This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Positive scalar curvature and homology cobordism invariants

Hokuto Konno Graduate School of Mathematical Sciences, the University of Tokyo, 3-8-1 Komaba, Meguro, Tokyo 153-8914, Japan [email protected]  and  Masaki Taniguchi 2-1 Hirosawa, Wako, Saitama 351-0198, Japan [email protected]
Abstract.

We determine the local equivalence class of the Seiberg–Witten Floer stable homotopy type of a spin rational homology 3-sphere YY embedded into a spin rational homology S1×S3S^{1}\times S^{3} with a positive scalar curvature metric so that YY generates the third homology. The main tool of the proof is a relative Bauer–Furuta-type invariant on a periodic-end 44-manifold. As a consequence, we give obstructions to positive scalar curvature metrics on spin rational homology S1×S3S^{1}\times S^{3}, typically described as the coincidence of various Frøyshov-type invariants. This coincidence also yields alternative proofs of two known obstructions by Jianfeng Lin and by the authors for the same class of 44-manifolds.

1. Introduction

1.1. Seiberg–Witten Floer stable homotopy type and local equivalence

Manolescu’s Seiberg–Witten Floer stable homotopy type [Ma03] is a space-valued Floer theoretic invariant of a rational homology 33-sphere equipped with a spinc structure, and recovers the monopole Floer homology defined by Kronheimer and Mrowka [KM07] for this class of 33-manifolds [LM18]. Therefore, in principle, the Seiberg–Witten Floer stable homotopy type contains all Floer-theoretic information from Seiberg–Witten theory for rational homology 33-spheres.

In this paper, we will consider a spin rational homology 33-sphere (Y,𝔱)(Y,\mathfrak{t}) embedded into a spin 44-manifold (X,𝔰)(X,\mathfrak{s}) with the rational homology of S1×S3S^{1}\times S^{3} so that the fundamental class of YY generates H3(X;)H_{3}(X;\mathbb{Z}). The main theorem of this paper states that, if XX admits a metric with positive scalar curvature (PSC), we can determine the Seiberg–Witten Floer stable homotopy type of such (Y,𝔱)(Y,\mathfrak{t}), denoted by SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}), up to the local equivalence relation explained below. This result gives a strong obstruction to PSC metrics of spin rational homology S1×S3S^{1}\times S^{3}, and this is the authors’ original motivation for this study. To the same class of 44-manifolds, there are two known obstructions based on Seiberg–Witten theory, Jianfeng Lin’s obstruction [Lin19] and the authors’ obstruction [KT20] explained later, and the main theorem of this paper recovers both of them.

To motivate to consider the local equivalence relation, let us recall several homology cobordism invariants from Seiberg–Witten theory. Applying various equivariant ordinary/generalized cohomologies to the Seiberg–Witten Floer stable homotopy type, many numerical homology cobordism invariants can be extracted, such as, the Frøyshov invariant [Fr96, Fr10], which we denote by δ\delta following [Ma16], Manolescu’s invariants α,β,γ,κ\alpha,\beta,\gamma,\kappa [Ma14, Ma16], and Stoffregen’s invariants δ¯,δ¯\overline{\delta},\underline{\delta} [Sto171]. These invariants have different applications, for example: The Frøyshov invariant δ\delta was used to extend Donaldson’s diagonalization theorem [Do83] to negative-definite 44-manifolds with boundary [Fr96, Fr10, Ma03]. Manolescu used the invariant β\beta to disprove the triangulation conjecture [Ma16], and used κ\kappa to extend Furuta’s 10/8-inequality [Fu01] to spin 44-manifolds with boundary [Ma14]. Stoffregen’s invariants δ¯,δ¯\overline{\delta},\underline{\delta} should correspond, respectively, to d¯,d¯\overline{d},\underline{d} in involutive Heegaard Floer homology [HM17], using /4\mathbb{Z}/4-equivariant ordinary cohomology.

These invariants α,β,γ,δ,δ¯,δ¯,κ\alpha,\beta,\gamma,\delta,\overline{\delta},\underline{\delta},\kappa are spin rational homology cobordism invariants, and obtained from SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}) described above. However, these invariants factor through a weaker invariant than SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}), the local equivalence class of SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}), defined by Stoffregen [Sto20]. The local equivalence is an equivalence relation on a certain class of spaces including SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}) for rational homology 33-spheres YY, and this is an abstraction of a relation between SWF(Y0,𝔱0)\mathrm{SWF}(Y_{0},\mathfrak{t}_{0}) and SWF(Y1,𝔱1)\mathrm{SWF}(Y_{1},\mathfrak{t}_{1}) for (Y0,𝔱0)(Y_{0},\mathfrak{t}_{0}) and (Y1,𝔱1)(Y_{1},\mathfrak{t}_{1}) which are spin rational homology cobordant to each other. To summarize this situation, let us denote by Θ3\Theta_{\mathbb{Z}}^{3} the 33-dimensional homology cobordism group, and denote by Θ,spin3\Theta_{\mathbb{Q},\rm spin}^{3} the 33-dimensional spin rational homology cobordism group. Namely, an element of Θ,spin3\Theta_{\mathbb{Q},\rm spin}^{3} is the equivalence class [(Y,𝔱)][(Y,\mathfrak{t})] of a spin rational homology 33-sphere, and the equivalence relation is given by a spin rational homology cobordism. Stoffregen [Sto20] introduced the local equivalence group \mathcal{LE}, which consists of the local equivalence classes of certain spaces modeled on SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}). Then one has group homomorphisms

Θ3Θ,spin3.\Theta_{\mathbb{Z}}^{3}\to\Theta_{\mathbb{Q},\rm spin}^{3}\to\mathcal{LE}.

For a spin rational homology 33-sphere (Y,𝔱)(Y,\mathfrak{t}), the local equivalence class [SWF(Y,𝔱)][\mathrm{SWF}(Y,\mathfrak{t})] is valued in \mathcal{LE}, and the above numerical invariants α,β,γ,δ,δ¯,δ¯,κ\alpha,\beta,\gamma,\delta,\overline{\delta},\underline{\delta},\kappa factor through \mathcal{LE}, such as α(Y,𝔱)=α([(Y,𝔱)])=α([SWF(Y,𝔱)])\alpha(Y,\mathfrak{t})=\alpha([(Y,\mathfrak{t})])=\alpha([\mathrm{SWF}(Y,\mathfrak{t})]):

Θ,spin3α,β,γ,δ,δ¯,δ¯,κ.\Theta_{\mathbb{Q},\rm spin}^{3}\to\mathcal{LE}\xrightarrow{\alpha,\beta,\gamma,\delta,\overline{\delta},\underline{\delta},\kappa}\mathbb{Q}.

1.2. Main theorem

As described, the local equivalence class [SWF(Y,𝔱)][\mathrm{SWF}(Y,\mathfrak{t})]\in\mathcal{LE} of SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}) is, so far at least, a candidate of the ‘universal’ Seiberg–Witten theoretic homology cobordism invariant of (Y,𝔱)(Y,\mathfrak{t}): it contains information of all known homology cobordism invariants obtained from Seiberg–Witten theory. The main theorem of this paper determines [SWF(Y,𝔱)][\mathrm{SWF}(Y,\mathfrak{t})] when YY is embedded into a spin rational homology S1×S3S^{1}\times S^{3} admitting a PSC metric so that YY generates H3(X;)H_{3}(X;\mathbb{Z}):

Theorem 1.1.

Let (X,𝔰)(X,\mathfrak{s}) be an oriented spin rational homology S1×S3S^{1}\times S^{3}, and (Y,𝔱)(Y,\mathfrak{t}) be an oriented spin rational homology 33-sphere. Suppose that (Y,𝔱)(Y,\mathfrak{t}) is a cross-section of (X,𝔰)(X,\mathfrak{s}), i.e. YY is embedded into XX so that it represents a fixed generator of H3(X;)H_{3}(X;\mathbb{Z}), and that 𝔰|Y\mathfrak{s}|_{Y} is isomorphic to 𝔱\mathfrak{t}. Assume that XX admits a PSC metric. Then the local equivalence class of SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}) is given by

(1) [SWF(Y,𝔱)]=[(S0,0,λSW(X,𝔰)2)].\displaystyle[\mathrm{SWF}(Y,\mathfrak{t})]=\left[\left(S^{0},0,-\frac{\lambda_{SW}(X,\mathfrak{s})}{2}\right)\right].

In particular, for an arbitrary spin rational homology cobordism invariant which factors through \mathcal{LE}, the invariant of (Y,𝔱)(Y,\mathfrak{t}) coincides with the invariant of the right-hand side of (1).

Here λSW(X,𝔰)\lambda_{SW}(X,\mathfrak{s}) is the Casson-type invariant defined by the Mrowka–Ruberman–Saveliev [MRS11] for an integral homology S1×S3S^{1}\times S^{3}, which was later generalized for a rational homology S1×S3S^{1}\times S^{3} by J. Lin–Ruberman–Saveliev [LRS17]. Recall that an element of \mathcal{LE} is expressed as the class of a triple (Z,m,n)(Z,m,n), where ZZ is a space of type SWF [Ma16, Definition 2.7], and mm\in\mathbb{Z}, nn\in\mathbb{Q}.

