Harmonic forms on manifolds with edges
Eugenie Hunsicker Partially supported by the NSF through an ROA supplement to grant DMS-0204730
Lawrence University
Rafe Mazzeo Supported by the NSF through the grant DMS-0204730
Stanford University
November 27, 2006
Abstract
Let
be a compact Riemannian stratified space with simple edge singularity.
Thus a neighbourhood of the singular stratum is a bundle of truncated cones over a lower dimensional compact smooth manifold. We calculate the various polynomially weighted de Rham cohomology spaces of
, as well as the associated spaces of harmonic forms. In the unweighted case, this is closely related to recent work of Cheeger and Dai [5] . Because the metric
is incomplete, this requires a consideration of the various choices of ideal boundary conditions at the singular set. We also calculate the space of
harmonic forms for any complete edge metric on the regular part of
.
1 Introduction
One of the early successes in the extension of Hodge theory to manifolds with singularities was the work of Cheeger in the early 1980's on manifolds with isolated conic singularities [3] . This provided the inspiration for, and one of the first corroborations of, conjectures made by him and Goresky and Macpherson relating Hodge theory on stratified spaces to intersection cohomology. Soon afterwards, Cheeger gave a general strategy to extend these results to singular spaces with an iterated stratified structure [4] .
In the intervening years substantial progress has been made in this subject. The goal of much of this work is the following: consider a particular class of noncompact or singular spaces, and a natural class of metrics on them, and find a relationship between the space of
harmonic forms and some purely topological invariants of the underlying space. Somewhat nonobviously, the dimension of this Hodge cohomology space is a quasi-isometry invariant of the metric, which makes this a feasible program. The problem is rather different for complete and for incomplete metrics since in the incomplete case one must also understand the contributions coming from the choice of boundary conditions at the singular locus.
Methods that have been successful for this encompass sheaf-theoretic techniques and both soft and hard analytic techniques. As an example of the state of the art of the former we refer to Saper's recent work on locally symmetric spaces [23] , [24] ; not surprisingly, combining these various tools can be very effective, cf. our recent work with Hausel [13] .
We do not attempt, however, to list the many other significant recent contributions to this area.
The specific problem we study here is to determine the topological meaning of the space of weighted
harmonic forms on a compact manifold with `incomplete edge' singularities. More precisely, let
be a compact stratified space with only two strata:
an open top-dimensional stratum
and a stratum
of dimension
in its closure. We assume that a tubular neighbourhood of
in
is diffeomorphic a bundle of cones over
with fibre a (truncated) cone
over a smooth compact manifold
.
The Riemannian metric
on
restricts to a standard conic metric on each fibre.
We recast this slightly differently. Let
be a smooth compact manifold with boundary
. Suppose that
is the total space of a fibration
with fiber
, and that
is a boundary defining function on
, so
and
there. Write
A metric
on the interior
is called an incomplete edge metric if in some collar neighborhood of the boundary
of
, it is quasi-isometric to one of the form
where
is the pullback to
of a metric
on
(via the projection
), and
is a symmetric two-tensor on
which restricts to a metric on each fiber
in
, i.e.
at
. The metric completion of
with respect to such a
is diffeomorphic in an appropriate sense to the stratified space
obtained by collapsing each fibre
at
to a point. Natural examples of incomplete edge metrics include hyperbolic cone metrics with unbranched singular set, cf. [14] . Hodge theory on incomplete edge spaces, subject to the condition that
is odd, is also the subject of the paper [5] by Cheeger and Dai. Their aim, however, is primarily directed at the study of the signature on cone bundles and its relationship with Dai's
-invariant for the bundle
. The present paper came into being because we realized that the methods developed in our previous work [13] adapt directly to this setting, and can be used to determine the individual Hodge cohomology spaces also when
is even; it requires little extra effort to study this problem on an entire scale of polynomially weighted spaces. This extended setting is quite natural and in [15] , the first author ties it to another interpretation of Dai's invariant
. Of course, we fully acknowledge the overlap of the material here with that in [5] , and are also grateful to X. Dai for several very useful conversations.
The main results in [13] identify the Hodge cohomology on a manifold
with the same differential topological structure as above, but endowed with a `fibred boundary' or `fibred cusp' metric, with certain intersection cohomology groups of the space
. These types of metrics are complete and occur frequently in interesting geometric contexts, e.g. as gravitational instanton metrics (the fibred boundary case) or locally symmetric metrics with
-rank
cusp ends (the fibred cusp case). The proofs there proceed by first showing that the appropriate intersection cohomology can be calculated in terms of weighted (conormal)
cohomology, and then showing that these are identified with
harmonic forms. The first step uses sheaf-theory and the second relies on analysis via a parametrix construction.
While the proofs here are similar, there are two important differences: first, the parametrix method in [13] draws on the fibred boundary pseudodifferential calculus from [19] , while here we use the edge pseudodifferential calculus from [18] . From the reader's point of view, this substitution is only formal, since the results we need appear quite similar (although the intricacies of the parametrix constructions and analytic phenomena in the two calculi are quite different). However, since the metrics in one of the classes we consider in this paper are incomplete, we must pay more careful attention to the whole question of choices of closed extension for
,
,
and
.
We now describe our results in more detail. We first consider the weighted de Rham complex
, where
is an incomplete edge metric. In general, if
is any function space on the Riemannian manifold
, then we denote by
the space of sections of the exterior bundle
with this regularity. When there is dependence on the metric, it is indicated explicitly by writing
. A form
is in
if
where
. Some results are presented in §3.3 concerning when
has a unique closed extension to these weighted spaces. In particular we prove the Proposition: (§3.3, Corollary 3) Suppose that
or else, if there exists an integer
then
. (Note that this is true when
and
is even, or else when
is odd and
.) Under either of these conditions, the operator
on
has a unique closed extension in all degrees.
This result is an analogue of, and generalizes, a result due to Cheeger in the conic case [4] . The proof involves the justification of a delicate integration by parts.
Although we state and prove this proposition separately, it is also a consequence of another result we prove later in this paper concerning closed extensions of the elliptic operator
, where
is the codifferential on
. The more technical proof in this case uses the ellipticity of
and the existence of a parametrix for it in the edge calculus, as described in §4. We prove the Proposition: (§4.3, Theorem 7): Let
satisfy the hypotheses in the preceding proposition relative to the weight
. If in addition
(the Laplacian on the fibres
with respect to any one of the family of metrics
) has no `small eigenvalues', as defined in §4.3, then
is essentially self-adjoint on
. If the first hypothesis is satisfied, then it is always possible to achieve this extra small eigenvalue hypothesis with a metric
which is quasi-isometric to
.
These results concern special situations where there is a unique closed extension, but in general, there are two different canonical procedures to extend
to a Hilbert complex, known as the maximal or minimal extension of
, and these may be lead to quite different complexes. The associated de Rham cohomologies are denoted
, respectively. In analogy with familiar terminology on manifolds with boundary (which is a special case of our setting when the fibre
is a point), we call the Hodge Laplacians associated to each of these complexes the absolute and relative Hodge Laplacians, and denote their nullspaces by
. It is almost a tautology that the maximal and minimal weighted de Rham cohomology spaces are identified with the absolute and relative weighted Hodge cohomology spaces. As for the topological interpretation, we prove the Theorem: (§3.5, Theorem 4)Let
be a manifold with incomplete edge metric. The maximal and minimal weighted Hodge cohomology spaces are canonically identified with intersection cohomology for the stratified space
by:
and
here
denotes the least integer strictly greater than
and
denotes the least integer greater than or equal to
, and
, respectively
, are the lower and upper middle perversities.
