April 5, 2005
Renormalizing Curvature Integrals on Poincaré-Einstein manifolds
Pierre Albin
Department of Mathematics, Stanford University E-mail address : pierre@math.stanford.edu
-
Abstract.
After analyzing renormalization schemes on a Poincaré-Einstein manifold, we study the renormalized integrals of scalar Riemannian invariants. The behavior of the renormalized volume is well-known, and we show any scalar Riemannian invariant renormalizes similarly. We consider characteristic forms and their behavior under a variation of the Poincaré-Einstein structure, and obtain, from the renormalized integral of the Pfaffian, an extension of the Gauss-Bonnet theorem.
1 Introduction
Let
be the interior of a compact manifold with boundary
. A boundary defining function (hereafter, a bdf ),
, is a smooth nonnegative function on
that vanishes precisely at the boundary, with non-degenerate differential there. A metric
on
is said to be conformally compact if there is a bdf
such that
extends to a nondegenerate metric on
. Conformally compact Einstein metrics, also known as Poincaré-Einstein or PE metrics, have been the object of many recent studies by geometers and physicists.
A conformally compact metric on the interior of a manifold determines a conformal class of metrics on the boundary but does not single out any particular metric within that class. In analogy to the well-known relation between the hyperbolic geometry of the ball and the conformal geometry of the sphere, Fefferman and Graham [12] proposed studying the conformal geometry of the boundary of a manifold through the geometry of a PE metric on the interior.
More recently, the string-theory community has been interested in these manifolds because of the role they play in the Anti-de-Sitter/Conformal field theory (AdS/CFT) correspondence conjectured by Maldacena. Witten [34] clarified the conjecture as an equality of partition functions (normalizing factors for the probabilities of states). Computing the partition function of the conformal field theory (in the massless case) involves evaluating the Einstein action. For an Einstein manifold, this is a multiple of the volume which, however, is not finite. Witten showed that the volume could be renormalized while preserving covariance. Hennington and Skenderis carried out explicit computations in low dimensions and verified that this notion of renormalized volume is consistent with computations from, for instance, superconformal SU(4) gauge theory. Later, Graham and Witten [16] , carried out a similar analysis of the area of (minimal) submanifolds of PE manifolds.
A mathematical survey of these ideas appeared in [13] . The definition of renormalized volume (see section
2.1 below) involves the choice of a bdf. One can restrict the choice of bdf to a natural class – the “special” bdfs – for which the metric near the boundary has the particularly nice form
|
(1.1)
|
The corresponding renormalized volume of an even dimensional PE manifold is independent of the choice of bdf from this smaller set. In odd-dimensions, even this restriction does not produce an invariant, and different choices of special bdfs yield different renormalized volumes producing the so-called “conformal anomaly”.
In sections
3 -
4.1 , we analyze the scalar Riemannian invariants of a PE manifold. These are all complete contractions of the curvature and its covariant derivatives. They include the volume and also the heat invariants. We show that one can define renormalized integrals of these invariants, and their dependence on the choice of bdf is just like that of the volume in the following sense (cf.
Theorem 4.2 below).
Theorem 1.1.
Let (M,g) be a Poincaré-Einstein manifold, and
a scalar Riemannian invariant of
. Using any given special bdf
, we may define the renormalized integral of
over
, denoted by
If
is even-dimensional, this renormalized integral is independent of the choice of special bdf. If
is odd-dimensional, and
is not integrable, its renormalized integralnecessarily depends on the choice of special bdf. In this case,
determines a “residue” integral on
which is independent of the choice of special bdf.
Of particular interest is the renormalized integral of the Pfaffian. Recall that the Pfaffian is a natural density that integrates to the Euler characteristic on closed even-dimensional manifolds.
On an even-dimensional PE manifold the integral of the Pfaffian diverges, but the renormalization process recovers the same topological information as in the compact setting:
Theorem 1.2.
On an even dimensional Poincaré-Einstein manifold, using a special bdf to renormalize the integral of the Pfaffian,
|
(1.2)
|
This generalizes the four-dimensional formula of Anderson [2] ,
|
(1.3)
|
where
is the Weyl curvature tensor and
is the renormalized volume. Notice that, by virtue of the Einstein condition, the Pfaffian is a polynomial in the Weyl curvature (see Lemma 4.4 for an explicit formula). Another particular case is the formula of Epstein [32] ,
|
(1.4)
|
valid for any convex cocompact hyperbolic manifold, which a fortiori is PE. In this context, we should mention the recent preprint of Chang, Qing, and Yang [8] . Their study of Branson's
-curvature allowed them, through a result of Alexakis, to show that
|
(1.5)
|
In this formula,
is a full contraction of the Weyl tensor and its covariant derivatives for the metric
(
a well-chosen special bdf ), and
is the renormalized volume. Like 1.2 this generalizes Anderson's four-dimensional Gauss Bonnet theorem, as well as Epstein's formula 1.4 .
The integrability issues in [8] are dealt with by passing to
;
is compact, so the integral of
needs no renormalization. A result of Fefferman and Graham, and the choice of
, produce the renormalized volume as a boundary term.
In section
4 , after rewriting the Pfaffian of an Einstein manifold in terms of the Weyl and scalar curvatures, we note that our formula 1.2 is similarly an integral of a complete contraction of the Weyl curvature plus a multiple of the renormalized volume (the same multiple as in 1.5 ). In contrast to 1.5 , we can explicitly identify the Weyl curvature integrand but its integral requires renormalization.
The proof of Theorems 1.1 and 1.2 depend only on the Fefferman-Graham expansion of the metric (see section
2.1 below), whereas that of 1.5 requires the Einstein condition to apply Alexakis' result.
It would be very interesting to better understand how these formulas are related.
