Monopoles over
–manifolds containing long necks, II
Kim A. Frøyshov
14 March 2005
Contents
1 Introduction
In this paper we prove a gluing theorem for monopoles suitable for the construction of Floer homology groups in the simplest cases and for establishing certain gluing formulae for Seiberg–Witten invariants of
–manifolds (to be discussed in [9] ).
There is now a large literature on gluing theory for instantons and monopoles. The theory was introduced by Taubes [17, 18] , who used it to obtain existence results for self-dual connections over closed
–manifolds. It was further developed in seminal work of Donaldson [2] , see also Freed–Uhlenbeck [8] .
General gluing theorems for instantons over connected sums were proved by Donaldson [3] and Donaldson–Kronheimer [5] . In the setting of instanton Floer theory there is a highly readable account in Donaldson [4] , see also Floer [7] and Fukaya [11] . Gluing with degenerate asymptotic limits was studied by Morgan–Mrowka [14] ; part of their work was adapted to the context of monopoles by Safari [16] . Nicolaescu [15] established gluing theorems for monopoles in certain situations, including one case involving gluing obstructions.
Marcolli–Wang [13] discuss gluing theory in connection with monopole Floer homology. For monopoles over closed
–manifolds split along certain tori, see Chen [1] . Finally, gluing theory is a key ingredient in a large programme of Feehan–Leness [6] (using ideas of Pidstrigach–Tyurin) for proving Witten's conjecture relating Donaldson and Seiberg–Witten invariants.
As should be evident from this brief survey, there are many different hypotheses under which one can consider the gluing problem. This paper does not aim at the utmost generality, but is an expositary account of gluing in what might be called the favourable cases. More precisely, we will glue precompact families of regular monopoles over
–manifolds with tubular ends, under similar general assumptions as in [10] . We do not consider gluing along degenerate critical points or obstructed gluing.
One source of difficulty when formulating a gluing theorem is that gluing maps are in general not canonical, but rather depend on various choices hidden in their construction. We have therefore chosen to express our gluing theorem as a statement about an ungluing map, which is explicitly defined in terms of data that appear naturally in applications.
If
is a
–manifold with tubular ends and
the glued manifold as in [10] , then the first component of the ungluing map involves restricting monopoles over
to some fixed compact subset
(which may also be regarded as a subset of
when each
is large). In the case of gluing along a reducible critical point, the ungluing map has an additional component which reads off the
gluing parameter by measuring the holonomy along a path running once through the corresponding neck in
.
Ungluing maps of a different kind were studied already in [2, 8] , but later authors have mostly formulated gluing theorems in terms of gluing maps, usually without characterizing these maps uniquely.
The proof of the gluing theorem is divided into two parts: injectivity and surjectivity of the ungluing map. In the first part the (quantitative) inverse function theorem is used to construct a smooth local right inverse
of an “extended monopole map”
. In the second part the inverse function theorem is applied a second time to show, essentially, that the image of
is not too small. There are many similarities with the proof of the gluing theorem in [5] , but also some differences. For instance, we do not use the method of continuity, and we handle gluing parameters differently.
It may be worth mentioning that the proof does not depend on unique continuation for monopoles (only for harmonic spinors), as we do not know if solutions to our perturbed monopole equations satisfy any such property.
(Unique continuation for genuine monopoles was used in [10,Proposition 4.3] in the discussion of perturbations, but this has little to do with gluing theory.) Therefore, in the injectivity part of the proof, we argue by contradiction, restricting monopoles to ever larger subsets
. This is also reflected in the statement of the theorem, which would have been somewhat simpler if unique continuation were available.
We will now outline the remainder of this paper. Section 2 contains the statement of the main theorem, which is proved in Sections 3 and 4 . Section 5 explains in a model case how the gluing theorem may be applied in conjunction with the compactnes results of [10] . This model case encompasses both the simplest gluing formulae for Seiberg–Witten invariants, and the formula
for the standard Floer differential as shown in Section 6 . The final two sections describe an extension of the gluing theorem to parametrized moduli spaces.
The reader may wish to skip the proof in Sections 3 , 4 on first reading and proceed directly from Section 2 to the model application in Section 5 .
The author gratefully acknowledges the hospitality of the Institut des Hautes Études Scientifiques, where the bulk of this work was carried out.
2 The gluing theorem
Consider the situation of [10,Subsection 1.4] , but without assuming any of the conditions (B1),(B2),(C). We now assume that every component of
contains an end
or
(ie an end that is being glued).
Fix non-degenerate monopoles
over
and
over
. (These should be smooth configurations rather than gauge equivalence classes of such.) Suppose
is reducible for
and irreducible for
, where
. We consider monopoles over
and
that are asymptotic to
over
and (in the case of
) asymptotic to
over
.
These monopoles build moduli spaces
Here
is a finite subset to be specified in a moment, and the subscript indicates that we only divide out by those gauge transformations that restrict to the identity on
, see [10,Subsection 3.4] . The ungluing map
will be a diffeomorphism between certain open subsets of
and
when
is large.
When gluing along the critical point
, the stabilizer of
in
appears as a “gluing parameter”. This stabilizer is a copy of
if
is reducible and trivial otherwise. When
is reducible we will read off the gluing parameter by means of the holonomy of the connection part of the glued monopole along a path
in
which runs once through the neck
. To make this precise, for
fix
and smooth paths
such that
for
. We require that these paths have different start-points. Let
denote the collection of all the start-points
. There is no loss of generality in assuming
. Note in passing that we then have
Define the smooth path
by
If
is a spin connection over
then what we need is not the holonomy along
of the induced connection in the determinant line bundle
, but rather the holonomy of the induced connection in a square root of
along
. To express this more conveniently, choose a reference configuration
over
with limit
over
and
over
. Let
denote the reference configuration over
obtained from
in the obvious way when gluing the ends. Precisely speaking,
is the unique smooth configuration over
which agrees with
over
(which can also be regarded as a subset of
).
Note that if
is any smooth manifold,
a smooth path,
a
connection over
, and
then
where
denotes holonomy along
. Motivated by this we define, for any spin connection
over
,
|
(1)
|
where as usual
is regarded as an imaginary valued
-form on
.