Remark 1.2.

In this paper, we developed Seiberg–Witten theory for 4-manifolds with periodic ends to prove Theorem 1.1. But we expect that an alternative proof of Theorem 1.1 without using Seiberg–Witten theory for 4-manifolds with periodic ends could be given by using Schoen–Yau’s argument [SY86] combined with a kind of gluing theorems for relative Bauer–Furuta invariants [KLS18, KLS'18, SS21].

1.3. Obstructions to PSC metrics

Now we regard Theorem 1.1 as an obstruction to PSC metrics on homology S1×S3S^{1}\times S^{3}, and compare this with known obstructions on PSC metrics for the same class of 44-manifolds. We can extract from Theorem 1.1 convenient obstructions to PSC metrics, and moreover that Theorem 1.1 provides a systematic way to recover prior results.

Recall that it is well-understood which rational homology 33-spheres admit PSC metrics: only connected sums of spherical 33-manifolds. Rational homology S1×S3S^{1}\times S^{3} is a class of 44-manifold that may be seen to be closed to rational homology 33-sphere, but it is not easy to rule out the existence of PSC metrics on such 44-manifolds. In dimension 44, the Seiberg–Witten invariant is known as a powerful obstruction to PSC metric, but it cannot be used to rational homology S1×S3S^{1}\times S^{3}, since the Seiberg–Witten invariant is not well-defined for such 44-manifolds. J. Lin recently made a breakthrough in this situation: he gave the first obstruction to PSC metric based on Seiberg–Witten theory for integral homology S1×S3S^{1}\times S^{3} in [Lin19], and later this result was generalized by himself with Ruberman and Saveliev to any rational homology S1×S3S^{1}\times S^{3} in [LRS17]. J. Lin’s obstruction is described as follows: under the same assumption with Theorem 1.1, one has the equality

(2) δ(Y,𝔱)=λSW(X,𝔰).\displaystyle\delta(Y,\mathfrak{t})=\lambda_{SW}(X,\mathfrak{s}).

Using Theorem 1.1, we can give an alternative proof of J. Lin’s formula (2), and further generalize it to various Frøyshov-type invariants:

Corollary 1.3.

Let (X,𝔰)(X,\mathfrak{s}) be an oriented spin rational homology S1×S3S^{1}\times S^{3}, and (Y,𝔱)(Y,\mathfrak{t}) be an oriented spin rational homology 33-sphere. Suppose that (Y,𝔱)(Y,\mathfrak{t}) is a cross-section of (X,𝔰)(X,\mathfrak{s}). Assume that XX admits a PSC metric. Then we have

(3) α(Y,𝔱)=β(Y,𝔱)=γ(Y,𝔱)=δ(Y,𝔱)=δ¯(Y,𝔱)=δ¯(Y,𝔱)=κ(Y,𝔱)=λSW(X,𝔰).\displaystyle\alpha(Y,\mathfrak{t})=\beta(Y,\mathfrak{t})=\gamma(Y,\mathfrak{t})=\delta(Y,\mathfrak{t})=\overline{\delta}(Y,\mathfrak{t})=\underline{\delta}(Y,\mathfrak{t})=\kappa(Y,\mathfrak{t})=\lambda_{SW}(X,\mathfrak{s}).
Proof.

By the definition of α,β,γ,δ,δ¯,δ¯,κ\alpha,\beta,\gamma,\delta,\overline{\delta},\underline{\delta},\kappa [Ma16, Ma14, Sto171], it is easy to see that the values of these invariants for the right-hand side of (1) are given by λSW(X,𝔰)\lambda_{SW}(X,\mathfrak{s}). Therefore the Corollary directly follows from Theorem 1.1. ∎

Note that, by Corollary 1.3, we can replace λSW(X,𝔰)\lambda_{SW}(X,\mathfrak{s}) in the right-hand side of (1) with various invariants of (Y,𝔱)(Y,\mathfrak{t}).

An obvious consequence of Corollary 1.3 is:

Corollary 1.4.

Let YY be an oriented homology 3-sphere. Suppose that at least two of α(Y),β(Y),γ(Y),δ(Y),δ¯(Y),δ¯(Y),κ(Y)\alpha(Y),\beta(Y),\gamma(Y),\delta(Y),\overline{\delta}(Y),\underline{\delta}(Y),\kappa(Y) do not coincide with each other. Then, for any homology cobordism WW from YY to itself, the homology S1×S3S^{1}\times S^{3} obtained from WW by gluing the boundary components does not admit a PSC metric.

Here we drop the unique spin structure from our notation for (integral) homology 33-spheres.

J. Lin [Lin19] and J. Lin–Ruberman–Saveliev [LRS17] used monopole Floer homology to establish the obstruction (2). Morally, our argument in this paper can be thought of as a stable cohomotopy version of J. Lin’s argument in [Lin19].

After J. Lin’s work, the authors [KT20] gave another obstruction based on a 10/8-type inequality, described in Corollary 1.5. Using Corollary 1.3 combined with Manolescu’s relative 10/8-inequality [Ma14], we can give an alternative proof of the authors’ previous result (with a minor change):

Corollary 1.5 ([KT20]).

Let (X,𝔰),(Y,𝔱)(X,\mathfrak{s}),(Y,\mathfrak{t}) be as in Theorem 1.1. Take a compact smooth spin 44-manifold MM bounded by (Y,𝔱)(Y,\mathfrak{t}). Suppose that (Y,𝔱)(Y,\mathfrak{t}) is a cross-section of (X,𝔰)(X,\mathfrak{s}). Assume that XX admits a PSC metric. Then we have

(4) b+(M)σ(M)8δ(Y,𝔱)1.\displaystyle b^{+}(M)\geq-\frac{\sigma(M)}{8}-\delta(Y,\mathfrak{t})-1.
Proof.

Manolescu’s relative 10/8-inequality, which is [Ma14, Theorem 1] generalized to a rational homology 33-sphere (see [Ma14, Remark 2]), implies that

b+(M)σ(M)8κ(Y,𝔱)1.b^{+}(M)\geq-\frac{\sigma(M)}{8}-\kappa(Y,\mathfrak{t})-1.

Combining this with (3), we obtain (4). ∎

Remark 1.6.

The inequality (4) is slightly weaker than the original inequality given in [KT20, Theorem 1.1]. The source of this difference is that, in [KT20], we used Furuta-Kametani’s 10/8-type inequality [FK05] based on KOKO-theory, whereas Manolescu’s inequality is based on KK-theory.

1.4. Outline of the proof of the main theorem

Here is an explanation of an outline of the proof of Theorem 1.1. The heart of this paper is, under the assumption of the existence of PSC metric on XX, to consider finite-dimensional approximations of the Seiberg–Witten equations on a periodic-end 44-manifold. More precisely, we shall construct a relative Bauer–Furuta-type invariant over a half-periodic-end 44-manifold

W[,0]=YWYWYW,W[-\infty,0]=\cdots\cup_{Y}W\cup_{Y}W\cup_{Y}W,

along the spirit of Furuta [Fu01], Bauer–Furuta [BF04], and Manolescu [Ma03]. Here WW is the 44-manifold defined by cutting XX open along YY, and the ‘left side’ end is equipped with a periodic PSC metric and a neighborhood of the ‘right side’ boundary is equipped with a product metric of the form [0,1]×Y[0,1]\times Y. Technically, the relative Bauer–Furuta invariant over such a non-compact 4-manifold is defined using the similar method given in [IT20] which defines the relative Bauer–Furuta invariant for a certain class of 4-manifolds with conical end.

The key observation is that W[,0]W[-\infty,0] with such a periodic PSC metric on the end looks like a homology cobordism from S3S^{3} to YY from Seiberg–Witten theoretic point of view. The relative Bauer–Furuta invariant over W[,0]W[-\infty,0] gives a local map from [(S0,0,λSW(X,𝔰)/2)]\left[\left(S^{0},0,-\lambda_{SW}(X,\mathfrak{s})/2\right)\right] to SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}). The quantity λSW(X,𝔰)\lambda_{SW}(X,\mathfrak{s}) emerges from the spin Dirac index over W[,0]W[-\infty,0], discussed in Subsection 3.3.

Similarly, by considering the relative Bauer–Furuta invariant over

W[0,]=WYWYWY,W[0,\infty]=W\cup_{Y}W\cup_{Y}W\cup_{Y}\cdots,

we get a local map from SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}) to [(S0,0,λSW(X,𝔰)/2)]\left[\left(S^{0},0,-\lambda_{SW}(X,\mathfrak{s})/2\right)\right], and we can conclude that SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}) is locally equivalent to [(S0,0,λSW(X,𝔰)/2)]\left[\left(S^{0},0,-\lambda_{SW}(X,\mathfrak{s})/2\right)\right].

1.5. Examples

In Section 8 we shall give examples of concrete 3-manifolds YY to which we can apply the obstructions given in Subsection 1.3. Here let us exhibit a part of those examples.

