Conformal Holonomy: a Classification
Stuart Armstrong
23 January 2005
Contents Acknowledgements: The Author would like to acknowledge the support of an EPSRC research studentship. He would like to thank his supervisor, Prof. Nigel Hitchin, for all the help provided, and Dr. Thomas Leistner for several illuminating exchanges.
1 Introduction
Conformal geometry is but one of the class of parabolic geometries, a group that includes, amongst others, projective, almost Grassmanian, almost quaternionic, and co-dimension one CR structures. Some of these geometries were studied as far back as E. Cartan [11] , whose technique of `moving frames' would ultimately develop into the concepts of principal bundles and Cartan connections.
Despite the work of T.Y. Thomas [26] , [27] who developed key ideas for Tractor calculus in the nineteen twenties and thirties, and Shiego Sasaki in 1943 [24] , [25] , the subject fell into abeyance until the work of N. Tanaka [23] in 1979. The subject was rediscovered and further developed by T.N. Bailey, M.G. Eastwood and R. Gover in 1994 [1] .
Since then, there have been a series of papers by A. Čap and R. Gover [9] , [8] , [14] , [10] , developing a lot of the techniques that will be used in the present paper.
Previous papers had focused on seeing the Cartan connection for conformal geometry as a property of a principal bundle
. More recently, the principal bundle is replaced by an associated vector bundle, the Tractor bundle
, and the Cartan connection by a connection form for
, the Tractor connection
.
With these tools, calculations are considerably simplified.
The purpose of this paper is to analyse one of the invariants of the Tractor connection, the holonomy group. There is an invariant metric of signature
on
, so this holonomy group must be a sub-group of
.
It is a well known fact that a parallel section of the Tractor bundle corresponds to the local existence of an Einstein metric in the conformal class of a manifold.
Beyond this, little was known about reductions of holonomy.
It is the purpose of this paper to classify all the possible local holonomy groups of
acting reducibly on
. In doing so, they must conserve a Lorentzian metric of signature
. Then a paper by A.J. Di Scala and C. Olmos [12] shows that we have the complete list: there exist no connected proper subgroups of
acting irreducibly on
. A very recent paper by Felipe Leitner, [20] , proves the same results as in this paper; but his methods, involving normal Killing Spinors, are different from those described here.
The classification comes in two main pieces; if a bundle of rank other than
or
is preserved, the manifold decomposes analogously to the De Rham decomposition:
Theorem 1.1.
Let
be a conformal,
-dimensional manifold, such that
has a holonomy preserved sub-bundle of rank
. Then there exists a metric
such that
splits locally into the direct product of two Einstein manifolds
,
of dimensions
and
. The Einstein constants
and
of
,
are related by
. Furthermore, the Tractor holonomy group of
is the direct product of those of
and
. This is possible as
and
are Einstein, hence their Tractor holonomies reduce.
The converse is also true. This decomposition, again, is a local result, and may become degenerate along some embedded sub-manifolds. The second result is to list all the possible Tractor holonomies for a conformally Einstein manifold. Using a metric cone construction, related to the Ambient Metric of [13] , [10] and [15] , the following list is established:
Theorem 1.2 (Einstein Classification).
The Tractor holonomy of
, for
conformal to an Einstein space of non-zero scalar curvature, is one of the following groups:
-
-
,
-
-
,
-
-
for
,
-
-
for
,
-
-
for
,
-
-
for
.
Moreover, all these actually occur as holonomy groups.
The Ricci-flat case must be treated differently; in fact, if
is Ricci-flat and conformally indecomposable, and
is the metric holonomy group of
, then
has Tractor holonomy
. Thus:
Theorem 1.3.
The possible indecomposable Tractor holonomy groups for the conformal manifold
, conformally Ricci-flat, are:
-
-
,
-
-
,
-
-
,
-
-
,
-
-
,
and all of these groups do occur.
This paper begins with defining and laying out the groundwork for the conformal Tractor Bundle and connection, in Section
. Furthermore, it will prove the equivalence of this (second order) point of view with the standard equivalence class of metric structures.
The brief Section 3 presents some known results, showing how an Einstein structure in the conformal class is equivalent to a parallel section of the Tractor bundle.
Section 4, the longest and most full of technical results, proves the decomposition theorem previously mentioned.
Section 5 then establishes the list for the Einstein spaces via the metric cone construction, with Section 6 complementing it to list the possible holonomies for conformally Ricci-flat manifolds, using different methods. A brief note on symmetric spaces ends the paper.
Remark.
In all the holonomy groups listed in this paper, the holonomy reduction corresponds to the existence of a particular metric in the conformal class. Hence we always have a canonical representative in the conformal class, whenever the holonomy reduces.
2 Cartan Connection: Theory
2.1 The Cartan Connection
With flat geometries, one deals with spaces
, for
a Lie group acting transitively and effectively on
and
a parabolic subgroup.
In the general case, we can use the Cartan connection to construct an infinitesimal analogue of this isomorphism, mapping the tangent space
into the quotient of Lie algebras
instead.
In order to do this, we will follow the exposition used in [9] . Let
be a semi-simple Lie group, with a
-grading of its Lie algebra, that is,
| |
with the additional conditions that no simple ideal of
is contained in
and the sub-algebra
is generated by
. In our case, we will in fact be dealing with
-gradings of a simple Lie algebra, which makes both of these conditions trivially true.
Then we define the groups
to be the subgroups of
that preserve the grading and the decreasing filtration respectively, i.e.
| |
| |
Obviously the Lie Algebra of
is
and that of
is
. Given a manifold
, the group
and the grading of
, we can define the Normal Cartan Connection.
Definition 2.1 (Normal Cartan Connection).
On
, given a principal
-bundle
, a normal Cartan connection
is a section of
, with the following properties:
-
1.
is invariant under the
-action (
acting by
on
),
-
2.
