Monopoles over 4   –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 4   –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 4   –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 [3and Donaldson–Kronheimer [5. In the setting of instanton Floer theory there is a highly readable account in Donaldson [4, see also Floer [7and 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 [15established gluing theorems for monopoles in certain situations, including one case involving gluing obstructions.
Marcolli–Wang [13discuss gluing theory in connection with monopole Floer homology. For monopoles over closed 3   –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 4   –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 X   is a 4   –manifold with tubular ends and X ( T )   the glued manifold as in [10, then the first component of the ungluing map involves restricting monopoles over X ( T )   to some fixed compact subset K X   (which may also be regarded as a subset of X ( T )   when each T j   is large). In the case of gluing along a reducible critical point, the ungluing map has an additional component which reads off the U ( 1 )   gluing parameter by measuring the holonomy along a path running once through the corresponding neck in X ( T )   .
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 ζ ^   of an “extended monopole map” Ξ ^   . In the second part the inverse function theorem is applied a second time to show, essentially, that the image of ζ ^   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 4.3in 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 K ~ X   . 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 d d = 0   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 X   contains an end R + × Y j   or R + × ( Y j )   (ie an end that is being glued).
Fix non-degenerate monopoles α j   over Y j   and α j   over Y j   . (These should be smooth configurations rather than gauge equivalence classes of such.) Suppose α j   is reducible for 1 j r 0   and irreducible for r 0 < j r   , where 0 r 0 r   . We consider monopoles over X   and X ( T )   that are asymptotic to α j   over R + × Y j   and (in the case of X   ) asymptotic to α j   over R + × ( ± Y j )   .
These monopoles build moduli spaces M b = M b ( X ; α , α , α ) , M ( T ) b = M b ( X ( T ) ; α ) .   Here b X   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 b   , see [10,Subsection 3.4. The ungluing map f   will be a diffeomorphism between certain open subsets of M ( T ) b   and M b   when T ˇ : = min ( T 1 , . . . , T r )   is large.
When gluing along the critical point α j   , the stabilizer of α j   in G Y j   appears as a “gluing parameter”. This stabilizer is a copy of U ( 1 )   if α j   is reducible and trivial otherwise. When α j   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 γ j   in X ( T )   which runs once through the neck [ T j , T j ] × Y j   . To make this precise, for 1 j r 0   fix y j Y j   and smooth paths γ j ± : [ 1 , ) X   such that γ j ± ( t ) = ι j ± ( t , y j )   for t 0   . We require that these paths have different start-points. Let b   denote the collection of all the start-points γ j ± ( 1 )   . There is no loss of generality in assuming b X : 0   . Note in passing that we then have M b * = M b .   Define the smooth path γ j : I j = [ T j 1 , T j + 1 ] X ( T )   by γ j ( t ) = { γ j + ( T j + t ) , T j 1 t T j , γ j ( T j t ) , T j t T j + 1 .   If A   is a spin connection over X ( T )   then what we need is not the holonomy along γ j   of the induced connection in the determinant line bundle X ( T )   , but rather the holonomy of the induced connection in a square root of X ( T )   along γ j   . To express this more conveniently, choose a reference configuration ( A o , Φ o )   over X   with limit α j   over R + × ( ± Y j )   and α j   over R × Y j   . Let ( A o , Φ o )   denote the reference configuration over X ( T )   obtained from ( A o , Φ o )   in the obvious way when gluing the ends. Precisely speaking, ( A o , Φ o )   is the unique smooth configuration over X ( T )   which agrees with ( A o , Φ o )   over X ( T ) \ j = 1 r { 0 } × Y j   (which can also be regarded as a subset of X   ).
Note that if Z   is any smooth manifold, γ : [ 0 , 1 ] Z   a smooth path, B   a U ( 1 )   connection over Z   , and b i Ω Z 1   then Hol γ ( B + b ) = Hol γ ( B ) exp ( [ 0 , 1 ] γ * b ) ,   where Hol γ ( )   denotes holonomy along γ   . Motivated by this we define, for any spin connection A   over X ( T )   ,
Hol j ( A ) = exp ( I j γ j * ( A A o ) ) , (1)
where as usual A A o   is regarded as an imaginary valued 1   -form on X ( T )   .
Then for gauge transformations u : X ( T ) U ( 1 )   we have Hol j ( u ( A ) ) = Hol j ( A ) if u | b 1 .   In particular, there is a natural map Hol : M ( T ) b U ( 1 ) r 0 , [ A , Φ ] ( Hol 1 ( A ) , . . . , Hol r 0 ( A ) ) .   If K X   is any compact codimension 0   submanifold containing b   then we define (using the L 1 p   configuration spaces etc introduced in [10,Subsection 2.5) ˘ b ( K ) = C ( K ) / G ˘ b ( K ) ,   where G ˘ b ( K )   is the subgroup of G b ( K )   consisting of those gauge transformations whose homotopy class, identified with an element of H 1 ( K ; Z )   , is the image of a class in H 1 ( X ( T ) ; Z )   whose restriction to each Y j   is zero. Here T ˇ   should be so large that K   may be regarded as a subset of X ( T )   , and the condition is then obviously independent of T   . Note that ˘ b * ( K ) = C b * ( K ) / G ˘ b ( K )   is a Hausdorff Banach manifold. Moreover, the group G b ( K ) / G ˘ b ( K )   acts freely and properly discontinuously on ˘ b * ( K )   with quotient b * ( K )   .
It is convenient here to agree once and for all that the Sobolev exponent p > 4   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 G M b   whose closure G ¯   is compact and contains only regular points.
Definition 1 By a kv-pair we mean a pair ( K , V )   where
  • K X   is a compact codimension 0   submanifold which contains b   and intersects every component of X   ,
  • V ˘ b ( K )   is an open subset containing R K ( G ¯ )   , where R K   denotes restriction to K   .
We define a partial ordering   on the set of all kv-pairs, by decreeing that ( K , V ) ( K , V )   if and only if K K   and R K ( V ) V   .
Now fix a kv-pair ( K , V )   which satisfies the following two additional assumptions: firstly, that V ˘ b * ( K )   ; secondly, that if X e   is any component of X   which contains a point from b   then X e K   is connected. The second condition ensures that the image of R K : M b = M b * ˘ b ( K )   lies in ˘ b * ( K )   .
Suppose we are given a smooth map
q : V M b (2)
such that q ( ω | K ) = ω   for all ω G   . (Such a map always exists, for a given G   , provided K   is sufficiently large, see the appendix. In concrete applications there is often a natural choice of q   , see Sections  5 ,  6 .) Let X #   and the forms η ~ j , η ~ j   be as in [10,Subsection 1.4, and choose λ j , λ j > 0   .
Theorem 1 Suppose there is class in H 2 ( X # )   whose restrictions to Y j   and Y j   are λ j η ~ j   and λ j η ~ j   , respectively, and suppose the perturbation parameters p , p   are admissible for α   . Then there exists a kv-pair ( K ~ , V ~ ) ( K , V )   such that if ( K , V )   is any kv-pair ( K ~ , V ~ )   then the following holds when T ˇ   is sufficiently large. Set G ( T ) = { ω M ( T ) b : ω | K V , q ( ω | K ) G } .   Then G ( T )   consists only of regular monopoles (hence is a smooth manifold), and the map f : G ( T ) G × U ( 1 ) r 0 , ω ( q ( ω | K ) , Hol ( ω ) )   is a diffeomorphism.
Note that when T ˇ   is large then K X   can also be regarded as a subset of X ( T )   , in which case the expression ω | K   in the definition of G ( T )   makes sense.
Concerning admissibility of perturbation parameters, see the remarks after [10,Theorem 1.4. Note that the assumption on λ j η ~ j   and λ j η ~ j   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 f   has a smooth local right inverse around every point in G × U ( 1 ) r 0   (Proposition  1 below). In the second part we will prove that f   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 X   and X ( T )   .
Let σ j , σ j 0   be small constants and w : X R   a smooth function which is equal to σ j t   on R + × ( ± Y j )   and equal to σ j t   on R + × Y j   . As usual, we require σ j > 0   if α j   is reducible (ie for j = 1 , . . . , r 0   ), and similarly for σ j   . For j = 1 , . . . , r   choose a smooth function w j : R R   such that w j ( t ) = σ j | t |   for | t | 1   . We will always assume T ˇ 4   , in which case we can define a weight function κ : X ( T ) R   by
κ = w on X ( T ) \ j [ T j , T j ] × Y j ,
κ ( t , y ) = σ j T j + w j ( t ) for ( t , y ) [ T j , T j ] × Y j .
Let C   denote the L p , w 1   configuration space over X   defined by the reference configuration ( A o , Φ o )   , and let C   denote the L p , κ 1   configuration space over X ( T )   defined by ( A o , Φ o )   . Let G b , G b   be the corresponding groups of gauge transformations and b , b   the corresponding orbit spaces.
Now fix ( ω 0 , z ) G ¯ × U ( 1 ) r 0   . In the case when ω 0 G   , our immediate goal is to construct a smooth local right inverse of f   around this point.
However, the following set-up will also be used in the injectivity part of the proof.
Choose a smooth representative S 0 C   for ω 0   which is in temporal gauge over each end [ 1 , ) × ( ± Y j )   . (This assumption is made in order to ensure exponential decay of S 0   over these ends.) Set d = dim M b   and let π : C b   be the projection. By the local slice theorem we can find a smooth map S : R d C   such that S ( 0 ) = S 0   and such that ϖ = π S   is a diffeomorphism onto an open subset of M b   .