We single out two important special cases:
Corollary: The maximal and minimal de Rham cohomologies when
correspond to upper and lower middle perversity intersection cohomology.
|
(1)
|
Moreover, when
is even, the maximal and minimal de Rham cohomologies at weights
coincide, and again correspond to upper and lower middle perversity intersection cohomology.
|
(2)
|
The notation
is somewhat nonstandard, and indicates a slight generalization of these spaces (so as to include, for example, the case where
is a point), which we discuss in §3.2.
These results are, to some extent, `soft' in that they do not require any serious use of elliptic theory, and for that reason we have separated them into the first few sections of the paper. The main ingredients in their proofs are some abstract functional analytic results involving Hilbert complexes from [2] , reviewed in §2, the sheaf-theoretic characterization of intersection cohomology from [6] , discussed in §3.2, and the appropriate Poincaré Lemmas (also known as `the local calculations'), which are developed in §3.4 and 3.5. Very helpful in our approach is the fact that we may restrict attention to conormal forms, but this is not strictly speaking necessary. This part of the paper is a recapitulation and extension of Cheeger's original work on the Hodge theory on cones, with an attempt to present the argument as cleanly as possible in this slightly more general context, but the results could all also be proved using the techniques in [4] .
On the other hand, we require more analytic information in order to study the minimal Hodge cohomology
, which is defined as the common nullspace of
and
. This terminology is slightly unfortunate, since
is not related to the minimal de Rham cohomology
discussed above, but rather corresponds to the nullspace of the minimal extension of
. In any case, we prove the Theorem: (§4.5, Theorem 8) Let
be a manifold with an incomplete edge metric,
.
The minimal weighted Hodge cohomology is given by
In particular, when
,
The proof requires two main analytic results: the conormal regularity of solutions in the minimal domain of the equation
, and the solvability (and regularity theory for the solution) of
for suitable
. For these we invoke the theory of pseudodifferential edge operators, as developed in [18] . The results from this theory which we require are reviewed in §4.1.
Refering to that section for the following terminology, we note that the specific computations we must make in order to apply this more general theory are the calculation of the indicial roots of
and the injectivity of the normal operator
on suitable weighted
spaces. As we show in §4.2.1, the calculation of indicial roots for
on manifolds with edges is essentially identical to that on cones, and this partially explains the similarity of the results in the two cases. However, the extra role played by the model operator
, see §4.2.2 as well as Proposition 8 in §4.1, is not required in the conic case, but is the key fact needed in the parametrix construction in the edge calculus.
We now turn to some applications and extensions of our results. The first is a Bochner-type vanishing result. Recall that the Weitzenböck formula for the Hodge Laplacian on
-forms on
states that
, where
is a curvature operator, acting by endomorphisms on
.
Theorem: Suppose that
admits an incomplete edge metric
such that
everywhere, and is strictly positive at some point of
. Then the minimal Hodge cohomology
(at weight
), and hence
both vanish. If in addition
is odd or else if
is even but
, then we may also deduce that
.
The proof is the usual one, and simply involves noting that when
is in the nullspace of
, then the integration by parts
is justified.
We can extend this type of analysis significantly further. For example, the generalization to second order natural geometric operators
is essentially straightforward. By definition, such an operator is one of the form
acting on sections of some subbundle
of the full tensor bundle over
, using the induced Levi-Civita connection (though we may also twist by any other bundle with connection); the symmetric endomorphism
on
is a generalized curvature operator.
is formally symmetric on
, and we may ask the same questions about its domain, mapping properties and nullspace as we have for the Hodge Laplacian. The point we wish to make is that answers to these questions are readily deduced, and direct consequences of the edge theory, once one has calculated the indicial roots of
and determined the mapping properties of the normal operator
. Furthermore, the calculation of indicial roots for such an operator reduces directly to the analogous computation on the cone
endowed with the metric
. The presence of an extra twisting bundle may change the arithmetic of the indicial root computation and the spaces on which the normal operator is injective, but makes no difference elsewise. We shall not develop these remarks further here, but shall return to them in greater detail elsewhere. (At that time we shall also give a more careful account of parametrices in the edge calculus for the Laplacian with relative and absolute boundary conditions.) There are many interesting geometric consequences of such results.
In particular, the infinitesimal rigidity of hyperbolic cone metrics with smooth singular set, as proved in [14] in
dimensions and in the recent paper [22] (for deformations amongst Einstein metrics rather than just hyperbolic metrics) in dimensions greater than
is a direct consequence. See [21] for further discussion.
Now consider a more topological application: When
is odd, the maximal and minimal cohomologies agree, as do the two middle perversities
and
for
. Our results in this case, when
, agree with those in [5] . In this special case, and assuming that
, there is a well-defined pairing on middle degree forms; Cheeger and Dai prove the corresponding signature theorem. We extend their result to cover also the case
even. Define the
-signature as the signature of the degenerate pairing on
induced from the map
and the nondegenerate pairing between these spaces. Similarly, one can also define a topological signature
as the signature of the degenerate pairing on
defined through its map to
and the nondegenerate pairing between these spaces. Finally, recall the invariant
defined by Dai [7] for the total space of the fibration
.
Theorem: The
-signature of the stratified Riemannian space
endowed with an incomplete edge metric
, is given by
Our final result concerns the Hodge cohomology of the manifold
endowed with a complete edge metric. By definition, and following the notation above,
is a complete edge metric if near
it has the form
The prototype would be the product of a hyperbolic space (or any conformally compact metric) and a compact manifold
. It is of interest, as a generalization of the main result of [17] , to calculate the Hodge cohomology for such manifolds. We obtain the Theorem: (§5, Theorem 9)Let
be a manifold with a complete edge metric. Let
be the associated compact stratified space. Suppose that
is not of the form
where
. Then
In this case, the
signature result is the same as above. In all other cases, where
does have this form,
is infinite dimensional.
In conclusion, let us remark that sorting out the detailed analysis of the Hodge Laplacian (and other natural geometric elliptic operators) for more general classes of stratified spaces, e.g. algebraic varieties, is a problem of great importance with many applications. Hodge theory on quite general real analytic manifolds is the subject of an ongoing project by D. Grieser and R. Melrose, using an approach closely related to (although more general than) the one used here. We appreciate their interest in and forebearance concerning the present paper.
2 Hodge-de Rham theory for Hilbert complexes
In this section we review some generalities about
cohomology, based on the formalism of Hilbert complexes from [2] , to which we refer in the interests of brevity for most of the proofs.
Consider a complex of the form
|
(3)
|
where each
is a separable Hilbert space,
is a closed operator with dense domain
such that
and
for all
. Under these conditions, ( 3 ) is called a Hilbert complex, and is denoted by
.
