold equipped with the data of is a closed pseudohermitian manifold. The corresponding conformal pseudohermitian structure was called a -structure in [EpMeMe91].
This Riemannian metric structure describes a non-compact incomplete manifold whose metric is asymptotic to complex hyperbolic space . A useful model of complex hyperbolic space is given by
with boundary sphere equal to a compactification of the -dimensional Heisenberg group,
This model of complex hyperbolic space realizes with the coordinates
foliating by a family of -hypersurfaces. Writing , in these coordinates we can also write the contact form at the boundary as
and the metric on complex hyperbolic is the Bergman metric,
The Heisenberg group is a Lie group of dimension . In these coordinates the group law is given by
which is abelian in the first components. Its Lie algebra has a basis , which satisfies the non-trivial bracket relations: for all and all brackets vanishing. This structure of a nilpotent Lie algebra gives an identification of the form
after which the group law can be written via Lie algebra elements as,
It is a consequence the nilpotence of that the group law is a finite order polynomial in the Lie algebra elements, rather than the asymptotic series given by the Baker-Campbell-Hausdorff formula, (see e.g. [Ei68]).
Finally, we explain how the complexified hyperbolic space arises as a semi-direct product: there is parabolic dilation on (consistent with the bracket relations of the Lie algebra ) given by . The group law on the semidirect product is given as
(2.2) |
The geometric picture described above of complex hyperbolic space being foliated by a family of Heisenberg groups as level-set hypersurfaces of is compatible with this group law: an open set in is related to the corresponding set in by pullback along .
Our reason for expressing the Lie group law of at the level of its Lie algebra is that the Lie algebra arises more naturally at the level of tangent spaces in our later analysis.
3. The wave kernel on asymptotically complex hyperbolic manifolds
In this section we begin the construction of a Fourier Integral Operator Calculus, which is adapted to the asymptotic geometry of the metric (2.1). Such a calculus will be comprised of operators whose Schwartz kernels have prescribed asymptotics on a manifold with corners, the -stretched product of [EpMeMe91].
Analogously to the 0-blow up, Epstein-Mendoza-Melrose defined the -blow up of an ACH manifold; this will be very similar to the zero-blow up of an AH manifold. The biggest distinction being the blow-up at the front face is non-isotropic, reflecting the different asymptotics in of boundary vector fields (namely those vector fields whose -duals span vs ).
Following [EpMeMe91], we next explain how we will modify the product to construct our algebra of Fourier integral operators. We begin with the notion of the -vector fields :
where is any smooth extension of to all of . It is shown in [EpMeMe91, §1] that this definition is dependent only on the choice of conformal class of . This is partly because a representative of determines a local frame by requiring
(3.1) |
in which we can express
and a different choice of bdf produces a frame as in (3.1) associated to a contact form conformal to .
Given this -module, we can define the -tangent bundle . This is a vector bundle over , with a bundle map , which is an isomorphism over such that
Next, we construct the -stretched product of [EpMeMe91, §8]. Notice first that in the product , the boundary of the diagonal is an embedded submanifold,
and is a clean submanifold in the sense of [DuGu75], since it is an embedded submanifold of the corner, and thus all differentials of bdfs vanish at . The 1-form on defines a line subbundle
spanned by
with denoting the projection onto the left and right factors respectively.

With this trivialization of the conormal bundle, we define the -blow up of the corner as the -parabolic blow-up (defined using the dilation structure on fibers given in (2.2)) of the boundary diagonal:
where the equivalence on fibers is defined using the decomposition with ,
This real unoriented blow-up replaces the submanifold with its inward-pointing parabolic-sphere bundle. This blow-up procedure furnishes a blow-down map
which is the identity on , and given by the bundle projection map of the parabolic-sphere bundle on . This is a manifold with corners, and has three new boundary faces:
By construction, the front face is a fiber bundle over with fiber a projective quotient of the inward pointing normal bundle ; the front face has fiber over given by
For more details and the proof of diffeomorphism invariance of this construction see [EpMeMe91, §5-7].
3.1. The -symplectic structure on
Similarly as in the [JoSá01], associated to the Lie algebra we can define the notion of a -Fourier integral operator, which will be operators whose Schwartz kernels have prescribed asymptotics on a resolution of the product , the -stretched product . A standard Fourier integral operator is characterized by its Schwartz kernel having singular support conormal to a Lagrangian inside ; to generalize this notion we must first understand how Lagrangians arise in the symplectic structure of .