We will require one more property of S   , involving holonomy. If a L p , w 1 ( X ; i R )   then we define Hol j ± ( A o + a ) U ( 1 )   by Hol j ± ( A o + a ) = exp ( [ 1 , ) ( γ j ± ) * a ) .   The integral exists because, by the Sobolev embedding L 1 p C B 0   in R 4   for p > 4   , we have
e w a C e w a L 1 p = C a L p , w 1 (3)
for some constant C   . It is clear that Hol j ±   is a smooth function on C   .
Because any smooth map R d U ( 1 )   factors through exp : R i U ( 1 )   , we can arrange, after perhaps modifying S   by a smooth family of gauge transformations that are all equal to 1   outside the ends R + × Y j   and constant on [ 1 , ) × Y j   , that
Hol j + ( S ( v ) ) ( Hol j ( S ( v ) ) ) 1 = z j (4)
for j = 1 , . . . , r 0   and every v R d   . Here Hol j ± ( S ( v ) )   denotes the holonomy, as defined above, of the connection part of the configuration S ( v )   , and the z j   are the coordinates of z   .
Lemma 1 Let E , F , G   be Banach spaces, S : E F   a bounded operator and T : E G   a surjective bounded operator such that S + T : E F G , x ( S x , T x )   is Fredholm. Then T   has a bounded right inverse.
Proof. Because S + T   is Fredholm there is a bounded operator A : F G E   such that ( S + T ) A I   is compact. Set A ( x , y ) = A 1 x + A 2 y   for ( x , y ) F G   . Then T A 2 I : G G   is compact, hence T A 2   is Fredholm of index 0   . Using the surjectivity of T   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 K : G E   (with finite-dimensional image) such that T ( A 2 + K )   is an isomorphism.
Let Θ : C L p , w   we be Seiberg–Witten map over X   . By assumption, every point in G ¯   is regular, so in particular ω 0   is regular, which means that D Θ ( S 0 ) : L p , w 1 L p , w   is surjective. Let Φ   be the spinor part of S 0   and define Φ   as in [10,Subsection 2.3. Then Φ * + D Θ ( S 0 ) : L p , w 1 L p , w   is Fredholm, so by Lemma  1  D Θ ( S 0 )   has a bounded right inverse Q   . (This can also be deduced from [10,Proposition 2.2 (ii).) Let r : R R   be a smooth function such that r ( t ) = 1   for t 0   and r ( t ) = 0   for t 1   . For τ 1   set r τ ( t ) = r ( t τ )   and let S v , τ   be the configuration over X   which agrees with S ( v )   away from the ends R + × ( ± Y j )   and satisfies S v , τ = ( 1 r τ ) α ̲ j + r τ S ( v )   over R + × ( ± Y j )   . Here α ̲ j   denotes, as before, the translationary invariant monopole over R × Y j   determined by α j   . For each v   we have S v , τ S ( v ) L p , w 1 0 as τ .   Therefore, when τ   is sufficiently large, the operator D Θ ( S 0 , τ ) Q : L p , w L p , w   will be invertible, and we set Q τ = Q ( D Θ ( S 0 , τ ) Q ) 1 : L p , w L p , w 1 ,   which is then a right inverse of D Θ ( S 0 , τ )   . It is clear that the operator norm Q τ Q 0   as τ   .
For the remainder of the proof of Theorem  1 , the term `constant' will always refer to a quantity that is independent of τ , T   , unless otherwise indicated. The symbols C 1 , C 2 , . . .   and c 1 , c 2 , . . .   will each denote at most one constant, while other symbols may denote different constants in different contexts.
Consider the configuration space C j = α ̲ j + L p , w j 1   over R × Y j   and the Seiberg–Witten map Θ j : C j L p , w j .   As explained in [10,Subsection 3.4there is an identification α ̲ * + D Θ j ( α ̲ j ) = d d t + P α .   By the results of [4the operator on the right hand side defines a Fredholm operator L p , w j 1 L p , w j   , and this must be surjective because of the choice of weight function w j   . In particular, D Θ j ( α ̲ j ) : L p , w j 1 L p , w j   is surjective, hence has a bounded right inverse P j   by Lemma  1 . (Here one cannot appeal to [10,Proposition 2.2 (ii).) Let Θ : C L p , κ   be the Seiberg–Witten map over X ( T )   . When T ˇ > τ + 1   then by splicing S v , τ   in the natural way one obtains a smooth configuration S v , τ , T   over X ( T )   . There is a constant C 0 0   such that if
T ˇ > τ + C 0 (5)
then we can splice the right inverses Q τ   and P 1 , . . . , P r   to obtain a right inverse Q τ , T   of D Θ ( S 0 , τ , T ) : L p , κ 1 L p , κ   which satisfies Q τ , T C ( Q τ + j P j )   for some constant C   . The construction of Q τ , T   is analogous to the construction of the left inverse of the operator D L   in the proof of [10,Lemma 5.4. Since Q τ   is bounded in τ   (ie as a function of τ   ), we see that Q τ , T   is bounded in τ , T   .
The inequality ( 5 ) will be assumed from now on.
We now introduce certain 1   -forms that will be added to the configurations S v , τ , T   in order to make small changes to the holonomies Hol j   . For any c = ( c 1 , . . . , c r 0 ) R r 0   define the 1   -form θ c , τ   over X ( T )   by θ c , τ = { 0 outside j = 1 r 0 [ T j , T j ] × Y j , i c j r τ + 1 T j d t on [ T j , T j ] × Y j , j = 1 , . . . , r 0 .   where r s ( t ) = d d t r s ( t )   . Set E = R d × R r 0 × L p , κ ( X ( T ) ; i Λ + S ) .   For 0 < ε < 1   let B ε E   be the open ε   -ball about 0   . Define a smooth map ζ : E C   by
ζ ( v , c , ξ ) = S v , τ , T + θ c , τ + Q τ , T ξ , (6)
where θ c , τ   is added to the connection part of S v , τ , T   .
When deciding where to add the perturbation 1   -form θ   one has to balance two concerns. One the one hand, because the weight function κ   increases exponentially as one approaches the middle of the necks [ T j , T j ] × Y j   , j = 1 , . . . , r 0   , it is desirable to add θ   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 S v , τ , T   needs to be “small” in the perturbation region. We have chosen to add θ   at the negative end of the cut-off region, where the spinor field is zero.
Although we will sometimes use the notation ζ ( x )   , we shall think of ζ   as a function of three variables v , c , ξ   , and D j ζ   will denote the derivative of ζ   with respect to the j   'th variable. Similarly for other functions on (subsets of ) E   that we will define later. Set σ = max ( 0 , σ 1 , . . . , σ r ) .   Notice that if r 0 = 0   , ie if we are not gluing along any reducible critical point, then we may take σ = 0   .
Lemma 2 There exists a constant C 1 > 0   such that for x E   the following hold:
  • (i) D 1 ζ ( x ) , D 2 ζ ( x ) < C 1   if x < 1   ,
  • (ii) D 2 ζ ( x ) < C 1 e σ τ   ,
  • (iii) D 3 ζ ( x ) < C 1   .
Proof. To prove (ii), note that if r 0 > 0   and c = ( c 1 , . . . , c r 0 )   then ζ ( v , c , ξ ) c j L p , κ 1 = const e σ j τ .   The other two statements are left to the reader.
Let C 1   be the set of all S C   such that [ S | K ] V   , q ( S | K ) ϖ ( R d )   , and Hol j ( S ) z j   for j = 1 , . . . , r 0   . Then C 1   is an open subset of C   , and there are unique smooth functions η j : C 1 ( π , π )   such that Hol j ( S ) = z j exp ( i η j ( S ) )   . Set η = ( η 1 , . . . , η r 0 )   and define Ξ ^ = ( ϖ 1 q R K , η , Θ ) : C 1 E .   A crucial point in the proof of Theorem  1 will be the construction of a smooth local right inverse of Ξ ^   , defined in a neighbourhood of 0   . The map ζ   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 τ   so large that K X : τ   and Hol j + ( S 0 , τ ) ( Hol j ( S 0 , τ ) ) 1 z j   for j = 1 , . . . , r 0   . Note that the left hand side of this equation is equal to Hol j ( S 0 , τ , T )   whenever T j > τ + 1   . There is then a constant ε > 0   such that ζ ( B ε ) C 1   , in which case we have a composite map Ξ = Ξ ^ ζ : B ε E .   Choose λ > 0   so that none of the operators H ~ α j   ( j = 1 , . . . , r   ) and H ~ α j   ( j = 1 , . . . , r   ) has any eigenvalue of absolute value λ   . (The notation H ~ α   was introduced in [10,Subsection 6.1.) Recall that we assume the σ j   are small and non-negative, so in particular we may assume 6 σ < λ   .
Lemma 3 There is a constant C 2 <   such that Ξ ( 0 ) C 2 e ( σ λ ) τ .  
Proof. The first two components of Ξ ( 0 ) = Ξ ^ ( S 0 , τ , T )   are in fact zero:
the first one because S 0 , τ , T = S 0   over K   , the second one because the d t   -component of S 0 , τ S 0 = ( 1 r τ ) ( α ̲ j S 0 )   vanishes on [ 1 , ) × ( ± Y j )   since S 0   and α ̲ j   are both in temporal gauge there.
The third component of Ξ ( 0 )   is Θ ( S 0 , τ , T )   . It suffices to consider τ   so large that the p   -perturbations do not contribute to Θ ( S 0 , τ , T )   , which then vanishes outside the two bands of length 1   in [ T j , T j ] × Y j   centred at t = ± ( T j τ 1 2 )   , j = 1 , . . . , r   . Our exponential decay result from [10says that for every k 0   there is a constant C k   such that for every ( t , y ) R + × ( ± Y j )   we have | k ( S 0 α ̲ j ) | ( t , y ) C k e λ t .   Consequently, Θ ( S 0 , τ , T ) const ( e λ τ + e 2 λ τ ) const e λ τ .   This yields Ξ ( 0 ) = Θ ( S 0 , τ , T ) L p , κ const e ( σ λ ) τ .  