, where
is the fundamental vector field of
,
-
3.
is a linear isomorphism for all
,
-
4.
The `curvature'
is such that
where
is the dual Spencer cohomology operator (see [3] for an exposition of Spencer cohomology).
Remark.
The last condition is what makes the connection normal ; this is a uniqueness condition for the Cartan connection of a particular parabolic geometry, similar to the torsion-free condition for a Levi-Civita connection. See [
9]
for proof of the existence of a normal Cartan connection in all cases. In the conformal case, an equivalent condition is that
for all
.
The bundle
and the form
together define the parabolic geometry.
Properties 1 and 2 make
analogous to a standard connection; property 3, however, implying a point-wise isomorphism, is very different. However, the inclusion
generates a principal bundle inclusion
, and we get the following result.
Proposition 2.2.
There is a unique
such that
is a standard connection form on
and
.
-
Proof.
At any point of
, define
for
, and
for
the fundamental vector field of
. These two formulas correspond whenever they are both defined (property 2 from definition 2.1 ), and completely define
on
. Then define
in the general case, for
.
Property 1 for
ensures this is well defined. It is then easy to see that
is a connection, and is unique.
Remark.
This
is hence a connection in the usual sense, equivalent with the Cartan connection that generated it; it is the so-called Tractor connection.
The Tractor connection
obviously generates a Cartan connection by pull-back to
. From now on, we shall use Cartan and Tractor connections interchangeably.
It is not the case that any
connection
will correspond to a Cartan connection via pull-back to
, as the isomorphism condition might be violated.
2.2 Conformal Geometry
A conformal manifold, in the standard sense, is a manifold
, where
is an equivalence class of metrics related by multiplication by a nowhere zero function. Alternatively, it is a reduction of the principal bundle of the tangent space of
to a principal
bundle which we shall call
. In this case,
.
From the Cartan connection point of view, however, the situation seems very different; later we shall see that both are equivalent.
In the classical, flat, case, Conformal Geometry is modelled on the sphere
. Taking the sphere as the collection of null-lines in
, the group
of conformal transformations is
. Then its Lie algebra has a 1-grading,
| |
where the conformal group
decomposes into the semisimple part
and the centre
, which is responsible for the conformal weight in representations of
.
Thus the data are
,
and
, on an
-dimensional space. Note that we have a natural action of
on
and hence an associated bundle to the
bundle mentioned previously:
Moreover, the action of
splits
, giving a corresponding splitting:
| |
This decomposition will be used extensively.
For later on, it will be important to explicate the Lie bracket of this algebra [9] . In fact,
, the Lie bracket on
is the natural commutator of endomorphisms, and
,
, for
sections of
and
respectively. The bracket between
and
is more complicated, and in fact
| |
with
the transpose operator, and
the identity element in (the centre of )
.
In their papers [9] and [8] , the authors demonstrate that the Cartan connection is equivalent to the standard conformal structure on a manifold
. This is an alternative treatment.
By definition, on a conformal manifold
, the structure group reduces to the principal
bundle
. This allows us to define the bundles
, the weighted line-bundles coming from the centre of
, i.e.
| |
From now on, we'll use the notation
for
. Then the conformal structure
can be more properly thought of as a section
of
. We will use
and
interchangeably from now on.
Then given a non-vanishing section
of
a conformal scale there is a corresponding metric in the conformal class
with a corresponding Levi-Civita connection
.
Another equivalent way of defining the conformal structure is to use the class of preferred connections:
Definition 2.3.
Given a conformal manifold
, a preferred connection
is simply a torsion-free
connection.
These preferred connections offer the first hint of how the Cartan connection ties in with the more conventional definitions of conformal structures, via the change of connection formula:
Theorem 2.4.
Let
be a section of any bundle associated to
, and let
be any vector field. Then if
and
are two preferred connections, there exists a one-form
such that
| |
where
is the Lie bracket for
previously described.
Note that if
were a metric connection, then
would be metric if and only if
were a closed form. In fact:
Proposition 2.6.
Let
and
be two metric, preferred connections, with
the corresponding conformal scales. Defining the function
as
, we have
| |
-
Proof.
By direct calculation, using the fact that
annihilates
while
annihilates
.
A variety of tensors connected with these preferred connections will be needed in subsequent chapters. To define them, we will use Penrose's abstract index notation, where
is understood as a section of the tangent bundle,
a section of the co-tangent bundle, and symmetrisation and anti-symmetrisation of indexes to be denoted by
and
respectively. This notation will be used intermittently throughout the paper.
Then if
is the curvature tensor of
, recall [14] :
|
(1)
|
with
the conformally invariant Weyl tensor, and the rho-tensor
:
| |
a particularly important tensor for the rest of the paper. In this case,
is the Ricci curvature, and
the scalar curvature
a section of
.
This is in the general case for a conformal connection; in the metric case, the picture is the same, except that
follows the simpler symmetric formula
| |
The last relevant tensor for
is the Cotton-York tensor:
|
(2)
|
It will be important to understand how the tensor
varies under a change of conformal structure, as this formula is the key to defining the Tractor bundle.
Letting
be the rho-tensor for
and
be that of
,
|
(3)
|
for
any vector field.
2.2.1 Equivalences
So far there are three ways of envisaging the conformal structure on a manifold
:
-
-
via a conformal metric
(a zero order structure),
-
-
via a class of preferred connections
(a first order structure) or
-
-
via a Cartan/Tractor Connection
/
(a second order structure),
and we will be using all three of them. This subsection will be dedicated to proving the important fact that all three viewpoints are equivalent. Though we will draw heavily on [8] for this exposition, we will use a slightly unconventional approach, which has the advantage of constructing the vital `Tractor Bundle' directly.
Remark.