Many familiar constructions in Hodge-de Rham theory carry over immediately to this setting, and we list in particular:
-
i)
There is a dual Hilbert complex
defined using the Hilbert space adjoints of the differentials,
;
-
ii)
The Laplacian
is a self-adjoint operator on
with domain
and nullspace
-
iii)
There is a weak Kodaira decomposition
-
iv)
The cohomology of
is defined by
if this space is finite dimensional, then
is necessarily closed and
The main case of interest here is when
is a (not necessarily complete) Riemannian manifold,
and
is the exterior derivative operator. Later we shall also consider the somewhat more general case
where
is some weight function, but for the remainder of this section, to be concrete, we shall assume that
; it will be clear that everything extends to the general case in a straightforward manner.
To turn the `core' de Rham complex
into a Hilbert complex, we must specify a closed extension of
, and there may be more than one way to do this.
Definition 1.
The two canonical closed extensions of
are:
-
∙
The maximal extension
; this is the operator
acting on the domain
| |
| |
| |
In other words,
is the largest set of forms
in
such that
, computed distributionally, is also in
.
-
∙
The minimal extension
; this is given by the graph closure of
on
, i.e.
| |
| |
in which case
.
Maximal and minimal extensions of
are defined in the same manner.
Clearly
. An old result due to Gaffney [8] shows that these domains are the same when
is complete, but in many other cases of interest (for example, on a manifold with boundary) they may differ.
In order to show that
are both Hilbert complexes, we require the
Lemma 1.
and
-
Proof.
The fact that
follows from the identity
on
, while the fact that
on test forms and the definition of distributional derivatives shows that
.
The cohomologies of these complexes are denoted
, respectively. Here and in the sequel we shall use notation like
in a hopefully self-explanatory manner to indicate statements which hold for each of the indicated extensions.
It is straightforward that the Hilbert complex adjoint of
is
, i.e.
There are three well-behaved weak Kodaira decompositions:
|
(4)
|
with summands mutually orthogonal in each case. The first summand on the right, called the absolute, relative or maximal Hodge cohomology, respectively, is defined as the orthogonal complement of the other two summands. Since
, etc., we see that
|
(5)
|
respectively. The third decomposition, incorporating both
and
, is the original one defined by Kodaira. The corresponding Hodge cohomology
is often infinite dimensional, though. We do not consider a fourth weak Kodaira decomposition involving the ranges of
and
, since these subspaces might not even be disjoint, let alone orthogonal. (This is due to the fact that
may not even be defined, let alone vanish.) Nonetheless we still define the minimal Hodge cohomology
|
(6)
|
Note also that
The operators
are both clearly quasi-isometry invariants, and the various Kodaira decompositions above then show that the minimal and maximal cohomologies
, and their `reduced' versions, the absolute and relative Hodge cohomologies
, are all quasi-isometry invariants. This invariance is also true for the maximal and minimal Hodge cohomologies
.
There are quite a few `Laplacians' one might consider, most prominent amongst which are the absolute and relative Laplacians
These are self-adjoint and satisfy
|
(7)
|
Furthermore, if
is finite dimensional, then the range of
is closed, and
. Consequently, these Hodge cohomology spaces may be computed using only tools from differential topology and general cohomology, e.g. sheaf theory, Mayer-Vietoris, etc.
One can also define
These are symmetric, but not necessarily self-adjoint, invariant under the Hodge star, and satisfy
Note that
so this is consistent with our prior definition of
.
We conclude this section by stating two more results, both true in the general Hilbert complex setting, but for simplicity we restrict to the setting of differential forms. The first concerns a Künneth-type theorem.
Proposition 1 ([2] Corollary 2.15).
Let
and
be two Hilbert complexes. Form the completed tensor product Hilbert complex
:
Suppose that
has closed range in all degrees. Then
The other result concerns the possibility of computing one of these cohomology groups using a `core subcomplex' of smooth (but not necessarily compactly supported) forms
consisting of all elements
which are in the domain of
for every
.
Proposition 2 ([2] Theorem 2.12).
The cohomology
is equal to the cohomology of the complex
.
For example, when
is compact without boundary, this is simply the well-known result that
can be computed using the complex of smooth forms. When
is compact with boundary, then as discussed carefully in [2] ,
is equal to the cohomology of the complex of smooth forms continuous to the boundary which satisfy absolute/relative boundary conditions.
3 De Rham theory and edges
The context in which we shall adapt and develop the material from the last section is the category of manifolds with edge singularities. After defining these we briefly review the intersection cohomology theory for such spaces, and then turn to an analysis of the maximal and minimal cohomologies of the Hilbert complexes of weighted
forms, and in particular the identification of different weighted de Rham cohomologies with intersection cohomologies with different perversities.
3.1 Manifolds with edge singularities
We now begin to develop some of the ideas in the last section in the concrete setting of manifolds with conic or edge singularities.
Definition 2.
A pseudomanifold
of dimension
has simple edge singularities if it has a dense open stratum
, which is a smooth manifold of top dimension, and the singular strata
are a disjoint union of closed smooth manifolds
(of possibly varying dimension) such that each
has a neighbourhood
which is diffeomorphic to a bundle with base
and fibre a truncated cone
over a smooth link
.
The boundary
of each cone bundle neighbourhood
is the total space of a bundle over
with fibre
. There are more complicated singular spaces with iterated edge singularities, which is why we call this class `simple'. However, for brevity, in this paper we shall refer to a space of this type as a manifold with edge singularities. Note that this class includes the case of manifolds with conic singularities, i.e. where some
are
-dimensional.
Definition 3.
A metric
on a space
with simple edge singularities is said to be if incomplete edge type if it is an ordinary smooth metric away from the singular strata
, while in each cone bundle neighbourhood
it is quasi-isometric to one of the form
here
is the polar distance on each cone
,
is a symmetric
-tensor on
which restricts to a metric on each fibre
,
is the projection, and
is a metric on
.
A metric
on the principal stratum
of such a space
is of complete edge type if in each
it has the form
where
,
,
and
are as above. We often drop the
in this notation and shall also frequently write
.
We next recall some analytic and geometric properties of Riemannian submersions from [13] and discuss their relevance to de Rham theory for edge metrics.
Let
be a fibration with fibre
, and suppose that it is endowed with a metric
of the form
, where
is a metric on
. We assume furthermore that
is a Riemannian submersion. The tangent bundle
splits into a vertical and horizontal subbundle,
, where
and
is its orthogonal complement (and also the subbundle annihilated by
). This induces a splitting of the form bundles on
, and thus every differential form has a (horizontal,vertical) bidegree, i.e.
The space of harmonic forms on
is finite dimensional, and we let
denote the natural orthogonal projectors; these extend naturally to each
.
Proposition 3.
The differential and codifferential on
decompose as
where
is the pullback of
to the fibre,
is the lift of
as a horizontal operator, and
and
are tensorial operators built from the second fundamental form of the fibres and the curvature of the bundle, respectively. These act as
| |
| |
Now consider the degenerating family of metrics
(
). If
is a
-form, then
Furthermore, as explained in [13] ,
,
and
; the
in the superscript signifies that the operator is to be calculated relative to the metric
, and the operator without a superscript is calculated relative to
. Hence
We can define the operator
this acts on the space of fibre-harmonic forms, or equivalently, we can think of this as acting on the space of forms on
with coefficients in the flat vector bundle of harmonic forms on
that comes from the fibre bundle
. In [13] we proved the following useful lemmas:
Lemma 2.