Elsewhere [1] , we consider the heat kernel on a PE manifold and apply renormalization to study index theory. Theorem 1.1 shows that the coefficients of the small-time asymptotics of the trace renormalize independently of the choice of special bdf. In fact, one can show that the trace of the heat kernel itself, for any fixed positive time, renormalizes in this way. Furthermore, using renormalizationas
, we are able to prove an index theorem for the de Rham operator. It is well-known (see [24] ) that the spaces of harmonic forms on an even-dimensional conformally compact manifold are finite-dimensional except for middle degree forms which always form an infinite-dimensional space.
We can use renormalization to define the renormalized index of the de Rham operator (a priori a real number) and show that:
|
(1.6)
|
Thus for PE manifolds, 1.2 shows that the renormalized index of the de Rham operator is the Euler characteristic. The proof of these results requires much more analysis than that of Theorems 1.1 and 1.2 and will not be presented here.
After considering the characteristic numbers of a PE manifold in section
4 , we study their behavior under variations of the PE structure in section
5 . A family of metrics
with Fefferman-Graham expansions (see section
2.1 below) imposes a similar expansion on
, the infinitesimal variation of
, as we show in Proposition 5.3 . We analyze the effect of such a variation on the Pontrjagin classes, the renormalized volume and the Pfaffian. For the renormalized volume we recover theorems of Anderson 1.7 and Graham-Hirachi 1.8 . We show in Proposition 5.6 , making strong use of the recent work of Labbi [21] on Weyl volume of tubes invariants, that the variation of the renormalized integral of the Pfaffian vanishes. The absence of boundary terms is due to the expansion of
, while the vanishing of the interior terms follows from a generalized Bach-Lanczos identity. Naturally, the vanishing of the variation of the Pfaffian is consistent with equation 1.2 .
Theorem 1.3 ([2] , [15] ).
A variation of the PE structure on
,
, that preserves the value of the scalar curvature induces a variation of the Pontrjagin characteristic numbers equal to the integral of the appropriate Chern-Simons form on the boundary.
The induced variation of the renormalized volume on an even-dimensional manifold is given by
|
(1.7)
|
while that of the residue of the volume on an odd-dimensional PE manifold is
|
(1.8)
|
In these formulas,
and
are the first odd term and first log term, respectively, in the expansion of the metric
, while
is the infinitesimal variation of
at
.
The proof of Theorem 1.2 readily generalizes to other asymptotically regular geometries. For example, in the context of asymptotically Euclidean metrics we obtain a formula for the defect of the topological Euler characteristic and the
-Euler characteristic involving the Weyl volume of tube invariants of the boundary at infinity. For this and other instances, as well as the proof of 1.6 , we refer the reader to [1] .
1.1 Acknowledgements
This work forms part of my thesis. I am very grateful to my advisor, Rafe Mazzeo, for sharing his great erudition and insight. Throughout this work, I received support from his NSF grant DMS-0204730. I would like to thank Tom Branson for his invitation to participate in the conformal geometry program at the Erwin Schrödinger Institute in Vienna in spring of 2004. I was fortunate to coincide there with him and Robin Graham, and I would like to thank them both for very interesting conversations. I am also grateful to the ETH in Zurich for its hospitality during the summer semester of 2004, and to the Starbucks branches in Zurich and Palo Alto where most of this work was carried out.
2 Renormalization
2.1 The Fefferman-Graham expansion
Suppose that
is a conformally compact manifold, and
a bdf. It is too optimistic to expect
to be a smooth metric on
, and indeed it is well known that
terms arise naturally in its expansion [12] . Functions with an expansion of the form
with
smooth functions independent of
, are known as “polyhomogeneous conormal” or phg, see
2A in [25] . We will always assume at least that the metric
is continuous and phg. For a discussion of the regularity of the metric, see the recent survey [4] .
Such metrics are always complete, and any non-trapped geodesic approaches a point on the boundary. The sectional curvatures along any geodesic approaching the boundary all converge to
, and, as a function on the boundary, this is independent of the choice of bdf. A metric is called asymptotically hyperbolic if
on the boundary. These metrics include, and are asymptotically modeled by, hyperbolic metrics. They were introduced in [26] and [24] where their resolvents and spectra were studied, respectively.
A choice of bdf determines a metric on the boundary with different choices yielding different but conformally equivalent metrics. Conversely, a boundary metric does not in itself determine a bdf. On an asymptotically hyperbolic manifold, boundary metrics within the conformal class are in one-to-one correspondence with “special” or “geodesic” boundary defining functions. A bdf
is special if
on a neighborhood of the boundary. The details can be found in Lemma 2.1 of [13] ; see also [14] .
A choice of special bdf induces, through the flow generated by
, an identification of a neighborhood of
with
with metric
where
is a family of metrics on
. Fefferman and Graham [12] showed that if
is a special bdf on an
-dimensional Poincaré-Einstein manifold and
is sufficiently regular, then the expansion of
below
is determined by the Einstein condition. Thus if
is even,
has an expansion of the form
|
(2.1)
|
where
are locally determined by
and
. For
odd, the analogous expansion is
|
(2.2)
|
where now
,
, and
are locally determined by
and furthermore
.
Following [13] , we write the volume form as
|
(2.3)
|
and note that the expansions above imply
|
(2.4)
|
where, for
,
is a locally determined function on
and
if
is even.
We will say that a phg expansion is
if there are no log terms or terms with odd exponents below
. Thus the metric on a Poincaré-Einstein manifold is, in suitable coordinates,
. Note that the product of two such expansions is again
. Graham observed that:
Theorem 2.1.
If
is a special bdf on a Poincaré-Einstein manifold, then
is even
, and the coefficients below
are locally determined.
We will extend this in Theorem 4.2 to full contractions of the curvature and its covariant derivatives. In the rest of this section, we study its implications for renormalization.