Lemma 4 There is a constant C 3 <   such that for sufficiently large τ   the following hold:
  • (i) D Ξ ( 0 ) C 3   ,
  • (ii) D Ξ ( 0 )   is invertible and D Ξ ( 0 ) 1 C 3   .
Proof. By construction, the derivative of Ξ   at 0   has the form D Ξ ( 0 ) = ( I 0 β 1 δ 2 I β 2 δ 3 0 I ) ,   where the k   'th column is the k   'th partial derivative and I   the identity map.
The middle top entry in the above matrix is zero because θ c , τ   vanishes on K   . The middle bottom entry is zero because S 0 , τ , T = α ̲ j   on the support of θ c , τ   and the spinor field of α ̲ j   is zero (for j = 1 , . . . , r 0   ). Adding θ c , τ   to S 0 , τ , T   therefore has the effect of altering the latter by a gauge transformation over [ T j + τ + 1 , T j + τ + 2 ] × Y j   , j = 1 , . . . , r 0   .
We claim that β k   is bounded in τ , T   for k = 1 , 2   . For k = 1   this is obvious from the boundedness of Q τ , T   . For k = 2   note that the derivative of η j : C 1 ( π , π )   at any S C 1   is
D η j ( S ) ( a , φ ) = i I j γ j * a (7)
where a   is an imaginary valued 1   -form and φ   a positive spinor. Because of the weights used in the Sobolev norms it follows that D η ( S )   is (independent of S   and) bounded in τ , T   (see ( 3 )). This together with the bound on Q τ , T   gives the desired bound on β 2   .
Note that, for k = 2 , 3   , δ k   is independent of T   when τ 0   , and routine calculations show that δ k 0   as τ   . (In the case of δ 2   this depends on the normalization ( 4 ) of the holonomy of S ( v )   .) Write D Ξ ( 0 ) = x y   , where x = ( I 0 β 1 0 I β 2 0 0 I ) , x 1 = ( I 0 β 1 0 I β 2 0 0 I ) .   When τ   is so large that y x 1 < 1   then of course y x 1 y x 1 < 1   , hence x y = ( I y x 1 ) x   is invertible. Moreover, ( x y ) 1 x 1 = x 1 [ ( I y x 1 ) 1 I ] = x 1 k = 1 ( y x 1 ) k ,   which gives ( x y ) 1 x 1 x 1 2 y 1 x 1 y 0 as τ .   We now record some basic facts that will be used in the proof of Lemma  6 below.
Lemma 5 If E 1 , E 2 , E 3   are Banach spaces, U j E j   an open set for j = 1 , 2   , and f : U 1 U 2   , g : U 2 E 3   smooth maps then the second derivate of the composite map g f : U 1 E 3   is given by
D 2 ( g f ) ( x ) ( y , z ) = D 2 g ( f ( x ) ) ( D f ( x ) y , D f ( x ) z )
+ D g ( f ( x ) ) ( D 2 f ( x ) ( y , z ) )
for x U 1   and y , z E 1   .
Proof. Elementary. It is also worth noting that embedding and multiplication theorems for L k q   Sobolev spaces on R 4   ( k 0   , 1 q <   ) carry over to X ( T )   , and that the embedding and multiplication constants are bounded functions of T   .
Furthermore, a differential operator of degree d   over X ( T )   which is translationary invariant over necks and ends induces a bounded operator L k + d q L k q   whose operator norm is a bounded function of T   .
Lemma 6 There is a constant C 4 > 0   such that D 2 Ξ ( x ) C 4   whenever x C 14   and τ C 4   .
Proof. We will say a quantity depending on x , τ   is s-bounded if the lemma holds with this quantity in place of D 2 Ξ   .
Let Ξ 1 , Ξ 2 , Ξ 3   be the components of Ξ   .
The assumption K X : τ   ensures that Ξ 1 ( v , c , ξ )   is independent of c   . It then follows from Lemma  5 and the bound on Q τ , T   that D 2 Ξ 1   is s-bounded.
When v , c , ξ   are small we have Ξ 2 ( v , c , ξ ) = η ( S v , τ , T + θ c , τ + Q τ , T ξ ) = c + η ( S v , τ , T + Q τ , T ξ ) .   Since D η   is constant, as noted above, we have D 2 η = 0   . From the bounds on D η   and Q τ , T   we then deduce that D 2 Ξ 2   is s-bounded.
To estimate Ξ 3   , we fix h 0   and consider only τ h   . It is easy to see that Ξ 3 ( x ) | X : h   is s-bounded. By restricting to small x   and choosing h   large we may arrange that the p   and o   -perturbations do not contribute to Ξ 3 , j : = Ξ 3 | [ T j + h , T j h ] × Y j   for j = 1 , . . . , r   . We need to show that each D 2 Ξ 3 , j   is s-bounded, but to simplify notation we will instead prove the same for D 2 Ξ 3   under the assumption that the p , o   perturbations are zero.
First observe that for any configuration ( A , Φ )   over X ( T )   and any closed, imaginary valued 1   -form a   we have Θ ( A + a , Φ ) = Θ ( A , Φ ) + ( 0 , a Φ ) .   Moreover, a Φ L p , κ = C a e κ Φ p C a 2 p e κ Φ 2 p C a 2 p Φ L p , κ 1   for some constants C , C <   . Taking ( A , Φ ) = S v , τ , T + Q τ , T ξ   and a = θ c , τ   we see that D k D 2 Ξ 3 ( x )   is s-bounded for k = 1 , 2 , 3   .
Next note that the derivative of the Seiberg–Witten map Θ : C L p , κ   at a point S o + s 1   has the form D Θ ( S o + s 1 ) s 2 = L s 2 + B ( s 1 , s 2 )   where B   is a pointwise bilinear operator, and L   a first order operator which is independent of s 1   and translationary invariant over necks and ends. This yields
D Θ ( S o + s ) const ( 1 + s 2 p ) . (8)
Moreover, D 2 Θ ( S ) = B   for all S   , hence there is a constant C <   such that D 2 Θ ( S ) C   for all T   .
Combining the above results on Θ   with Lemma  5 we see that D j D k Ξ 3   is s-bounded also when j , k 2   .
Lemma 7 There exist c 5 > 0   and C 6 <   such that if 0 < ε < c 5 ε < c 5 2   then for sufficiently large τ   the following hold:
  • (i) Ξ : B ε E   is injective,
  • (ii) There is a (unique) smooth map Ξ 1 : B ε B ε   such that Ξ Ξ 1 = I   ,
  • (iii) D ( Ξ 1 ) ( x ) C 6   for all x B ε   ,
  • (iv) D 2 ( Ξ 1 ) ( x ) C 6   for all x B ε   ,
  • (v) Ξ 1 ( 0 ) C 6 e ( σ λ ) τ   .
Proof. For sufficiently large τ   we have ε + Ξ ( 0 ) < c 5 ε   by Lemma  3 . Statements (i)–(iv) now follow from the inverse function theorem, [10,Proposition B.1, applied to the function x Ξ ( x ) Ξ ( 0 )   , together with Lemmas  4 and  6 . To prove (v), set h = Ξ 1   , x = Ξ ( 0 )   and take τ   so large that x B ε   . Since Ξ   is injective on B ε   we must have h ( x ) = 0   , so h ( 0 ) = h ( x ) h ( 0 ) x sup y ε D h ( y ) .   Now (v) follows from (iii) and Lemma  3 .
From now on we assume that ε , ε , τ   are chosen so that the conclusions of the lemma are satisfied. Define ζ ^ = ζ Ξ 1 : B ε C 1 .   Then clearly Ξ ^ ζ ^ = I .   Thus, ( v , c ) ζ ^ ( v , c , 0 )   is a “gluing map”, ie it solves the problem of gluing the monopoles S ( v )   over X   (for v   small) to get a monopole over X ( T )   with prescribed holonomy z j e i c j   along the path γ j   for j = 1 , . . . , r 0   .
Lemma 8 There is a constant C 7 <   such that for x B ε   one has D ζ ^ ( x ) , D 2 ζ ^ ( x ) C 7 e σ τ .  
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 ( K , V )   is any kv-pair ( K , V )   then the following holds when T ˇ   is sufficiently large. Define M ( T )   and f   as in Theorem  1 . Then G × U ( 1 ) r 0   is a union of finitely many open sets U   such that for each U   there exists a smooth map h : U G ( T )   whose image consists only of regular points, and which satisfies f h = I   .
Here we do not need any assumptions on η ~ j , η ~ j   or on p j , p j   .
Proof. Let ( ω 0 , z ) G ¯ × U ( 1 ) r 0   and consider the set-up above, with τ   so large that K X : τ   and ε   so small that
ζ ( x ) | K V for every x B ε . (9)
Note that taking ε   small may require taking τ   (and hence T ˇ   ) large, see Lemma  7 . For any sufficiently small open neighbourhood W M b × U ( 1 ) r 0   of ( ω 0 , z )   we can define a smooth map ν : W C   by the formula ν ( ω , a ) = ζ ^ ( ϖ 1 ( ω ) , i log a z , 0 ) .   Here log e b = b   for any complex number b   with | Im b | < π   , and i log a z R r 0   denotes the vector whose j   'th component is i log a j z j   . Because Ξ ^ ζ ^ = I   and the Seiberg–Witten map is the third component of Ξ ^   , the image of ν   consists of regular monopoles. Set U = W ( G × U ( 1 ) r 0 )   and h = π ν | U   , where π : C b   is the projection. Unravelling the definitions involved and using ( 9 ) one finds that h   maps into G ( T )   , and that f h = I   .