The operator
and its adjoint
are differentials, i.e.
.
Corollary 1.
Let
, and suppose that
for some fibre-harmonic form
. Then
, and so the terms
of pure bidegree also satisfy
.
3.2 Intersection cohomology
Let
be a pseudomanifold which is a smoothly stratified space of real dimension
, with no codimension one stratum. For this subsection only, we allow
to be more singular than was considered earlier in this paper; namely, around any point
contained in the stratum
of codimension
is a neighbourhood
diffeomorphic to
, where
is a Euclidean ball and
is the cone over a link
, which itself is a stratified space (of dimension
).
A perversity
is an
-tuple of natural numbers,
satisfying
and
for all
. Associated to such a space
and perversity
is the intersection complex
; roughly speaking, the integer
regulates the dimension of the intersection of chains (in general position) with the stratum of codimension
. The homology of this complex is the intersection homology
; the cohomology of the dual cochain complex is the intersection cohomology
.
The following result, which asserts that the cohomology of any fine sheaf over
is equal to the intersection cohomology of
(with respect to some perversity
, so long as the local sheaf cohomology in any sufficiently small neighbourhood equals the intersection cohomology of that neighbourhood.
Proposition 4 ([12] ).
Let
be a stratified space and let
be a complex of fine sheaves on
with cohomology
. Suppose that if
is a neighbourhood in the principal (smooth) stratum of
, then
, while if
lies in a stratum of codimension
, and
as above, then
|
(8)
|
Then there is a natural isomorphism between the hypercohomology
associated to this complex of sheaves and
, the intersection cohomology of perversity
.
The details and proof of this theorem can be found in [6] and [1] . We refer to equation ( 8 ) by saying that
satisfies the correct local calculation for intersection cohomology with perversity
. This result will be one of our primary tools below.
On any pseudomanifold
, there are two distinguished perversity functions:
the lower middle, and
the upper middle perversity. When all strata of
are even dimensional, the two corresponding intersection cohomologies are the same and satisfy Poincaré duality. When not all the strata are all even dimensional, then
in general, but these two cohomologies are Poincaré dual to one another.
In the case of an
-dimensional manifold with edge singularities, there is only one relevant value for the perversity function
in the neighborhood of any of the components of the singular stratum, namely, its value on the codimension of that stratum,
.
If the fibre
for this stratum has dimension
then this codimension is
. If
is odd, then the upper and the lower middle perversities satisfy
. If
is even, then
and
. We extend the standard definition of perversity and intersection cohomology in this situation to include perversities with
and
. We see from the local calculation above that these perversities give cohomologies
| |
| |
When we extend the definition this way, we retain the Poincaré duality property of intersection cohomology:
but we lose the property that the intersection cohomology groups are independent of stratification, which was the original reason for restricting the values of the perversity function. Since there is a natural stratification in the case of manifolds with edge singularities, this does not cause us trouble here.
3.3 Maximal and minimal domains for weighted complexes
From now on, we assume that
is a space with simple edge singularities and
is a metric on it which is of incomplete edge type. Let us fix a smooth function
on the principal stratum
which equals the polar distance function in a neighbourhood
of each singular stratum
.
For any
, consider the complex of weighted
forms
As already explained in §2, there are (at least) two ways to turn this into a Hilbert complex, using either the maximal or minimal extensions of
. The corresponding complexes,
are called the maximal and minimal weighted cohomologies (with weight
), respectively, and have cohomologies and spaces of harmonic forms
and
. Note that the formal adjoint of
with respect to the
inner product has a term of order
depending on
.
It will be possible to simplify many of the calculations in the computations of these spaces below by applying Proposition 2 . As we show later, it will ultimately suffice to consider forms which are `tangentially regular' or conormal, as we now define. Let
be a compact manifold with boundary. The space
of
-vector fields on
is, by definition, the space of all smooth vector fields which are constrained at
to be tangent to the boundary. Thus, in any local coordinate system
, where
is a boundary defining function and
is a local coordinate system on the boundary,
is generated over
by the basis sections
,
.
Definition 4.
Let
. The space of conormal functions on
of (
) order
,
, is defined as
We write
and
. If
is any smooth vector bundle over
, then
is defined using any system of smooth trivializations of
; in particular, the space of differential forms conormal of weight
is denoted
.
Note in particular that if
, then
along with all its
derivatives, and is
in the interior of
. In the following, we frequently make use of the fact that if
and
(with respect to any smooth nonsingular Hermitian metric on
), then
uniformly in
as
. The proof is elementary and left to the reader.
We remark that
is dense in
, and similarly
is dense in
. The former of these is immediate from the definition, while the latter follows (in the cone or edge setting) using standard mollification arguments.
Let us now consider the problem of when
equals
on the truncated cone
, with exact warped product conic metric
. This is already contained in Cheeger [4] (when
), but we present this argument to make it accessible for later generalization. By the remark above, it suffices to consider only conormal forms, and this simplifies the discussion somewhat.
Lemma 3.
The form
lies in
if and only if
(or equivalently,
) when
. In particular, if
is in this range, and if
and
, then
if and only if
.
Before commencing with the proof, we state the immediate and important consequence.
Corollary 2.
Suppose either that the interval
contains no integer value, or else that
and
. Then the extensions
and
coincide.
-
Proof.
Following the definitions, to show that the minimal and maximal domains are equal it suffices to show that for any conormal forms
and
, with
vanishing near
, we have
; in other words, we must show that when
, the boundary term in the integration by parts vanishes if and only if
.
Define
by
. Then
(depending on the parity of the degree), and we find that
and in addition, if
, then
.
Now
and so, writing
, integrating by parts produces the boundary term
|
(9)
|
Hence
if and only if this boundary term vanishes for all
.
Next, the weighted
conditions
and conormality yield
but this gives only that ( 9 ) is
, which is not enough.
To proceed further, use the Hodge decomposition on
|
(10)
|
Inserting these expressions into ( 9 ), many terms vanish and we are left with
|
(11)
|
Since
, we can estimate
,
; in addition,
,
. Hence the first two terms here are
, and it remains only to analyze the third.
Now decompose
and
, where
is an orthonormal basis for
, and where each
and
is conormal on
. The boundary integral now reduces to the finite sum
, and we must show that each
. However,
and these imply that
,
, and
,
. The improvement comes by writing
,
as integrals of
,
, respectively, where the other limit of integration is taken at
or
depending on
or
; this gives
(In particular,
,
exist when
.) Hence, if either
or
, we get
. If
then
and we reach the same conclusion.
If both
, which occurs precisely when
, then we obtain a well-defined, but not necessarily vanishing, boundary term
.
This completes the proof.
We wish to extend this result to incomplete edge metrics. Using the quasi-isometry invariance and localizability (proved in §3.4 below) of the maximal and minimal domains, it suffices to consider forms on
,
, with warped product metric
, and we may as well assume that
is the Euclidean metric too.