As we mentioned above, physical considerations suggested the possibility of renormalizing the volume. This is accomplished by making use of the expansion of the volume form. Indeed, in even dimensions
and in odd dimensions
where the coefficients
and
are integrals over
of local curvature expressions of the metric
.
is known as the renormalized volume, it depends a priori on the choice of special bdf.
2.2 Renormalization Schemes
The renormalization carried out above,
is known as Hadamard regularization. It coincides with the “b-integral” used in Melrose's proof of the Atiyah-Patodi-Singer index theorem [28] .
In a subsequent work, Melrose and Nistor use an alternate method known as Riesz regularization to renormalize integrals ([29] ; see also [33] ). Given a bdf, Riesz regularization of the integral of a density
is defined by meromorphically extending
|
(2.5)
|
and taking the finite part at
; symbolically:
Both of these approaches make heavy use of the expansion of the integrand; they are only defined on phg densities. As a preliminary step we assume that the volume form has been written as in 2.3 and the boundary integral has been carried out. Thus we only need to consider one-dimensional integrals. Furthermore, making use of linearity, we can localize to a neighborhood of zero, say
.
Thus localized, it is easy to compare Hadamard and Riesz renormalizations directly on phg densities. Once the integral along the boundary has been carried out,
while
We can trace this difference back to the fundamental theorem of calculus. Assume that
|
(2.6)
|
Then
We define a renormalization scheme
to be a linear functional on phg densities such that, for
as in 2.6 ,
where
is a linear function of
, hence depends on the choice of bdf. Given any two renormalization schemes,
and
, it is easy to see that
|
(2.7)
|
The difficulties in comparing the same renormalization scheme for two different bdfs come from the transformation of the coefficients
. The situation is greatly simplified by assuming that there are no singular
terms in the expansion of
. Note that the absence of singular
terms is independent of the choice of smooth bdf.
Proposition 2.2.
Let
be density on
, phg with respect to
, and with no singular log terms. The coefficient
of
in the expansion of
is given by
and is independent of the choice of bdf.
If
and
are any two renormalization schemes, then
is independent of the choice of bdf.
-
Proof.
It is easy to see that
. To see the independence from the choice of bdf, let
be any other bdf. Let
,
be the zeta functions as defined in 2.5 . Note that
Due to the absence of singular
terms, the meromorphic continuation of
has at most a simple pole at
. Hence
extends to be holomorphic at
, and the residues must be equal.
The final statement follows directly from 2.7 . □
In what follows, we will only seek to renormalize integrals of densities without singular
terms, and thus we shall unabashedly consider only Riesz renormalization, allowing for much simpler computations. A case in point is the proof of Proposition 2.2 which yields the following proposition, taken from [29] .
Proposition 2.3.
Let
be a density without singular log terms, and let
and
be two bdfs. The difference between the renormalized integrals with respect to
and
is given by
|
(2.8)
|
where
denotes to the term in the expansion of
of homogeneity
in
.
-
Proof.
Simply note from the above computation that
|
(2.9)
|
□
2.3 Renormalization on Poincaré-Einstein manifolds
In general, for a density to admit a renormalization independent of the choice of bdf, it needs to actually be integrable. Nevertheless, we will see that on Poincaré-Einstein manifolds there is a rich class of densities renormalizing independently of the choice of special bdf. The reason is twofold.
On the one hand, the metric has the Fefferman-Graham expansion 2.1 , 2.2 . On the other hand, we have the following lemma from [14] (see also [18] ).
Lemma 2.4.
Let
be an asymptotically hyperbolic manifold, and
a special bdf. For any odd number
, the expansion of
is even below
if and only if for any other special bdf
the expansion of
is even below
.
-
Proof.
The condition
imposes
|
(2.10)
|
In terms of the expansion of
in
, this does not restrict
. On the other hand, it is easy to see that if
is even below
, then the first term in 2.10 is odd below
, while the other terms are odd below
, hence for
,
. Thus the first odd term is
, and it is given by:
This is zero for arbitrary
precisely when
. □
Recall from 2.3 and 2.4 , that the volume form of an even-dimensional PE manifold has an expansion of the form
Note that
has no residue, and that the lemma guarantees that changing to another special bdf will not produce a residue. In view of Proposition 2.3 , this means that the renormalized integral of the volume is independent of the choice of special bdf. This has the following immediate generalization.
Theorem 2.5.
Let
be a Poincaré-Einstein manifold,
a special bdf, and
a density of the form
where
is
. Then, if
is even,
is independent of the choice of special bdf. If
is odd, and
is another special bdf, then
Though in any case,
is independent of the choice of bdf.
One might initially expect the renormalized volume, coming from a conformal theory, to depend only on the conformal class. Its definition involves the choice of a bdf and this breaks the conformal invariance, explaining the appearance of a conformal anomaly in odd dimensions. In even dimensions the conformal invariance is restored after renormalization. To apply this theorem to the scalar Riemannian invariants of a PE manifold, we need to understand the expansions of their curvature tensors.
3 Curvature of an Asymptotically Hyperbolic Manifold
In this section we develop some of the geometry of an asymptotically hyperbolic manifold. Though our applications in the next section will all be to PE manifolds, we will not assume that the metric is Einstein in this section. We begin by introducing the coordinates in which our computations will be carried out.