Then for gauge transformations
we have
In particular, there is a natural map
If
is any compact codimension
submanifold containing
then we define (using the
configuration spaces etc introduced in [10,Subsection 2.5] )
where
is the subgroup of
consisting of those gauge transformations whose homotopy class, identified with an element of
, is the image of a class in
whose restriction to each
is zero. Here
should be so large that
may be regarded as a subset of
, and the condition is then obviously independent of
. Note that
is a Hausdorff Banach manifold. Moreover, the group
acts freely and properly discontinuously on
with quotient
.
It is convenient here to agree once and for all that the Sobolev exponent
is to be an even integer. This ensures that our configuration spaces admit smooth partitions of unity, which is needed in Section 5 and in the appendix (but not in the proof of Theorem 1 ).
Fix an open subset
whose closure
is compact and contains only regular points.
Definition 1
By a kv-pair we mean a pair
where
-
∙
is a compact codimension
submanifold which contains
and intersects every component of
,
-
∙
is an open subset containing
, where
denotes restriction to
.
We define a partial ordering
on the set of all kv-pairs, by decreeing that
if and only if
and
.
Now fix a kv-pair
which satisfies the following two additional assumptions: firstly, that
; secondly, that if
is any component of
which contains a point from
then
is connected. The second condition ensures that the image of
lies in
.
Suppose we are given a smooth map
such that
for all
. (Such a map always exists, for a given
, provided
is sufficiently large, see the appendix. In concrete applications there is often a natural choice of
, see Sections 5 , 6 .) Let
and the forms
be as in [10,Subsection 1.4] , and choose
.
Theorem 1
Suppose there is class in
whose restrictions to
and
are
and
, respectively, and suppose the perturbation parameters
are admissible for
. Then there exists a kv-pair
such that if
is any kv-pair
then the following holds when
is sufficiently large. Set
Then
consists only of regular monopoles (hence is a smooth manifold), and the map
is a diffeomorphism.
Note that when
is large then
can also be regarded as a subset of
, in which case the expression
in the definition of
makes sense.
Concerning admissibility of perturbation parameters, see the remarks after [10,Theorem 1.4] . Note that the assumption on
and
in the theorem above is weaker than either of the conditions (B1) and (B2) in [10] .
However, in practice the gluing theorem is only useful in conjunction with a compactness theorem, so one may still have to assume (B1) or (B2).
The proof of Theorem 3 has two parts. The first part consists in showing that
has a smooth local right inverse around every point in
(Proposition 1 below). In the second part we will prove that
is injective (Proposition 2 below).
3 Surjectivity
The next two sections are devoted to the proof of Theorem 1 . Both parts of the proof make use of the same set-up, which we now introduce.
We first choose weight functions for our Sobolev spaces over
and
.
Let
be small constants and
a smooth function which is equal to
on
and equal to
on
. As usual, we require
if
is reducible (ie for
), and similarly for
. For
choose a smooth function
such that
for
. We will always assume
, in which case we can define a weight function
by
| |
| |
Let
denote the
configuration space over
defined by the reference configuration
, and let
denote the
configuration space over
defined by
. Let
be the corresponding groups of gauge transformations and
the corresponding orbit spaces.
Now fix
. In the case when
, our immediate goal is to construct a smooth local right inverse of
around this point.
However, the following set-up will also be used in the injectivity part of the proof.
Choose a smooth representative
for
which is in temporal gauge over each end
. (This assumption is made in order to ensure exponential decay of
over these ends.) Set
and let
be the projection. By the local slice theorem we can find a smooth map
such that
and such that
is a diffeomorphism onto an open subset of
.
We will require one more property of
, involving holonomy. If
then we define
by
The integral exists because, by the Sobolev embedding
in
for
, we have
|
(3)
|
for some constant
. It is clear that
is a smooth function on
.
Because any smooth map
factors through
, we can arrange, after perhaps modifying
by a smooth family of gauge transformations that are all equal to
outside the ends
and constant on
, that
|
(4)
|
for
and every
. Here
denotes the holonomy, as defined above, of the connection part of the configuration
, and the
are the coordinates of
.
Lemma 1
Let
be Banach spaces,
a bounded operator and
a surjective bounded operator such that
is Fredholm. Then
has a bounded right inverse.
Proof. Because
is Fredholm there is a bounded operator
such that
is compact. Set
for
. Then
is compact, hence
is Fredholm of index
. Using the surjectivity of
and the fact that any closed subspace of finite dimension or codimension in a Banach space is complemented, it is easy to see that there is a bounded operator
(with finite-dimensional image) such that
is an isomorphism.
Let
we be Seiberg–Witten map over
. By assumption, every point in
is regular, so in particular
is regular, which means that
is surjective. Let
be the spinor part of
and define
as in [10,Subsection 2.3] . Then
is Fredholm, so by Lemma 1
has a bounded right inverse
. (This can also be deduced from [10,Proposition 2.2 (ii)] .) Let
be a smooth function such that
for
and
for
. For
set
and let
be the configuration over
which agrees with
away from the ends
and satisfies
over
. Here
denotes, as before, the translationary invariant monopole over
determined by
. For each
we have
Therefore, when
is sufficiently large, the operator
will be invertible, and we set
which is then a right inverse of
. It is clear that the operator norm
as
.
For the remainder of the proof of Theorem 1 , the term `constant' will always refer to a quantity that is independent of
, unless otherwise indicated. The symbols
and
will each denote at most one constant, while other symbols may denote different constants in different contexts.
Consider the configuration space
over
and the Seiberg–Witten map
As explained in [10,Subsection 3.4] there is an identification
By the results of [4] the operator on the right hand side defines a Fredholm operator
, and this must be surjective because of the choice of weight function
. In particular,
is surjective, hence has a bounded right inverse
by Lemma 1 . (Here one cannot appeal to [10,Proposition 2.2 (ii)] .) Let
be the Seiberg–Witten map over
. When
then by splicing
in the natural way one obtains a smooth configuration
over
. There is a constant
such that if
then we can splice the right inverses
and
to obtain a right inverse
of
which satisfies
for some constant
. The construction of
is analogous to the construction of the left inverse of the operator
in the proof of [10,Lemma 5.4] . Since
is bounded in
(ie as a function of
), we see that
is bounded in
.