As before,
denotes the orthogonal projection onto fibre-harmonic forms, and we let
.
Lemma 4.
Let
, where
are the decompositions into fibre degree. Then
if
and
whenever
.
In particular, if
is fibre harmonic and
is conormal, then
if and only if
.
-
Proof.
Following the same proof as above, the form
is also in
if and only if, for every
with
, the boundary term
vanishes. Decomposing into fibre degree, the boundary term becomes:
We have
so
only at this stage. Continuing as before, we can reduce to the case where each
and
are fibre-harmonic, and write
where
. In order to improve the order of vanishing of
and
as before, we observe that, for example, assuming each
is fibre-harmonic, then
However, conormality already gives that
and so we conclude that
A similar argument applies to
. The rest of the proof is now the same as in the conic case.
Corollary 3.
Suppose that either
or else if
then
. (This is true in particular when
either when
is even, or else when
is odd but
.) Then
, and hence
in every degree.
3.4 Preparations for the Poincaré lemma
We now prove several preliminary results which will be used in the computation of weighted de Rham cohomology on manifolds with edge singularities. Our arguments closely follow those in [4] , but with simplifications since by Propositions 1 and 2 it suffices to work only with conormal forms.
To begin, define two complexes of sheaves,
; these are induced by the presheaves obtained by restricting the forms in
to the cofinal sequence of coverings of
of the form
, where each
is either disjoint from all of the singular strata
or else, if
, then
is a product neighbourhood
; here
and
is the truncation to
of the cone
.
Lemma 5.
The sheaves in each of the two complexes
over
are fine.
-
Proof.
We define a partition of unity
subordinate to
in such a way that each
is bounded. Indeed, away from the singular strata this is obvious, whereas if
then we can let
be a product of cutoff functions in each factor. It is now standard that if
, then
as well. This gives the result.
Now form the associated spectral sequence for either of the double deRham/Čech complexes
. Its hypercohomology may be computed taking either the Čech or the de Rham differential first. In the former case, at the first stage this becomes the standard definition of
; for the latter case, however, we must calculate the weighted de Rham cohomology of each neighbourhood
, which is the main goal of this section. We shall see that these are the same as the local intersection cohomology with respect to a certain perversity function
depending on
, and this will prove the equality of weighted de Rham and intersection cohomologies.
By a slight abuse of notation, we denote the local cohomology of these sheaves by
, respectively. If
is disjoint from all singular strata, then
Neither the weight function
nor the metric
play a role. On the other hand, first note that
Furthermore,
so the range of
on
is closed. Hence by Proposition 1 and quasi-isometry invariance,
| |
| |
We have now reduced the computation of weighted de Rham cohomology on
to that of the truncated cone. For the next step we use the following regularity result
Lemma 6.
-
Proof.
The proof relies on the mapping properties of a parametrix for
; this parametrix is an element
, the calculus of
-pseudodifferential operators on
, such that
for some residual element
which satisfies
. (Both
and
depend on
and the choice of extension.) Suppose that
, so in particular
. Then
, hence
. Induction on
gives the result. We refer to [18] and [16] for more details.
Now we prove three lemmas taken from [4] . In the following,
on
.
Proposition 5.
Let
be the canonical retraction map. Then
is well-defined and bounded if and only if
.
-
Proof.
Let
; then
and this is finite if and only if
. Since
, the image of
lies in
.
Proposition 6.
There exists a
such that for any
there exists some
for which
-
Proof.
If not, then for any
, there exists some (nontrivial)
such that
for all
. However, if this is the case, then for this
,
This is a contradiction when
is large.
Proposition 7.
If
, then for any
, the map
is bounded.
-
Proof.
First,
|
(12)
|
Next, for any real number
,
Setting
and using this in ( 12 ) gives
Since
is bounded away from
, both of these coefficients on the right are uniformly bounded when
.
3.5 Poincaré lemma
We now compute the weighted cohomologies of the truncated cone.
Lemma 7.
and
-
Proof.
By Lemma 6 , we may work exclusively with conormal forms. First let
.
We wish to define a map
and show that it is an isomorphism. Let
, and choose a conormal representative
. For any
, set
.
To check that
is defined independently of all choices, first note that
, so
for any
. Next, if
is another conormal representative of
, then there exists
with
. This implies in particular that
, so
.
Similarly,
and hence
implies that
thus
.
is certainly linear; it is bounded by Proposition 6 and surjective by Proposition 5 , so we must only show that it is injective. Suppose
and
.
Then
for some
. By Proposition 5 ,
, while Proposition 7 gives that
. Thus
Since
, this implies that
, hence
is exact in the maximal complex, i.e.
, as desired.
We next show that when
, any
is trivial. Thus, for any representative
, we must find a
-form
with
. Assume
for some
. The condition
gives that
. Furthermore, by assumption,
, so
is defined. Using
, we deduce that the integral
is finite. Now,
, so
where
, i.e.
, so
as desired. We have now shown that
and hence
in
. This completes the computation of
.
The computation of
proceeds identically when
or
, but the remaining cases are treated slightly differently.
Claim 1.
If
and
is closed, then
for some
. As a consequence,
.
-
Proof.
As above,
, so if
,
Since
,
, so
for all
. This gives that
for some
. Thus
.
As for its domain, it suffices by Lemma 3 to show that
since
. But on the one hand,
; furthermore,
since
, again because
, and
for some
, so
. This proves the claim.
Claim 2.
The map
is injective when
; hence, for
in this range,
.
-
Proof.
If
represents a class in
, then we already know that
for some
. So we must show that we can arrange for
to lie in
as well. Since
,
for some
, and so
is defined. We can choose
which solves
by specifying that
. Now let
. As before,
and
. Furthermore,
and
for
, so
. This means that
.
This completes the calculation of the local cohomology for the sheaves
on
. By Proposition 4 , we now obtain one of our main results:
Theorem 4.
If
is a manifold with an incomplete edge metric and
is the associated stratified space, then
and
here
denotes the least integer strictly greater than
and
denotes the least integer greater than or equal to
.
There are two important special cases which we single out:
Corollary 5.
The maximal and minimal de Rham cohomologies at weight zero correspond to upper and lower middle perversity intersection cohomology.
|
(13)
|
Moreover, when
is even, the maximal and minimal de Rham cohomologies at weights
coincide, and again correspond to upper and lower middle perversity intersection cohomology.
|
(14)
|
From §2, these weighted de Rham cohomology spaces are identified with the nullspaces of the associated (absolute and relative) Laplacians, and we conclude that the nullspaces of
are (finite dimensional and) identified with particular intersection cohomology spaces. Note that we are not asserting anything about the nullspaces of the `ordinary' Laplacians
on the weighted spaces
; indeed, these nullspaces are either infinite dimensional when
or vanish identically when
, cf. [18] .
4 Elliptic edge operators and minimal Hodge cohomology
To proceed further in the study of these weighted de Rham complexes, we must use elliptic methods. More specifically, we still wish to study the question of when there is a unique closed extension for
on
for incomplete edge metrics, and we also wish to compute the minimal Hodge cohomology. We shall study both of these questions using the formally symmetric operator
on
. The proper context for this analysis is the calculus of pseudodifferential edge operators, and in the next subsection we review the generalities of this theory. After that we show how it applies to the specific problems at hand.