Let
be an asymptotically hyperbolic manifold, and
a special bdf. Pick a point on the boundary
, and choose coordinate vector fields
for
by exponentiating first on the boundary, then into the manifold. That is,
form a normal coordinate chart for
centered at
and are extended into the interior of
along geodesics normal to
. In this way, with
as the
coordinate, throughout the chart we have
The Christoffel symbols and components of the curvature tensor of
in this coordinate chart will be denoted by
and
, respectively. We shall consider the structure of
using the frame
. We will use the letters
,
,
, and
as indices varying between
and
and
,
,
, and
to denote indices between
and
. Note that
for all
, but
|
(3.2)
|
Also, in this frame
|
(3.3)
|
The Levi-Civita connection of
is closely related to that of
. Indeed, Theorem 1.159 of [6] specializes to:
|
(3.4)
|
In terms of the frame
above, in analogy with the Christoffel symbols, we define
by the equation
.
Lemma 3.1.
With
as above,
|
(3.5)
|
Furthermore, recalling 3.1 ,
Moving on to the curvature, define
by the equation
and, directly from
, obtain
|
(3.6)
|
An expansion of
in
induces expansions of
and
. Note that the expansion of
in
below
uses the expansion of
below
. An advantage of using
instead of
is that it does not lower order of homogeneity and hence, for example, the expansion of
below any
depends only on the expansion of
below
. Indeed, this is still true of the curvature and even its covariant derivatives. We shall prove this with an eye towards Poincaré-Einstein metrics.
Theorem 3.2.
Let
be an asymptotically hyperbolic manifold, and
a special bdf. If
has a phg expansion that extends continuously to
(thus
can occur for
, but not for
), then so does the curvature and any of its covariant derivatives.
Furthermore, the terms below
in the expansion of any covariant derivative of the curvature are determined by the terms below
in the expansion of
.
Remark. A consequence of using a frame for
built from a frame for
, i.e.,
is that we raise and lower indices using
instead of
. For instance,
| |
Hence the conclusion of the theorem does not change if we raise or lower indices.
-
Proof.
As stated before the theorem, this is a simple consequence of using
instead of
. Note that
and that
, so the formulas 3.5 , 3.6 show that the theorem holds for
and
. For the covariant derivatives of the curvature we use induction and the formula
|
(3.7)
|
□
For the PE context, one is interested in a metric with an expansion containing only even powers of
below some term, say
. To organize our discussion of such metrics, define
of a function to be
if the function is
below
and
if it is
below
.
is clearly a multiplicative homomorphism among functions with such expansions.
Corollary 3.3.
In the above context, assume that the expansion of
below
consists entirely of even exponents. Then the expansion of each
is either even or odd below
. Indeed, if
is the number of
's among the indices of
, it satisfies
|
(3.8)
|
If
is a full contraction of the curvature and its covariant derivatives (i.e., a scalar Riemannian invariant), then
.
-
Proof.
Because of Theorem 3.2 , we may truncate the expansion of
below
so that it consists entirely of even powers of
. Our formulas for
in Lemma 3.1 show that
For any function
,
so a close look at the formula for the curvature 3.6 shows that the parity of
is just the parity of the number of
's among
. As we have chosen a chart where
,
which proves the theorem in the case
. Induction and the formula 3.7 finish the proof of 3.8 . Note that once we know 3.8 it is clear that raising or lowering indices in the curvature or any of its covariant derivatives does not change the value of
.
Consider now
, a full contraction of the curvature and its covariant derivatives. We know that
of any such contraction is the product of
of the individual factors, each of which is
raised to the number of derivatives plus the number of
's among the indices. Hence
of the full contraction is
raised to the total number of
's appearing among the indices plus the total number of derivatives among all of the factors. The total number of derivatives is necessarily even, as is the total number of
's among the indices, since the indices are paired together to form the contraction. This proves that any full contraction has only even exponents below
. □
4 Invariants of a Poincare-Einstein Manifold
4.1 Scalar Riemannian Invariants
Throughout this section
is a Poincaré-Einstein manifold, and
will always denote a special bdf. Recall from section
2.1 that, in this context, the expansion of
is even below
(regardless of the parity of
). Let
be any scalar Riemannian invariant (i.e., a full contraction of the curvature and its covariant derivatives). Corollary 3.3 guarantees that
is also even below
. Our first task is to understand the effect of the
term in the expansion of
on the
term in expansion of
.
Recall that the frame
is centered at a point
. We may assume that, at
,
is the identity and
is diagonal, say
. We will work under these assumptions and compute the expansion of
at the point
.
Lemma 4.1.
Let
be an asymptotically hyperbolic manifold, and
a special bdf. Assume the metric has an expansion of the form
and let
be a full contraction of the curvature and its covariant derivatives. Then It has a similar expansion
wherein the dependence of
and
on
and
is only through
and
.
-
Proof.
For our purpose, we may consider
given by:
Indeed, note that we do not need to compute the
term in the expansion of
, only its contribution from the
term in
, and interactions between say
and
do not occur below
.
We will compute at the point
as discussed before the statement of the lemma. Say that a function,
, is respectable if the
term is a constant and the contribution of the
term in the expansion of
to the
term in the expansion of
is only through a (constant) linear combination of the components of
and
. It is easy to see from the formulas in Lemma 3.1 that
is respectable.
Note that if
is respectable, then so are
and
. Also, the product of respectable functions is again respectable. It follows that any component of the curvature computed in such a frame is respectable, and similarly, components of covariant derivatives of the curvature are respectable.
Thus in the contraction we are interested in, the individual terms are respectable. We finish by noting that contracting a linear combination of components of
and
yields a linear combination of
and
. □
We can now state our generalization of Theorem 2.1 .
Theorem 4.2.
If
is a special bdf on a PE manifold, then any full contraction of the curvature and its covariant derivatives has an even expansion
, and the coefficients below
are locally determined. The behavior of scalar Riemannian invariants under renormalization is thus given by Theorem 2.5 . In particular, on an even-dimensional PE manifold, all scalar invariants renormalize independently of the choice of special bdf.
-
Proof.