The inequality ( 5 ) will be assumed from now on.
We now introduce certain
-forms that will be added to the configurations
in order to make small changes to the holonomies
. For any
define the
-form
over
by
where
. Set
For
let
be the open
-ball about
. Define a smooth map
by
|
(6)
|
where
is added to the connection part of
.
When deciding where to add the perturbation
-form
one has to balance two concerns. One the one hand, because the weight function
increases exponentially as one approaches the middle of the necks
,
, it is desirable to add
as close to the boundaries of these necks as possible. On the other hand, in order for Lemma 4 below to work, the spinor field of
needs to be “small” in the perturbation region. We have chosen to add
at the negative end of the cut-off region, where the spinor field is zero.
Although we will sometimes use the notation
, we shall think of
as a function of three variables
, and
will denote the derivative of
with respect to the
'th variable. Similarly for other functions on (subsets of )
that we will define later. Set
Notice that if
, ie if we are not gluing along any reducible critical point, then we may take
.
Lemma 2
There exists a constant
such that for
the following hold:
-
(i)
if
,
-
(ii)
,
-
(iii)
.
Proof. To prove (ii), note that if
and
then
The other two statements are left to the reader.
Let
be the set of all
such that
,
, and
for
. Then
is an open subset of
, and there are unique smooth functions
such that
. Set
and define
A crucial point in the proof of Theorem 1 will be the construction of a smooth local right inverse of
, defined in a neighbourhood of
. The map
is a first approximation to such a local right inverse. The construction of a genuine local right inverse will involve an application of the quantitative inverse function theorem (see Lemma 7 below). From now on we will take
so large that
and
for
. Note that the left hand side of this equation is equal to
whenever
. There is then a constant
such that
, in which case we have a composite map
Choose
so that none of the operators
(
) and
(
) has any eigenvalue of absolute value
. (The notation
was introduced in [10,Subsection 6.1] .) Recall that we assume the
are small and non-negative, so in particular we may assume
.
Lemma 3
There is a constant
such that
Proof. The first two components of
are in fact zero:
the first one because
over
, the second one because the
-component of
vanishes on
since
and
are both in temporal gauge there.
The third component of
is
. It suffices to consider
so large that the
-perturbations do not contribute to
, which then vanishes outside the two bands of length
in
centred at
,
. Our exponential decay result from [10] says that for every
there is a constant
such that for every
we have
Consequently,
This yields
Lemma 4
There is a constant
such that for sufficiently large
the following hold:
-
(i)
,
-
(ii)
is invertible and
.
Proof. By construction, the derivative of
at
has the form
where the
'th column is the
'th partial derivative and
the identity map.
The middle top entry in the above matrix is zero because
vanishes on
. The middle bottom entry is zero because
on the support of
and the spinor field of
is zero (for
). Adding
to
therefore has the effect of altering the latter by a gauge transformation over
,
.
We claim that
is bounded in
for
. For
this is obvious from the boundedness of
. For
note that the derivative of
at any
is
|
(7)
|
where
is an imaginary valued
-form and
a positive spinor. Because of the weights used in the Sobolev norms it follows that
is (independent of
and) bounded in
(see ( 3 )). This together with the bound on
gives the desired bound on
.
Note that, for
,
is independent of
when
, and routine calculations show that
as
. (In the case of
this depends on the normalization ( 4 ) of the holonomy of
.) Write
, where
When
is so large that
then of course
, hence
is invertible. Moreover,
which gives
We now record some basic facts that will be used in the proof of Lemma 6 below.
Lemma 5
If
are Banach spaces,
an open set for
, and
,
smooth maps then the second derivate of the composite map
is given by
| |
| |
for
and
.
Proof. Elementary. It is also worth noting that embedding and multiplication theorems for
Sobolev spaces on
(
,
) carry over to
, and that the embedding and multiplication constants are bounded functions of
.
Furthermore, a differential operator of degree
over
which is translationary invariant over necks and ends induces a bounded operator
whose operator norm is a bounded function of
.
Lemma 6
There is a constant
such that
whenever
and
.
Proof. We will say a quantity depending on
is s-bounded if the lemma holds with this quantity in place of
.
Let
be the components of
.
The assumption
ensures that
is independent of
. It then follows from Lemma 5 and the bound on
that
is s-bounded.
When
are small we have
Since
is constant, as noted above, we have
. From the bounds on
and
we then deduce that
is s-bounded.
To estimate
, we fix
and consider only
. It is easy to see that
is s-bounded. By restricting to small
and choosing
large we may arrange that the
and
-perturbations do not contribute to
for
. We need to show that each
is s-bounded, but to simplify notation we will instead prove the same for
under the assumption that the
perturbations are zero.
First observe that for any configuration
over
and any closed, imaginary valued
-form
we have
Moreover,
for some constants
. Taking
and
we see that
is s-bounded for
.
Next note that the derivative of the Seiberg–Witten map
at a point
has the form
where
is a pointwise bilinear operator, and
a first order operator which is independent of
and translationary invariant over necks and ends. This yields
|
(8)
|
Moreover,
for all
, hence there is a constant
such that
for all
.
Combining the above results on
with Lemma 5 we see that
is s-bounded also when
.
Lemma 7
There exist
and
such that if
then for sufficiently large
the following hold:
-
(i)
is injective,
-
(ii)
There is a (unique) smooth map
such that
,
-
(iii)
for all
,
-
(iv)
for all
,
-
(v)
.
Proof. For sufficiently large
we have
by Lemma 3 . Statements (i)–(iv) now follow from the inverse function theorem, [10,Proposition B.1] , applied to the function
, together with Lemmas 4 and 6 . To prove (v), set
,
and take
so large that
. Since
is injective on
we must have
, so
Now (v) follows from (iii) and Lemma 3 .