4.1 Edge operators
We now review the general theory of elliptic edge operators. This is the correct context to study
for an incomplete edge metric (and also the corresponding operator for a complete edge metric). This theory is developed fully in [18] , and we refer there for more details Fix a local coordinate system
on
and
on
, so that
is a local coordinate system in some neighbourhood of a singular stratum in
.
By including the hypersurface
, we are blowing up
in
; the resulting manifold with boundary is denoted
and its interior is denoted
.
A differential operator
on
is called an edge operator of order
if it can be expressed in the form
|
(15)
|
where the (scalar or matrix-valued) coefficients are bounded. We shall assume that these coefficients are smooth in these variables, down to
. For example, if
is a complete edge metric, then the scalar or Hodge Laplacian is an operator of this type; similarly, if
is an incomplete edge metric, then its Laplacian is of the form
, where
is an edge operator of order
.
4.1.1 Ellipticity and model operators
There is a well-defined symbol in this setting:
and we say that
is elliptic in the edge calculus provided
is invertible when
.
Ellipticity alone does not guarantee that
is Fredholm between appropriate function spaces; one must also require that certain model operators for
also be invertible. There are two such operators:
-
∙
The normal operator of
is defined by
here
enters only parametrically and the operator acts on functions on
. This operator can be regarded as
with its coefficients frozen (in an appropriate sense) at
, acting on functions (or sections of an appropriate bundle) on the space
.
-
∙
The indicial operator of
is defined by
For example, the normal and indicial operators associated to the scalar Laplacian for the complete edge metric
are
The indicial operator captures some fundamental invariants associated to
:
Definition 5.
The number
is said to be an indicial root of
at
if there exists a function
on
such that (in local coordinates where
corresponds to
)
Indicial roots may often be calculated in terms of eigenvalues for an induced elliptic operator on the fibre
, and might depend on
.
The operator
acts naturally on weighted Sobolev spaces. Let
be a manifold with complete edge metric
. For
and
, define
(By interpolation and duality, these spaces can be defined for any
.) Clearly, if
is any edge operator of order
, then
|
(16)
|
for any
. Further hypotheses, beyond the ellipticity of
are required to ensure that this mapping is well-behaved.
The first instance of this is that the indicial roots of
yield weights
for which ( 16 ) does not have closed range; these are precisely the weights
for which an indicial root
`just fails' to lie in
near
, i.e. where
for any
but
.
We denote this critical weight
associated to a given indicial root
as
. With respect to the measure
,
. However, the measure appearing in our application below is
, and
near
if and only if
, so that we shall define
.
Even when
is not equal to one of these critical values, the behaviour of the normal operator at weight
plays another very important role.
Proposition 8.
Let
be an elliptic differential edge operator of order
.
Fix
such that
for any indicial root
,
. Suppose also that
is surjective (for all
). Then ( 16 ) is essentially surjective, in the sense that its range is closed and of finite codimension. On the other hand, if
is injective on
, then any element of the nullspace of
is necessarily conormal.
There are many more refined statements one can make about the mapping properties of
. For later applications, we state only one very special result. We shall restrict to a special setting, which is what arises in our applications below. The hypothesis that the the normal operator
is surjective is equivalent to the injectivity of the normal operator for the adjoint
. This adjoint depends on the choice of measure, and we shall assume (as in our applications) that the adjoint of
on
corresponds to the same operator
on a `dually weighted' space
, for some
. The fact that
has closed range implies the existence of a generalized inverse
which satisfies
, where
is the orthogonal projector onto the cokernel. By duality, elements of this cokernel are identified with elements of the nullspace of
on
, and by the result above, these are conormal.
Proposition 9.
Let
satisfy the special assumptions of the preceding paragraph. Suppose furthermore that the interval
contains a finite set of indicial roots
,
, all of which are constant in
. Let
. Then
satisfies
where
,
, and
; where each
solves the indicial equation
, and the error term
.
The proofs rely on the construction of a pseudodifferential parametrix
for
, depending on
. This is an element in the calculus of pseudodifferential edge operators
. We do not define this calculus here, but remark only that these operators are described by specifying the precise asymptotic behaviour of their Schwartz kernels, near the diagonal and also near the boundaries and corner of
. We refer as before to [18] . (We should note also that the results stated here are slightly more general than what is written explicitly in that source because we are allowing the possibility of variable indicial roots outside the critical interval; however, these can be derived easily from the same techniques.)
4.2 Edge analysis of
We now proceed to apply the methods of the last subsection to the analysis of
.
The first tasks are the calculation of the indicial roots of
and the analysis of the normal operator
.
4.2.1 Indicial roots of
For simplicity, first consider the calculation of the indicial roots for
for the metric
on the cone
. To do this, we decompose this operator as much as possible. Thus, first regard
as a
matrix acting on pairs
, where
for each
; normalize by writing the
-form part of
as
, and similarly for
. A short calculation then shows that, acting on pairs
,
the full indicial operator is the direct sum over
of these matrices. Similarly, the indicial family is the direct sum of matrices
|
(17)
|
This can be reduced further using the eigendecomposition for
. In particular, we see that the operator in ( 17 ) is noninvertible if and only if for some eigenvalue
for
,
is singular, or equivalently
Hence the indicial roots come in pairs:
|
(18)
|
The extension of these calculations to incomplete edge metrics requires only some mild alterations. Write
, where
is a degenerating family of metrics on
. When
is a
-form on
, its pointwise norm satisfies
Denote by
the operator
on
for the metric
. By Proposition 3, with respect to the metric
on
,
hence,
where
A quick review of the definitions shows that
where all the components on the right are the corresponding operators at
.
Now let
be the bundle with fibre
obtained from
, with metric
. Decompose any form
on
as
, where
is of type
on
, and similarly for
. Thus the pointwise norms satisfy
. In terms of these decompositions and normalizations, the restriction of the operator
to pairs of
-forms on
is given by
|
(19)
|
From this expression, we see that neither
nor
appear in the indicial operator
. Hence the computation of the indicial roots is exactly the same as in the conic case; in other words, all indicial roots are of the form ( 18 ). Note, however, that the eigenvalues
may depend on
, hence the same may be true of these indicial roots.
Notice that if
and
, then
(near
) if and only if
. The indicial roots which lie near to (and above) this `
cutoff ' are the ones which cause the difference between minimal and maximal domains. We explain this later, but for now record the
Corollary 6.
The indicial roots of the operator
contained in the interval
correspond to the eigenvalues
of
on
-forms such that
. In order for this condition to be nonvacuous, it is necessary that
. Note in particular that
precisely when
and
.
4.2.2 The normal operator of
According to the discussion in the final paragraphs of §4.1, we must also study the mapping properties of the normal operator
. Before doing so, we address some `duality' issues. The main point is that
is formally symmetric on
, hence the adjoint of the (closed range) operator
|
(20)
|
is identified with
|
(21)
|
Recall also that, according to the computations of the preceding subsection, there are at most two indicial roots
in the interval
, and these are symmetric around the midpoint
. We shall assume that the metric
is such that
has no small nonzero eigenvalues, so that no other indicial roots intersect the closed interval
.