Recalling the expansions 2.1 and 2.2 , Theorem 3.2 establishes the theorem below
, and we use Lemma 4.1 to deal with the
term. □
As is clear from the proof of the theorem, the Einstein condition only enters through the Fefferman-Graham expansion and the result holds whenever
is even below
as long as
.
Also recall, from Proposition 2.3 , that changing from one special bdf
to another one,
, changes the volume by
. For a PE metric, this term vanishes due to the Fefferman-Graham expansion of the metric, but if the metric were “any less even” this integrand would certainly not be zero in general. Even expansions of asymptotically hyperbolic (not necessarily Einstein) metrics are related in [18] to the domain on which the resolvent is meromorphic with poles of finite rank. It would be interesting to find a direct connection between the extent of this domain and the renormalization of the volume, for example, without going through the expansion of the metric.
4.2 Pontrjagin Classes
We start by reviewing the conformal invariance of the Pontrjagin classes. In dealing with these, it is useful to think of the curvature as an endomorphism-valued 2-form. On any manifold, define the 2-form
by
where
is computed with respect to a frame
with dual frame
. In the same way, we define the 2-form
replacing the components of
above with the corresponding components of the Weyl curvature tensor. It is easy to see that the forms
are conformally invariant 2-forms.
The following theorem was first shown by Avez [5] , though the conformal invariance of the Pontrjagin classes had already been established by Chern and Simons.
Theorem 4.3.
On any Riemannian manifold
In particular, on a conformally compact manifold, any characteristic number built up from Pontrjagin classes is integrable and equals the corresponding characteristic number of
.
4.3 The Pfaffian
On an even-dimensional manifold, we have one more characteristic class, the Pfaffian. As a PE scalar Riemannian invariant, we know that its integral renormalizes independently of the choice of special bdf, and in fact (Theorem 4.5 below)
|
(4.1)
|
As with any scalar Riemannian invariant on an Einstein manifold, the Pfaffian is a polynomial in the scalar and Weyl curvatures. We will identify this polynomial in terms of the “volume of tubes” invariants of the Weyl curvature and begin by reviewing these.
Weyl (see [17] ) derived a formula for the volume of a tube of small radius
around a
-dimensional submanifold
of
:
where
is the volume of the unit ball in
. Weyl showed that the coefficients are intrinsic to
by identifying them with integrals over
of complete contractions of the curvature and its covariant derivatives. Donnelly [11] later proved that they are universal linear combinations of the heat invariants of
.
Weyl's theorem was used in the first proofs of the Gauss Bonnet theorem for a general even-dimensional compact manifold (independently by Allendoerfer and Fenchel for submanifolds of
, then by Allendoerfer and Weil in general). The link is
One expression for
comes from considering the curvature as an endomorphism valued 2-form as in
4.2 . Indeed, the coefficients are multiples of the Lipschitz-Killing curvatures (see [9] )
|
(4.2)
|
To obtain another expression for
, it is convenient to work in the formalism of double forms as set out in [17] and especially [19] . In this way we will obtain the aforementioned polynomial for the Pfaffian. Define
We consider the metric and the curvature as (symmetric) elements of
,
, respectively. They are given by
There is a natural operation on
, the Kulkarni-Nomizu product given by
For instance the curvature tensor with constant sectional curvature,
, is given by
. Note that this is sometimes, e.g. Definition 1.110 of [6] , denoted
. We also introduce the contraction map
. Let
be any locally defined orthonormal frame; then for
,
Thus
is the Ricci curvature and
is the scalar curvature.
The volume of tube invariants are given by
|
(4.3)
|
Lemma 5.5 in [17] expresses the Pfaffian of an
-dimensional manifold as
|
(4.4)
|
The Weyl curvature is defined by
so on an Einstein manifold
Note that since
and
are in
, a commutative algebra, the binomial theorem applies to give
|
(4.5)
|
Furthermore, for any
we have the following formula from [19] :
Iterating we get
|
(4.6)
|
Finally, plugging 4.5 and 4.6 back into 4.4 we obtain the following formula.
Lemma 4.4.
On an
dimensional Einstein manifold, the Pfaffian is given by
Note that the factors
are (multiples of ) the Weyl volume of tubes invariants, except that they are evaluated in the Weyl curvature instead of the full curvature tensor. Note the scalar curvature factor in each summand with
; these terms are not conformally invariant.
Returning to our PE manifold, we may replace
, and this lemma gives us a formula for the Pfaffian in terms of the successive contractions of the Weyl tensor. Although the Weyl tensor itself has trivial contraction, this is not true of its higher powers. For instance, from [17] ,
.
Theorem 4.5.
Let
be a special bdf on an even dimensional asymptotically hyperbolic manifold
, and assume that
where
and
. Then
-
Proof.
This follows from Chern's Gauss-Bonnet Theorem for the manifold with boundary
,
|
(4.7)
|
Here
is a polynomial in the curvature and the second fundamental form of
in
. In terms of the curvature and connection forms it is given in [10] by
|
(4.8)
|
for some constants
whose precise value we will not need. Note that, for
small,
.
Since the right hand side of 4.7 does not depend on
, neither does the left, and thus
So we need only show that the second term vanishes.
If we denote the second fundamental form of
as a double form by
, 4.8 shows that
is a linear combination of terms of the form
|
(4.9)
|
It is easy to see that, for any
,
where
denotes contracting via a local orthonormal frame of
instead of
. Recall that
and
have even expansions mod
, hence so do
and
, and the constant term in 4.9 comes from the
terms in the expansion of these coefficients. Now, since
| |
we may conclude that the constant term in the expansion of 4.9 is a multiple of
, which vanishes as
. □
It is instructive to use Lemma 4.4 and write out the renormalized Gauss-Bonnet theorem 4.5 in four dimensions:
|
(4.10)
|
and in six dimensions:
|
(4.11)
|
Equation 4.10 is due to Anderson [2] (note that he uses a different convention for
, see the comment after equation (1.15) in [2] ). It is only in dimension four that the integrand requires no further renormalization than the volume. Generally, from Lemma 4.4 we know that
|
(4.12)
|
where
is a polynomial in the Weyl curvature, and
is the Pfaffian evaluated in the Weyl curvature instead of the full curvature. Equation 4.11 shows that
is not zero in general.