How large T ˇ   must be for this to work might depend on ( ω 0 , z )   . But G ¯ × U ( 1 ) r 0   is compact, hence it can be covered by finitely many such open sets W   . If T ˇ   is sufficiently large then the above construction will work for each of these W   .

4 Injectivity

We now continue the discussion that was interrupted by Proposition  1 .
Set S ~ = S 0 , τ , T , S ^ = ζ ^ ( 0 ) .  
Lemma 9 There is a constant C 8 <   such that for sufficiently large τ   one has S ^ S ~ L p , κ 1 C 8 e ( 2 σ λ ) τ , S ^ S o L p , κ 1 C 8 .  
Proof. Set Ξ 1 ( 0 ) = ( v , c , ξ ) B ε   . For sufficiently large τ   we have
S ^ S ~ L p , κ 1 S v , τ , T S 0 , τ , T L p , κ 1 + θ c , τ + Q τ , T ξ L p , κ 1
const ( v | | + e σ τ c + ξ )
const e ( 2 σ λ ) τ ,
where we used Lemma  7 (v) to obtain the last inequality. Because S ~ S o L p , κ 1   is bounded in τ , T   , and we assume 3 σ < λ   , the second inequality of the lemma follows as well.
For positive spinors Φ   on X ( T )   it is convenient to extend the definition of Φ   to complex valued functions on X ( T )   : Φ f = ( d f , f Φ ) .   (However, Φ *   will always refer to the formal adjoint of Φ   acting on imaginary valued functions.) When Φ   is the spinor part of S o , S ~ , S ^   then the corresponding operators Φ   will be denoted o , ~ ,   , respectively. (We omit the ^ on   to simplify notation.) As in [10,Subsection 2.2we define = { f L 2 , loc p ( X ( T ) ; C ) : o f L p , κ 1 } .   We can take the norm to be f = o f L p , κ 1 + x b | f ( x ) | .  
Lemma 10 There is a constant C 9 <   such that if I   is any of the operators o , ~ ,   then for all f   one has f C 9 ( I f L p , κ + x b | f ( x ) | ) .  
Proof. We first prove the inequality for I = ~   (the case of o   is similar, or easier). If X e   is any component of X   and 0 τ ¯ τ   then for some constant C τ ¯ <   one has f const f L 1 p C τ ¯ ( ~ f p + x b X e | f ( x ) | )   for all L 1 p   functions f : ( X e ) : τ ¯ C   . Here the Sobolev inequality holds because p > 4   , whereas the second inequality follows from [10,Lemma 2.1. We use part (i) of that lemma if the spinor field of S 0   is not identically zero on X e   , and part (ii) otherwise. (In the latter case b X e   is non-empty.) When τ ¯ , τ   are sufficiently large we can apply part (i) of the same lemma in a similar fashion to the band [ t , t + 1 ] × Y j   provided t τ ¯   and α j   is irreducible, and to the band [ t 1 , t + 1 ] × Y j   provided | t | T j τ ¯ 1   and α j   is irreducible. To estimate | f |   over these bands when α j   resp. α j   is reducible one can use [10,Lemma 2.2(ii). This proves the lemma for I = ~   (and for I = o   ). We now turn to the case I =   . Let φ   denote the spinor part of S ^ S ~   . Then
~ f L p , κ f L p , κ + f φ L p , κ
f L p , κ + const ( ~ f L p , κ + x b | f ( x ) | ) φ L p , κ .
By Lemma  9 we have φ L p , κ 1 0   as τ 0   , so for sufficiently large τ   we get ~ f L p , κ const ( f L p , κ + x b | f ( x ) | ) .   Therefore, the lemma holds with I =   as well.
Lemma 11 There is a constant C 10 <   such that for all f , g   and φ L p , κ 1 ( X ( T ) ; S + )   one has
  • (i) f g C 10 f g   ,
  • (ii) f φ C 10 f φ   ,
where we use the L p , κ 1   norm on spinors and the   norm on elements of   .
Proof. By routine calculation using Lemma  10 with I = o   one easily proves (ii) and the inequality d ( f g ) L p , κ 1 const f g .   Now observe that by definition g Φ o L p , κ 1   , where as before Φ o   denotes the spinor field of the reference configuration S o   . Applying (ii) we then obtain f g Φ o L p , κ 1 const f g Φ o L p , κ 1 const f g ,   completing the proof of (i).
Recall from [10,Subsection 2.4that the Lie algebra L G b   is the space of imaginary valued functions in   that vanish on b   .
Lemma 12 There is a constant C 11 > 0   such that for τ > C 11   and all f L G b   one has C 111 o f L p , κ 1 f L p , κ 1 C 11 o f L p , κ 1 .  
Proof. Let ψ   denote the spinor part of S ^ S o   . Then
f ψ L p , κ 1 const ( f ψ L p , κ 1 + d f L 2 p , κ ψ 2 p )
const f L p , κ 1 ψ L p , κ 1 ,
and similarly with o   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
Π : L G b × B ε C 1 ,
( f , x ) exp ( f ) ( ζ ^ ( x ) ) .
contains a “not too small” neighbourhood of S ^   . The derivative of Π   at ( 0 , 0 )   is
D Π ( 0 , 0 ) : L G b E L p , κ 1 ,
( f , x ) f + D ζ ^ ( 0 ) x .
To be concrete, let L G b E   have the norm ( f , x ) = f + x E   .
Lemma 13 D Π ( 0 , 0 )   is a linear homeomorphism.
Proof. By [10,Proposition 2.2, * : L G b L p , κ   is a Fredholm operator with the same kernel as   . Now,   is injective on L G b   , because ζ ^   maps into C 1   and therefore [ S ^ | K ] V b *   . Since W = * ( L G b )   is a closed subspace of L p , κ   of finite codimension, we can choose a bounded operator π : L p , κ W   such that π | W = I   . Set # = π * : L p , κ 1 W .   Then # : L G b W   is an isomorphism. Furthermore, index ( # + D Θ ( S ^ ) ) = dim M ( T ) b = dim M b + r 0 ,   where `dim' refers to expected dimension (which in the case of M b   is equal to the actual dimension of G   ), and the second equality follows from the addition formula for the index (see [4). Consequently, index ( # + D Ξ ^ ( S ^ ) ) = 0 .   We now compute
( # + D Ξ ^ ( S ^ ) ) D Π ( 0 , 0 ) = ( # B 0 I ) : L G b E W E , (10)
where B : E W   . The zero in the matrix above is due to the fact that D Ξ ^ ( S ^ ) f = d d t | 0 Ξ ^ ( e t f ( S ^ ) ) = 0 ,   which holds because Ξ ^ 1 , Ξ ^ 2   are G b   invariant, Ξ ^ 3   is G b   equivariant, and Ξ ^ ( S ^ ) = 0   .
Since the right hand side of ( 10 ) is invertible, it follows that # + D Ξ ^ ( S ^ )   is a surjective Fredholm operator of index 0   , hence invertible. Of course, this implies that D Π ( 0 , 0 )   is also invertible.
Lemma 14 There is a constant C 12 <   such that for sufficiently large τ   , D Π ( 0 , 0 ) 1 C 12 e σ τ .  
Proof. In this proof all unqualified norms are L p , κ 1   norms. It follows from ( 7 ), ( 8 ) and Lemma  9 that D Ξ ^ ( S ^ )   is bounded in τ , T   . Therefore there exists a constant C <   such that x E = D Ξ ^ ( S ^ ) ( f + D ζ ^ ( 0 ) x ) C D Π ( 0 , 0 ) ( f , x )   for all f L G b   and x E   . From Lemma  12 and Lemma  8 we get
C 111 o f f
D Π ( 0 , 0 ) ( f , x ) + D ζ ^ ( 0 ) x
D Π ( 0 , 0 ) ( f , x ) + C 7 e σ τ x E
( 1 + C C 7 e σ τ ) D Π ( 0 , 0 ) ( f , x ) .
This yields f + x E const e σ τ D Π ( 0 , 0 ) ( f , x ) .  
Lemma 15 There is a constant C 13 <   such that for sufficiently large τ   one has D 2 Π ( f , x ) C 13 e σ τ   for all f L G b   , x E   such that f < 1   and x < ε   .
Proof. For the purposes of this proof it is convenient to rescale the norm on   so that we can take C 10 = 1   in Lemma  11 .
If f , g   then e f g   , and from Lemma  11 we obtain e f g n = 0 1 n ! f n g e f g ,   and similarly with a spinor φ L p , κ 1   instead of g   .
The first two derivatives of exp : 1 +   are D exp ( f ) g = g exp ( f ) , D 2 exp ( f ) ( g , h ) = g h exp ( f ) ,   so D exp ( f ) , D 2 exp ( f ) exp ( f ) .   Let ζ ^ 1 , ζ ^ 2   be the connection and spinor parts of ζ   , respectively, and define Π 1 , Π 2   similarly. Then Π ( f , x ) = ( ζ ^ 1 ( x ) d f , e f ζ ^ 2 ( x ) ) .   We regard Π ( f , x )   as a function of the the two variables f , x   . Let D j Π   denote the derivative of Π   with respect to the j   'th variable. Similarly for the second derivatives D j D k Π   .
Applying Lemmas  8 and  11 we now find that
D 2 2 Π 1 ( f , x ) = D 2 ζ ^ 1 ( x ) const e σ τ , D j D k Π 2 ( f , x ) const e σ τ , j , k = 1 , 2  
for f < 1   and x < ε   . Since D j D 1 Π 1 = 0   for j = 1 , 2   , the lemma is proved.
In the following, B ( x ; r )   will denote the open r   -ball about x   (both in various Banach spaces and in C   ).