From now on we assume that
are chosen so that the conclusions of the lemma are satisfied. Define
Then clearly
Thus,
is a “gluing map”, ie it solves the problem of gluing the monopoles
over
(for
small) to get a monopole over
with prescribed holonomy
along the path
for
.
Lemma 8
There is a constant
such that for
one has
Proof. This follows from Lemmas 2 and 7 and the chain rule.
We will now prove roughly one half of Theorem 1 .
Proposition 1
In the situation of Section 2 , if
is any kv-pair
then the following holds when
is sufficiently large. Define
and
as in Theorem 1 . Then
is a union of finitely many open sets
such that for each
there exists a smooth map
whose image consists only of regular points, and which satisfies
.
Here we do not need any assumptions on
or on
.
Proof. Let
and consider the set-up above, with
so large that
and
so small that
|
(9)
|
Note that taking
small may require taking
(and hence
) large, see Lemma 7 . For any sufficiently small open neighbourhood
of
we can define a smooth map
by the formula
Here
for any complex number
with
, and
denotes the vector whose
'th component is
. Because
and the Seiberg–Witten map is the third component of
, the image of
consists of regular monopoles. Set
and
, where
is the projection. Unravelling the definitions involved and using ( 9 ) one finds that
maps into
, and that
.
How large
must be for this to work might depend on
. But
is compact, hence it can be covered by finitely many such open sets
. If
is sufficiently large then the above construction will work for each of these
.
4 Injectivity
We now continue the discussion that was interrupted by Proposition 1 .
Set
Lemma 9
There is a constant
such that for sufficiently large
one has
Proof. Set
. For sufficiently large
we have
| |
| |
| |
where we used Lemma 7 (v) to obtain the last inequality. Because
is bounded in
, and we assume
, the second inequality of the lemma follows as well.
For positive spinors
on
it is convenient to extend the definition of
to complex valued functions on
:
(However,
will always refer to the formal adjoint of
acting on imaginary valued functions.) When
is the spinor part of
then the corresponding operators
will be denoted
, respectively. (We omit the ^ on
to simplify notation.) As in [10,Subsection 2.2] we define
We can take the norm to be
Lemma 10
There is a constant
such that if
is any of the operators
then for all
one has
Proof. We first prove the inequality for
(the case of
is similar, or easier). If
is any component of
and
then for some constant
one has
for all
functions
. Here the Sobolev inequality holds because
, whereas the second inequality follows from [10,Lemma 2.1] . We use part (i) of that lemma if the spinor field of
is not identically zero on
, and part (ii) otherwise. (In the latter case
is non-empty.) When
are sufficiently large we can apply part (i) of the same lemma in a similar fashion to the band
provided
and
is irreducible, and to the band
provided
and
is irreducible. To estimate
over these bands when
resp.
is reducible one can use [10,Lemma 2.2(ii)] . This proves the lemma for
(and for
). We now turn to the case
. Let
denote the spinor part of
. Then
| |
| |
By Lemma 9 we have
as
, so for sufficiently large
we get
Therefore, the lemma holds with
as well.
Lemma 11
There is a constant
such that for all
and
one has
-
(i)
,
-
(ii)
,
where we use the
norm on spinors and the
norm on elements of
.
Proof. By routine calculation using Lemma 10 with
one easily proves (ii) and the inequality
Now observe that by definition
, where as before
denotes the spinor field of the reference configuration
. Applying (ii) we then obtain
completing the proof of (i).
Recall from [10,Subsection 2.4] that the Lie algebra
is the space of imaginary valued functions in
that vanish on
.
Lemma 12
There is a constant
such that for
and all
one has
Proof. Let
denote the spinor part of
. Then
| |
| |
and similarly with
instead of
. The lemma now follows from Lemma 9 .
We are going to use the inverse function theorem a second time, to show that the image of the smooth map
| |
| |
contains a “not too small” neighbourhood of
. The derivative of
at
is
| |
| |
To be concrete, let
have the norm
.
Lemma 13
is a linear homeomorphism.
Proof. By [10,Proposition 2.2] ,
is a Fredholm operator with the same kernel as
. Now,
is injective on
, because
maps into
and therefore
. Since
is a closed subspace of
of finite codimension, we can choose a bounded operator
such that
. Set
Then
is an isomorphism. Furthermore,
where `dim' refers to expected dimension (which in the case of
is equal to the actual dimension of
), and the second equality follows from the addition formula for the index (see [4] ). Consequently,
We now compute
|
(10)
|
where
. The zero in the matrix above is due to the fact that
which holds because
are
invariant,
is
equivariant, and
.
Since the right hand side of ( 10 ) is invertible, it follows that
is a surjective Fredholm operator of index
, hence invertible. Of course, this implies that
is also invertible.
Lemma 14
There is a constant
such that for sufficiently large
,
Proof. In this proof all unqualified norms are
norms. It follows from ( 7 ), ( 8 ) and Lemma 9 that
is bounded in
. Therefore there exists a constant
such that
for all
and
. From Lemma 12 and Lemma 8 we get
| |
| |
| |
| |
This yields
Lemma 15
There is a constant
such that for sufficiently large
one has
for all
,
such that
and
.
Proof. For the purposes of this proof it is convenient to rescale the norm on
so that we can take
in Lemma 11 .
If
then
, and from Lemma 11 we obtain
and similarly with a spinor
instead of
.
The first two derivatives of
are
so
Let
be the connection and spinor parts of
, respectively, and define
similarly. Then
We regard
as a function of the the two variables
. Let
denote the derivative of
with respect to the
'th variable. Similarly for the second derivatives
.
Applying Lemmas 8 and 11 we now find that
for
and
. Since
for
, the lemma is proved.
In the following,
will denote the open
-ball about
(both in various Banach spaces and in
).
Lemma 16
There exist constants
such that for sufficiently large
the image of
contains the ball of radius
about
in
; more precisely one has
Proof. We wish to apply the inverse function theorem [10,Proposition B.1] to the map
restricted to a ball
, where
is to be chosen. For the time being let
have the same meaning as in that proposition. By Lemma 15 we can take
, and by Lemma 14 we have
. We need
This will hold if
When
is large we can take
to be the right hand side of this inequality.