In a neighborhood of the boundary , if we use coordinates , where , then
the induced map on the dual bundles is given by
(3.2) |
In these coordinates the canonical 1-form
pulls back to the 1-form
and hence we have a symplectic form,
(3.3) |
With this symplectic structure on we can explore the many ways to create Lagrangian submanifolds on this rescaled bundle.
Following [JoSá01], we can define extendible Lagrangian submanifolds. Set
the double of the -stretched product across the front face. We say that a smooth conic closed Lagrangian submanifold is extendible, if it intersects transversely. This implies there exists a smooth conic Lagrangian such that
One reason for the interest in extendible Lagrangians is that their intersection with the cotangent bundle over the front face is again a Lagrangian submanifold.
Lemma 3.1.
If is extendible then is a Lagrangian submanifold of
Proof.
Fix coordinates of valid near , and with dual variables . Then give local coordinates for near . By transversality, , thus and some subset of must give local coordinates for . Since is Lagrangian, the canonical 2-form
must vanish on ; hence it vanishes on as well. From the overall vanishing of this symplectic form, and the non-vanishing of on , we must have that restricted to is a multiple of . This implies existence of a function satisfying
In particular, and on . ∎
Having introduced extendible Lagrangians we immediately explain their relation to our the class of distributions we will ultimately be concerned with. We define a Lagrangian distribution associated to an extendible Lagrangian, (either or ), to be the restriction to of a distribution which is Lagrangian with respect to an extension of across . As usual we denote the set of order distributions which are Lagrangian with respect to by (resp. ).
Now that we have introduced Lagrangians in this setting we can see some ways they arise naturally. If are two ACH manifolds, a -canonical relation between them is a -map
defined on an open conic subset such that . Certain -canonical relations will define Lagrangian submanifolds in , by associating to its graph relation
and we denote such Lagrangians by . Particularly relevant Lagrangians will arise from liftable canonical transformations; these are homogeneous canonical transformations , whose projections to the base is the identity over .
Using the left and right projections we can define a symplectic form on by
(3.4) |
Further, the dual to the differential of the blow-down map induces a smooth map
(3.5) |
which is an isomorphism over between and the standard symplectic form on .
Lemma 3.2.
(Liftable Canonical Transformations induce Extendible Lagrangians)
Let be a liftable canonical transformation. The map (3.5), combined with the identification (over ) gives a smooth map
(3.6) |
which, restricted to the graph of , extends by continuity to the boundary and embeds into it as a smooth Lagrangian of , denoted . Further intersects the boundary of only over , it is extendible across the front face, and this intersection
defines a Lagrangian submanifold of .
Proof.
On the two copies of in the product we consider respectively coordinates , and valid near the boundary. These induce corresponding local coordinates on the cotangent bundles, which we denote by
(3.7) |
and
(3.8) |
on the left and right copies of the respective cotangent bundles. We fix
as coordinates valid near the front face , away from . The map (3.6) gives an identification between the 1-forms
and
defined on and respectively. We will first determine how the coefficients of these 1-forms are related under the map (3.6), in this neighborhood of . Since we have
and so the canonical 1-form in is given by,
where
Now, using the fact that is a -canonical relation (and thus ), and the fact that restricts to the identity over . To determine in the coordinates (3.8) we observe first that and are both bdfs on and thus conformal: . Further, we have that , and , hence for some smooth functions on . Finally, we use the relation between the fundamental 1-forms to observe that
The final bracketed term will only contribute terms which are or after computing their -differential (e.g. ), thus after grouping such terms we obtain
where are smooth functions of . Taken together, these computations imply that its graph is of the form
From this we can see that
Since is smooth and positive on , the map (3.6) (defined over the interior ), extends by continuity to the boundary when restricted to , thus defining
as the image of under the map (3.6). Further, this shows that intersects the boundary of only over and does so transversely. Thus it is an extendible Lagrangian, and we have by the previous lemma that this intersection is a Lagrangian submanifold of . ∎
This lemma elucidates the name liftable canonical transformation as they provide examples of canonical transformation with “good” lifts to as the associated Lagrangian meets the diagonal only in the front face .
Given we define its -Hamiltonian vector field by the relation . In local coordinates in which is given by (3.3), is given by
And observe that this vector field has the special property that the projection of the vector field to the base vanishes when restricted to .