Lemma 16 There exist constants r 1 , r 2 > 0   such that for sufficiently large τ   the image of Π   contains the ball of radius r 2 e 3 σ τ   about S ^   in C   ; more precisely one has B ( S ^ ; r 2 e 3 σ τ ) Π ( B ( 0 ; r 1 e 2 σ τ ) ) .  
Proof. We wish to apply the inverse function theorem [10,Proposition B.1to the map Π   restricted to a ball B ( 0 ; R 1 )   , where R 1 ( 0 , ε ]   is to be chosen. For the time being let M , L , κ   have the same meaning as in that proposition. By Lemma  15 we can take M = C 13 e σ τ   , and by Lemma  14 we have L 1 C 12 e σ τ   . We need 0 κ = L 1 1 R 1 M .   This will hold if R 1 ( C 12 C 13 ) 1 e 2 σ τ .   When τ   is large we can take R 1   to be the right hand side of this inequality.
By [10,Proposition B.1, Π ( B ( 0 ; R 1 ) )   contains the ball B ( S ^ ; R 2 )   where R 2 = 1 2 R 1 C 112 e σ τ = 1 2 C 2 12 C 113 e 3 σ τ .   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 ( K , V )   has the effect of restricting the map f   to a perhaps smaller domain. It therefore suffices to prove that for some kv-pair ( K , V ) ( K , V )   the map f   is injective for all sufficiently large T ˇ   .
For any natural number m   which is so large that K X : m   , let V m   be the set of all ω ˘ b ( X : m )   such that there exist a representative S   of ω   , and a configuration S ¯ = ( A ¯ , Φ ¯ )   over X   representing an element of G ¯   , such that d m ( S , S ¯ ) : = X : m | S ¯ S | p + | A ¯ ( S ¯ S ) | p < 1 m .   Note that d m ( u ( S ) , u ( S ¯ ) ) = d m ( S , S ¯ )   for any gauge transformation u   over X : m   .
Lemma 17 Let ω n V m n   for n = 1 , 2 , . . .   , where m n   . Then there exists for each n   a representative S n   of ω n   such that a subsequence of S n   converges locally in L 1 p   over X   to a smooth configuration representing an element of G ¯   .
Proof. By assumption there exist for each n   a representative S n   of ω n   and a configuration S ¯ n   over X   representing an element of G ¯   such that
d m n ( S n , S ¯ n ) < 1 m n . (11)
After passing to a subsequence we may assume (since G ¯   is compact) that [ S ¯ n ]   converges in G ¯   to some element [ S ¯ ]   , and we can choose S ¯   smooth. Since M b = M b *   , the local slice theorem guarantees that for large n   we can find u n G b   such that S ¯ n = u n ( S ¯ n )   satisfies S ¯ n S ¯ L p , w 1 0 .   Set S n = u n ( S n )   , which is again a representative of ω n   . Let A ¯ , A ¯ n   be the connection parts of S ¯ , S ¯ n   , respectively. Then ( 11 ) implies that S ¯ n S n 0   and A ¯ n ( S ¯ n S n ) 0   locally in L p   over X   , hence also S n S ¯   locally in L p   over X   . Now A ¯ ( S n S ¯ ) = A ¯ n ( S n S ¯ n ) + A ¯ ( S ¯ n S ¯ ) + ( A ¯ A ¯ n ) ( S n S ¯ n ) ,   and each of the three terms on the right hand side converges to 0   locally in L p   over X   (the third term because of the continuous multiplication L 1 p × L p L p   in R 4   for p > 4   ). Hence S n S ¯   locally in L 1 p   over X   .
Corollary 1 For sufficiently large n   one has that R K ( V n ) V   .
Lemma 18 Let ω n V m n   for n = 1 , 2 , . . .   , where m n   . Suppose q ( ω n | K )   converges in M b   to an element g   as n   . Then g G ¯   , and there exists for each n   a representative S n   of ω n   such that the sequence S n   converges locally in L 1 p   over X   to a smooth configuration representing g   .
Proof. Let S n , S ¯ n   be as in the proof of Lemma  17 . First suppose that [ S ¯ n ]   converges in G ¯   to some element [ S ¯ ]   , where S ¯   is smooth. Choosing S n , S ¯ n   as in that proof we find again that S n S ¯   locally in L 1 p   over X   , hence g = lim n q ( S n | K ) = q ( S ¯ | K ) = [ S ¯ ] .   We now turn to the general case when [ S ¯ n ]   is not assumed to converge.
Because G ¯   is compact, every subsequence of [ S ¯ n ]   has a convergent subsequence whose limit must be g   by the above argument. Hence [ S ¯ n ] g   .
Suppose we are given a sequence { m n } n = 1 , 2 , . . .   of natural numbers tending to infinity, and for each n   an r   -tuple T ( n )   of real numbers such that T ˇ ( n ) : = min j T j ( n ) > m n .   Define G ( T ( n ) )   and f n   as in Theorem  1 , with K = X : m n   and V = V m n   .
Lemma 19 For n = 1 , 2 , . . .   suppose S n   is a smooth configuration over X ( T ( n ) )   representing an element ω n G ( T ( n ) )   , and such that f n ( ω n ) ( ω 0 , z ) G ¯ × U ( 1 ) r 0   as n   . There exists a constant C 14 <   such that for sufficiently large τ   the following holds for sufficiently large n   . Let the map ζ ^ = ζ ^ n   be defined as above and set S ^ n = ζ ^ n ( 0 )   . Then there exists a smooth gauge transformation u n G b   such that u n ( S n ) S ^ n L p , κ 1 C 14 e ( 3 σ λ ) τ .  
Note: This constant C 14   depends on ( ω 0 , z )   but not on the sequence S n   .
Before proving the lemma, we will use it to show that f n   is injective for some n   . This will prove Proposition  2 . Suppose ω n , ω n G ( T ( n ) )   and f n ( ω n ) = f n ( ω n )   , n = 1 , 2 , . . .   . After passing to a subsequence we may assume that f n ( ω n )   converges to some point ( ω 0 , z ) G ¯ × U ( 1 ) r 0   .
Combining Lemmas  16 ,  19 and the assumption 6 σ < λ   we conclude that if τ   is sufficiently large then for sufficiently large n   we can represent ω n   and ω n   by configurations ζ ^ ( x n )   and ζ ^ ( x n )   , respectively, where x n , x n B ε   .
Now recall that Ξ ^ ζ ^ = I   , and that the components Ξ ^ 1 , Ξ ^ 2   are G b   invariant whereas Ξ ^ 3   is the Seiberg–Witten map. Comparing the definitions of f n   and Ξ ^   we conclude that x n = Ξ ^ ( ζ ^ ( x n ) = Ξ ^ ( ζ ^ ( x n ) = x n ,   hence ω n = ω n   for large n   . 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 S n   (as well as of τ   as before).
By Lemma  18 we can find for each n   an L 2 , loc p   gauge transformation v n   over X   with v n | b = 1   such that S n = v n ( S n )   converges locally in L 1 p   over X   to a smooth configuration S   representing ω 0   . A moment's thought shows that we can choose the v n   smooth, and we can clearly arrange that S = S 0   .
Then for any t 0   we have
limsup n S n S ^ n L p , κ 1 ( X : t ) = limsup n S 0 S ^ n L p , κ 1 ( X : t ) const e ( 2 σ λ ) τ (12)
when τ   is so large that Lemma  9 applies.
For t 0   and any smooth configurations S   over X : t   consider the functional
E ( S , t ) = j = 1 r λ j ( ϑ ( S | { t } × Y j ) + ϑ ( S | { t } × ( Y j ) ) ) + j = 1 r λ j ( ϑ ( S | { t } × Y j ) ϑ ( α j ) ) ,  
where in this formula { t } × ( ± Y j )   has the boundary orientation inherited from X : t   . (Recall that the Chern-Simons-Dirac functional ϑ   changes sign when the orientation of the 3   -manifold in question is reversed.) The assumption on λ j , λ j   and η ~ j , η ~ j   in Theorem  1 implies that E ( S , t )   depends only on the gauge equivalence class of S   . Since ϑ   is a smooth function on the L 1 / 2 2   configuration space by [10,Lemma 3.1, we obtain E ( S n , t ) = E ( S n , t ) E ( S 0 , t )   as n   . By our exponential decay results (see the proof of [10,Theorem 6.1), E ( S 0 , t ) < const e 2 λ t for t 0 .   It follows that E ( S n , t ) < const e 2 λ t for n > N ( t )   for some positive function N   . By assumption the perturbation parameters p , p   are admissible, hence there is a constant C <   such that when T ˇ ( m n ) > C   , each of the ( r + r )   summands appearing in the definition of E ( S n , t )   is non-negative. Explicitly, this yields
0 ϑ ( S n | { T j ( n ) + t } × Y j ) ϑ ( S n | { T j ( n ) t } × Y j ) < const e 2 λ t , 0 ϑ ( S n | { t } × Y j ) ϑ ( α j ) < const e 2 λ t ,  
where the first line holds for 0 t T j ( n )   and j = 1 , . . . , r   , the second line for t 0   and j = 1 , . . . , r   , and in both cases we assume T ˇ ( n ) > C   and n > N ( t )   .
In the following we will ignore the ends R + × Y j   of X   , ie we will pretend that X #   is compact. If α j   is irreducible then the argument for dealing with the end R + × Y j   is completely analogous to the one given below for a neck [ T j , T j ] × Y j   , while if α j   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 [ T j ( n ) , T j ( n ) ] × Y j   where 1 j r   . To simplify notation we will therefore mostly omit j   from notation and write T ( n ) , Y , α   etc instead of T j ( n ) , Y j , α j   .