As a consequence of his formula (2), J. Lin proved in [Lin19, Corollary 1.3] that a homology S1×S3S^{1}\times S^{3} having a cross-section YY with μ(Y)δ(Y)mod2\mu(Y)\neq\delta(Y)\mod 2 does not admit a PSC metric. Here μ(Y)/2\mu(Y)\in\mathbb{Z}/2\mathbb{Z} denotes the Rohlin invariant. For Seifert homology 3-spheres, we can get an ‘integer-valued lift’ of this result by J. Lin. Moreover, also for linear combinations of Seifert homology 3-spheres of certain type, we can get an obstructions described in terms of some integer-valued invariants of certain 3-manifolds:

Theorem 1.7.

The following statements hold:

  • (i)

    Let YY^{\prime} be a Seifert homology 3-sphere such that

    μ¯(Y)δ(Y),-\overline{\mu}(Y^{\prime})\neq\delta(Y^{\prime}),

    where μ¯\overline{\mu} is the Neumann–Siebenmann invariant for graph homology 3-spheres, introduced in [N80, Si80]. Let YY be an oriented homology 3-sphere which is homology cobordant to YY^{\prime}. Then, for any homology cobordism WW from YY to itself, the 4-manifold obtained from WW by gluing the boundary components does not admit a PSC metric.

  • (ii)

    Let Y1,,YnY_{1},\cdots,Y_{n} be negative Seifert homology 3-spheres of projective type. Suppose that δ(Y1)δ(Yn)\delta(Y_{1})\leq\cdots\leq\delta(Y_{n}). Set δ~i:=δ(Yi)+μ¯(Yi)\widetilde{\delta}_{i}:=\delta(Y_{i})+\overline{\mu}(Y_{i}). Suppose that at least two of following four integers are distinct:

    i=1nδ(Yi),2i=1nδ~i+12i=1nμ¯(Yi),\displaystyle\sum_{i=1}^{n}{\delta}(Y_{i}),\quad 2\lfloor\frac{\sum_{i=1}^{n}\widetilde{\delta}_{i}+1}{2}\rfloor-\sum_{i=1}^{n}\overline{\mu}(Y_{i}),
    2i=1n1δ~i+12i=1nμ¯(Yi),2i=1n2δ~i+12i=1nμ¯(Yi).\displaystyle 2\lfloor\frac{\sum_{i=1}^{n-1}\widetilde{\delta}_{i}+1}{2}\rfloor-\sum_{i=1}^{n}\overline{\mu}(Y_{i}),\quad 2\lfloor\frac{\sum_{i=1}^{n-2}\widetilde{\delta}_{i}+1}{2}\rfloor-\sum_{i=1}^{n}\overline{\mu}(Y_{i}).

    Let YY be an oriented homology 3-sphere which is homology cobordant to Y1##YnY_{1}\#\cdots\#Y_{n}. Then, for any homology cobordism WW from YY to itself, the 4-manifold obtained from WW by gluing the boundary components does not admit a PSC metric.

For the definition of projective Seifert homology 3-spheres, see Section 8.

1.6. Outline of this paper

We finish off this introduction with an outline of the contents of this paper. The contents until Section 5 are devoted to construct the relative Bauer–Furuta invariant on the periodic-end 4-manifold W[,0]W[-\infty,0]. In Section 2 we give several notations related to infinite cyclic covering spaces of a 44-manifold. We also review Fredholm theory for infinite cyclic covering spaces, Seiberg–Witten Floer homotopy types and notion of local equivalence. In Section 3 we ensure Fredholm properties of elliptic operators on certain 4-manifolds with periodic end and boundary. We calculate cohomologies of the Atiyah–Hitchin–Singer operator on such non-compact 4-manifolds. We also calculate the Dirac index on W[,0]W[-\infty,0] in Subsection 3.3. In Section 4 we show a boundedness result which is needed to construct the relative Bauer–Furuta invariant. In Section 5 we construct the relative Bauer–Furuta invariant for the 4-manifolds W[,0]W[-\infty,0] with periodic end and boundary. In Section 6 we prove Theorem 1.1 along the idea explained in Subsection 1.4. In Section 7 we give a generalization of Theorem 1.1, which is stated as an obstruction of embeddings of 33-manifolds into 44-manifolds admitting PSC metrics. In Section 8 we provide several families of examples of homology S1×S3S^{1}\times S^{3}’s which cannot admit PSC metrics using Theorem 1.1.

Acknowledgment.

The authors would like to express their gratitude to the organizers and participants of Gauge Theory Virtual for giving them an opportunity to reconsider their past work [KT20]. The authors also wish to thank Nobuo Iida for discussing Remark 1.2 with us. The first author was partially supported by JSPS KAKENHI Grant Numbers 17H06461, 19K23412, and 21K13785. The second author was supported by JSPS KAKENHI Grant Number 20K22319 and RIKEN iTHEMS Program.

2. Preliminaries

2.1. Notations

In this Subsection we introduce several notations on periodic 44-manifolds. Let (X,𝔰)(X,\mathfrak{s}) be an oriented spin rational homology S1×S3S^{1}\times S^{3}, i.e. a spin 44-manifold whose rational homology is isomorphic to that of S1×S3S^{1}\times S^{3}. Fix a Riemannian metric gXg_{X} on XX and a generator of H3(X;)H_{3}(X;\mathbb{Z}), denoted by 1H3(X;)1\in H_{3}(X;\mathbb{Z}). Note that H3(X;)H_{3}(X;\mathbb{Z}) is isomorphic to H1(X;)H^{1}(X;\mathbb{Z}), and hence to \mathbb{Z}. Let YY be an oriented rational homology 33-sphere, and assume that YY is embedded into XX so that [Y]=1[Y]=1. We call such YY a cross-section of XX. Let W0W_{0} be the rational homology cobordism from YY to itself obtained by cutting XX open along YY. The manifold W0W_{0} is equipped with an orientation and a spin structure induced by those of XX. For (m,n)({})×({})(m,n)\in(\{-\infty\}\cup\mathbb{Z})\times(\mathbb{Z}\cup\{\infty\}) with m<nm<n, we define the periodic 44-manifold

W[m,n]:=WmYWm+1YYWn,W[m,n]:=W_{m}\cup_{Y}W_{m+1}\cup_{Y}\dots\cup_{Y}W_{n},

where WiW_{i} is a copy of W0W_{0} for each ii\in\mathbb{Z}. This 44-manifold W[m,n]W[m,n] is also equipped with an orientation and a spin structure as well as W0W_{0}. The element of H1(X;)H^{1}(X;\mathbb{Z}) corresponding to 1H3(X;)1\in H_{3}(X;\mathbb{Z}) via the Poincaré duality gives the isomorphism class of an infinite cyclic covering

(5) p:X~X\displaystyle p:\widetilde{X}\to X

and an identification

(6) X~W[,].\displaystyle\widetilde{X}\cong W[-\infty,\infty].

Via the identification (6), let us think of pp as a map from W[,]W[-\infty,\infty] to XX. Define the map p:W[,0]Xp_{-}:W[-\infty,0]\to X as the restriction of pp. We call an object defined on W[,0]W[-\infty,0], such as connection, metric, bundle, and differential operator, a periodic object if the restriction of the object to W[,0]W[-\infty,0] can be identified with the pull-back of an object on XX under pp_{-}. Considering the pull-back under pp_{-}, the Riemannian metric gXg_{X} on XX induces a Riemannian metric, denoted by gW[,0]g_{W[-\infty,0]}, on W[,0]W[-\infty,0]. Let S+,SS^{+},S^{-} be the positive/negative spinor bundles respectively over W[,0]W[-\infty,0] with respect to the metric and the spin structure above. Fixing a trivialization of the determinant line bundle of the spin structure on W[,0]W[-\infty,0], we obtain the canonical reference connection A0A_{0} on W[,0]W[-\infty,0] corresponding to the trivial connection.

To consider the weighted Sobolev norms on W[,0]W[-\infty,0], fix a function

τ:X~\tau:\widetilde{X}\to\mathbb{R}

with Tτ=τ+1T^{*}\tau=\tau+1, where T:X~X~T:\widetilde{X}\to\widetilde{X} is the deck transform determined by T(Wi)=Wi1T(W_{i})=W_{i-1}. Note that dτd\tau defined a cohomology class [dτ]H1(X;)[d\tau]\in H^{1}(X;\mathbb{Z}) which is equal to 1H1(X;)1\in H^{1}(X;\mathbb{Z}) corresponding to 1H3(X;)1\in H_{3}(X;\mathbb{Z}) via the Poincaré duality.

Definition 2.1.

Let EE be a periodic vector bundle on W[,0]W[-\infty,0] with a periodic inner product. For a fixed k>0k>0 and δ\delta\in\mathbb{R}, we define the weighted Sobolev norm by

fLk,δ2(W[,0]):=eδτfLk2(W[,0]).\|f\|_{L^{2}_{k,\delta}(W[-\infty,0])}:=\|e^{\delta\tau}f\|_{L^{2}_{k}(W[-\infty,0])}.

for a smooth comactly supported section ff of EE. Here we used a periodic metric and a periodic connection on EE to define the Lk2L^{2}_{k}-norm. Let Lk,δ2(E)L^{2}_{k,\delta}(E) denote the Lk,δ2L^{2}_{k,\delta}-completion of compactly supported smooth sections of EE.

Note that the equivalence class of norms Lk,δ2(W[,0])\|-\|_{L^{2}_{k,\delta}(W[-\infty,0])} does not depend on the choices of a periodic metric and a periodic connection on EE.