Because our focus is the wave equation we are most interested in the Hamiltonian associated to our ACH metric. Since our metric satisfies
we can conclude its dual metric on has the form
or in the coordinates on our dual metric is given by
(3.9) |
This function on will be the Hamiltonian of interest in our study of the wave equation.
Lemma 3.3.
-canonical flowouts
Let be the dual metric associated to the metric , and let be its -Hamilton vector field. For all , the canonical transformation , given as the flow-out of the Hamiltonian,
is a liftable canonical transformation. Thus the graph of defines a smooth extendible Lagrangian submanifold of . Further, the intersection
is a smooth Lagrangian submanifold of given by
where , the restriction to the front face of the lift of to .
Proof.
Since the flow-out of a Hamilton vector field is always a canonical transformation, the first claim follows from the fact that only the projection onto the base vanishes. Thus we only to check the claim regarding . We can study the graph of after viewing as a function depending only on the second copy of .
On this space, we can write our canonical 1-form in the coordinates
(3.10) |
thus we can write the Hamilton vector field of a function on this space with respect to this 1-form, with the same formula as we calculated above. In this case is the flow-out of the diagonal in along the vector field . In these coordinates our length function is given by
and we can consider local coordinates near the front face, projective with respect to the left face:
with blow-down map
The pullback of (3.10) by is,
and in these coordinates we have that and the interior lift of the diagonal is given by . The lift of to is given by
where the functions are -homogeneous of order 2 in the fiber variables.
Now we lift our symplectic form (3.4) to , and denote it by , lifts to . In the coordinates
our lifted Hamilton vector field has the form
thus is smooth all the way down to . Further, with respect to our coordinate transformation on induced by the blow-down map , we have that the diagonal
lifts to
Thus transversely intersects at . Finally we see that projects down to as
which is precisely the Hamilton vector field of . ∎
Remark 1.
Notice that because
the projection of the Hamilton vector field of to is precisely the Hamilton vector field of the restriction of to , with respect to the induced symplectic form on .
In other words, for the Hamiltonian given by our length functional (3.9), we have that: , the twisted graph of defines a Lagrangian submanifold . Further this Lagrangian intersects the boundary only over , and it does so transversely. The transversal intersection is itself a Lagrangian flow-out
which is the flow-out by the Hamilton vector field of , the principal symbol of the normal operator at the front face.
3.2. -FIOs and the Wave Kernel
Here we construct the calculus of operators that our wave group will lie in. These shall be restricted to the subclass of Lagrangian distributions whose support does not meet the left or right faces, , and respectively. Due to the finite speed of propagation, initial data supported in the interior of which evolves according to the wave equation,
remains supported away from the left and right faces, . In particular, when considering our calculus of FIOs, we can ignore the complement of the front face in the corner, and restrict ourselves to Lagrangians which meet the boundary only at .
Since the canonical relation of the wave group will be a Lagrangian in , we mildly extend our class of Lagrangians from the last section. The canonical 1-form on is given by
With this 1-form, we can define a canonical relation
and this canonical relation in turn defines a Lagrangian of given by
where is an extendible Lagrangian associated to the graph of the liftable canonical transformation
and is the lift of from the second copy of . In particular, this Lagrangian intersects the boundary only over the front face , and
where is the restriction of to .
Now, given a liftable canonical transformation we define our -Fourier Integral Operators associated to to be the linear operators whose Schwartz kernels lie in the space of distributions
where is the extendible Lagrangian submanifold of associated to by lemma 3.2. Similarly, for the canonical relation defined above, we say that -Fourier Integral Operators associated to are the linear operators whose Schwartz kernels lie in the space of distributions
In both cases, such operators are those whose Schwartz kernels are Lagrangian distributions with respect to , ( resp.), and vanish to order at the front face . Such operators carry two different principal symbol mappings: one is the usual symbol of a Lagrangian distribution, in the interior; the second operator is obtained by the principal symbol of the normal operator (resp. ) associated to the Lagrangian in (resp. ).
This second symbol is again the symbol of a Lagrangian distribution from the fact that our Lagrangian (resp. ) has transversal intersection with (resp. ), thus the restriction of Lagrangian distribution to is again a Lagrangian distribution with respect to (resp. ).
We now take a moment again to highlight the normal operator. If , then , and is a Lagrangian distribution with respect to (resp. ). Further the normal operator satifies an analogue of the short exact sequence for principal symbols of operators:
Proposition 3.4.