For 0 t T ( n )   set B t = [ T ( n ) + t , T ( n ) t ] × Y ,   regarded as a subset of X ( T ( n ) )   . By the above discussion there is a constant t 1 > 0   such that when n   is sufficiently large, S n   will restrict to a genuine monopole over the band B t 1 + 3   by [10,Lemmas 4.1,4.2,4.3and will have small enough energy over this band for [10,Theorem 6.2to apply. That theorem then provides a smooth v ~ n : B t 1 U ( 1 )   such that S n = v ~ n ( S n | B t 1 )   is in temporal gauge and S n α ̲ L p , κ 1 ( B t ) const e ( σ λ ) t , t t 1 .   Writing S n S ^ n = ( S n α ̲ ) + ( α ̲ S ~ ) + ( S ~ S ^ n )   we get
limsup n S n S ^ n L p , κ 1 ( B t ) const ( e ( 2 σ λ ) τ + e ( σ λ ) t ) (13)
when t t 1   and τ   is so large that Lemma  9 applies.
To complete the proof of the lemma we interpolate between v n   and v ~ n   in the overlap region O τ = X : τ B τ 1   . (This requires τ t 1 + 1   .) The choice of this overlap region is somewhat arbitrary but simplifies the exposition.
Define w n = v ~ n v 1 n : O τ U ( 1 ) .   Then w n ( S n ) = S n on O τ .   Set x ± = γ ( ± ( T τ ) )   , where γ = γ j   is the path introduced in the beginning of this section. If α   is reducible then by multiplying each v ~ n   by a constant and redefining w n , S n   accordingly we can arrange that w n ( x + ) = 1   for all n   .
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 C 15 <   such that if τ C 15   then limsup n w n 1 L 2 p ( O τ ) C 15 e ( 2 σ λ ) τ .  
Proof of sublemma: If α   is irreducible then the sublemma follows from inequalities ( 12 ), ( 13 ) and [10,Lemmas 6.9,6.11. (In this case the sublemma holds with C 15 e ( σ λ ) τ   as upper bound.) Now suppose α   is reducible. We will show that
limsup n | w n ( x ) 1 | const e ( 2 σ λ ) τ (14)
for large τ   . Granted this, we can prove the sublemma by applying [10,Lemma 6.9and [10,Lemma 6.10(ii)to each component of O τ   .
In the remainder of the proof of the sublemma we will omit n   from subscripts. To prove ( 14 ), define intervals J 0 = [ T 1 , T + τ ] , J 1 = [ T + τ , T τ ] , J 2 = [ T τ , T + 1 ]   and for k = 0 , 1 , 2   set γ ( k ) = γ | J k   . Let Hol ( k )   denote holonomy along γ ( k )   in the same sense as ( 1 ), ie Hol ( k )   is the result of replacing the domain of integration I j   in that formula with J k   . Define δ ( k ) C   by
Hol ( k ) ( S ^ ) = Hol ( k ) ( S ) ( 1 + δ ( k ) ) , k = 0 , 2 ,
Hol ( 1 ) ( S ^ ) = Hol ( 1 ) ( S ) ( 1 + δ ( 1 ) )
where as usual we mean holonomy with respect to the connection parts of the configurations. For large τ   the estimates ( 12 ) and ( 13 ) give | δ ( k ) | const e ( 2 σ λ ) τ   when n   is sufficiently large.
Writing b = k = 0 2 ( 1 + δ ( k ) )   we have z = Hol ( S ^ ) = 2 k = 0 Hol ( k ) ( S ^ ) = b Hol ( 0 ) ( S ) Hol ( 1 ) ( S ) Hol ( 2 ) ( S ) .   Now, by the definition of holonomy, Hol ( 1 ) ( S ) = v ~ ( x + ) v ~ ( x ) Hol ( 1 ) ( S ) ,   and there are similar formulas for Hol ( k ) ( S )   . Because w ( x + ) = 1   we obtain z = b Hol ( S ) w ( x ) 1 .   Setting a = Hol ( S ) z 1   we get w ( x ) 1 = a b 1 = ( a 1 ) b + b 1 .   Since by assumption a 1   as n   , we have | w ( x ) 1 | const ( | a 1 | + k | δ ( k ) | ) const e ( 2 σ λ ) τ   for large n   , 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 d d = 0   for the standard Floer differential.
Recall that the Seiberg–Witten invariant of a closed spin c   4   –manifold (with b 2 + > 1   ) 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 E ˘ * ( K )   with section s   , where K X   . In the proof of Theorem  2 below we will see how the section s   gives rise in a natural way to a map q   as in Theorem  1 . Thus, the section s   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 X   be as in [10,Subsection 1.4with r = 1   and r 0   , and set Y = Y 1   . In other words, we will be gluing one single pair of ends R + × ( ± Y )   of X   , but X   may have other ends R + × Y j   not involved in the gluing. We assume X #   is connected, which means that X   has one or two connected components.
For j = 1 , . . . , r   fix a critical point α j ~ Y j   . Let μ   be a 2   -form and p   a perturbation parameter for Y   , and let μ j , p j   be similar data for Y j   . Let each p , p j   have small C 1   norm. To simplify notation we write, for α , β ~ Y   , M α , β = M ( X ; α , β , α ) , M ( T ) = M ( X ( T ) ; α ) .   We make the following assumptions:
  • (Compactness) At least one of the conditions (B1), (B2) of [10holds,
  • (Regularity) All moduli spaces over R × Y   , R × Y j   , and X   contain only regular points, and
  • (No reducibles) Given α 1 , α 2 ~ Y   and α j ~ Y j   , if there exist a broken gradient line over R × Y   from α 1   to α 2   and for each j   a broken gradient line over R × Y j   from α j   to β j   then M ( X ; α 1 , α 2 , α )   contains no reducible. (It then follows by compactness that M ( T )   contains no reducible when T   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 Y   and each Y j   are rational homology spheres; in the remaining cases the discussion here is therefore somewhat theoretical at this time.
Let K X   be a compact codimension 0   submanifold which intersects every component of X   . When T 0   then K   may also be regarded as a submanifold of X ( T )   , and we have restriction maps R α , β : M α , β * ˘ * ( K ) , R : M ( T ) ˘ * ( K ) .   These take values in ˘ * ( K )   rather than just in ˘ ( K )   because of the unique continuation property of harmonic spinors.
Suppose E ˘ * ( K )   is an oriented smooth real vector bundle whose rank d   is equal to the (expected) dimension of M ( T )   . Choose a smooth section s   of E   such that the pull-back section s α , β = R α , β * s   is transverse to the zero-section of the pull-back bundle E α , β = R α , β * E   over M α , β *   for each pair α , β   . (Here the Sobolev exponent p > 4   should be an even integer to ensure the existence of smooth partitions of unity.) Set s = ( R ) * s   , which is a section of E = ( R ) * E   . We write M α = M α , α = M α , α *   and s α = s α , α   etc.
Let M ^ α   , M ^ ( T )   denote the zero-sets of s α   , s   respectively. By index theory we have 0 = dim M ^ ( T ) = dim M ^ α + n α ,   where n α = 0   if α   is irreducible and n α = 1   otherwise. Thus, M ^ α   is empty if α   is reducible.
Lemma 20 If ω n M ^ ( T ( n ) )   for n = 1 , 2 , . . .   , where T ( n )   , then a subsequence of ω n   chain-converges to an element of M ^ α   for some α ~ Y *   .
Moreover, if ω n = [ S n ]   chain-converges to [ S ] M ^ α   then there exists for each n   a smooth u n : X ( T ( n ) ) U ( 1 )   whose restriction to each end R + × Y j   is null-homotopic and such that the sequence u n ( S n )   c-converges over X   to S   .
Proof. The statement of the first sentence follows from [10,Theorem 1.4by dimension counting. Such maps u n   exist in general for chain-convergent sequences when the ω n   all have the same asymptotic limits over the ends R + × Y j   , see [10. Let J H 1 ( Y ; Z )   be the subgroup consisting of elements of the form z | Y   where z   is an element of H 1 ( X # ; Z )   satisfying z | Y j = 0   for j = 1 , . . . , r   .
This group J   acts on the disjoint union M ^ u = α ~ Y * M ^ α ,   permuting the summands.
Lemma 21 The quotient M ^ = M ^ u / J   is a finite set.
Proof. By [10,Theorem 1.3any sequence ω n M ^ α n   , n = 1 , 2 , . . .   has a chain-convergent subsequence, and for dimensional reasons the limit (well-defined up to gauge equivalence) must lie in some moduli space M ^ β   . Furthermore, if ω n   chain-converges to an element in M ^ β   then M ^ α n   is contained in the orbit J M ^ β   for n 0   . Therefore, each M ^ α   is a finite set, and only finitely many orbits J M ^ α   are non-empty. This is equivalent to the statement of the lemma.
Note that J   is the largest subgroup of H 1 ( Y ; Z )   which acts on M ^ u   in a natural way. On the other hand, if M ^ u   is non-emtpy then, since H 1 ( Y ; Z )   acts freely on ~ Y   , only subgroups J J   of finite index have the property that M ^ u / J   is finite.
Lemma 22 There is a compact codimension 0   submanifold K 0 X   such that the restriction map M ^ ( K 0 )   is injective.
Proof. Let [ S j ] M ^ β j   , j = 1 , 2   , where each S j   is in temporal gauge over the ends of X   (and therefore decays exponentially). Suppose there exists a sequence of smooth gauge transformations u n : X : t n U ( 1 )   where t n   , such that u n ( S 1 ) = S 2   over X : t n   . After passing to a subsequence we can arrange that u n   c-converges over X   to some gauge transformation u   with u ( S 1 ) = S 2   . If t 0   then u | { t } × ( ± Y )   will both be homotopic to a smooth v : Y U ( 1 )   with v ( α 1 ) = α 2   . Hence M ^ α 1 , M ^ α 2   lie in the same J   –orbit, and S 1 , S 2   represent the same element of M ^   by [10,Proposition 2.5 (iii).