2.2. Fredholm theory on X~\widetilde{X}

In this Subsection we review the Fredholm property of periodic elliptic operators on the infinite cyclic covering X~\widetilde{X} developed by C. Taubes [T87]. He showed that a periodic elliptic operator is Fredholm under some condition with respect to Lk,δ2L^{2}_{k,\delta}-norms for generic δ\delta\in\mathbb{R}. For the details, see [T87], or [KT20, Subsection 2.1].

Let 𝔻=(Di,Ei)\mathbb{D}=(D_{i},E_{i}) be a periodic elliptic complex on X~\widetilde{X}, i.e. the complex

(7) 0Γ(X~;EN)DNΓ(X~;EN1)D1Γ(X~;E0)0\displaystyle 0\to\Gamma(\widetilde{X};E_{N})\xrightarrow{D_{N}}\Gamma(\widetilde{X};E_{N-1})\to\cdots\xrightarrow{D_{1}}\Gamma(\widetilde{X};E_{0})\to 0

consisting of first order periodic linear differential operators DiD_{i} between periodic vector bundles EiE_{i} on X~\widetilde{X} with exact symbol sequence. Here, for a vector bundle EE, the notation Γ(X~,E)\Gamma(\widetilde{X},E) denotes the set of compactly supported smooth sections of EE. As well as Definition 2.1, define the weighted Sobolev norm on X~\widetilde{X} by

fLk,δ2(X~):=eτδfLk2(X~)\|f\|_{L^{2}_{k,\delta}(\widetilde{X})}:=\|e^{\tau\delta}f\|_{L^{2}_{k}(\widetilde{X})}

using a periodic connection and a periodic metric. The complex (7) gives rise to the complex of bounded operators

(8) Lk+N+1,δ2(X~;EN)DNLk+N,δ2(X~;EN1)D1Lk,δ2(X~;E0)\displaystyle L^{2}_{k+N+1,\delta}(\widetilde{X};E_{N})\xrightarrow{D_{N}}L^{2}_{k+N,\delta}(\widetilde{X};E_{N-1})\to\cdots\xrightarrow{D_{1}}L^{2}_{k,\delta}(\widetilde{X};E_{0})

for each k>0k>0 and δ\delta\in\mathbb{R}.

Note that, since the operators in (8) are periodic differential operators, there exist differential operators 𝔻^=(D^i,E^i)i=0,,N\hat{\mathbb{D}}=(\hat{D}_{i},\hat{E}_{i})_{i=0,\cdots,N} on XX such that 𝔻\mathbb{D} is given as the pull-back p𝔻^p_{-}^{*}\hat{\mathbb{D}}.

Definition 2.2.

For zz\in\mathbb{C}, define the complex 𝔻^(z)\hat{\mathbb{D}}(z) by

0Γ(X;E^N)D^N(z)Γ(X;E^N1)D^1(z)Γ(X;E^0)0,0\to\Gamma(X;\hat{E}_{N})\xrightarrow{\hat{D}_{N}(z)}\Gamma(X;\hat{E}_{N-1})\to\cdots\xrightarrow{\hat{D}_{1}(z)}\Gamma(X;\hat{E}_{0})\to 0,

where the operators D^i(z):Γ(X;E^i)Γ(X;E^i1)\hat{D}_{i}(z):\Gamma(X;\hat{E}_{i})\to\Gamma(X;\hat{E}_{i-1}) are defined by

D^i(z)(f):=eτzD^i(eτzf).\hat{D}_{i}(z)(f):=e^{-\tau z}\hat{D}_{i}(e^{\tau z}f).
Theorem 2.3 ([T87, Lemmas 4.3 and 4.5]).

Suppose that there exists z0z_{0}\in\mathbb{C} where the complex 𝔻^(z0)\hat{\mathbb{D}}(z_{0}) is acyclic. Then there exists a discrete subset 𝒟\mathcal{D} in \mathbb{R} with no accumulation points such that the complex (8) is an acyclic complex for all δ\delta in 𝒟\mathbb{R}\setminus\mathcal{D}.

Definition 2.4 ([RS07]).

We call gXg_{X} an admissible metric on XX if the kernel of

DA0++fdθ:Lk2(X;S+)Lk12(X;S)D^{+}_{A_{0}}+f^{*}d\theta:L^{2}_{k}(X;S^{+})\to L^{2}_{k-1}(X;S^{-})

is zero, where the map f:XS1f:X\to S^{1} is a smooth classifying map of (5).

The admissibility condition does not depend on the choice of classifying map ff. One can show that every PSC metric on XX is an admissible metric (See (2) in [RS07]).

In [KT20], we confirmed that Theorem 2.3 can be used for differential operators appearing as the linearization of the Seiberg–Witten equations:

Lemma 2.5 ([KT20, Lemma 2.6]).

The assumption of Theorem 2.3 is satisfied for the following operator/complexes:

  • The Dirac operator DA0+:Lk,δ2(X~;S+)Lk1,δ2(X~;S)D^{+}_{A_{0}}:L^{2}_{k,\delta}(\widetilde{X};S^{+})\to L^{2}_{k-1,\delta}(\widetilde{X};S^{-}) with respect to the pull-back of an admissible metric gXg_{X} on XX.

  • The Atiyah–Hitchin–Singer complex

    0Lk+1,δ2(iΛ0(X~))𝑑Lk,δ2(iΛ1(X~))d+Lk1,δ2(iΛ+(X~))0.0\to L^{2}_{k+1,\delta}(i\Lambda^{0}(\widetilde{X}))\xrightarrow{d}L^{2}_{k,\delta}(i\Lambda^{1}(\widetilde{X}))\xrightarrow{d^{+}}L^{2}_{k-1,\delta}(i\Lambda^{+}(\widetilde{X}))\to 0.
  • The de Rham complex

    0Lk+1,δ2(iΛ0(X~))𝑑Lk,δ2(iΛ1(X~))𝑑𝑑Lk3,δ2(iΛ4(X~))0.0\to L^{2}_{k+1,\delta}(i\Lambda^{0}(\widetilde{X}))\xrightarrow{d}L^{2}_{k,\delta}(i\Lambda^{1}(\widetilde{X}))\xrightarrow{d}\cdots\xrightarrow{d}L^{2}_{k-3,\delta}(i\Lambda^{4}(\widetilde{X}))\to 0.
Remark 2.6.

Since the subset 𝒟\mathcal{D} of \mathbb{R} given in Theorem 2.3 has no accumulation points, we can take a sufficiently small δ0>0\delta_{0}>0 so that for any δ(0,δ0)\delta\in(0,\delta_{0}) the operators in Lemma 2.5 are Fredholm. Henceforth we fix the notation δ0\delta_{0}.

2.3. Seiberg–Witten Floer stable homotopy type

In the proof of Theorem 1.1, we use a variant of the relative Bauer–Furuta invariant for 4-manifolds with periodic end. In this subsection we review several notions of Manolescu’s Seiberg–Witten Floer stable homotopy type, which is necessary to describe the relative Bauer–Furuta invariant. The main references of this subsection are Manolescu [Ma03] and Khandhawit [Kha15].

Let YY be an oriented rational homology 33-sphere with a Riemannian metric gYg_{Y}. Let 𝔱\mathfrak{t} be a spinc structure on YY, and SS be the spinor bundle of 𝔱\mathfrak{t}. We fix a flat spinc reference connection a0a_{0} of the determinant line bundle of SS.

Definition 2.7.

For an integer k>2k>2, we define the configuration space by

𝒞k(Y,𝔱):=(a0+Lk122(iΛY1))Lk122(S).\mathcal{C}_{k}(Y,\mathfrak{t}):=(a_{0}+L^{2}_{k-\frac{1}{2}}(i\Lambda^{1}_{Y}))\oplus L^{2}_{k-\frac{1}{2}}(S).

The Chern–Simons–Dirac functional CSD:𝒞k(Y,𝔱)CSD:\mathcal{C}_{k}(Y,\mathfrak{t})\to\mathbb{R} is deined by

CSD(a,ϕ):=12(Yada+Yϕ,a0+a+ϕdvolY),CSD(a,\phi):=\frac{1}{2}\left(-\int_{Y}a\wedge da+\int_{Y}\left<\phi,\cancel{\partial}^{+}_{a_{0}+a}\phi\right>\text{dvol}_{Y}\right),

where a0+a+\cancel{\partial}^{+}_{a_{0}+a} is the spinc\text{spin}^{c} Dirac operator with respect to the connection a0+aa_{0}+a.

The gauge group 𝒢k(Y)\mathcal{G}_{k}(Y) and a subgroup 𝒢~k(Y)\widetilde{\mathcal{G}}_{k}(Y) of 𝒢k(Y)\mathcal{G}_{k}(Y) are defined by

𝒢k(Y):=Lk+122(Y,S1)\mathcal{G}_{k}(Y):=L^{2}_{k+\frac{1}{2}}(Y,S^{1})

and

𝒢~k(Y):={g𝒢k(Y)|g=eif,YfvolY=0}.\widetilde{\mathcal{G}}_{k}(Y):=\Set{g\in\mathcal{G}_{k}(Y)}{g=e^{if},\ \int_{Y}f\text{vol}_{Y}=0}.