By ( 19 ),
|
(22)
|
where the first operator on the right is the analogous weighted operator on the complete cone
and the second is on Euclidean space. In fact, the identification of
with
may also be seen by naturality, since the operator on the left must equal, at
, the Hodge-de Rham operator on
with respect to
.
Proposition 10.
is injective, and hence ( 21 ) has a finite dimensional nullspace consisting of conormal forms
-
Proof.
Suppose that
is in the nullspace of this operator. Take the Fourier transform in the
direction; denoting the dual variable by
, then whenever
we can rescale, setting
,
. Then
where
is Clifford multiplication
. Apply
to this equation to deduce that
It is not hard to show, cf. [18] , that any solution of this equation either grows or decays exponentially as
, and the
hypothesis prohibits the former. Furthermore, solutions are polyhomogeneous as
, and hence decay at some indicial weight
with
. Hence both
and
decay like
, and in particular are still in
. This means we can integrate by parts to obtain
all boundary terms vanish. This gives
, as desired.
When
, the problem reduces to showing that
has no nullspace in
on the entire cone
, which is even more easily verified to be true (e.g. by separation of variables).
Following the discussion from the end of §4.1, if
contains no indicial roots for
,
is injective on
.
There are no forms in the nullspace of the normal operator which lie in
or
. From this it is possible to show that
, i.e. such harmonic forms are conormal. Unfortunately, these results rely on a slightly more elaborate parametrix construction than is available in the literature, so at present we are only asserting this informally. At the end of the next subsection, however, we show that forms in
are conormal.
4.3 Closed extensions of
It is a general fact that closed extensions of
on
are in bijective correspondence with the self-adjoint extensions of
on this space. To see this, first note that if
is any closed extension of
, and if
is its Hilbert space adjoint, then
is a self-adjoint extension of
. Conversely, any self-adjoint extension of
determines an associated closed extension for
, cf. [2,Lemma2.3] . We summarize this in the
Proposition 11.
If
has more than one closed extension on
, then
has more than one self-adjoint extension on this space; equivalently, if
is essentially self-adjoint, then
.
Notice that since
always has closed extensions,
always has self-adjoint extensions.
However, if
is not essentially self-adjoint, then it will have closed extensions which are not necessarily self-adjoint; the relationship between these and the closed extensions of
is somewhat more complicated, and we shall not attempt to describe it.
Theorem 7.
The symmetric operator
is essentially self-adjoint on
if and only if there exists no indicial root for
in the interval
. As explained earlier, this is equivalent to the nonexistence of small eigenvalues
for
on
-forms such that
. In this case
also has a unique closed extension.
-
Proof.
Assume that there are no small eigenvalues, as described in the statement of the theorem. Fix a parametrix
for
relative to the space
. This is an element of order
in the edge calculus. If
, then
, and applying
gives that in fact
. We recall that in general,
would be the sum of two terms, the first corresponding to these small indicial roots and the second an error term in
, but by our hypothesis, the former of these is absent. It is now straightforward to check that
may be smoothly approximated in the
-graph norm, i.e. that there exists a sequence of smooth compactly supported forms
such that
,
in
. This shows that
.
Conversely, if there do exist indicial roots in the critical range, then these may be used to construct nontrivial elements in
, and by the general abstract theory, there will be more than one self-adjoint extension of
. Because this is not central to our discussion, we leave details to the reader (and refer to [10] for a thorough discussion of the conic case).
In the conic case (when
) this result is due to Cheeger [3] ; cf. also [16] . The analysis needed in that case is simpler than the edge analysis used here, though this is not apparent `on the surface'.
We turn now to a description of
in the more general case where this vanishing condition is no longer satisfied. By definition, this domain is the set of all
such that
. We have already remarked that the mapping ( 20 ) has closed range, which we denote
. Its cokernel is finite dimensional since, by Proposition ( 10 ), the adjoint mapping ( 21 ) has a finite dimensional nullspace. Hence we can choose a generalized inverse
this is a pseudodifferential edge operator of order
which satisfies
on
, where
is the orthogonal projector onto the cokernel. The condition
is equivalent to
for all
in the nullspace of ( 21 ). Notice also that
is dense in
. Therefore,
where
is the nullspace of ( 20 ).
Finally, if
, then
for all
, i.e.
for all
. Using the density statement above, this shows that
lies in the nullspace of ( 21 ), and is thus conormal.
We note in conclusion that any
has a `weak' asymptotic expansion of the form
where
, again in a suitable weak sense. We refer to [18] for more details.
4.4 The minimal Hodge cohomology
We are now in a position to prove the
Theorem 8.
Let
be a manifold with an incomplete edge metric
. Then
In particular, when
,
-
Proof.
Recalling that
is quasi-isometry invariant, we may as well assume that
has no small nonzero eigenvalues, in the sense of the preceding subsections. If there are no small eigenvalues at all, i.e. either when
(which holds, for example, when
and
is odd), or else if there exists
but
, then the result follows directly from what we have already done, since then
and and
are all equal, cf. Corollary 3 . Thus we suppose that there exists
such that
.
According to Theorem 4 , the space appearing on the right in the statement of this theorem, for
even or odd, is identified with
. For simplicity, we denote it as
.
We claim first that there is a natural injective map
To see this, recall that any form
naturally represents a class in
. If
in
, then
for some
. But this would imply that
since
.
This proves the claim.
The issue, then, is to prove that any class
is represented by an element of
. Choose a representative
,
, for this class. Now use a generalized inverse
for
(acting on its maximal domain). According to Propositions 8 and 9, this gives
and an element of the cokernel, which by duality corresponds to an element
, such that
. (Recall that
is the lower of the two indicial roots in the critical interval.) The theorem will follow once we show that
. Now
Integrating by parts formally this should vanish, so it remains to show that each integration by parts is valid. First,
since both terms are conormal and
. Similarly
since
. To show that the remaining term vanishes, observe that
where
,
. Writing
, then
and in the nullspace of
. A closer inspection of the equation
shows that
(The operators
,
are the differential and codifferential for the (fibre-harmonic) projected complex for the Riemannian submersion metric
.) In any case, the identity
is now immediate. This proves the remaining assertion, and hence the theorem.
5 Hodge theory for complete edge metrics
We are also able to determine the dimensions of the spaces of the spaces
when
is a manifold with a complete edge metric. Unlike the incomplete case, in certain degrees this Hodge cohomology may be infinite dimensional, i.e. there is an infinite dimensional space of
harmonic forms. The simplest example of this is when
is the
-dimensional hyperbolic space, or indeed any conformally compact manifold, and
, cf. [20] .
Theorem 9.
Let
be a manifold with a complete edge metric. Let
be the compact stratified space defined in §3. Suppose that
is not of the form
where
. Then
In this case, the
signature theorem for
is the same as the
signature theorem for
endowed with the conformally equivalent incomplete edge metric
. In all other cases, where
does have this form,
is infinite dimensional.
-
Proof.