For a variety of reasons to do mainly with conventional notation and ease of calculations, we will be working with the Tractor bundle
in the rest of the paper. However, to get a better understanding of what this bundle actually is, we need to start by defining the dual bundle
.
Consider the two-jet prolongation of
of the weighted bundle
.
By definition, we have the short exact sequences
| |
| |
The conformal structure
contracts
to
. Hence
splits as
, where the first space is the kernel of the contraction. Then we define the dual Tractor bundle
as the quotient:
| |
It is actually possible to realise
as a sub-bundle of
rather than a quotient bundle; we shall not be needing this result, though. Let
be the second order operator
given by composing the projection
with the two-jet operator
.
Proposition 2.7.
Given a preferred connection
,
any section of
and
any point on the manifold, the map
| |
generates an isomorphism
.
-
Proof.
This formula clearly generates a bundle map
. All that remains is to prove that
is the kernel of this map.
Assume
.
Then obviously
at
, implying that
is the (well-defined) section of
that corresponds to the second derivative of
at
. Thus
, or equivalently
.
Notice that we have not used the tensor
yet; the next proposition shows what we need it for.
Proposition 2.8.
Under a change of preferred connection
, the isomorphism of Proposition 2.7 changes as
| |
-
Proof.
Direct computation from the transformation properties of
and
.
The first component obviously stays the same. As
, we have the transformation law for the second component. For the third component:
| |
We need to remember that
acts on a section
of
as
.
Then recall the transformation law for
in equation 3
:
| |
Thus in total:
| |
| |
The last term in brackets is anti-symmetric, so disappears upon taking the contraction with the symmetric
. Hence, for
at
,
| |
| |
However, for reasons of convenience and notation, we will be working not with the bundle
but with its dual. Define the Tractor Bundle as
.
The previous results carry through to the dual of
; any preferred connection
defines a splitting
, and under a change of connection, this splitting changes via
| |
where
is the dual to
using the conformal metric
.
This particularly nice change of splitting formula implies the next vital lemma:
Lemma 2.9.
There is a natural metric
, of type
, on
.
-
Proof.
Given a preferred connection
and two sections of
,
and
, we define the metric by
| |
Direct calculation then shows this formula is invariant under a change of splitting.
Since
came about as a quotient bundle of a jet-bundle, we have invariant subspaces of
:
| |
and invariant projections:
Call
the sub-bundle of
that is the inclusion of
. Note that
is null under
.
Lemma 2.10.
There is a
-bundle
which is a principal bundle for
.
-
Proof.
The metric
shows that the structure group of
reduces to
. The invariant null sub-bundle
further reduces the structure group to
, i.e. to
.
Then we define
to be the bundle of orthonormal frames of
preserving
.
Let us review what we have so far. Starting from the conformal metric
and the class of preferred connections, we have constructed, via a tensor
dependent on the connections, a bundle
. And this bundle generates a principal bundle
, where it is natural to suppose the Cartan connection living.
Define the Lie Algebra bundle
. Then given a preferred connection
we have a splitting of
, and hence a corresponding splitting:
| |
In order to finish the construction of this Cartan connection, we will start by building a
connection on
and then prove that it is a Tractor connection.
Definition 2.11.
Given a preferred connection
, we have a splitting of
. Each of these bundles is a
bundle, so
ascends to a connection on
. Then we define the Tractor connection
as
| |
with the vector
and the one-form
seen as sections of the Lie algebra bundle
.
Then since
is a
connection,
is a
connection. Of course, this definition makes no sense without:
Proposition 2.12.
This definition is independent of the choice of
.
-
Proof.
The formula for the change of splitting of
(deduced directly from that of
) is:
| |
Then a direct calculation proves the result.
Thus for any bundle
associated to
, we have an invariant connection form:
| |
In the case of the Tractor bundle
, the detailed expression is:
|
(4)
|
Now we get to the result that ties all the structures together:
Theorem 2.13.
The connection
is a normal Tractor connection.
-
Proof.
Let
be the inclusion
,
projection
, and let
be the one-form associated with the connection
.
We need to prove that
is an isomorphism
for all points
; then
will be the Cartan connection generating the Tractor connection
.
Reasoning by contradiction, assume that
is not an isomorphism at some point
, so there exists a vector
such that
. As vertical vectors in
are mapped isomorphically onto
,
is a non zero vector in
.
Then define a local section
of
in
, such that
. This also gives us a section
of
. Then in the frame bundle determined by this section, the connection
is of the form
| |
This shows that the bundle
is infinitesimally conserved at
in the
direction, or, switching to the associated bundle
, that there is a section
of
, non-zero at
, such that
at
.
However, the connection on the tractor bundle is given by equation ( 4 ):
| |
which is a contradiction as
(hence
) is non-zero at
. So
is indeed a Tractor connection.
And finally, to complete the circle:
Lemma 2.14.
The Cartan connection generated by
is normal.
-
Proof.
In order to demonstrate this result, it suffices to show that the curvature of
lies in the Lie algebra bundle of
or, equivalently, that it preserves the canonical bundle
.
In abstract index notation, the expression for
is:
| |
| |
Here we have used the connection
on
to define the second covariant derivative; however, we could have used any other connection, as we are about to anti-symmetrise
and
. Upon doing this, the terms in brackets vanish.
Moreover,
and similarly for one-forms, meaning that:
| |
Looking back at equations 1
and 2
, we see that this expression is the sum of the Weyl tensor and the Cotton York one. Or, expressed in more conventional notation, in the splitting of
determined by
:
| |
This curvature expression obviously preserves the canonical bundle
.
3 Conformally Einstein Manifolds
3.1 Important Note
In most of the proofs in the remainder of this paper, it will be assumed that for a certain holonomy preserved sub-bundle
used in the proof, one has
. This will not be the case everywhere, of course; however:
Proposition 3.1.
Let
be a holonomy preserved sub-bundle. Then
on
, an open, dense subset of
.