The gauge group 𝒢k(Y)\mathcal{G}_{k}(Y) naturally acts on 𝒞k(Y,𝔱)\mathcal{C}_{k}(Y,\mathfrak{t}) and the functional CSDCSD is invariant under the action. The global slice of the action of 𝒢~k(Y)\widetilde{\mathcal{G}}_{k}(Y) on 𝒞k(Y,𝔱)\mathcal{C}_{k}(Y,\mathfrak{t}) is given by

Vk(Y,𝔰)=(Kerd:Lk122(ΛY1)Lk322(ΛY0))Lk122(S),V_{k}(Y,\mathfrak{s})=(\mathop{\mathrm{Ker}}\nolimits d^{*}:L^{2}_{k-\frac{1}{2}}(\Lambda_{Y}^{1})\to L^{2}_{k-\frac{3}{2}}(\Lambda_{Y}^{0}))\oplus L^{2}_{k-\frac{1}{2}}(S),

on which we still have the remaining S1S^{1}-action. We often drop kk and/or (Y,𝔰)(Y,\mathfrak{s}) from our notation to denote Vk(Y,𝔰)V_{k}(Y,\mathfrak{s}). The S1S^{1}-equivariant formal gradient flow on V(Y,𝔰)V(Y,\mathfrak{s}) of CSDCSD with respect to the Coulomb projection of the L2L^{2}-metric can be written as the sum of the linear term

l=(d,a0):Vk(Y,𝔰)Vk1(Y,𝔰)l=(*d,\cancel{\partial}_{a_{0}}):V_{k}(Y,\mathfrak{s})\to V_{k-1}(Y,\mathfrak{s})

and some quadratic term, denoted by c:Vk(Y,𝔰)Vk1(Y,𝔰)c:V_{k}(Y,\mathfrak{s})\to V_{k-1}(Y,\mathfrak{s}).

For λ<0<μ\lambda<0<\mu, we define Vλμ(Y)V_{\lambda}^{\mu}(Y) as the direct sum of eigenspaces of ll, regarded as an unbounded operator on V1/2(Y,𝔰)V_{1/2}(Y,\mathfrak{s}), whose eigenvalues belong to (λ,μ](\lambda,\mu]. Here we think of Vλμ(Y)V_{\lambda}^{\mu}(Y) as a subspace of Vk(Y,𝔰)V_{k}(Y,\mathfrak{s}). We denote by

pλμ:Vk(Y,𝔰)Vλμ(Y)p_{\lambda}^{\mu}:V_{k}(Y,\mathfrak{s})\to V_{\lambda}^{\mu}(Y)

the L2L^{2}-projection of Vk(Y,𝔰)V_{k}(Y,\mathfrak{s}) onto Vλμ(Y)V_{\lambda}^{\mu}(Y). We often abbreviate Vλμ(Y)V_{\lambda}^{\mu}(Y) as VλμV_{\lambda}^{\mu}. Since ll is the sum of a real operator and a complex operator, VλμV_{\lambda}^{\mu} decomposes into a real vector space and a complex vector space, denoted by

(9) Vλμ=Vλμ()Vλμ().\displaystyle V_{\lambda}^{\mu}=V_{\lambda}^{\mu}(\mathbb{R})\oplus V_{\lambda}^{\mu}(\mathbb{C}).

Let us use basic terms of Conley index theory following [Ma03, Section 5]. Manolescu proved some compactness result [Ma03, Proposition 3], and as a consequence, it turns out that a closed ball in VλμV_{\lambda}^{\mu} of sufficiently large radius centered at the origin is an isolating neighborhood of the invariant part of the ball. Precisely, the flow on VλμV_{\lambda}^{\mu} considered here is a flow obtained from (l+pλμc)(l+p_{\lambda}^{\mu}c) by cutting off outside a larger ball (see [Ma03, page 907]). We denote by IλμI_{\lambda}^{\mu} the S1S^{1}-equivariant Conley index of the invariant part. The Seiberg–Witten Floer homotopy type SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}) is defined as the triple (ΣVλ0Iλμ,0,n(Y,𝔱,gY))(\Sigma^{-V^{0}_{\lambda}}I_{\lambda}^{\mu},0,n(Y,\mathfrak{t},g_{Y})), which is symbolically denoted by

SWF(Y,𝔱)=Σn(Y,𝔱,gY)Vλ0Iλμ.\mathrm{SWF}(Y,\mathfrak{t})=\Sigma^{-n(Y,\mathfrak{t},g_{Y})\mathbb{C}-V^{0}_{\lambda}}I_{\lambda}^{\mu}.

The triple is regarded as an object a certain suspension category \mathfrak{C}. In general an object of \mathfrak{C} is given as a triple (Z,m,n)(Z,m,n), where ZZ is a pointed topological S1S^{1}-space, mm\in\mathbb{Z}, and nn\in\mathbb{Q}. The quantity n(Y,𝔱,gY)n(Y,\mathfrak{t},g_{Y})\in\mathbb{Q} is defined to be

(10) n(Y,𝔱,gY):=indD++σ(W)8,\displaystyle n(Y,\mathfrak{t},g_{Y}):=\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}+\frac{\sigma(W)}{8},

where (W,𝔱)(W,\mathfrak{t}^{\prime}) is a compact spinc\text{spin}^{c} 4-manifold satisfying (W,𝔱)=(Y,𝔱)\partial(W,\mathfrak{t}^{\prime})=(Y,\mathfrak{t}) and indD+\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+} means the index of the Dirac operator with APS boundary condition. For the meaning of formal desuspensions, see [Ma03].

Here let us consider the case when the spinc structure 𝔱\mathfrak{t} comes from a spin structure. In this case, the formal gradient flow of CSDCSD admits a larger symmetry of the group Pin(2)\operatorname{Pin}(2) defined by

Pin(2):=S1jS1Sp(1).\operatorname{Pin}(2):=S^{1}\cup jS^{1}\subset Sp(1).

This group Pin(2)\operatorname{Pin}(2) acts on Vk(Y,𝔰)V_{k}(Y,\mathfrak{s}) for any non-negative integer kk as follows: the Pin(2)\operatorname{Pin}(2)-action on spinors given as the restriction of the natural Sp(1)Sp(1)-action on spinor bundles, and the Pin(2)\operatorname{Pin}(2)-action on ΩY1\Omega^{1}_{Y} is given via the non-trivial homomorphism Pin(2)O(1)\operatorname{Pin}(2)\to O(1). We denote by ~\widetilde{\mathbb{R}} the real 11-dimensional representation of Pin(2)\operatorname{Pin}(2), and by \mathbb{H} the space of quaternions, on which Pin(2)\operatorname{Pin}(2) naturally acts. Thus we have decompositions

Vk(Y,𝔰)=V()V()V_{k}(Y,\mathfrak{s})=V({\mathbb{R}})\oplus V(\mathbb{H})

and

(11) Vλμ=Vλμ()Vλμ().\displaystyle V_{\lambda}^{\mu}=V_{\lambda}^{\mu}(\mathbb{R})\oplus V_{\lambda}^{\mu}(\mathbb{H}).

Considering Pin(2)\operatorname{Pin}(2)-equivariant Conley index instead, we obtain a stable homotopy type of a pointed Pin(2)\operatorname{Pin}(2)-space

SWF(Y,𝔱)=Σn(Y,𝔱,g)2Vλ0Iλμ,\mathrm{SWF}(Y,\mathfrak{t})=\Sigma^{-\frac{n(Y,\mathfrak{t},g)}{2}\mathbb{H}-V^{0}_{\lambda}}I^{\mu}_{\lambda},

which lies in a suspension category \mathfrak{C}^{\prime}. An object of \mathfrak{C}^{\prime} is given as a triple (Z,m,n)(Z,m,n), where ZZ is a pointed topological Pin(2)\operatorname{Pin}(2)-space, mm\in\mathbb{Z}, and nn\in\mathbb{Q}.

Let us recall the definition of local equivalence.

Definition 2.8 ([Sto20]).

For two objects (Z1,m1,n1)(Z_{1},m_{1},n_{1}) and (Z2,m2,n2)(Z_{2},m_{2},n_{2}) in \mathfrak{C}^{\prime}, a local map is a Pin(2)\operatorname{Pin}(2)-equivariant map

f:Σ(Nn1)Σ(Mm1)~Z1Σ(Nn2)Σ(Mm2)~Z2f:\Sigma^{(N-n_{1})\mathbb{H}}\Sigma^{(M-m_{1})\widetilde{\mathbb{R}}}Z_{1}\to\Sigma^{(N-n_{2})\mathbb{H}}\Sigma^{(M-m_{2})\widetilde{\mathbb{R}}}Z_{2}

for some MM\in\mathbb{Z} and NN\in\mathbb{Q} such that the S1S^{1}-invariant part fS1f^{S^{1}} is a Pin(2)\operatorname{Pin}(2)-homotopy equivalence. Two objects (Z1,m1,n1)(Z_{1},m_{1},n_{1}) and (Z2,m2,n2)(Z_{2},m_{2},n_{2}) are locally equivalent if there exist local maps f:(Z1,m1,n1)(Z2,m2,n2)f:(Z_{1},m_{1},n_{1})\to(Z_{2},m_{2},n_{2}) and g:(Z2,m2,n2)(Z1,m1,n1)g:(Z_{2},m_{2},n_{2})\to(Z_{1},m_{1},n_{1}).