By [10,Proposition B.1] ,
contains the ball
where
Theorem 1 is a consequence of Proposition 1 and the following proposition:
Proposition 2
Theorem 1 holds if in the last sentence of that theorem the words “a diffeomorphism” are replaced by “an injective map”.
The proof of Proposition 2 occupies the remainder of this section.
Note first that shrinking
has the effect of restricting the map
to a perhaps smaller domain. It therefore suffices to prove that for some kv-pair
the map
is injective for all sufficiently large
.
For any natural number
which is so large that
, let
be the set of all
such that there exist a representative
of
, and a configuration
over
representing an element of
, such that
Note that
for any gauge transformation
over
.
Lemma 17
Let
for
, where
. Then there exists for each
a representative
of
such that a subsequence of
converges locally in
over
to a smooth configuration representing an element of
.
Proof. By assumption there exist for each
a representative
of
and a configuration
over
representing an element of
such that
|
(11)
|
After passing to a subsequence we may assume (since
is compact) that
converges in
to some element
, and we can choose
smooth. Since
, the local slice theorem guarantees that for large
we can find
such that
satisfies
Set
, which is again a representative of
. Let
be the connection parts of
, respectively. Then ( 11 ) implies that
and
locally in
over
, hence also
locally in
over
. Now
and each of the three terms on the right hand side converges to
locally in
over
(the third term because of the continuous multiplication
in
for
). Hence
locally in
over
.
Corollary 1
For sufficiently large
one has that
.
Lemma 18
Let
for
, where
. Suppose
converges in
to an element
as
. Then
, and there exists for each
a representative
of
such that the sequence
converges locally in
over
to a smooth configuration representing
.
Proof. Let
be as in the proof of Lemma 17 . First suppose that
converges in
to some element
, where
is smooth. Choosing
as in that proof we find again that
locally in
over
, hence
We now turn to the general case when
is not assumed to converge.
Because
is compact, every subsequence of
has a convergent subsequence whose limit must be
by the above argument. Hence
.
Suppose we are given a sequence
of natural numbers tending to infinity, and for each
an
-tuple
of real numbers such that
Define
and
as in Theorem 1 , with
and
.
Lemma 19
For
suppose
is a smooth configuration over
representing an element
, and such that
as
. There exists a constant
such that for sufficiently large
the following holds for sufficiently large
. Let the map
be defined as above and set
. Then there exists a smooth gauge transformation
such that
Note: This constant
depends on
but not on the sequence
.
Before proving the lemma, we will use it to show that
is injective for some
. This will prove Proposition 2 . Suppose
and
,
. After passing to a subsequence we may assume that
converges to some point
.
Combining Lemmas 16 , 19 and the assumption
we conclude that if
is sufficiently large then for sufficiently large
we can represent
and
by configurations
and
, respectively, where
.
Now recall that
, and that the components
are
invariant whereas
is the Seiberg–Witten map. Comparing the definitions of
and
we conclude that
hence
for large
. To complete the proof of Proposition 2 it therefore only remains to prove Lemma 19 .
Proof of Lemma 19 : In this proof, constants will be independent of the sequence
(as well as of
as before).
By Lemma 18 we can find for each
an
gauge transformation
over
with
such that
converges locally in
over
to a smooth configuration
representing
. A moment's thought shows that we can choose the
smooth, and we can clearly arrange that
.
Then for any
we have
|
(12)
|
when
is so large that Lemma 9 applies.
For
and any smooth configurations
over
consider the functional
where in this formula
has the boundary orientation inherited from
. (Recall that the Chern-Simons-Dirac functional
changes sign when the orientation of the
-manifold in question is reversed.) The assumption on
and
in Theorem 1 implies that
depends only on the gauge equivalence class of
. Since
is a smooth function on the
configuration space by [10,Lemma 3.1] , we obtain
as
. By our exponential decay results (see the proof of [10,Theorem 6.1] ),
It follows that
for some positive function
. By assumption the perturbation parameters
are admissible, hence there is a constant
such that when
, each of the
summands appearing in the definition of
is non-negative. Explicitly, this yields
where the first line holds for
and
, the second line for
and
, and in both cases we assume
and
.
In the following we will ignore the ends
of
, ie we will pretend that
is compact. If
is irreducible then the argument for dealing with the end
is completely analogous to the one given below for a neck
, while if
is reducible it is simpler. (Compare the proof of [10,Proposition 6.3(ii)] .) For the remainder of the proof of this lemma we will focus on one particular neck
where
. To simplify notation we will therefore mostly omit
from notation and write
etc instead of
.
For
set
regarded as a subset of
. By the above discussion there is a constant
such that when
is sufficiently large,
will restrict to a genuine monopole over the band
by [10,Lemmas 4.1,4.2,4.3] and will have small enough energy over this band for [10,Theorem 6.2] to apply. That theorem then provides a smooth
such that
is in temporal gauge and
Writing
we get
|
(13)
|
when
and
is so large that Lemma 9 applies.
To complete the proof of the lemma we interpolate between
and
in the overlap region
. (This requires
.) The choice of this overlap region is somewhat arbitrary but simplifies the exposition.
Define
Then
Set
, where
is the path introduced in the beginning of this section. If
is reducible then by multiplying each
by a constant and redefining
accordingly we can arrange that
for all
.
These changes have no effect on the estimates above.
Lemma 19 is a consequence of the estimates ( 12 ), ( 13 ) together with the following sublemma (see the proof of [10,Proposition 6.3(ii)] .)
Sublemma 1
There is a constant
such that if
then
Proof of sublemma: If
is irreducible then the sublemma follows from inequalities ( 12 ), ( 13 ) and [10,Lemmas 6.9,6.11] . (In this case the sublemma holds with
as upper bound.) Now suppose
is reducible. We will show that
|
(14)
|
for large
. Granted this, we can prove the sublemma by applying [10,Lemma 6.9] and [10,Lemma 6.10(ii)] to each component of
.