As we mentioned in the introduction, Chang, Qing, and Yang [8] have recently established a very similar formula 1.5 . It would be interesting to consolidate these formulas.
5 Varying the Poincare-Einstein metric
In this section we will consider the variation of the characteristic forms of a Poincaré-Einstein manifold when we allow the metric to vary along a family
of PE metrics with the same scalar curvature. After describing the expansion of
, we recover results of Anderson [2] and Graham-Hirachi [15] on the variation of the renormalized volume. Finally, we verify directly that the variation of the renormalized integral of the Pfaffian vanishes.
5.1 Variation of the Pontrjagin Forms
As before, Pontrjagin forms and numbers on Poincaré-Einstein manifolds are easily dealt with due to conformal invariance. Indeed, we know (Theorem 4.3 ) that
so we need only consider variations of
. Given a family of metrics
on
with connection and curvature forms
,
and a homogeneous invariant polynomial in the curvature,
, of degree
with complete polarization
, and denoting the derivative at
by an overdot, we have
The derivative of a product of Pontrjagin forms is exact, being the product of closed and exact forms. Note that the boundary of
is totally geodesic, so restricting the Levi-Civita connections of
yields the Levi-Civita connections of the induced metrics. Hence
when restricted to the boundary is the derivative of the connection 1-forms of the boundary metrics. We conclude with the following well-known result.
Proposition 5.1.
Let
a smooth family of metrics on
such that, for every
,
is a compact manifold with totally geodesic boundary. For any polynomial
, the variation of the characteristic number of
corresponding to
is the integral over the boundary of the Chern-Simons number corresponding to
on
.
5.2 Variation of the Renormalized Volume and the Pfaffian
Naturally, the renormalized Gauss-Bonnet theorem discussed previously shows that the variation of the renormalized integral of the Pfaffian vanishes. We verify this directly by studying the variation of the renormalized volume and the renormalized Weyl volume of tubes invariants.
Let
be a family of Poincaré-Einstein metrics on
with the same scalar curvature,
.
We use
, a special boundary defining function for
, to define a corresponding family of metrics on the boundary
. The metrics
in turn determine bdfs
, special with respect to
(as in Lemma 2.1 of [13] ). We shall compute in a frame
as in previous sections, and denote derivatives with respect to
at
by an overdot. Set
.
Just as one needs to break conformal invariance by choosing a bdf in order to study the renormalized volume, the study of the Einstein equation requires breaking gauge invariance. The approach followed in [2] (see also [15] ) is to use the Fefferman-Graham expansion of the metrics
. A second approach is to impose the so called Bianchi gauge. One defines the operator
and demands that
satisfy the Einstein equation and
. This approach was first espoused in [7] , though we shall follow [27] (see also [3] and [22] ).
The following result appears as Lemma 3.4 in [3] .
Lemma 5.2.
.
-
Proof.
We differentiate the relation
:
| |
to get
As
, we conclude that
□
The linearized Einstein equation,
, is given by (1.179 in [6] )
|
(5.1)
|
wherein
As the scalar curvature is the same for all
, we also have
, which by Theorem 1.174(e) of [6] , means
|
(5.2)
|
It is easy to compute the
-th order term in 5.1 , for instance:
so
implies
, and we gain no more information from the other
-th order terms.
Computing higher order terms will allow us to show that any perturbation of the metric through PE metrics has an expansion like that of the metric.
Proposition 5.3.
Assume that
, and that for some natural number
both:
|
(5.3)
|
hold, then
|
(5.4)
|
Consequently, once gauge invariance is broken, either through the Fefferman-Graham expansion or by imposing the Bianchi gauge, we have
|
(5.5)
|
(
as in Corollary 3.3 above), though in the odd-dimensional case log terms can occur with
.
-
Proof.
The proof of 5.4 is a straightforward computation using 5.3 in 5.1 .
We first prove 5.5 in the Fefferman-Graham expansion approach. We know that
A computation like that of Lemma 5.2 gives
. This implies that
and, together with Lemma 5.2 , that for
|
(5.6)
|
This sets up an iterative scheme. In the first step, we use
to see that the conditions 5.3 are satisfied with
. Then 5.4 shows that
, and this in turn, by 5.6 , that
.
Plugging this back into 5.3 with
gives us the next iterative step. This works until
, which establishes 5.5 .
For the Bianchi gauge approach, consider the operator
Clearly,
is in the kernel of
and
in that of
. We recall from Proposition 3 of [27] that
has no indicial roots between
and
. This will yield 5.5 through the following two observations.
First, we split
into two pieces,
, by requiring that
have an even expansion in
when there are an even number of
's in
and have an odd expansion in
otherwise. It is not hard to see that
is even or odd in
when
is even or odd, and similarly
. Secondly,
and
(see 5.2 above) together imply that
and
.
An easy computation shows that
implies
.
These observations together show that
vanishes to first order at the boundary, and that
vanishes to order
. We conclude that
vanishes to order
at the boundary. Thus
and this is 5.5 . □
We observe that 5.5 shows that the conditions 5.3 are satisfied with
; this implies by 5.4 that
.
Once we know the structure of
, it is a simple matter to compute the variation of the renormalized volume in the even-dimensional case (Theorem 2.2 in [2] ) and the variation of the residue in the odd-dimensional case (Theorem 1.1 in [15] ).
Theorem 5.4.