There are several viable ways to proceed: one could use a parametrix construction based on the edge calculus to do a global Hodge theoretic argument as in [20] ; one could also, as in the incomplete case, use sheaf theory, calculations of local cohomologies, etc., ab initio; we take a shorter intermediate route, reducing to the incomplete edge case using the conformal invariance of the space of middle degree
harmonic forms.
It suffices to consider forms of degree
. Define
with
.
For later use, we also set
,
. Now, since
,
is the middle degree on the manifold
. Endow
with the product metric
(where
is the standard metric on
); in a neighbourhood of
In other words,
is still a complete edge metric with the same base
as
, but with fibre
. There is a corresponding incomplete edge metric
on
. We denote its compactification, obtained by pinching the fibres
at the boundary, by
.
We first claim that
|
(23)
|
The verification is straightforward; by conformal invariance of the
condition and the operator
in the middle degree, harmonic forms on the complete manifold are in the maximal domains of
and
(and indeed their nullspaces) on the incomplete manifold, and conversely.
By the
Künneth theorem,
On the other hand, while we have not shown how to compute
in general, and indeed have noted that it is sometimes infinite dimensional, we now show that under certain hypotheses, it equals
; thus we can then apply Theorem 3 to calculate the right side of ( 23 ) as
. (Under these hypotheses, the intersection cohomologies with upper or lower middle perversity are the same, so we just write
.) So, let us suppose that either
is even, or else if
is odd then
(and hence
) is trivial; note this last condition is automatic when
. We claim that under these conditions, the minimal and maximal de Rham complexes on
coincide, and thus
. Indeed, this follows directly from Corollary 3 : we have
even; if
is even, then
is odd, while if
is odd, then
Taking these facts together, and assuming this vanishing of the fibre cohomology when
is odd, we have proved that
|
(24)
|
It remains to compute the final term on the right of ( 24 ). We decompose
where
is the cone bundle over
with fibre
and boundary
. Of course,
On the other hand,
To prove this, we return to the sheaf-theoretic description. For product neighbourhoods
, we have
Note that we have combined the conditional inequality in this last step, which should depend on the parity of
, into one condition. This condition is correct as stated if
is odd; if
is even, then one would expect the condition
or
, depending on whether one was using upper or lower middle perversity. However, the hypothesis
guarantees that we get the same result in either case. Thus we see that the spectral sequence whose hypercohomology computes the intersection cohomology of
decouples into two noninteracting pieces.
We have now proved that for a fixed
, assuming the hypotheses above,
We would, of course, like the summands to be equal separately; this can be seen simply by noting that the correspondence takes place on the level of forms, and we can separate out the terms with like degree. It remains to show that in the remaining cases, i.e. when
is odd and
,
,
is infinite dimensional. This follows from two assertions:
that
is in the essential spectrum of
, and that there is a spectral gap at
for
. We content ourselves with sketching the proofs briefly. The first step relies on the observation that up to quasiisometry, some neighbourhood of infinity looks like the product of half of a hyperbolic space and a compact manifold, specifically
. Since
is even, there is an infinite dimensional family of
harmonic forms on the first factor. Since
, we can take suitable truncations of these, coupled with harmonic forms of degree
on
, to produce a Weyl sequence on
. For the second step, we construct a parametrix in the edge calculus for
. Its normal operator is given by
Both of these operators have spectral gaps at zero (even when
), so it is possible to construct a parametrix for
with compact remainder when
is small but nonzero. We refer to [17] and [20] for more complete descriptions of such proofs in a slightly simpler context.
References
-
A. Borel et al., Intersection Cohomology, Birkhauser (1994).
-
J. Brüning and M. Lesch, Hilbert Complexes, J. Func. Anal. 108, 88-132, (1992).
-
J. Cheeger, On the spectral geometry of spaces with cone-like singularities, Proc. Nat. Acad. Sci. 76 (1979), 2103-2106.
-
J. Cheeger, On the Hodge theory of Riemannian pseudomanifolds in Geometry of the Laplace operator Proc. Symp. Pure Math Vol. XXXVI, Amer. Math. Soc., Providence, (1980),91-146.
-
J. Cheeger and X. Dai,
cohomology of a non-isolated conical singularity and nonmultiplicativity of the signature. Preprint.
-
J. Cheeger, M. Goresky and R. McPherson,
-cohomology and intersection homology of singular algebraic varieties in Seminar on Differential Geometry, Annals of Mathematical Studies 102, Princeton Univ. Press, Princeton, N.J., (1982), 303–340.
-
X. Dai, Adiabatic limits, nonmultiplicativity of signature, and the Leray spectral sequence, J. Amer. Math. Soc. 4 (1991), no. 2, 265–321.
-
G. deRham, Differentiable manifolds. Grundlehren der Mathematischen Wissenschaften, 266, Springer-Verlag, Berlin, (1984).
-
M. P. Gaffney, A special Stokes's theorem for complete Riemannian manifolds, Ann. Math. 60 no. 2, (1954), 140–145.
-
J. Gil and G. Mendoza, Adjoints of elliptic cone operators, Amer. J. Math 125 (2003), 357–408.
-
M. Goresky and R. MacPherson, Intersection homology theory, Topology, 19 (1980), no. 2, 135–162.
-
M. Goresky and R. MacPherson, Intersection homology II, Invent. Math. 72 no. 1 (1983), 77–129.
-
T. Hausel, E. Hunsicker, and R. Mazzeo, Hodge cohomology of gravitational instantons. Duke Math. J. 122 no. 3 (2004), 485–548.
-
C. Hodgsen, S. Kerckhoff Rigidity of hyperbolic cone-manifolds and hyperbolic Dehn surgery, J. Diff. Geom. 48 (1998), 1–59.
-
E. Hunsicker, Hodge and signature theorems for a class of manifolds with fibration boundary, in preparation.
-
M. Lesch, Differential Operators of Fuchs Type, Conical Singularities, and Asymptotic Methods. Teubner Texte für Mathematik 136, Teubner–Verlag, Leipig (1997).
-
R. Mazzeo, The Hodge cohomology of a conformally compact metric, J. Diff. Geom. 28 (1988), 309–339.
-
R. Mazzeo Elliptic theory of differential edge operators I, Comm. PDE. 16 (1991), no. 10, 1616–1664.
-
R. Mazzeo and R. Melrose, Pseudodifferential operators on manifolds with fibred boundaries in “Mikio Sato: a great Japanese mathematician of the twentieth century.”, Asian J. Math. 2 (1998), no. 4, 833–866.
-
R. Mazzeo and R. S. Phillips, Hodge theory on hyperbolic manifolds, Duke Math. J. 60 (1990), no. 2, 509–559.
-
R. Mazzeo and H. Weiss, The Teichmüller theory of conic surfaces, in preparation.
-
G. Montcouquiol, Rigidité infinitésimale de cônes-variétés Einstein à courbure négative. arXiv:math.DG/0503195
-
L. Saper
cohomology of locally symmetric spaces I, arXiv:math.RT/0412353
-
L. Saper On the cohomology of locally symmetric spaces and their compactifications, in Current developments in mathematics 2002, Eds. D. Jerison et al., International Press (2003), 219-289.