In the remainder of the proof of the sublemma we will omit
from subscripts. To prove ( 14 ), define intervals
and for
set
. Let
denote holonomy along
in the same sense as ( 1 ), ie
is the result of replacing the domain of integration
in that formula with
. Define
by
| |
| |
where as usual we mean holonomy with respect to the connection parts of the configurations. For large
the estimates ( 12 ) and ( 13 ) give
when
is sufficiently large.
Writing
we have
Now, by the definition of holonomy,
and there are similar formulas for
. Because
we obtain
Setting
we get
Since by assumption
as
, we have
for large
, proving the sublemma and hence also Lemma 19 .
This completes the proof of Proposition 2 and thereby also the proof of Theorem 1 .
5 A model application
In this section we will show in a model case how the gluing theorem may be applied in combination with the compactness results of [10] . Here we only consider gluing along irreducible critical points. Examples of gluing along reducible critical points will appear in [9] . The main result of this section, Theorem 2 , encompasses both the simplest gluing formulae for Seiberg–Witten invariants (in situations where reducibles are not encountered) and, as we will see in Section 6 , the formula
for the standard Floer differential.
Recall that the Seiberg–Witten invariant of a closed
–manifold (with
) can be defined as the number of points (counted with sign) in the zero-set of a generic section of a certain vector bundle over the moduli space. To obtain a gluing formula, this vector bundle and its section should be expressed as the pull-back of a vector bundle
with section
, where
. In the proof of Theorem 2 below we will see how the section
gives rise in a natural way to a map
as in Theorem 1 . Thus, the section
is being incorporated into the equations that the gluing map is required to solve. (We owe this idea to [4,p 99] .) We will now describe the set-up for our model application. Let
be as in [10,Subsection 1.4] with
and
, and set
. In other words, we will be gluing one single pair of ends
of
, but
may have other ends
not involved in the gluing. We assume
is connected, which means that
has one or two connected components.
For
fix a critical point
. Let
be a
-form and
a perturbation parameter for
, and let
be similar data for
. Let each
have small
norm. To simplify notation we write, for
,
We make the following assumptions:
-
∙
(Compactness) At least one of the conditions (B1), (B2) of [10] holds,
-
∙
(Regularity) All moduli spaces over
,
, and
contain only regular points, and
-
∙
(No reducibles) Given
and
, if there exist a broken gradient line over
from
to
and for each
a broken gradient line over
from
to
then
contains no reducible. (It then follows by compactness that
contains no reducible when
is large.)
The regularity condition is stronger than necessary, because there are energy constraints on the moduli spaces that one may encounter in the situation to be considered, but we will not elaborate on this here.
Note that we have so far only developed a full transversality theory in the case when
and each
are rational homology spheres; in the remaining cases the discussion here is therefore somewhat theoretical at this time.
Let
be a compact codimension
submanifold which intersects every component of
. When
then
may also be regarded as a submanifold of
, and we have restriction maps
These take values in
rather than just in
because of the unique continuation property of harmonic spinors.
Suppose
is an oriented smooth real vector bundle whose rank
is equal to the (expected) dimension of
. Choose a smooth section
of
such that the pull-back section
is transverse to the zero-section of the pull-back bundle
over
for each pair
. (Here the Sobolev exponent
should be an even integer to ensure the existence of smooth partitions of unity.) Set
, which is a section of
. We write
and
etc.
Let
,
denote the zero-sets of
,
respectively. By index theory we have
where
if
is irreducible and
otherwise. Thus,
is empty if
is reducible.
Lemma 20
If
for
, where
, then a subsequence of
chain-converges to an element of
for some
.
Moreover, if
chain-converges to
then there exists for each
a smooth
whose restriction to each end
is null-homotopic and such that the sequence
c-converges over
to
.
Proof. The statement of the first sentence follows from [10,Theorem 1.4] by dimension counting. Such maps
exist in general for chain-convergent sequences when the
all have the same asymptotic limits over the ends
, see [10] . Let
be the subgroup consisting of elements of the form
where
is an element of
satisfying
for
.
This group
acts on the disjoint union
permuting the summands.
Lemma 21
The quotient
is a finite set.
Proof. By [10,Theorem 1.3] any sequence
,
has a chain-convergent subsequence, and for dimensional reasons the limit (well-defined up to gauge equivalence) must lie in some moduli space
. Furthermore, if
chain-converges to an element in
then
is contained in the orbit
for
. Therefore, each
is a finite set, and only finitely many orbits
are non-empty. This is equivalent to the statement of the lemma.
Note that
is the largest subgroup of
which acts on
in a natural way. On the other hand, if
is non-emtpy then, since
acts freely on
, only subgroups
of finite index have the property that
is finite.
Lemma 22
There is a compact codimension
submanifold
such that the restriction map
is injective.
Proof. Let
,
, where each
is in temporal gauge over the ends of
(and therefore decays exponentially). Suppose there exists a sequence of smooth gauge transformations
where
, such that
over
. After passing to a subsequence we can arrange that
c-converges over
to some gauge transformation
with
. If
then
will both be homotopic to a smooth
with
. Hence
lie in the same
–orbit, and
represent the same element of
by [10,Proposition 2.5 (iii)] .
Thus we can take
for
.
Now fix
as in Lemma 22 and with
. Let
be the image of the restriction map
. Choose disjoint open neighbourhoods
of the points
. If
then
by Lemma 20 . For such
we get a natural map
It is clear that if
is the map corresponding to a different choice of
and neighbourhoods
then
for
sufficiently large.
Theorem 2
For sufficiently large
the following hold:
-
(i)
Every element of
is a regular point in
and a regular zero of
.
-
(ii)
is a bijection.
Proof. If
is empty then, by Lemma 20 ,
is empty as well for
, and there is nothing left to prove.
We now fix
and for the remainder of the proof omit
from notation.
(Thus
,
etc.) We will show that for
the set
consists of precisely one element, and that this element is regular in the sense of (i). This will prove the theorem.