-
Proof.
is open because of the
condition.
Let
, and
be a non-zero element of
. Then extend
locally to a section
of
by parallel transport along `rays' from
. This implies that
at
. Then picking any nowhere-zero section
of
, we can define the function
by
The derivative of
is
. At
, this is just
, and, since
at
:
| |
If
, this is non-zero for some
, so
is non-zero arbitrarily close to
.
If
, then the first derivative is zero, and the second derivative is thus:
| |
with
as
. Then the second derivative is non-zero for
, for instance, forcing
to be non-zero arbitrarily close to
.
This implies that
arbitrarily close to
, proving the result.
In fact, if the first derivative vanishes,
must be an isolated point.
The classic examples of this are the various conformally Einstein metrics on the sphere
. The sphere is conformally flat, so there are many holonomy preserved sections of its Tractor bundle.
A preserved section
of negative norm corresponds to the Spherical metric
on the whole space. In this case,
is never zero.
A preserved section
of zero norm corresponds to the Euclidean metric
on
. In this case,
for
.
A preserved section
of positive norm corresponds to the Hyperbolic metric
on two half spheres of
. In this case
is zero only on the boundary
cutting
into two.
3.2 Einstein Spaces
Though it is well known in general that any conformally Einstein space corresponds to a parallel section of the tractor bundle
, what follows is a direct proof of this fact using the Tractor connection approach.
Remark.
This is a first instance of a holonomy reduction of
.
Theorem 3.2.
If
has an Einstein metric
in its conformal class then there exists a parallel section
of its tractor bundle
.
-
Proof.
Let
be the Einstein metric,
. Then the
-tensor is
Hence
| |
where
is the section of
corresponding to
.
To prove the converse of this theorem, we need the following lemma:
Lemma 3.3.
If a conformal connection
has a symmetric Ricci tensor, then
is actually a metric connection.
-
Proof.
Let
be the curvature of
. Then
acts on the determinant bundle
via its trace
. However, by the first Bianci identity,
| |
| |
the anti-symmetric part of the Ricci tensor. So if
has a symmetric Ricci tensor, its curvature must vanish on
, so locally
must preserve a section
of the determinant bundle. Then
is a preserved conformal scale and
a metric preserved by
.
Theorem 3.4.
If a line bundle
of
is holonomy preserved, then a section
of
is preserved, and
has an Einstein metric
in its conformal class, wherever
.
-
Proof.
The line bundle
defines a connection on
, and hence a torsion free connection on
, in the following way. Let
be any section of
, and let
be the section of
such that
. Then define
; it is easy to see that this is indeed a connection. A connection on
naturally defines a torsion-free
connection on
using the conformal metric
and the same formula as for the Levi-Civita connection.
Using
, we split
in the usual way. Then equation ( 4 ) implies that
. Thus, by the very definition of
,
for any section of
.
If
is not null, then a section
of constant norm, is preserved.
This generates a metric
.
If
is null, then
.
But
.
Consequently, if
is not null, then
, so
for
.
Alternatively, if
is null, then
. Hence
has a symmetric Ricci tensor, implying that it is actually a metric connection. Let
be the section corresponding to this metric. Then the section
, is parallel.
Finally, the metric
is Ricci-flat.
Remark.
Note that the sign of
is the opposite of the sign of the Einstein constant
.
4 Decomposition Theorem
This section presents the decomposition theorem for Tractor connections, similar to the De Rham decomposition for Riemannian connections.
Remark.
Related terminology may be found in [
18 ]
.
The theorem to be proved is the following:
Theorem 4.1.
Let
be a conformal,
-dimensional manifold, such that
has a holonomy-preserved sub-bundle of rank
. Then there exists a metric
such that
splits locally into the direct product of two Einstein manifolds
,
of dimensions
and
. The Einstein constants
and
of
,
are related by
. Furthermore, the Tractor holonomy group of
is the direct product of those of
and
.
This theorem also has a converse.
4.1 Preparatory Results
The contents of this section deal with various geometric results for submanifolds that will be needed in the second part of this chapter. The proofs are technical geometric manipulations.
Let
be a Riemannian manifold, with submanifold
, and let
be the Levi-Civita connection of
.
The second fundamental form on
is a section of
, defined as
for
and
sections of
. As
is torsion free,
is symmetric.
The submanifold
is called minimal if
is trace free. It is called umbilical if
, for
a section of
. It is totally geodesic if
. Of these three, only umbilical is a conformally invariant concept.
Proposition 4.2.
Given a submanifold
, there is a canonical inclusion of the tractor bundle of
,
.
Note that it is still important to distinguish
and
, as, for instance, they generally behave very differently under their various tractor connections.
To get any further, we need to place some restrictions on
:
Definition 4.4.
A manifold
admits a crossfoliation if it admits two separate foliations, with tangent bundles
and
, such that
-
-
-
-
If the two sets of leaves are umbilical, we call this crossfoliation umbilical. If
and
are orthogonal, they form an orthogonal crossfoliation.
Definition 4.5.
A metric
on a manifold
with an orthogonal crossfoliation is said to be split at a point
, if, for
and
the two leaves intersecting orthogonally at
and
,
we have:
This last definition is equivalent with the fact that
| |
This gives us the main result of this section:
Lemma 4.6 (Metric Splitting).
Let the submanifold
be a leaf in an umbilical orthogonal crossfoliation, and let
be a point in
. Then there exists a metric
such that
is totally geodesic in
, and
splits at
.
-
Proof.
Pick local coordinates
so that
and
are the leaves of the foliation, with
being the leaf for
. Then as
is totally umbilical,
, for some vector field
. Thus
, with
functions of
only.
Then, as in the previous theorem, defining
for
gives us a new second fundamental form
. Hence
is totally geodesic.