The normal operator participates in a short exact sequence
such that for any -differential operator and any -Fourier integral operator we have
Proof.
The injectivity portion of the statement of exactness is immediate from the definition. Since we have is in and defines an operator on for each fixed. In particular, since the kernels of these operators are smooth up to the front face, it makes sense to consider their Taylor series on . The surjectivity of thus arises from a version of Borel’s lemma for the Taylor series of in local coordinates for .
To prove the composition formula, we can use the structure of the Normal operator at , and the fact that we are not blowing up in the variable, so it commutes with the normal operator.
We observe first that such a can be written with respect to our frame for :
As usual we choose to identify this as acting on -densities: if we choose coordinates , these induce a trivialization of the square root of the -density bundle
and acts on by . Of course this is simply for -differential operators. More generally -FIOs will act on -densities via their normal operator:
In particular, this implies that the normal operator of
∎
Having proven this lemma, we arrive at a short time parametrix for the wave group.
Proposition 3.5.
For each , for the canonical relation
the wave group is -Fourier integral operator of the class
Proof.
Given the normal sequence, the argument reduces to a purely local one: using proposition 3.4, and the fact that
we can take as ansatz the wave group in this fiber :
here corresponds to the identity element in the group. Note also that the specific form of the model Laplacian
means we can also construct the model wave group, and study its asymptotics via analyzing those of the wave group in .
Since , it does not meet the corners of . Similarly does not meet the corners in finite time, so we can follow the argument of Duistermaat-Guillemin prop 1.1 to conclude
Now we iterate. Choose a such that . Then
and where is a defining function for the left face. (This is well-defined since is supported away from the left face, as was, and the wave operator preserves this support due to the condition on wave front of , via [Hör71, Thm 2.5.15]). Now we solve the inhomogeneous wave equation to find a solving
solving as before we obtain such a . We now have .
Proceeding iteratively we obtain such that vanishes to infinite order at . The error term also has infinite order vanishing at in the Cauchy data from the construction. Finally, after extending this error term to be identically zero across the front face, we can use Hörmander’s transverse intersection calculus to remove this error term (see e.g. [Hör71, Thm 2.5.15]). ∎
Unfortunately, this is a short time parametrix, as this construction is only valid for finite . If we allow , our Lagrangian flow-out will meet the corners of , which would require a more sophisticated composition formula.
4. Wave Trace Asymptotics
Now that we know the wave group is a -Fourier integral operator we can ask whether its trace can be studied, as in the case of the wave trace on a compact manifold without boundary. This presents some technical difficulties, since the operator is not trace class, so we need to introduce a regularization of its trace.
Heuristically, our goal is to study the trace,
(4.1) |
using appropriate maps , to define this integral via pullback and pushforward. An analysis of the wavefront sets of these maps will permit an analysis of their associated operators, and prove that the resulting object is well-defined distribution on , with wavefront set to be determined.
First, notice that for all , the restriction of to the diagonal is well-defined. To see this we proceed as in [DuGu75, §1] by introducing the map,
of the inclusion of the diagonal. Pullback along this map is a Fourier integral operator of order , defined by the canonical relation
Now, using the fact that (as defined in proposition 3.5), assuming , then whenever we have , thus at such points (where is the set of normals of the map). Thus we can apply [Hör71, Thm ] to conclude that is a well-defined distribution on with wavefront set
Duistermaat-Guillemin next study the wavefront set of the projection . In our case we now introduce the regularization procedure. For , define for our bdf . Consider the cutoff projection
for which integration over the range is equal to the pushforward along (the transpose of the operator ). This map thus defines a Fourier integral operator of order defined by the canonical relation
Again applying Hörmander’s Theorem [Hör71, Thm ] we can conclude that the cutoff wave trace
is a well-defined distribution on satisfying
We obtain as a corollary
Corollary 4.1.
For , the singular support of is contained in the set of periods of closed geodesics in . Moreover, there exists such that all closed geodesics of with period greater than zero are contained in .
In particular for all , the singular support of is contained in the set of period of closed geodesics of .
Proof.
Only the claim regarding closed geodesics remaining in remains to be proven. This is a statement about strict convexity of the geodesic flow in a neighborhood of infinity (see e.g. [JoSá01, Proposition 4.1], [DaVa12, Lemma 4.1]). We show that if sufficiently small, any geodesic which intersects cannot be closed. Introducing coordinates with corresponding dual coordinates , such that is a boundary defining function for .