On an even-dimensional PE manifold, if
to first order with
as in Proposition 5.3 , the variation of the renormalized volume is
|
(5.7)
|
the inner product and the integral being taken with respect to
. Similarly, on an odd-dimensional PE manifold, the variation of the residue of the volume is given by
|
(5.8)
|
Remark In [15] , 5.8 is expressed in terms of Branson's
curvature and the Fefferman-Graham obstruction tensor
as
-
Proof.
Using Riesz renormalization, we have
|
(5.9)
|
The second equality is 5.2 , the third follows from integrating by parts, and the last uses Lemma 5.2 and the observation made before the statement of the theorem.
Now as in Proposition 5.3 , we can use the structure of
to compute
|
(5.10)
|
which gives the theorem in the even-dimensional case.
In odd dimensions the same proof works. The reason is that the
, being the first
term, behaves like the first odd term in the even-dimensional case. Specifically, it satisfiesconditions 5.3 with
which allows us to compute as in 5.9 and 5.10 replacing
by
. □
To deal with the other terms in the expression for the Pfaffian in section
4.3 we use the formalism of double forms, e.g., we interpret
as an element of
and denote it by
. We benefit greatly from a recent description of the variation of Weyl's volume of tubes invariants by Labbi [21] , [20] . We denote the integrand in the definition of
by
Labbi defines the tensors
|
(5.11)
|
and establishes that, on any closed Riemannian manifold,
|
(5.12)
|
We will show that a very similar formula holds, on Einstein manifolds, for the variation of the Pfaffian. Our definitions consistently differ from those of Labbi by a factor of
but agree with those of [17] .
The tensors 5.11 were first introduced by Lovelock in [23] . His interest was to find all divergence-free symmetric (0,2)-tensors built from the metric and its first two covariant derivatives. His theorem is that they are given by arbitrary linear combinations of the
and the metric. These properties are sometimes used to motivate the Einstein field equation in four dimensions (e.g., 3.7(i) in [6] ).
Lovelock's generalization plays an analogous role in higher dimensional GB gravity theories (see [30] and references therein). In the physics literature, a linear combination of the
is known as a Lovelock Lagrangian; an arbitrary linear combination of the
is known as a Lovelock tensor. Note that
is the usual stress energy tensor,
. Recall that any surface is automatically Einstein and that any higher dimensional Einstein manifold has constant scalar curvature. The following results appear in [31] .
Lemma 5.5.
On any
dimensional manifold, a) (Bach-Lanczos identity) If
is even,
b) If
and
for some constant
, then
is constant.
-
Proof.
a) Any double form in
is of the form
for some function
, and it is easy to see that
b) We always have
so if in addition we know that
and
, we find that
is constant. □
Labbi's proof of 5.12 follows from this formula, Lemma 4.2 of [21] (
):
|
(5.13)
|
where
extends to double forms via
and
is a second order differential operator; all we will need to know here is that its adjoint is twice the Hessian (see comments after equation (10) in [21] ), so that
with
the second fundamental form of
.
For a variation of Einstein metrics preserving the scalar curvature, a very similar formula holds.
The following proposition is proved by adapting Labbi's approach from
to
.
Proposition 5.6.
Consider a family
of PE metrics on an even-dimensional manifold, with
to first order with
as in Proposition 5.3 .
a) For
and for
,
b) For these variations, the variation of the renormalized integral of the Pfaffian vanishes.
Remark. Notice that
because
. In four dimensions, the functional
is well understood. Its gradient is the Bach tensor which vanishes for metrics conformal to Einstein metrics. This is reflected in the vanishing of the interior integral in the formula above for
, since by the Bach-Lanczos formula
in four dimensions. It would be interesting to understand the behavior of the
under an arbitrary variation of the metric.
-
Proof.
Define
| |
Lemma 4.1 of [21] establishes
for some differential operators
. Simple manipulations using that the metrics
all have the same scalar curvature yield
|
(5.14)
|
Labbi's computations in the proof of Lemma 4.2 of [21] go through with
replaced with
and
replaced by 5.14 because, on Einstein manifolds, Weyl curvature satisfies the second Bianchi identity. This gives
|
(5.15)
|
Consider the renormalized integral of the last term in equation 5.15 ,
| |
So this term produces residues, and we need to find the
term in the expansions. We revisit Lemma 3.1 and ascertain that
and
Thus the coefficients of
computed in this frame have no constant term, and the residue vanishes if
. To compute the residue for
, denote the double form corresponding to
by
and similarly
. Note that replacing the coefficients of
by their
-th order part produces the double form
and replacing the coefficients of
by their
-th order part produces the double form
We can use the following formulas from [20]
to see that
|
(5.16)
|
Putting these observations together, we see that the variation of the
-th renormalized volume of tube invariant evaluated in the Weyl curvature
is given by
| |
The second residue has been computed above for
and vanishes otherwise. For the first residue, we know that
(e.g., [19] ) hence neither
nor
have a constant term, and their product does not have an
term in its expansion, so that residue also vanishes. This establishes part (a) of the proposition.
Using the expression in Lemma 4.4 for the Pfaffian of an Einstein metric, its variation can be read off from part (a) of the proposition and the variation of the volume in Theorem 5.4 . The interior integrals telescope
| |
where we have abbreviated the scalar curvature factor to
. We are left only the
term, which is zero by Lemma 5.5 . There are two terms with residues, one coming from the volume and the other from
. The corresponding summands in the formula for the Pfaffian (in Lemma 4.4 ) are
so from 5.16 and Theorem 5.4 , we see that these residues cancel each other out. □
References
-
Albin, P. The Gauss-Bonnet theorem and Index theory on conformally compact manifolds. Ph.D. Dissertation, Stanford University, in preparation.