Thus we can take K 0 = X : t   for t 0   .
Now fix K 0   as in Lemma  22 and with K K 0   . Let { b 1 , . . . , b m }   be the image of the restriction map R K 0 : M ^ ˘ ( K 0 )   . Choose disjoint open neighbourhoods W j ˘ ( K 0 )   of the points b j   . If T 0   then R K 0 ( M ^ ( T ) ) j W j   by Lemma  20 . For such T   we get a natural map g : M ^ ( T ) M ^ .   It is clear that if g   is the map corresponding to a different choice of K 0   and neighbourhoods W j   then g = g   for T   sufficiently large.
Theorem 2 For sufficiently large T   the following hold:
  • (i) Every element of M ^ ( T )   is a regular point in M ( T )   and a regular zero of s   .
  • (ii) g   is a bijection.
Proof. If M ^   is empty then, by Lemma  20 , M ^ ( T )   is empty as well for T 0   , and there is nothing left to prove.
We now fix b j   and for the remainder of the proof omit j   from notation.
(Thus b = b j   , W = W j   etc.) We will show that for T 0   the set B ^ ( T ) = { ω M ^ ( T ) : ω | K 0 W }   consists of precisely one element, and that this element is regular in the sense of (i). This will prove the theorem.
By definition, b   is the restriction of some ω 0 M ^ α   . Choose an open neighbourhood V ˘ * ( K )   of b | K   and a smooth map π : E | V R d   which restricts to a linear isomorphism on every fibre. Choose an open neighbourhood V 0 W   of b   such that R K ( V 0 ) V   . Let G + M α   be a precompact open neighbourhood of ω 0   such that R K 0 ( G ¯ + ) V 0   . The assumption that ω 0   be a regular zero of s α   means that the composite map G + R K V π s R d   is a local diffeomorphism at ω 0   . We can then find an injective smooth map p : R d M α   such that p π s R K = id   in some open neighbourhood G G +   of ω 0   .
In particular, p 1 ( ω 0 ) = { 0 }   and p   is a local diffeomorphism at 0   . Set q = p π s : V M α .   By Theorem  1 there is a kv-pair ( K , V ) ( K 0 , V 0 )   such that if T 0   then G ( T )   consists only of regular monopoles and f = q R K : G ( T ) G   is a diffeomorphism. By Lemma  20 one has B ^ ( T ) = G ( T ) ( s ) 1 ( 0 ) = f 1 ( ω 0 )   for T 0   . For such T   the set B ^ ( T )   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 p   of small C 1   norm has been chosen for which all moduli spaces M ( α 1 , α 2 )   over R × Y   are regular. (This is possible at least when Y   is a rational homology sphere, see [10.) Fix α 1 , α 2 ~ Y *   with dim M ( α 1 , α 2 ) = 2 .   We will show that the disjoint union M ˇ : = β ~ Y * \ { α 1 , α 2 } M ˇ ( α 1 , β ) × M ˇ ( β , α 2 )   is the boundary of a compact 1   –manifold. (In other words, the standard Floer differential d   satisfies d d = 0   at least with Z / 2   coefficients.) To this end we will apply Theorem  2 to the case when X   consists of two copies of R × Y   , say X = R × Y × { 1 , 2 } ,   and we glue R + × Y × { 1 }   with R × Y × { 2 }   . Thus r = 1 , r = 2   . We take K = K 1 K 2   , where K j = [ 0 , 1 ] × Y × { j }   . In this case, ˘ * ( K )   is the quotient of C * ( K )   by the null-homotopic gauge transformations. The bundle E   over ˘ * ( K )   will be the product bundle with fibre R 2   . To define the section s   of E   , choose δ 1 , δ 2 > 0   such that ϑ   has no critical value in the set ( ϑ ( α 2 ) , ϑ ( α 2 ) + δ 2 ] [ ϑ ( α 1 ) δ 1 , ϑ ( α 1 ) ) .   This is possible because we assume Condition (O1) of [10. For any configuration S   over [ 0 , 1 ] × Y   set s j ( S ) = 0 1 ϑ ( S t ) d t ϑ ( α j ) ( 1 ) j δ j .   Note that s j ( S )   does not change if we apply a null-homotopic gauge transformation to S   .
A configuration over K   consists of a pair ( S 1 , S 2 )   of configurations over [ 0 , 1 ] × Y   . Define a smooth function s : ˘ * ( K ) R 2   (ie a section of E   ) by s ( [ S 1 ] , [ S 2 ] ) = ( s 1 ( S 1 ) , s 2 ( S 2 ) ) .   If [ S ]   belongs to some moduli space M ( β 1 , β 2 )   over R × Y   with β 1 β 2   in ~ Y   then d d t ϑ ( S t ) < 0   for all t   by choice of p   . Since J = 0   , the natural map M ^ M ˇ   is therefore a bijection. Let s j   be the pull-back of s j   to M ( T )   . Here M ( T )   is defined using [10,Equation 17with q = 0   , and so can be identified with M ( α 1 , α 2 )   . By [10,Theorem 1.3the set Z ( T ) = { ω M ( T ) : s 1 ( ω ) = 0 , s 2 ( ω ) 0 }   is compact for all T > 0   . If T 0   then, by Theorem  2 , Z ( T )   is a smooth submanifold of M ( T )   , and the composition of the two bijections Z ( T ) = M ^ ( T ) g M ^ M ˇ   yields the desired identification of M ˇ   with the boundary of a compact 1   -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 4   -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 1.3let X   be a spin c   4   -manifold with Riemannian metric g ¯   and tubular ends R ¯ + × Y j   , j = 1 , . . . , r   . Let W   be a finite-dimensional Euclidean vector space and g = { g w } w W   a smooth family of Riemannian metrics on X   all of which agree with g ¯   outside X : 0   . We then have a principal SO ( 4 )   bundle P SO ( g ) X × W   whose fibre over ( x , w )   consists of all positive g w   -orthonormal frames in T x X   .
In the notation of [10,Subsection 3.1let P GL c P GL +   be the spin c   structure on X   . Denote by P Spin c ( g )   the pull-back of P SO ( g )   under the projection P GL c × W P GL + × W   . Then P Spin c ( g )   is a principal Spin c ( 4 )   bundle over X × W   .
For j = 1 , . . . , r   let α j C ( Y j )   be a non-degenerate smooth monopole.
Let C ( X , g w ; α )   denote the L p , w 1   configuration space for the metric g w   and limits α j   , where p , w   are as in [10,Subsection 3.4. We will provide the disjoint union C ( X , g ; α ) = w W C ( X , g w ; α ) × { w }   with a natural structure of a (trivial) smooth fibre bundle over W   . Let
v : P Spin c ( g ) P Spin c ( g 0 ) × W (15)
be any isomorphism of Spin c ( 4 )   bundles which covers the identity on X × W   and which outside X : 1   is given by the identification P Spin c ( g w ) = P Spin c ( g 0 )   .
There is then an induced isomorphism of SO ( 4 )   bundles P SO ( g ) P SO ( g 0 ) × W   , since these are quotients of the corresponding Spin c ( 4 )   bundles by the U ( 1 )   action. Such an isomorphism v   can be constructed by means of the holonomy along rays of the form { x } × R ¯ + w   where ( x , w ) X × W   , with respect to any connection in P Spin c ( g )   which outside X : 1   is the pull-back of a connection in P Spin c ( g 0 )   . Then v   induces a G ( X ; α )   equivariant diffeomorphism
C ( X , g w ; α ) C ( X , g 0 ; α ) (16)
for each w   , 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 v   . Putting together the maps ( 16 ) for all w   yields a bijection v * : C ( X , g ; α ) C ( X , g 0 ; α ) × W .   If v ~   is another isomorphism as in ( 15 ) then v * ( v ~ * ) 1   is smooth, hence we have obtained the desired structure on C ( X , g ; α )   . Furthermore, because of the gauge equivariance of v *   we also get a similar smooth fibre bundle structure on
b * ( X , g ; α ) = w W b * ( X , g w ; α ) × { w } (17)
for any finite subset b X   .
The monopole equations on C ( X , g ; α )   that we will consider will involve an additional class of perturbations which are useful for handling deformations of p   -perturbations. Such a perturbation will depend on the choice of an isomorphism v   as above, a non-negative constant t   , and a smooth map o : C ( X : t , g 0 ) × W L p ( X : t , g 0 ; Λ 2 )   whose image consists of smooth 2   -forms supported in the interior of X : t   .
We make two further assumptions on o   . Firstly, o   should be gauge invariant in the restricted sense that o ( u ( S ) ) = o ( S )   for all S C ( X : t , g 0 ) × W   and all smooth u : X : t U ( 1 )   that restrict to 1   on X : t b   . Here u   acts trivially on W   . Secondly, for every non-negative integer k   the image of o   should be a bounded set with respect to the C k   norm.