Then given any point
, shift the
by a constant so that
at
. Then let
be (locally) the orthogonal submanifold to
at
. Then, as
is itself umbilical by definition,
for
as
is in
at
. Thus we may change
by a function
; this makes
totally geodesic at
, and since
is a section of
,
remains totally geodesic.
Then with
totally geodesic at
, for any section
of
,
and
, as
is metric. As
is also totally geodesic at
, the previous formulas are also true for
a section of
, hence for any
at
.
Remark.
It is possible to get much stronger results from the conditions of the lemma; however, they will not be needed, and will drop out quite naturally from later considerations.
Theorem 4.7.
If
is a leaf of an umbilical orthogonal crossfoliation, then the sub-bundle
is conserved under parallel transport along
.
-
Proof.
Using the previous lemma ( 4.6 ), let us pick
so that
is totally geodesic and
splits at
. Then
is a fortiori minimal at each point, so
is given in the splitting
by
.
Then for
,
| |
where the splitting of
ensures that that
at
. This metric splitting further makes the curvature and Ricci tensors split at
,
. Then the
-tensor also splits, implying that
.
In consequence,
at
, and as
is arbitrary, on all points of
.
Remark.
Note that this does not imply that the Tractor connection
acts on
the same way as
acts on
. Indeed, though
splits into
and
components, it is not generally equal to
, the various coefficients being different, dependent as they are on the dimensions of the different (sub)manifolds.
4.2 Decomposition Theorem
Let
be a conformal manifold of dimension
, with tractor bundle
, and let
of rank
be a holonomy preserved sub-bundle.
Let
be the canonical null-line, and assume furthermore that
and
(this will always be the case on an open, dense subset of
, by Proposition 3.1 ).
Let
be the kernel of
. So
, in any splitting we can choose.
Then define
, and
.
The conditions on
and
imply that neither of these two bundles is contained in
, since
. As rank
rank
, we get rank
and rank
.
Lemma 4.8.
-
Proof.
As
, then any vector in
is null.
The tractor metric on
is given by
. Hence the only null vectors in
lie in
. Thus as
, we must have
.
Lemma 4.9.
-
Proof.
Let
for
and
. Then
and
. Then as
, implying that
, hence
, implying that
and proving the lemma.
Remark.
This lemma only works in the case of definite signature, and is the reason that the decomposition theorem cannot be extended to the indefinite case.
These two results together prove:
Proposition 4.10.
As
is the kernel of
,
maps
injectively onto
. As scale is irrelevant, let
, via any
.
Let
and
. Then
, and
.
Theorem 4.11.
and
are both integrable, giving an orthogonal foliation. Moreover, this foliation is umbilical.
-
Proof.
Let the
be an orthogonal frame of
. Pick any metric
.
Let
.
Then
for some
. As
is conserved under holonomy,
. As the first component is zero,
is a section of
, hence
. Thus
is a section of
, and hence so is
.
So for
,
, and, a fortiori,
. Since
are an orthogonal frame, both these properties generalize to arbitrary sections of
. Thus
is integrable.
Lemma 4.12.
The corresponding leaves of
are umbilical.
-
Proof.
Let
be an orthonormal frame in the metric
. We know that
for
, so
where
is the second fundamental form of
.
Now
, implying that
| |
Hence
for
. As both
and
are bilinear, this equality extends to any sections of
, making the leaves umbilical.
The same proof works for
.
Let
and
be these leaves. As we are in the orthogonal umbilical crossfoliation case, Theorem 4.7 shows that there are sub-tractor bundles
and
that are conserved along
and
respectively.
The rank of
is
, hence has a non-zero intersection with
. This sub-bundle is hence conserved along
. The same is true for
and
along
.
Proposition 4.13.
is a line bundle.
-
Proof.
In this case, we need to use a specific metric. Given a leaf
, let
be a metric such that
is totally geodesic in
.
Then along
, in the splitting given by
,
.
Reasoning by contradiction, assume there exist
, sections of
, linearly independent. Then in the span of
and
there exists
. Hence
, so
, with
. But since
is in
as well, this gives
. So
, which, by the definition of
, is a contradiction.
The same proof shows that
is a line bundle as well.
4.2.1 Split Metrics in the Conformal Class
Let
and
. Then
and
are orthogonal line bundles, and we are going to use them to construct a special metric on
. From now on, let
be sections of
,
sections of
.
For
, let
be the leaves that cross at
. Then choose a metric
that splits at
(Lemma 4.6 ).
Hence at
, in the splitting of
defined by
,
and
.
Lemma 4.14.
In this splitting,
and
at
.
-
Proof.
Let
be a section of
. Then for some
,
is a section of
.
So
is a section of
. However, at
,
is a section of
as well; hence
at
.
The same proof implies the same result for
.
Lemma 4.15.
In this splitting, at
,
for some
, and
.
-
Proof.
At
,
. Hence, as
is a line bundle,
. The same proof shows that
. Moreover,
and
are orthogonal, hence
.
We notice two things at
: firstly that
. Secondly for sections
of these bundles,
implies that
.
If
is non-null, then the second implication is an equivalence.
However,
is a generic point on the manifold, and these two properties are conformally invariant; hence they are true in general, over the whole manifold.
If
is null, then
and
is holonomy preserved over the whole manifold.
Then by Theorem 3.4 , there exists a Ricci-flat metric
.
Conversely, let
(and hence
) be non-null. Without loss of generality, let
be the positive definite one of the two. Then for any constant
, the section
such that
, defines a metric
on the manifold via
.
Lemma 4.16.
The metric
splits at every point of the manifold.
-
Proof.
In the splitting defined by
,
. Then there is a corresponding section
of
, with
covariantly constant along
, and
covariantly constant along
. However, the first component of
is
, and the first component of
is
. Hence
must be zero, and thus
, and
if
is non-null, and
in the null-case.
Letting
be a section of
, then
is a section of
. Hence
is a section of
, making
into a section of
. The metric nature of
then implies that
is a section of
.