Typical examples of local maps are obtained as the relative Bauer–Furuta invariants for negative definite spin cobordisms between rational homology 3-spheres.

2.4. The Seiberg–Witten equations on W[,0]W[-\infty,0]

In this subsection we describe the Seiberg–Witten equations on W[,0]W[-\infty,0], mainly to fix notations. We use the double Cloumb gauge condition introduced in [Kha15].

Definition 2.9.

Let kk be a positive integer with k4k\geq 4 and δ\delta a positive real number. We first define the configuration space 𝒞k,δ(W[,0])\mathcal{C}_{k,\delta}(W[-\infty,0]) by

𝒞k,δ(W[,0]):=(A0,0)+Lk,δ2(iΛW[,0]1)Lk,δ2(SW[,0]+).\mathcal{C}_{k,\delta}(W[-\infty,0]):=(A_{0},0)+L^{2}_{k,\delta}(i\Lambda_{W[-\infty,0]}^{1})\oplus L^{2}_{k,\delta}(S^{+}_{W[-\infty,0]}).

The gauge group 𝒢k+1,δ(W[,0])\mathcal{G}_{k+1,\delta}(W[-\infty,0]) is given by

(12) 𝒢k+1,δ(W[,0]):={u:W[,0]||u(x)|=1(xW[,0]),1uLk+1,δ2(¯)}.\displaystyle\mathcal{G}_{k+1,\delta}(W[-\infty,0]):=\left\{u:W[-\infty,0]\to\mathbb{C}~{}\middle|~{}|u(x)|=1~{}(\forall x\in W[-\infty,0]),~{}1-u\in L^{2}_{k+1,\delta}(\underline{\mathbb{C}})\right\}.

Here ¯\underline{\mathbb{C}} denotes the trivial bundle over W[,0]W[-\infty,0] with fiber \mathbb{C}. The action of 𝒢k+1,δ(W[,0])\mathcal{G}_{k+1,\delta}(W[-\infty,0]) on 𝒞k,δ(W[,0])\mathcal{C}_{k,\delta}(W[-\infty,0]) is given by

u(A,Φ):=(Au1du,uΦ).u\cdot(A,\Phi):=(A-u^{-1}du,u\Phi).

The double Coulomb slice introduced in [Kha15] is defined by

𝒰k,δ(W[,0]):=Lk,δ2(iΛW[,0]1)CCLk,δ2(SW[,0]+),\mathcal{U}_{k,\delta}(W[-\infty,0]):=L^{2}_{k,\delta}(i\Lambda_{W[-\infty,0]}^{1})_{CC}\oplus L^{2}_{k,\delta}(S^{+}_{W[-\infty,0]}),

where

Lk,δ2(iΛW[,0]1)CC:={aLk,δ2(iΛW[,0]1)|dδa=0,d𝐭a=0}.L^{2}_{k,\delta}(i\Lambda_{W[-\infty,0]}^{1})_{CC}:=\Set{a\in L^{2}_{k,\delta}(i\Lambda_{W[-\infty,0]}^{1})}{d^{*_{\delta}}a=0,d^{*}{\bf t}a=0}.

Here 𝐭{\bf t} denotes the restriction of 11-forms as differential forms and dδd^{*_{\delta}} is the formal adjoint of dd with respect to Lδ2L^{2}_{\delta}.

We will prove that 𝒰k,δ(W[,0])\mathcal{U}_{k,\delta}(W[-\infty,0]) gives a global slice with respect to the action of 𝒢k+1,δ(W[,0])\mathcal{G}_{k+1,\delta}(W[-\infty,0]) on 𝒞k,δ(W[,0])\mathcal{C}_{k,\delta}(W[-\infty,0]). Note that, on 𝒞k,δ(W[,0])\mathcal{C}_{k,\delta}(W[-\infty,0]), the ‘full gauge group’

{u:W[,0]||u(x)|=1(xW[,0]),duLk,δ2(¯)}\left\{u:W[-\infty,0]\to\mathbb{C}~{}\middle|~{}|u(x)|=1~{}(\forall x\in W[-\infty,0]),~{}du\in L^{2}_{k,\delta}(\underline{\mathbb{C}})\right\}

also acts. Thus we have an additional S1S^{1}-symmetry on 𝒰k,δ(W[,0])\mathcal{U}_{k,\delta}(W[-\infty,0]) coming from the limits with respect to the end of gauge transformations.

Based on the Sobolev embedding 𝒢k+1,δ(W[,0])C0(W[,0],S1)\mathcal{G}_{k+1,\delta}(W[-\infty,0])\to C^{0}(W[-\infty,0],S^{1}), we can naturally define the group structure on 𝒢k+1,δ(W[,0])\mathcal{G}_{k+1,\delta}(W[-\infty,0]) by pointwise multiplication.

On W[,0]W[-\infty,0], one can define the Seiberg–Witten map

(13) W[,0]:𝒞k,δ(W[,0])Lk1,δ2(iΛW[,0]+SW[,0])\displaystyle\mathcal{F}_{W[-\infty,0]}:\mathcal{C}_{k,\delta}(W[-\infty,0])\to L^{2}_{k-1,\delta}(i\Lambda_{W[-\infty,0]}^{+}\oplus S^{-}_{W[-\infty,0]})

by

(14) W[,0](A,Φ):=(12FAt+ρ1(ΦΦ)0,DA+Φ).\displaystyle\mathcal{F}_{W[-\infty,0]}(A,\Phi):=\left(\frac{1}{2}F^{+}_{A^{t}}-\rho^{-1}(\Phi\Phi^{*})_{0},D^{+}_{A}\Phi\right).

When we write (a,ϕ)=(A,Φ)(A0,0)(a,\phi)=(A,\Phi)-(A_{0},0), we often decompose the Seiberg–Witten map W[,0]\mathcal{F}_{W[-\infty,0]} as the sum of the linear part

(15) LW[,0](a,ϕ):=(d+a,DA0+ϕ),\displaystyle L_{W[-\infty,0]}(a,\phi):=\left(d^{+}a,D^{+}_{A_{0}}\phi\right),

the quadratic part

CW[,0](a,ϕ):=((ϕϕ)0,ρ(a)ϕ).C_{W[-\infty,0]}(a,\phi):=(-(\phi\phi^{*})_{0},\rho(a)\phi).

We regard LW[,0]L_{W[-\infty,0]} also as an operator with domain 𝒰k,δ(W[,0])\mathcal{U}_{k,\delta}(W[-\infty,0]) by the restriction. The quadratic part is a compact operator by [Lin19, Proposition 2.13] for a positive δ\delta. The differential equation

(16) W[,0](A,Φ)=0\displaystyle\mathcal{F}_{W[-\infty,0]}(A,\Phi)=0

is called the Seiberg–Witten equation for W[,0]W[-\infty,0]. The linearlization of W[,0]\mathcal{F}_{W[-\infty,0]} is given by LW[,0]L_{W[-\infty,0]}.

3. Linear analysis on W[,0]W[-\infty,0]

Fix a Riemann metric gW[,0]g_{W[-\infty,0]} on W[,0]W[-\infty,0] such that

  • gW[,0]|W[,1]g_{W[-\infty,0]}|_{W[-\infty,-1]} is periodic and PSC, and

  • gW[,0]g_{W[-\infty,0]} is product metric near W[,0]=Y\partial W[-\infty,0]=Y.

3.1. Fredholm theory on W[,0]W[-\infty,0]

In this subsection, we prove certain Fredholm properties which will be used in the proof of Theorem 1.1. For a fixed periodic spin structure on W[,0]W[-\infty,0], the spinor bundles are written as S+S^{+} and SS^{-}. In this section, we use the following completions:

Lk,δ2(iΛW[,0]1),Lk,δ2(iΛW[,0]+), and Lk,δ2(S±).L^{2}_{k,\delta}(i\Lambda^{1}_{W[-\infty,0]}),\ L^{2}_{k,\delta}(i\Lambda^{+}_{W[-\infty,0]}),\text{ and }L^{2}_{k,\delta}(S^{\pm}).