By definition,
is the restriction of some
. Choose an open neighbourhood
of
and a smooth map
which restricts to a linear isomorphism on every fibre. Choose an open neighbourhood
of
such that
. Let
be a precompact open neighbourhood of
such that
. The assumption that
be a regular zero of
means that the composite map
is a local diffeomorphism at
. We can then find an injective smooth map
such that
in some open neighbourhood
of
.
In particular,
and
is a local diffeomorphism at
. Set
By Theorem 1 there is a kv-pair
such that if
then
consists only of regular monopoles and
is a diffeomorphism. By Lemma 20 one has
for
. For such
the set
consists of precisely one point, and this point is regular in the sense of (i).
6 The Floer differential
Consider the situation of [10,Subsection 1.2] . Suppose a perturbation parameter
of small
norm has been chosen for which all moduli spaces
over
are regular. (This is possible at least when
is a rational homology sphere, see [10] .) Fix
with
We will show that the disjoint union
is the boundary of a compact
–manifold. (In other words, the standard Floer differential
satisfies
at least with
coefficients.) To this end we will apply Theorem 2 to the case when
consists of two copies of
, say
and we glue
with
. Thus
. We take
, where
. In this case,
is the quotient of
by the null-homotopic gauge transformations. The bundle
over
will be the product bundle with fibre
. To define the section
of
, choose
such that
has no critical value in the set
This is possible because we assume Condition (O1) of [10] . For any configuration
over
set
Note that
does not change if we apply a null-homotopic gauge transformation to
.
A configuration over
consists of a pair
of configurations over
. Define a smooth function
(ie a section of
) by
If
belongs to some moduli space
over
with
in
then
for all
by choice of
. Since
, the natural map
is therefore a bijection. Let
be the pull-back of
to
. Here
is defined using [10,Equation 17] with
, and so can be identified with
. By [10,Theorem 1.3] the set
is compact for all
. If
then, by Theorem 2 ,
is a smooth submanifold of
, and the composition of the two bijections
yields the desired identification of
with the boundary of a compact
-manifold.
7 Parametrized moduli spaces
We will now introduce certain parametrized moduli spaces that are useful for the construction of Floer homology groups. A natural setting here would involve certain fibre bundles whose fibres are
-manifolds. We feel, however, that gauge theory for such bundles in general deserves a separate treatment, and will therefore limit ourselves, at this time, to the case of a product bundle over a vector space.
As in [10,Subsection 1.3] let
be a
-manifold with Riemannian metric
and tubular ends
,
. Let
be a finite-dimensional Euclidean vector space and
a smooth family of Riemannian metrics on
all of which agree with
outside
. We then have a principal
bundle
whose fibre over
consists of all positive
-orthonormal frames in
.
In the notation of [10,Subsection 3.1] let
be the
structure on
. Denote by
the pull-back of
under the projection
. Then
is a principal
bundle over
.
For
let
be a non-degenerate smooth monopole.
Let
denote the
configuration space for the metric
and limits
, where
are as in [10,Subsection 3.4] . We will provide the disjoint union
with a natural structure of a (trivial) smooth fibre bundle over
. Let
|
(15)
|
be any isomorphism of
bundles which covers the identity on
and which outside
is given by the identification
.
There is then an induced isomorphism of
bundles
, since these are quotients of the corresponding
bundles by the
action. Such an isomorphism
can be constructed by means of the holonomy along rays of the form
where
, with respect to any connection in
which outside
is the pull-back of a connection in
. Then
induces a
equivariant diffeomorphism
|
(16)
|
for each
, where the map on the spin connections is obtained by identifying these with connections in the respective determinant line bundles and applying the isomorphism between these bundles induced by
. Putting together the maps ( 16 ) for all
yields a bijection
If
is another isomorphism as in ( 15 ) then
is smooth, hence we have obtained the desired structure on
. Furthermore, because of the gauge equivariance of
we also get a similar smooth fibre bundle structure on
|
(17)
|
for any finite subset
.
The monopole equations on
that we will consider will involve an additional class of perturbations which are useful for handling deformations of
-perturbations. Such a perturbation will depend on the choice of an isomorphism
as above, a non-negative constant
, and a smooth map
whose image consists of smooth
-forms supported in the interior of
.
We make two further assumptions on
. Firstly,
should be gauge invariant in the restricted sense that
for all
and all smooth
that restrict to
on
. Here
acts trivially on
. Secondly, for every non-negative integer
the image of
should be a bounded set with respect to the
norm.
Having made these choices, for any
we denote by
the set of all elements
that satisfy the perturbed monopole equations
|
(18)
|
where
Here
and the term involving
should be of the same kind as in [10,Equation 18] , and of course the spin bundles and Dirac operators are those corresponding to
. Taking the disjoint union of these moduli spaces for all
yields the parametrized moduli space
To get a more concrete description of the parametrized monopole equations note that the left hand side of ( 18 ) defines a fibre-preserving smooth map from
into a Banach vector bundle over
whose fibre over
is
. If we conjugate this monopole map by the diffeomorphisms induced by
we obtain the smooth
equivariant map
Here
denotes the isomorphism that
induces from the Clifford bundle of
to the Clifford bundle of
, and
is a local
-orthonormal frame on
. Finally, if we temporarily let
denote the
-Riemannian connection in the tangent bundle of
then
Note that
is supported in
.
By construction,
induces a homeomorphism
A point in
is called regular if the corresponding zeros of
are regular (a regular zero being one where the derivative of
is surjective).
This notion is independent of
. By the local slice theorem, the set of regular points in
is a smooth submanifold of
.
8 The gluing theorem in the parametrized case
We continue the discussion of the previous section, but we now specialize to the case when the ends of
are
,
and
,
, with non-degenerate limits
over
and
over
, as in Section 2 . The family of metrics
on
defines, in a natural way, a smooth family of metrics
on
for any
. To simplify notation we write
and
. The corresponding orbit spaces will be denoted
, respectively. The parametrized moduli space over
is denoted
We digress briefly to point out that the notions of chain-convergence and admissible perturbations have natural generalizations to the parametrized situation, and that the compactness theorem [10,Theorem 1.4] carries over to sequences
provided the sequence
is bounded (and similarly for [10,Theorem 1.3] ). The only new ingredient in the proof is the following simple fact: Suppose
is a Banach space,
vector bundles over a compact manifold,
differential operators of order
, and
a linear operator. If
satisfies an inequality
and
are sufficiently close in the sense that
for some constant
with
, then
obeys the inequality
This ends our digression, and we continue the main discussion.