The same relationship applied to sections of
completes the proof.
Theorem 4.17.
In
, the leaves
and
are totally geodesic, and hence,
is a direct product.
-
Proof.
By the previous lemma,
and
are both zero.
Theorem 4.18.
If
, then the leaf
is Einstein. The same result holds for the
leaves. Furthermore, the two Einstein constants
and
are related by
, where
.
-
Proof.
Along
,
along
. Similarly,
along
, so:
This result is also valid in the null case, with
.
But
| |
and
splits as
, and
.
Let
dim
. Then there exist constants
and
such that
and
, and
.
This gives the simultaneous equations:
| |
| |
Solving these, and remembering that we can freely rescale
,
So
is non-positive Einstein, and
is non-negative Einstein.
Finally, to complete the decomposition result, one needs,
Theorem 4.19.
The tractor holonomy of
is the direct product of those of
and
.
-
Proof.
The tractor expression on
in the split metric is
| |
Since
is totally geodesic,
. Thus the tractor expression on
is the same as above, with
replaced by
. Then the result is proved along
by the following lemma:
Lemma 4.20.
In the circumstances as above,
=
.
-
Proof.
By direct calculation. As
is the Einstein constant on
, a manifold of dimension
,
On the other hand, the scalar curvature of
is
, so
| |
| |
by the relation between
and
.
If any of the leaves are oneor two-dimensional, the situation is different, as there is no intrinsic tractor connection on the tractor bundle
. The restriction of
to
, along
, does define a restricted Tractor connection however. If the leaf
is one dimensional, then this connection must be flat. If not . . .
Lemma 4.21.
If the leaf
is two dimensional, then the restricted connection on it is flat.
-
Proof.
By the relation between the Einstein structures,
. Then the curvature of
along
is:
| |
where
is the metric curvature of
.
But we know that
, and since
takes values in
, a rank one bundle, for the trace of
to be what we have found,
itself must be
| |
So the curvature of
restricted to
, is zero.
Then to prove the whole theorem, one needs only to note the next obvious result:
Lemma 4.22.
The tractor connection on
,
, makes
parallel along
.
-
Proof.
Let
and
be sections of
and
respectively, extended to
trivially. Then
as
is trivially extended, and since the eigenspaces of
are orthogonal,
. So
| |
Obviously, the same result holds for
as well, so the holonomy splits.
Remark.
As is well known, if an Einstein manifold is three dimensional, it is conformally flat. Adding this to the flatness results along one or two dimensional leaves, it is revealed that groups of the form
,
and
, can never appear as elements of a group product in the holonomy of any conformal manifold.
4.2.2 Converse to the Decomposition
Theorem 4.23.
Let
be locally a direct product of the Einstein manifolds
and
.
Furthermore, let the Einstein constants
of
and
obey the relations
. Then the tractor bundle
of
has a holonomy-conserved sub-bundle
.
-
Proof.
Let
and
, and let's work in the
metric. Then since the foliation is totally geodesic,
is of the form
, and the converse holds for
.
Theorem 4.18 reverses, giving
for some
. This gives a section
, covariantly constant along
and
, covariantly constant along
. Note that these definitions work in the Ricci-flat case as well, for
and using theorem 3.4 . Define
span
,
span
.
Moreover, let
, and
; then in this metric,
, and similarly for
.
Finally, let
and
.
Lemma 4.24.
.
-
Proof.
and
are certainly orthogonal.
is orthogonal to
and
to
by components, and they are mutually orthogonal as well. Thus
and a consideration of rank proves equality.
Now to the main theorem.
is conserved along
, as both
and
are. As
is pseudo-metric,
is thus conserved along
as well. Reversing the argument, we see that
, and hence
are conserved along
, consequently along any path.
A final technicality exists, for the case of oneor two-dimensional manifolds, where the Tractor connection is not defined. We may still define
formally for dim
, as before. This implies Theorem 4.23 , meaning that the converse still applies, with the caveat that the Einstein coefficients in the two dimensional case must be constant.
Example.
To illustrate these proofs, we can see that
does not have any holonomy-conserved sub-bundles in its tractor connection (in fact it has full holonomy), while
is conformally flat, for
the hyperbolic 4-space.
Remark.
Some old results of H. W. Brinkmann [
5]
, [
6]
can be proved directly using this decomposition theorem. For instance, the fact that any 4-manifold with two distinct Einstein structures in the conformal class is conformally flat. In our setting, the preserved sub-bundle spanned by the two Einstein vectors decomposes the manifold into a direct product of 3and 1-dimensional Einstein spaces. But both these spaces are conformally flat, so our original manifold has trivial holonomy; in other words, it is conformally flat.
Remark.
Analogously to the previous remark, we can see that if not conformally flat, a five dimensional manifold can have up to two linearly independent Einstein structures, a six dimensional manifold can have three, an
dimensional manifold
.
5 Einstein Spaces: Metric Cones
In this section we will give a full classification of the possible Tractor holonomies of the non Ricci-flat Einstein spaces, using to this effect the construction of a metric cone, whose Levi-Civita holonomy corresponds to the Tractor holonomy of the original manifold.
Remark.
As we mentioned in the introduction, this metric cone construction is related to the Ambient Metric construction of [
13]
and [
10]
, for conformally Einstein manifolds. The actual relation is slightly subtle. This also provides a direct proof of a result of [
15]
, namely that the Ambient Metric construction always exists if the manifold is conformally Einstein.
Definition 5.1.
A conformal manifold
is said to be indecomposable if they cannot be decomposed into Einstein spaces as in the previous section. In other words,
may preserve a single line bundle (and its orthogonal complement), but nothing else.
Remark 5.1.
In the non Ricci-flat Einstein case, indecomposable implies that the tractor holonomy acts irreducibly on
or
(since the only preserved line bundle is positive or negative definite).