In these coordinates, we write the metric in a collar neighborhood of the boundary as
and we write
for the bilinear form on induced by the dual metric of . In these coordinates the geodesic Hamiltonian is given by
where , and . The Hamilton vector field of this function is given by
where is a local -orthonormal frame dual to . Computing the change in these vector fields with respect to the change of coordinates gives
Thus the Hamilton vector field can be re-expressed as
where we have defined , the infinitesimal generator of the Heisenberg dilation action on . Using the facts that
and writing the vector field , we can re-express this formula as
Thus, along integral curves of the vector field we have . Thus, at a critical point of along the flow which is an interior point of we have
hence at such points we have
Now, using the fact that is positive definite, thus for sufficiently small this quantity is negative. Thus we have shown that for all geodesic curves which intersect satisfy,
Now, assuming for the sake of contradiction that is closed. Then there exists such that intersects in at least two points. Therefore there exists a with where has a minimum. However at such a minimum we have and , contradicting our convexity statement. ∎
Using this corollary, we can now begin an analysis of the renormalized wave trace. If we denote by be the operators defined in the proof of proposition 3.5. The same arguments used above can be used to show that the distribution
is well-defined, with singular support satisfying the conclusions of corollary 4.1. Since and intersect transversally, only the density factor implicit in this operator can obstruct the convergence of as . Since this density, a trivialization of the -bundle, diverges at the rate at , the integrals converges for any . Applying Taylor’s Theorem to as , we see that there exists constants such that the limit
exists, which we call the renormalized wave trace. From corollary 4.1, we immediately obtain
Proposition 4.2.
The singular support of is contained in the set of periods of closed geodesics of .
Finally, we can begin our analysis of the renormalized wave trace as (in fact its inverse Fourier transform). First we choose a cutoff function , with the appropriate support to study the transform of the cutoff wave trace. If we denote the first non-zero period of a closed geodesic on as , then choose such that for and for .
Now, using the arguments of [Hö68], (which are purely local, applying to any paracompact manifold), or alternatively the proof of [DuGu75, Prop 2.1], we immediately obtain
Proposition 4.3.
There exists coefficients such that the cutoff wave trace satisfies,
(4.2) |
as and rapidly decaying as . The leading term,
Given this result for the asymtotics of the cutoff wave trace we can then conclude similarly for the full wave trace 4.1 that
Theorem 4.4.
There exists coefficients such that the renormalized trace satisfies,
as and rapidly decaying as . The leading term, , is called the renormalized volume, and can be computed as
(4.3) |
where the are the unique real numbers such that this limit exists.
References
- [AlBaNa20] Spyridon Alexakis, Tracey Balehowsky, and Adrian Nachman. Determining a Riemannian Metric from Minimal Areas. Advances in Mathematics. 2020 ; Vol. 366.
- [AlMa10] Spyridon Alexakis and Rafe Mazzeo. Renormalized Area and Properly Embedded Minimal Surfaces in Hyperbolic 3-Manifolds. Comm. Math. Phys. 297 (2010) 621-651.
- [An10] Michael Anderson. Boundary regularity, uniqueness and non-uniqueness for AH Einstein metrics on 4-manifolds. Advances in Mathematics. Volume 179, Issue 2, (2003), Pages 205-249.
- [Br95] Thomas Branson. Sharp inequalities, the functional determinant, and the complementary series. Trans. AMS 347 (1995), 3671-3742.
- [ChDeLeSk05] Piotr Chrusciel, Erwann Delay, John M. Lee, and Dale N. Skinner. Boundary Regularity of Conformally Compact Einstein Metrics. J. Differential Geometry. 69 (2005) 111-136.
- [DaVa12] Kiril Datchev and Andras Vasy. Gluing semiclassical resolvent estimates via propagation of singularities. International Mathematics Research Notices, Volume 2012, Issue 23, 2012, Pages 5409-5443.
- [DuGu75] J.J. Duistermaat, Viktor Guillemin. The spectrum of positive elliptic operators and periodic bicharacteristics. Invent. Math. 29, 39-79 (1975).
- [DuHö72] J. J. Duistermaat, L. Hörmander. Fourier integral operators, II. Acta Math., 128(3-4):183-269, 1972.