We prove the Fredholm properties of the following two types of operators on W[,0]W[-\infty,0]:

  • the Atiyah–Hitchin–Singer operator with APS-boundary condition:

    (17) dδ+d++p^0r^:Lk,δ2(iΛW[,0]1)Lk1,δ2(iΛW[,0]0ΛW[,0]+)V^0(Y;),\displaystyle\begin{split}d^{*_{\delta}}+d^{+}+\widehat{p}^{0}_{-\infty}\circ\widehat{r}:L^{2}_{k,\delta}(i\Lambda_{W[-\infty,0]}^{1})\\ \to L^{2}_{k-1,\delta}(i\Lambda_{W[-\infty,0]}^{0}\oplus\Lambda_{W[-\infty,0]}^{+})\oplus\widehat{V}^{0}_{-\infty}(Y;\mathbb{R}),\end{split}

    where

    • (i)

      the space V^0(Y;)\widehat{V}^{0}_{-\infty}(Y;\mathbb{R}) is the Lk122L^{2}_{k-\frac{1}{2}}-completion of the negative eigenspaces of the operator

      l^:=(0ddd):ΩY0ΩY1ΩY0ΩY1,\widehat{l}:=\begin{pmatrix}0&-d^{*}\\ -d&*d\\ \end{pmatrix}:\Omega^{0}_{Y}\oplus\Omega^{1}_{Y}\to\Omega^{0}_{Y}\oplus\Omega^{1}_{Y},
    • (ii)

      the map r^:Lk,δ2(iΛW[,0]1)Lk122(ΛY0ΛY1)\widehat{r}:L^{2}_{k,\delta}(i\Lambda_{W[-\infty,0]}^{1})\to L^{2}_{k-\frac{1}{2}}(\Lambda^{0}_{Y}\oplus\Lambda^{1}_{Y}) is the restriction,

    • (iii)

      the operator

      p^0:Lk122(ΛY0ΛY1)V^0(Y;)\widehat{p}^{0}_{-\infty}:L^{2}_{k-\frac{1}{2}}(\Lambda^{0}_{Y}\oplus\Lambda^{1}_{Y})\to\widehat{V}^{0}_{-\infty}(Y;\mathbb{R})

      is the L2L^{2}-projection to V^0(Y;)\widehat{V}^{0}_{-\infty}(Y;\mathbb{R}).

  • the Dirac operator with APS-boundary condition:

    (18) DA0++p^0r^:Lk,δ2(SW[,0]+)Lk1,δ2(SW[,0])V^0(Y,),\displaystyle D^{+}_{A_{0}}+\widehat{p}^{0}_{-\infty}\circ\widehat{r}:L^{2}_{k,\delta}(S^{+}_{W[-\infty,0]})\to L^{2}_{k-1,\delta}(S^{-}_{W[-\infty,0]})\oplus\widehat{V}^{0}_{-\infty}(Y,\mathbb{C}),

    where

    • (i)

      the space V^0(Y,)\widehat{V}^{0}_{-\infty}(Y,\mathbb{C}) is the Lk122L^{2}_{k-\frac{1}{2}}-completion of the negative eigenspaces of the operator

      B0:Γ(S)Γ(S).\cancel{\partial}_{B_{0}}:\Gamma(S)\to\Gamma(S).
    • (ii)

      the map r^:Lk,δ2(SW[,0]+)Lk122(S)\widehat{r}:L^{2}_{k,\delta}(S^{+}_{W[-\infty,0]})\to L^{2}_{k-\frac{1}{2}}(S) is the restriction,

    • (iii)

      the operator

      p^0:Lk122(S)V^0(Y,)\widehat{p}^{0}_{-\infty}:L^{2}_{k-\frac{1}{2}}(S)\to\widehat{V}^{0}_{-\infty}(Y,\mathbb{C})

      is the L2L^{2}-projection to V^0(Y)\widehat{V}^{0}_{-\infty}(Y).

We first prove the following proposition:

Proposition 3.1.

The following facts hold:

  • (i)

    For any δ\delta\in\mathbb{R}, the operator (18) is Fredholm.

  • (ii)

    Let δ0\delta_{0} be a positive real number given in Remark 2.6. For any δ(0,δ0)\delta\in(0,\delta_{0}), the operator (17) is Fredholm.

Proof.

Both statements follow from the standard patching argument of parametrixes of these operators.

  • First, we prove (i). By Lemma 2.5, since positive scalar curvature metrics are admissible, we see that the Dirac operator

    DA0+:Lk,δ2(X~;S+)Lk1,δ2(X~;S)D^{+}_{A_{0}}:L^{2}_{k,\delta}(\widetilde{X};S^{+})\to L^{2}_{k-1,\delta}(\widetilde{X};S^{-})

    is an isomorphism for any δ\delta\in\mathbb{R}, and we get a continuous inverse Pδ:Lk1,δ2(X~;S)Lk,δ2(X~;S+)P_{\delta}:L^{2}_{k-1,\delta}(\widetilde{X};S^{-})\to L^{2}_{k,\delta}(\widetilde{X};S^{+}). By patching a local parametrix of (18) near the boundary YY and PδP_{\delta}, we obtain a parametrix of (18). This implies the conclusion.

  • Next, we prove (ii). By Lemma 2.5,

    0Lk+1,δ2(iΛ0(X~))𝑑Lk,δ2(iΛ1(X~))d+Lk1,δ2(iΛ+(X~))00\to L^{2}_{k+1,\delta}(i\Lambda^{0}(\widetilde{X}))\xrightarrow{d}L^{2}_{k,\delta}(i\Lambda^{1}(\widetilde{X}))\xrightarrow{d^{+}}L^{2}_{k-1,\delta}(i\Lambda^{+}(\widetilde{X}))\to 0

    is an acyclic complex for δ𝒟\delta\in\mathbb{R}\setminus\mathcal{D}, where 𝒟\mathcal{D} is a discrete subset of \mathbb{R} given in Theorem 2.3. This implies that

    d++dδ:Lk,δ2(iΛ1(X~))Lk1,δ2(iΛ+(X~))Lk1,δ2(iΛ0(X~))d^{+}+d^{*_{\delta}}:L^{2}_{k,\delta}(i\Lambda^{1}(\widetilde{X}))\to L^{2}_{k-1,\delta}(i\Lambda^{+}(\widetilde{X}))\oplus L^{2}_{k-1,\delta}(i\Lambda^{0}(\widetilde{X}))

    is an isomorphism for δ𝒟\delta\in\mathbb{R}\setminus\mathcal{D}. Since 𝒟\mathcal{D} does not have accumulation points, there exists a small positive real number δ0\delta_{0} such that

    (0,δ0)𝒟=.(0,\delta_{0})\cap\mathcal{D}=\emptyset.

    Then the remaining part is the same as the proof of (i).

Set

W(Y):=ib0(Y)dLk1/22(iΛY0)W(Y):=i\mathbb{R}^{b_{0}(Y)}\oplus dL^{2}_{k-1/2}(i\Lambda^{0}_{Y})

and consider the operators

LW[,0](p0r):𝒰k,δLk1,δ2(iΛ+S)V0,\displaystyle L_{W[-\infty,0]}\oplus(p^{0}_{-\infty}\circ r):\mathcal{U}_{k,\delta}\to L^{2}_{k-1,\delta}(i\Lambda^{+}\oplus S^{-})\oplus V^{0}_{-\infty},
L^W[,0](p^0r^):Lk,δ2(iΛ1S+)Lk1,δ2(iΛ0iΛ+S)V^0\displaystyle\widehat{L}_{W[-\infty,0]}\oplus(\widehat{p}^{0}_{-\infty}\circ\hat{r}):L^{2}_{k,\delta}(i\Lambda^{1}\oplus S^{+})\to L^{2}_{k-1,\delta}(i\Lambda^{0}\oplus i\Lambda^{+}\oplus S^{-})\oplus\widehat{V}^{0}_{-\infty}

over W[,0]W[-\infty,0]. Here LW[,0]L_{W[-\infty,0]} is defined in (15), and L^W[,0]\widehat{L}_{W[-\infty,0]} is defined by

L^W[,0](a,ϕ):=(dδa,d+a,DA0+ϕ).\widehat{L}_{W[-\infty,0]}(a,\phi):=(d^{\ast_{\delta}}a,d^{+}a,D^{+}_{A_{0}}\phi).

It follows from Proposition 3.1 that the operator L^W[,0](p^0r^)\widehat{L}_{W[-\infty,0]}\oplus(\widehat{p}^{0}_{-\infty}\circ\hat{r}) is Fredholm for all δ(0,δ0)\delta\in(0,\delta_{0}).

Proposition 3.2.

Let δ0\delta_{0} be the positive real number given in Remark 2.6. For any δ(0,δ0)\delta\in(0,\delta_{0}), we obtain

{Ker(LW[,0](p0r))Ker(L^W[,0](p^0r^)),Coker(LW[,0](p0r))Coker(L^W[,0](p^0r^)),\begin{cases}\mathop{\mathrm{Ker}}\nolimits(L_{W[-\infty,0]}\oplus(p^{0}_{-\infty}\circ r))\cong\mathop{\mathrm{Ker}}\nolimits(\widehat{L}_{W[-\infty,0]}\oplus(\widehat{p}^{0}_{-\infty}\circ\hat{r})),\\ \operatorname{Coker}(L_{W[-\infty,0]}\oplus(p^{0}_{-\infty}\circ r))\cong\operatorname{Coker}(\widehat{L}_{W[-\infty,0]}\oplus(\widehat{p}^{0}_{-\infty}\circ\hat{r})),\end{cases}

where Coker\mathop{\mathrm{Coker}}\nolimits denotes the algebraic cokernel. In particular, LW[,0](p0r)L_{W[-\infty,0]}\oplus(p^{0}_{-\infty}\circ r) is Fredholm and the index of LW[,0](p0r)L_{W[-\infty,0]}\oplus(p^{0}_{-\infty}\circ r) coincides with that of L^W[,0](p^0r^)\widehat{L}_{W[-\infty,0]}\oplus(\widehat{p}^{0}_{-\infty}\circ\hat{r}).

Proof.