As all Einstein manifolds of dimension 3 are conformally flat, we shall assume
.
Let
be an Einstein manifold,
,
.
Theorem 5.2 (Einstein Classification).
The Tractor holonomy of
is one of the following,
:
-
-
,
-
-
,
-
-
for
,
-
-
for
,
-
-
for
,
-
-
for
.
Moreover, all these holonomy groups actually occur.
Remark.
It is interesting to note that there is only a single holonomy possible for an indecomposable Einstein manifold with negative constant.
The remainder of this chapter will be dedicated to proving Theorem 5.2 .
Definition 5.3.
Given an Einstein manifold
, we define the metric cone on
as
with
and
.
Note that
is of definite signature if and only if
has positive Einstein constant. In the negative case, we call
a Lorentzian cone.
Then defining
as the Levi-Civita connection of
, and remembering the formula:
| |
| |
we can calculate the following equalities. For
, and
a local basis of vector fields of
, extended trivially to
:
| |
with
the Levi-Civita connection of
.
Given a path
in
, with tangent vector field
, let
be the parallel transport of a vector along the path, thus
. Split
as
, with
. Then we get the following result:
Lemma 5.4.
Extend
and
in the
direction, with
, and
| |
Then
and
on
.
-
Proof.
The function
is independent of
, so
. Hence
.
Furthermore,
| |
| |
so
.
We can expand out the original equation
at
, giving:
| |
By linearity, this is equivalent to the two equations
and
.
Then similarly expanding
for varying
:
| |
| |
The previous result shows that when we're computing the holonomy of
, we only need to consider paths in
.
We can now turn to the tractor connection
on
, for the conformal structure
. Using the splitting given by the metric
, we can see the formal similarities with
at
.
As
is Einstein, with coefficient
, then
. Furthermore, for
, then:
| |
Hence under the formal identification of
with
and
with
, we get
for
at
. Then by the previous lemma and its implication for the holonomy of
, the next theorem is proved:
Theorem 5.5.
The holonomy groups of
and
are isomorphic.
Hence the holonomy of
is metric, and irreducible by Remark 5.1 , and must be one of those classified by Merkulov and Schwachhöfer in [22] . In the negative Einstein case, a look at the table shows that the only possible holonomy is the full
group itself. For the positive Einstein, we need the following result:
Proposition 5.6.
The metric cone
is Ricci-flat.
-
Proof.
By the definition of
and the corresponding
, the curvature
is
| |
with
the curvature of
.
Then taking traces,
| |
| |
So the possible holonomies reduce to those corresponding to metrics which are Ricci-flat, namely
and
.
The
case is generic. The
holonomy on the cone corresponds to Sasaki-Einstein manifolds, the
to 3-Sasakian ones,
and
to weak holonomy manifolds [2] ; all of which can be realised on compact manifolds.
Remark.
In [
3]
R.J. Baston presents a local twistor theory, which, in the case of conformal manifolds, is just given by the spin representation of
and the extension of
to this new context. A parallel section of this bundle is equivalent with the existence of a spinor
solving the twistor equation for all vector field
:
| |
with
the Dirac operator. Paper [
16]
by Katharina Habermann analyses solutions to this twistor equation; she shows that these imply that the manifold is conformally Einstein, of non-negative scalar curvature.
So the Tractor holonomy groups
and
actually correspond to the existence of twistor-spinors on the manifold.
Remark.
The concept of a twistor-spinor is a generalisation of that of a Killing spinor. A Killing spinor is a spinor
solving the equation
for all vector fields
and some constant
. In [
2]
, C. Bär showed that having a Killing spinor is equivalent with having a parallel spinor on the metric cone. So the cases of weak holonomy SU(3) and nearly Kählerian structures are covered by the Tractor connection; in fact in his paper [
7]
constructing manifolds of exceptional holonomy, R.L. Bryant produces manifolds of holonomy
and
as metric cones on
and
respectively. Thus all the holonomies listed actually occur.
We can now turn to the Ricci-flat case, which is actually simpler than the general Einstein case, but with an added subtlety.
6 Ricci-Flat Spaces
Let
be a Ricci-flat space of indecomposable tractor holonomy. As
is Ricci-flat, its Tractor holonomy is contained within
. Fix a point
for calculating the holonomy groups, and let
be the metric holonomy of
,
its Tractor holonomy,
,
their Lie algebras.
Then:
Lemma 6.1.
, or, equivalently,
.
-
Proof.
Let
be the parallel transport of a vector along a path
with tangent field
; in other words
, for
the metric connection on
. Then
| |
which is zero for
, proving that every metric holonomy element is a tractor holonomy element. This argument also works in reverse, showing that
, where
is the projection of
onto its first component.
This demonstrates that
. But first:
Lemma 6.2.
The representation of
on
is irreducible.
-
Proof.
If a bundle
is preserved by
, then the bundle
is preserved by
. Thus, since we assume our Tractor holonomy to be indecomposable, then
must act irreducibly on
.
Then since the Lie bracket on
is trivial, the adjoint representation of
on the second component of
is the usual, irreducible one. Accordingly this adjoint representation splits
into two irreducible representations, isomorphic to
and
.
As a consequence,
or
. We now claim that
-
Proof.
Reasoning by contradiction, we assume that
, and go on to show that this violates our indecomposability assumption.
Express
as
, the sum of the irreducible representations of
. Then, as
acts irreducibly on
, there is, at
, a new splitting of
corresponding to the splitting
. This splitting is
| |
Then
preserves the new vectors
and
. This shows that
preserves a rank two sub-bundle, contradicting indecomposability.
Putting this together, we can now claim the following theorem:
Theorem 6.4.