Let the paths
and
be as in Section 2 , so that
. As before, a choice of reference configuration in
gives rise to a reference configuration in
and a holonomy map
Composing this with the map
defined by the chosen isomorphism
yields a holonomy map
Fix an open subset
whose closure is compact and contains only regular points.
By a kv-pair we mean as before a pair
, where
is a compact codimension
submanifold which contains
and intersects every component of
, and
is an open neighbourhood of
in
Now fix a kv-pair
satisfying similar additional assumptions as before: firstly, that
; secondly, that if
is any component of
which contains a point from
then
is connected.
Suppose
is a smooth map such that
for all
. Choose
.
Theorem 3
Theorem 1 holds in the present situation if one replaces
by
.
Proof. The proof of Theorem 1 carries over without any substantial changes.
A Existence of maps
Let
be as in Section 2 . In this appendix we will show that there is always a map
as in ( 2 ) provided
is sufficiently large. It clearly suffices to prove the same with
in place of
.
Throughout this appendix
will be as in Theorem 1 , ie a compact codimension
submanifold of
containing
such that if
is any component of
which contains a point from
then
is connected. The restriction map
then takes values in
.
Proposition 3
If
is sufficiently large then there exist an open neighbourhood
of
and a smooth map
such that
for all
.
We first prove two lemmas. Choose a precompact open subset
containing
.
Lemma 23
If
is sufficiently large then
restricts to an immersion on an open neighbourhood of
.
By `immersion' we mean the same as in [12] . Since a finite-dimensional subspace of a Banach space is always complemented, the condition in our case is simply that the derivative of the map be injective at every point.
Proof. Fix
. We will show that
is an immersion at
(hence in a neighbourhood of
) when
is large enough. Since
is compact, this will prove the lemma.
Let
be the ambient configuration space containing
as defined in the beginning of Section 3 .
Let
be a linear subspace such that the derivative at
of the projection
is a linear isomorphism onto the tangent space of
at
. Let
denote that derivative. For
so large that
let
be the derivative at
of the natural map
. We claim that
is injective for
. For suppose
is a sequence in
such that
and
for each
, where
. Set
.
Then
for some
, where
is the spinor field of
. After passing to a subsequence we may assume that
converges in
to some
.
By [10,Lemma 2.1] there exists for each
a constant
such that for all
one has
It follows that
converges in
over compact subsets of
, with limit
, say. We obviously have
, hence
and
. But this is impossible, since
has norm
. This proves the lemma.
Lemma 24
If
is sufficiently large then the restriction map
is a smooth embedding.
Proof. Argue as in the proof of Lemma 22 and apply Lemma 23 .
Proof of Proposition 3 : It suffices to show that
has a tubular neigbourhood in
. According to [12,p96] , if
is a Banach manifold admitting partitions of unity then any closed submanifold
possesses a tubular neigbourhood. However, the condition that
be closed is in fact redundant. To see this, note that
is by definition locally closed, hence
is closed in
. But then
is a closed submanifold of
.
Since a tubular neigbourhood of
in
is also a tubular neigbourhood of
in
, the assertion follows.
References
-
W. Chen. Casson's invariant and Seiberg–Witten gauge theory. Turkish J. Math., 21:61–81, 1997.
-
S. K. Donaldson. An application of gauge theory to four dimensional topology. J. Differential Geometry, 18:279–315, 1983.
-
S. K. Donaldson. Connections, cohomology and the intersection forms of
–manifolds. J. Differential Geometry, 24:275–341, 1986.
-
S. K. Donaldson. Floer Homology Groups in Yang–Mills Theory. Cambridge University Press, 2002.
-
S. K. Donaldson and P. B. Kronheimer. The Geometry of Four-Manifolds. Oxford University Press, 1990.
-
P. M. N. Feehan and T. G. Leness. A general
–monopole cobordism formula relating Donaldson and Seiberg–Witten invariants. math.DG/0203047.
-
A. Floer. An instanton invariant for 3–manifolds. Comm. Math. Phys., 118:215–240, 1988.
-
D. S. Freed and K. K. Uhlenbeck. Instantons and Four-Manifolds. MSRI Publications. Springer-Verlag, second edition, 1991.
-
K. A. Frøyshov. Monopole Floer homology for rational homology
–spheres. In preparation.
-
K. A. Frøyshov. Monopoles over
–manifolds containing long necks, I. Geometry & Topology, 9:1–93, 2005.
-
K. Fukaya. Floer homology for oriented
–manifolds. In Y. Matsumoto, editor, Aspects of low dimensional manifolds, volume 20 of Adv. Stud. Pure Math., pages 1–92. Kinokuniya Company Ltd, 1992.
-
S. Lang. Differential Manifolds. Springer-Verlag, 1985.
-
M. Marcolli and B-L. Wang. Equivariant Seiberg–Witten Floer homology. Comm. Anal. and Geom., 9:451–639, 2001.
-
J. W. Morgan and T. S. Mrowka. On the gluing theorem for instantons on manifolds containing long cylinders, 1994. Preprint.
-
L. I. Nicolaescu. Notes on Seiberg–Witten theory. American Mathematical Society, 2000.
-
P. Safari. Gluing Seiberg–Witten monopoles. math.DG/0311329.
-
C. H. Taubes. Self-dual Yang–Mills connections on non-self-dual 4-manifolds. J. Differential Geometry, 17:139–170, 1982.
-
C. H. Taubes. Self-dual connections on
–manifolds with indefinite intersection matrix. J. Differential Geometry, 19:517–560, 1984.
Fakultat fur Mathematik, Universitat Bielefeld, Postfach 100131, D-33501 Bielefeld, Germany. E-mail: froyshov@mathematik.uni-bielefeld.de