-
Anderson, M. T.
curvature and volume renormalization of AHE metrics on 4-manifolds. Math. Res. Lett. 8 (2001), no. 1-2, 171–188.
-
Anderson, M. T. Some results on the structure of conformally compact Einstein metrics. preprint, 2005. Available online at arXiv:math.DG/0402198.
-
Anderson, M. T. Topics in conformally compact Einstein metrics. preprint, 2005. Available online at arXiv:math.DG/0503243.
-
Avez, A. Characteristic classes and Weyl tensor: Applications to general relativity. Proc. Nat. Acad. Sci. U.S.A. 66 1970 265–268.
-
Besse, A. L. Einstein manifolds. Ergebnisse der Mathematik und ihrer Grenzgebiete (3), 10. Springer-Verlag, Berlin, 1987. xii+510 pp. ISBN: 3-540-15279-2
-
Biquard, O. Métriques d'Einstein asymptotiquement symétriques. Astérisque No. 265 (2000), vi+109 pp.
-
Chang, S. Y. A.; Qing, J.; Yang, P. On the Renormalized Volumes for Conformally Compact Einstein Manifolds. Vienna, preprint ESI 1513 (2004), available at www.esi.ac.at/preprints/ESI-Preprints.html.
-
Cheeger, J.; Muller, W.; Schrader, R. Kinematic and tube formulas for piecewise linear spaces. Indiana Univ. Math. J. 35 (1986), no. 4, 737–754.
-
Chern, S. S. On the curvatura integra in a Riemannian manifold. Ann. of Math. (2) 46, (1945). 674–684.
-
Donnelly, H. Heat equation and the volume of tubes. Invent. Math. 29 (1975), no. 3, 239–243.
-
Fefferman, C.; Graham, C. R.
-curvature and Poincaré metrics. Math. Res. Lett. 9 (2002), no. 2-3, 139–151.
-
Graham, C. R. Volume and area renormalizations for conformally compact Einstein metrics. The Proceedings of the 19th Winter School ”Geometry and Physics” (Srni, 1999). Rend. Circ. Mat. Palermo (2) Suppl. No. 63 (2000), 31–42.
-
Graham, C. R.; Lee, J. M. Einstein metrics with prescribed conformal infinity on the ball. Adv. Math. 87 (1991), no. 2, 186–225.
-
Graham, C. R.; Hirachi, K. The Ambient Obstruction Tensor and Q-Curvature. to appear in a proceeding volume, available online at arXiv:math.DG/0405068.
-
Graham, C. R.; Witten, E. Conformal anomaly of submanifold observables in AdS/CFT correspondence. Nuclear Phys. B 546 (1999), no. 1-2, 52–64.
-
Gray, A. Tubes. Second edition. Progress in Mathematics, 221. Birkhuser Verlag, Basel, 2004. xiv+280 pp. ISBN: 3-7643-6907-8
-
Guillarmou, C. Meromorphic properties of the resolvent on asymptotically hyperbolic manifolds. to appear in Duke Math. J., available online at arXiv:math.SP/0311424.
-
Kulkarni, R. S. On the Bianchi Identities. Math. Ann. 199 (1972), 175–204.
-
Labbi, M. L. Double forms, curvature structures and the
-curvatures. preprint, 2004, available online at arXiv:math.DG/0404081.
-
Labbi, M. L. On a variational formula for the H. Weyl curvature invariants. preprint, 2004, available online at arXiv:math.DG/0406548.
-
Lee, J. M. Fredholm operators and Einstein metrics on conformally compact manifolds. preprint, 2001, available online at arXiv:math.DG/0105046.
-
Lovelock, D. The Einstein tensor and its generalizations. J. Mathematical Phys. 12 1971 498–501.
-
Mazzeo, R. The Hodge cohomology of a conformally compact metric. J. Differential Geom. 28 (1988), no. 2, 309–339.
-
Mazzeo, R. Elliptic theory of differential edge operators. I. Comm. Partial Differential Equations 16 (1991), no. 10, 1615–1664.
-
Mazzeo, R. R.; Melrose, R. B. Meromorphic extension of the resolvent on complete spaces with asymptotically constant negative curvature. J. Funct. Anal. 75 (1987), no. 2, 260–310.
-
Mazzeo, R. R.; Pacard, F. Maskit combinations of Poincaré-Einstein metrics. preprint, 2002, available online at arXiv:math.DG/0211099.
-
Melrose, R. B. The Atiyah-Patodi-Singer index theorem. Research Notes in Mathematics, 4. A K Peters, Ltd., Wellesley, MA, 1993. xiv+377 pp. ISBN 1-56881-002-4
-
Melrose, R. B.; Nistor, V. Homology of pseudodifferential operators I. Manifolds with boundary. preprint, 1996, available online at arXiv:funct-an/9606005.
-
Padilla, A. Surface terms and the Gauss-Bonnet Hamiltonian. Classical Quantum Gravity 20 (2003), no. 14, 3129–3149.
-
Patterson, E. M. A class of critical Riemannian metrics. J. London Math. Soc. (2) 23 (1981), no. 2, 349–358.
-
Patterson, S. J.; Perry, P. A. The divisor of Selberg's zeta function for Kleinian groups. Appendix A by Charles Epstein. Duke Math. J. 106 (2001), no. 2, 321–390.
-
Paycha, S. From heat-operators to anomalies; a walk through various regularization techniques in mathematics and physics Emmy Noether Lectures delivered at the University of Goettingen, April-May 2003 (available online at http://wwwlma.univ-bpclermont.fr/ paycha/publications.html)
-
Witten, E. Anti de Sitter space and holography. Adv. Theor. Math. Phys. 2 (1998), 253-291, arXiv:hep-th/9802150.
Department of Mathematics, Stanford University E-mail address : pierre@math.stanford.edu