The possible indecomposable Tractor holonomy groups for the conformal manifold
, conformally Ricci-flat, are:
-
-
(generic),
-
-
(Calabi-Yau),
-
-
(Hyper-Kähler),
-
-
(see [7] ),
-
-
(see [7] ),
and all of these groups do occur.
Remark.
This result offers an alternative proof for the theorems in Mario Listing's paper [
21]
, in the special case of conformally Ricci-flat manifolds.
6.1 Addendum: Symmetric Spaces
A symmetric space
is a manifold such that
for
the full curvature tensor. It is quite easily to show, using the infinitesimal holonomy developed by S. Kobayashi and K. Nomizu [17] , that any indecomposable conformal manifold that is conformal to a symmetric space has the maximal holonomy in its category -
if the symmetric space is not Einstein, and
or
if it is (Ricci-flat symmetric spaces are flat).
Example.
These results give an independent proof to the results of F. Leitner [
19]
, that the conformal holonomy of
, locally isomorphic to
, is
. The group
is a positive Einstein Symmetric space, not conformally flat (consider the fate of the tractor vector
under parallel translation), so the result follows.
Example.
Note that the same argument shows that the manifold
where
are the radii of the spheres, has full holonomy
.
Two very similar results also are implied:
Example.
If a manifold
is conformal to an Einstein Symmetric Space, then it cannot be conformal to any other Einstein space, or any other Symmetric Space, unless it is conformally flat.
And:
Example.
If
is conformal to a Symmetric space in two different ways, then its tractor holonomy is full or null.
References
-
T.N. Bailey, M.G. Eastwood and R. Gover: Thomas's Structure Bundle for Conformal, Projective and Related Structures, Rocky Mountain J., 24 (1994), 1191-1217. MR 95h:53064
-
C. Bär: Real Killing Spinors and Holonomy, Comm. Math. Phys., 154 (1993), No. 3, 509-521.
-
R.J. Baston: Almost Hermitian Symmetric Manifolds I Local Twistor Theory, Duke Mathematical Journal, 63 (1991), No. 4, 81-112.
-
T. Branson, A. Čap, M. Eastwood and R. Gover: Prolongations of Geometric Overdetermined Systems, arXiv:math.DG/ 0402100 v2 (2004).
-
H.W. Brinkmann: Riemmann Spaces Conformal to Einstein Spaces, Math. Ann., 91 (1924), 269-278.
-
H.W. Brinkmann: Einstein Spaces which are Mapped Conformally on Each Other, Math. Ann., 94 (1925), 119-145.
-
R.L. Bryant: Metrics with Exceptional Holonomy, Annals of Mathematics, 126 (1987), No. 3, 525-576.
-
A. Čap and A.R. Gover: Tractor Calculi For Irreducible Parabolic Geometries, Global analysis and harmonic analysis (Marseille-Luminy, 1999), Sémin. Congr., 4 (2000), Soc. Math. France, Paris, 129-154.
-
A. Čap and A.R. Gover: Tractor Calculi For Parabolic Geometries, Transactions of the American Mathematical Society, 354 (2001), No. 4, 1511-1548.
-
A. Čap and R. Gover: Standard Tractors and the Conformal Ambient Metric Construction, Annals of Global Analysis and Geometry, 24 (2003), No. 3, 231-259.
-
E. Cartan: Les Espaces à Connexion Conform, Ann. Soc. Pol. Math., 2 (1923), 171-202.
-
A.J. Di Scala, C. Olmos: The Geometry of Homogeneous Submanifolds of Hyperbolic Space, Mathematische Zeitschrift, 237 (2001), 199-209.
-
C. Fefferman and C.R. Graham: Conformal Invariants, Astérisque, hors série: `Elie Cartan et les Mathématiques d'Aujourd'hui' (1985), 95-116.
-
A.R. Gover: Invariant Theory and Calculus for Conformal Geometries, Advances in Mathematics, 163 (2001), 206-257.
-
R. Gover and L.J. Peterson: The Ambient Obstruction Tensor and the Conformal Deformation Complex, arXiv:math.DG/0408229 v1 (2004).
-
K. Habermann: The Twistor Equation on Riemannian Manifolds, Journal of Geometry and Physics, 7 (1990), No. 4, 469-488.
-
S. Kobayashi and K. Nomizu: Foundations of Differential Geometry, 1 (1996), Wiley-Interscience Publication.
-
N. Koike: Totally Umbilic Orthogonal Nets and Decomposition Theorems, Saitama Mathematical Journal, 10 (1992), 1-19.
-
F. Leitner: Conformal Holonomy of Bi-Invariant Metrics, arXiv:math.DG/ 0406299 v1 (2004).
-
F. Leitner: Normal Conformal Killing Forms, arXiv:math.DG/ 0406316 (2004).
-
M. Listing: Conformal Einstein Spaces in
-Dimensions, Annals of Global Analysis and Geometry, 20 (2001), 183-197.
-
S. Merkulov, L. Schwachhöfer: Classification of Irreducible Holonomies of Torsion-free Affine Connections, Annals of Mathematics, 150 (1999), 77-150.
-
N. Tanaka: On the Equivalence Problem Associated with Simple Graded Lie Algebras, Hokkaido Math. J., 8 (1979), 23-84. MR 80h:53034
-
S. Sasaki: On the Spaces with Normal Conformal Connexions whose Groups of Holonomy fixes a Point or a Hypersphere. I. II. III., Jap. Journ. Math., 18 (1943), 615-622, 623-633, 634-795.
-
S. Sasaki and K. Yano: On the Structure of Spaces with Normal Conformal Connection whose Holonomy Group leaves Invariant a Sphere of Arbitrary Dimension, Sugaku (Mathematics), 1 (1947), 18-28.
-
T.Y. Thomas: On Conformal Geometry, Proc. N.A.S., 12 (1926), 352-359.
-
T.Y. Thomas: Conformal Tensors, Proc. N.A.S., 18 (1931), 103-189.