- [Ei68] Martin Eichler. A new proof of the Baker-Campbell-Hausdorff formula. Journal of the Mathematical Society of Japan. 20 (1968): 23-25.
- [EpMeMe91] Charles Epstein, Richard Melrose, and Gerardo Mendoza. Resolvent of the Laplacian on Strictly Pseudoconvex Domains. Acta. Math., 167 (1991), 1-106
- [FeGr85] Charles Fefferman and C. Robin Graham. Conformal invariants. Élie Cartan et les Mathématiques d’Aujourdui, Asterisque (1985), 95-116.
- [FeHi03] Charles Fefferman and Kengo Hirachi. Ambient metric construction of Q-curvature in conformal and CR geometries Math. Res. Lett. 10 (2003), no. 5-6, 819–832
- [GrWi99] C. Robin Graham and Edward Witten. Conformal anomaly of submanifold observables in AdS/CFT correspondence. Nuclear Physics B. Vol. 546, Issues 1-2, (1999), 52-64
- [GrZw01] C. Robin Graham and Maciej Zworski. Scattering Matrix in Conformal Geometry. Séminaire Équations aux dérivées partielles (Polytechnique) (2000-2001), Talk no. 22, 14 p.
- [Gr01] Daniel Grieser, Basics of the -calculus, Approaches to singular analysis (Berlin, 1999), Oper. Theory Adv. Appl., vol. 125, Birkhäuser, Basel, 2001, pp. 30–84.
- [Gu05] Colin Guillarmou. Absence of resonance near the critical line on asymptotically hyperbolic spaces. Asymptot. Anal. 42 (2005), no. 1-2, 105–121
- [GuSa06] Colin Guillarmou and Antônio Sá Barreto. Scattering and Inverse Scattering on ACH Manifolds. Journal fü die reine und angewandte Mathematik (Crelles Journal) 622 (2006)
- [HeSk98] Mans Henningson and Kostas Skenderis. The holographic Weyl anomaly. J. High Ener. Phys. 07 (1998), 023, hep-th/9806087
- [HMM17] Kengo Hirachi, Taiji Marugame, Yoshihiko Matsumoto. Variation of Total Q-prime Curvature. Adv. Math. Vol. 306 (2017) 1019-145.
- [Hö68] Lars Hörmander. The spectral function of an elliptic operator. Acta Math. 121 (1968), 193218.
- [Hör71] by same author. Fourier integral operators I. Acta Math. 127 (1971), 79183.
- [JoSá00] Mark Joshi, Antônio Sá Barreto. Scattering and Inverse Scattering on Asymptotically Hyperbolic Manifolds. Acta. Math., 184 (2000), 41-86.
- [JoSá01] by same author. Wave Group on Asymptotically Hyperbolic Manifolds. Journal of Functional Analysis. 184, (2001) 291-312.
- [Ma16] Yoshihiko Matsumoto. GJMS operators, Q-curvature, an obstruction of partially integrable CR manifolds. Diff. Geom. and Appl. Vol 45,
- [Ma18] Yoshihiko Matsumoto. Einstein metrics on strictly pseudoconvex domains from the viewpoint of bulk-boundary correspondence. In: Byun J., Cho H., Kim S., Lee KH., Park JD. (eds) Geometric Complex Analysis. Springer Proceedings in Mathematics & Statistics, vol 246. Springer, Singapore. https://doi.org/10.1007/978-981-13-1672-2_18
- [MaMe87] Rafe Mazzeo and Richard Melrose. Meromorphic extension of the resolvent on complete spaces with asymptotically constant negative curvature. J. Funct. Anal. 75 (1987), 260310.
- [Me94] by same author, Spectral and scattering theory for the Laplacian on asymptotically Euclidian spaces, in Spectral and scattering theory (M. Ikawa, ed.), Marcel Dekker, 1994.
- [Mel96] by same author, Differential analysis on manifolds with corners, Available online at http://www-math.mit.edu/~rbm/book.html, 1996.
- [PeHiTa08] Peter Perry, Peter Hislop, and Siu-Hung Tang. CR-Invariants and Scattering Operator for Complex Manifolds with Boundary. Analysis & PDE. Vol 1, No. 2 (2008)
- [Va17] Andras Vasy. Microlocal Analysis of Asymptotically Hyperbolic Spaces and High Energy Resolvent Estimates. Adv. Math. Vol. 306 (2017) 1019-145