Having made these choices, for any w W   we denote by b ( X , g w ; α )   the set of all elements [ A , Φ ] b ( X , g w ; α )   that satisfy the perturbed monopole equations
( F ^ A + m ( v * ( A , Φ , w ) ) ) + Q ( Φ ) = 0 D A Φ = 0 (18)
where m ( A , Φ , w ) = i μ + i q j = 1 r h A , Φ , p j + o ( A , Φ , w ) .   Here μ   and the term involving p j   should be of the same kind as in [10,Equation 18, and of course the spin bundles and Dirac operators are those corresponding to P Spin c ( g w )   . Taking the disjoint union of these moduli spaces for all w   yields the parametrized moduli space b ( X , g ; α ) = w W b ( X , g w ; α ) × { w } .   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 C ( X , g ; α )   into a Banach vector bundle over W   whose fibre over w   is L p , w ( X , g w ; i Λ + S )   . If we conjugate this monopole map by the diffeomorphisms induced by v   we obtain the smooth G b ( X ; α )   equivariant map
Θ : C ( X , g 0 ; α ) × W L p , w ( X , g 0 ; i Λ + S ) , ( A , Φ , w ) ( ( v w ( F ^ A + m ( A , Φ , w ) ) ) + Q ( Φ ) , j v w ( e j ) e j A + a w ( Φ ) ) .  
Here v w   denotes the isomorphism that v   induces from the Clifford bundle of ( X , g w )   to the Clifford bundle of ( X , g 0 )   , and { e j }   is a local g w   -orthonormal frame on X   . Finally, if we temporarily let ( w )   denote the g w   -Riemannian connection in the tangent bundle of X   then a w = v w ( ( w ) ) ( 0 ) .   Note that a w   is supported in X : 1   .
By construction, v *   induces a homeomorphism b ( X , g ; α ) Θ 1 ( 0 ) / G b ( X ; α ) .   A point in b ( X , g ; α )   is called regular if the corresponding zeros of Θ   are regular (a regular zero being one where the derivative of Θ   is surjective).
This notion is independent of v   . By the local slice theorem, the set of regular points in b * ( X , g ; α )   is a smooth submanifold of b * ( X , g ; α )   .

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 X   are R + × ( ± Y j )   , j = 1 , . . . , r   and R + × Y j   , j = 1 , . . . , r   , with non-degenerate limits α j   over R + × ( ± Y j )   and α j   over R + × Y j   , as in Section  2 . The family of metrics g   on X   defines, in a natural way, a smooth family of metrics g ( T ) = { g ( T , w ) } w W   on X ( T )   for any T   . To simplify notation we write C = C ( X , g 0 ; α , α , α )   and C = C ( X ( T ) , g ( T , 0 ) ; α )   . The corresponding orbit spaces will be denoted b , b   , respectively. The parametrized moduli space over X ( T )   is denoted ( T ) b = b ( X ( T ) , g ( T ) ; α ) .   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 1.4carries over to sequences [ A n , Φ n , w n ] b ( X ( T ( n ) ) , g ( T ( n ) ) ; α n )   provided the sequence w n   is bounded (and similarly for [10,Theorem 1.3). The only new ingredient in the proof is the following simple fact: Suppose B   is a Banach space, E , F   vector bundles over a compact manifold, L , L : Γ ( E ) Γ ( F )   differential operators of order d   , and K : Γ ( E ) B   a linear operator. If L   satisfies an inequality f L k p C ( L f L k d p + K f B )   and L , L   are sufficiently close in the sense that ( L L ) f L k d p ε f L k p   for some constant ε > 0   with ε C < 1   , then L   obeys the inequality f L k p ( 1 ε C ) 1 C ( L f L k d p + K f B ) .   This ends our digression, and we continue the main discussion.
Let the paths γ j ± , γ j   and b X   be as in Section  2 , so that ( T ) b = ( ( T ) b ) *   . As before, a choice of reference configuration in C   gives rise to a reference configuration in C   and a holonomy map b U ( 1 ) r 0 .   Composing this with the map ( T ) b b   defined by the chosen isomorphism v   yields a holonomy map Hol : ( T ) b U ( 1 ) r 0 .   Fix an open subset G b   whose closure is compact and contains only regular points.
By a kv-pair we mean as before a pair ( K , V )   , where K X   is a compact codimension 0   submanifold which contains b   and intersects every component of X   , and V   is an open neighbourhood of R K ( G ¯ )   in ˘ b ( K , g ) = w W ˘ b ( K , g w ) × { w } .   Now fix a kv-pair ( K , V )   satisfying similar additional assumptions as before: firstly, that V ˘ b * ( K , g )   ; secondly, that if X e   is any component of X   which contains a point from b   then X e K   is connected.
Suppose q : V b   is a smooth map such that q ( ω | K ) = ω   for all ω G   . Choose λ j , λ j > 0   .
Theorem 3 Theorem  1 holds in the present situation if one replaces M ( T ) b   by ( T ) b   .
Proof. The proof of Theorem  1 carries over without any substantial changes.

A Existence of maps q  

Let G M b   be as in Section  2 . In this appendix we will show that there is always a map q   as in ( 2 ) provided K   is sufficiently large. It clearly suffices to prove the same with b * ( K )   in place of ˘ b * ( K )   .
Throughout this appendix K   will be as in Theorem  1 , ie a compact codimension 0   submanifold of X   containing b   such that if X e   is any component of X   which contains a point from b   then X e K   is connected. The restriction map R K : M b b ( K )   then takes values in b * ( K )   .
Proposition 3 If K   is sufficiently large then there exist an open neighbourhood V b * ( K )   of R K ( G ¯ )   and a smooth map q : V M b   such that q ( ω | K ) = ω   for all ω G   .
We first prove two lemmas. Choose a precompact open subset G 0 M b   containing G ¯   .
Lemma 23 If K   is sufficiently large then R K : M b b * ( K )   restricts to an immersion on an open neighbourhood of G ¯ 0   .
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 ω = [ S ] G ¯ 0   . We will show that R K   is an immersion at ω   (hence in a neighbourhood of ω   ) when K   is large enough. Since G ¯ 0   is compact, this will prove the lemma.
Let b *   be the ambient configuration space containing M b   as defined in the beginning of Section  3 .
Let W L p , w 1   be a linear subspace such that the derivative at S   of the projection S + W b *   is a linear isomorphism onto the tangent space of M b   at ω   . Let δ   denote that derivative. For t 0   so large that b X : t   let δ t   be the derivative at S   of the natural map S + W b * ( X : t )   . We claim that δ t   is injective for t 0   . For suppose { w n }   is a sequence in W   such that w n L p , w 1 = 1   and δ t n ( w n ) = 0   for each n   , where t n   . Set K n = X : t n   .
Then w n | K n = Φ f n   for some f n L G b ( K n )   , where Φ   is the spinor field of S   . After passing to a subsequence we may assume that w n   converges in L p , w 1   to some w W   .
By [10,Lemma 2.1there exists for each n   a constant C n <   such that for all h L G b ( K n )   one has h L 2 p C n Φ h L 1 p .   It follows that f n   converges in L 2 p   over compact subsets of X   , with limit f   , say. We obviously have Φ f = w   , hence f L G b   and δ ( w ) = 0   . But this is impossible, since w   has norm 1   . This proves the lemma.
Lemma 24 If K   is sufficiently large then the restriction map G 0 b * ( K )   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 R K ( G 0 )   has a tubular neigbourhood in b * ( K )   . According to [12,p96, if B   is a Banach manifold admitting partitions of unity then any closed submanifold Z B   possesses a tubular neigbourhood. However, the condition that Z   be closed is in fact redundant. To see this, note that Z   is by definition locally closed, hence C = Z ¯ \ Z   is closed in B   . But then Z   is a closed submanifold of B \ C   .
Since a tubular neigbourhood of Z   in B \ C   is also a tubular neigbourhood of Z   in B   , the assertion follows.
References

  1. W. Chen. Casson's invariant and Seiberg–Witten gauge theory. Turkish J. Math., 21:61–81, 1997.
  2. S. K. Donaldson. An application of gauge theory to four dimensional topology. J. Differential Geometry, 18:279–315, 1983.
  3. S. K. Donaldson. Connections, cohomology and the intersection forms of 4   –manifolds. J. Differential Geometry, 24:275–341, 1986.
  4. S. K. Donaldson. Floer Homology Groups in Yang–Mills Theory. Cambridge University Press, 2002.
  5. S. K. Donaldson and P. B. Kronheimer. The Geometry of Four-Manifolds. Oxford University Press, 1990.
  6. P. M. N. Feehan and T. G. Leness. A general S O ( 3 )   –monopole cobordism formula relating Donaldson and Seiberg–Witten invariants. math.DG/0203047.
  7. A. Floer. An instanton invariant for 3–manifolds. Comm. Math. Phys., 118:215–240, 1988.
  8. D. S. Freed and K. K. Uhlenbeck. Instantons and Four-Manifolds. MSRI Publications. Springer-Verlag, second edition, 1991.
  9. K. A. Frøyshov. Monopole Floer homology for rational homology 3   –spheres. In preparation.
  10. K. A. Frøyshov. Monopoles over 4   –manifolds containing long necks, I. Geometry & Topology, 9:1–93, 2005.
  11. K. Fukaya. Floer homology for oriented 3   –manifolds. In Y. Matsumoto, editor, Aspects of low dimensional manifolds, volume 20 of Adv. Stud. Pure Math., pages 1–92. Kinokuniya Company Ltd, 1992.
  12. S. Lang. Differential Manifolds. Springer-Verlag, 1985.
  13. M. Marcolli and B-L. Wang. Equivariant Seiberg–Witten Floer homology. Comm. Anal. and Geom., 9:451–639, 2001.
  14. J. W. Morgan and T. S. Mrowka. On the gluing theorem for instantons on manifolds containing long cylinders, 1994. Preprint.
  15. L. I. Nicolaescu. Notes on Seiberg–Witten theory. American Mathematical Society, 2000.
  16. P. Safari. Gluing Seiberg–Witten monopoles. math.DG/0311329.
  17. C. H. Taubes. Self-dual Yang–Mills connections on non-self-dual 4-manifolds. J. Differential Geometry, 17:139–170, 1982.
  18. C. H. Taubes. Self-dual connections on 4   –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