The proof is essentially the same as the proof in [Khan15]. First, by the choice of δ\delta, Proposition 3.1 implies that dδ+d++p^0r^d^{*_{\delta}}+d^{+}+\widehat{p}^{0}_{-\infty}\circ\widehat{r} is Fredholm. Set

V^(Y)=V^=iΩ0(Y)iΩ1(Y),\widehat{V}(Y)=\widehat{V}=i\Omega^{0}(Y)\oplus i\Omega^{1}(Y),

and let

ϖ:V^(Y)W(Y)\varpi:\widehat{V}(Y)\to W(Y)

be the L2L^{2}-orthogonal projection, and consider an operator

(19) L^W[,0]((p0ϖ)r^):Lk,δ2(iΛW[,0]1S+)Lk1,δ2(iΛW[,0]0iΛW[,0]+S)V0(Y;)W(Y)\displaystyle\begin{split}\widehat{L}_{W[-\infty,0]}\oplus((p^{0}_{-\infty}\oplus\varpi)\circ\hat{r}):L^{2}_{k,\delta}(i\Lambda^{1}_{W[-\infty,0]}\oplus S^{+})\\ \to L^{2}_{k-1,\delta}(i\Lambda^{0}_{W[-\infty,0]}\oplus i\Lambda^{+}_{W[-\infty,0]}\oplus S^{-})\oplus V^{0}_{-\infty}(Y;\mathbb{R})\oplus W(Y)\end{split}

as an intermediary between the two operators in the statement of the Proposition.

We first show that

(20) {Ker(L^W[,0]((p0ϖ)r^))Ker(L^(p^0r^))Coker(L^W[,0]((p0ϖ)r^))Coker(L^(p^0r^)).\displaystyle\begin{cases}\mathop{\mathrm{Ker}}\nolimits(\widehat{L}_{W[-\infty,0]}\oplus((p^{0}_{-\infty}\oplus\varpi)\circ\hat{r}))\cong\mathop{\mathrm{Ker}}\nolimits(\widehat{L}\oplus(\widehat{p}^{0}_{-\infty}\circ\widehat{r}))\\ \operatorname{Coker}(\widehat{L}_{W[-\infty,0]}\oplus((p^{0}_{-\infty}\oplus\varpi)\circ\hat{r}))\cong\operatorname{Coker}(\widehat{L}\oplus(\widehat{p}^{0}_{-\infty}\circ\widehat{r})).\end{cases}

Set

V=V(Y)=iΩ0(Y)idΩ0(Y),V^{\perp}=V^{\perp}(Y)=i\Omega^{0}(Y)\oplus id\Omega^{0}(Y),

and let

l:VVl^{\perp}:V^{\perp}\to V^{\perp}

be the operator defined by

l=[0dd0].l^{\perp}=\begin{bmatrix}0&-d^{*}\\ -d&0\end{bmatrix}.

We denote the Lk1/22L^{2}_{k-1/2}-completion of ll^{\perp} by the same notation. Then we have

V^=VV\widehat{V}=V\oplus V^{\perp}

and

l^=ll.\widehat{l}=l\oplus l^{\perp}.

Let (V)0(V^{\perp})^{0}_{-\infty} be the span of non-positive eigenvectors of ll^{\perp}. As shown in [Khan15], the projection ϖ:(V)0W(Y)\varpi:(V^{\perp})^{0}_{-\infty}\to W(Y) is an isomorphism, and hence so is

idV0ϖ:V^0=V0(V)0V0W(Y).id_{V^{0}_{-\infty}}\oplus\varpi:\widehat{V}^{0}_{-\infty}=V^{0}_{-\infty}\oplus(V^{\perp})^{0}_{-\infty}\to V^{0}_{-\infty}\oplus W(Y).

Thus we obtain the following commutative diagram between functional spaces over W[,0]W[-\infty,0]:

Lk,δ2(iΛ1S+)L^W[,0]p^0r^Lk1,δ2(iΛ0iΛ+S)V^0idϖLk,δ2(iΛ1S+)L^W[,0]((p0ϖ)r^)Lk1,δ2(iΛ0iΛ+S)V0W(Y).\begin{CD}L^{2}_{k,\delta}(i\Lambda^{1}\oplus S^{+})@>{\widehat{L}_{W[-\infty,0]}\oplus\widehat{p}^{0}_{-\infty}\circ\widehat{r}}>{}>L^{2}_{k-1,\delta}(i\Lambda^{0}\oplus i\Lambda^{+}\oplus S^{-})\oplus\widehat{V}^{0}_{-\infty}\\ \Big{\|}@V{id\oplus\varpi}V{\cong}V\\ L^{2}_{k,\delta}(i\Lambda^{1}\oplus S^{+})@>{\widehat{L}_{W[-\infty,0]}\oplus((p^{0}_{-\infty}\oplus\varpi)\circ\hat{r})}>{}>L^{2}_{k-1,\delta}(i\Lambda^{0}\oplus i\Lambda^{+}\oplus S^{-})\oplus V^{0}_{-\infty}\oplus W(Y).\end{CD}

From this diagram we obtain the isomorphisms (20). Moreover, as noted, it follows from Proposition 3.1 that the operator L^W[,0](p^0r^)\widehat{L}_{W[-\infty,0]}\oplus(\widehat{p}^{0}_{-\infty}\circ\hat{r}) is Fredholm. Therefore this diagram implies that L^W[,0]((p0ϖ)r^)\widehat{L}_{W[-\infty,0]}\oplus((p^{0}_{-\infty}\oplus\varpi)\circ\hat{r}) is also Fredholm.

The remaining task is to show that

(21) {Ker(LW[,0](p0r))Ker(L^W[,0]((p0ϖ)r^))Coker(LW[,0](p0r))Coker(L^W[,0]((p0ϖ)r^)).\displaystyle\begin{cases}\mathop{\mathrm{Ker}}\nolimits(L_{W[-\infty,0]}\oplus(p^{0}_{-\infty}\circ r))\cong\mathop{\mathrm{Ker}}\nolimits(\widehat{L}_{W[-\infty,0]}\oplus((p^{0}_{-\infty}\oplus\varpi)\circ\hat{r}))\\ \operatorname{Coker}(L_{W[-\infty,0]}\oplus(p^{0}_{-\infty}\circ r))\cong\operatorname{Coker}(\widehat{L}_{W[-\infty,0]}\oplus((p^{0}_{-\infty}\oplus\varpi)\circ\hat{r})).\end{cases}

The assertion of the Proposition immediately follows from this and (20). But applying the snake lemma to the following commutative diagram between functional spaces over W[,0]W[-\infty,0], we can obtain (21):

00Lk,δ2(iΛ1S+)CCLW[,0]p0rLk1,δ2(iΛ+S)V0Lk,δ2(iΛ1S+)L^W[,0]((p0ϖ)r^)Lk1,δ2(iΛ0iΛ+S)V0W(Y)dδϖr^Lk1,δ2(iΛ0)W(Y)=Lk1,δ2(iΛ0)W(Y)00.\begin{CD}00\\ @V{}V{}V@V{}V{}V\\ L^{2}_{k,\delta}(i\Lambda^{1}\oplus S^{+})_{CC}@>{L_{W[-\infty,0]}\oplus p^{0}_{-\infty}\circ r}>{}>L^{2}_{k-1,\delta}(i\Lambda^{+}\oplus S^{-})\oplus V^{0}_{-\infty}\\ @V{{}}V{}V@V{}V{}V\\ L^{2}_{k,\delta}(i\Lambda^{1}\oplus S^{+})@>{\widehat{L}_{W[-\infty,0]}\oplus((p^{0}_{-\infty}\oplus\varpi)\circ\hat{r})}>{}>L^{2}_{k-1,\delta}(i\Lambda^{0}\oplus i\Lambda^{+}\oplus S^{-})\oplus V^{0}_{-\infty}\oplus W(Y)\\ @V{{d^{*_{\delta}}\oplus\varpi\circ\widehat{r}}}V{}V@V{}V{}V\\ L^{2}_{k-1,\delta}(i\Lambda^{0})\oplus W(Y)@=L^{2}_{k-1,\delta}(i\Lambda^{0})\oplus W(Y)\\ @V{}V{}V@V{}V{}V\\ 00.\\ \end{CD}

We consider a Riemannian manifold

W^[,0]:=W[,0](0×Y)\hat{W}[-\infty,0]:=W[-\infty,0]\cup(\mathbb{R}^{\geq 0}\times Y)

obtained by gluing the half-cylinder (0×Y,dt2+gY)(\mathbb{R}^{\geq 0}\times Y,dt^{2}+g_{Y}) with W[,0]W[-\infty,0] along their boundary. We will compare formal adjoints dd^{*} for several weights, and would like to introduce a family of weight functions

τδ,δ:W^[,0]0\tau_{\delta,\delta^{\prime}}:\hat{W}[-\infty,0]\to\mathbb{R}_{\geq 0}

such that

(τδ,δ)|W[,1]=δτ and (τδ,δ)|[1,)×Y=δt.(\tau_{\delta,\delta^{\prime}})|_{{W}[-\infty,-1]}=\delta\tau\text{\quad and\quad}(\tau_{\delta,\delta^{\prime}})|_{[1,\infty)\times Y}=\delta^{\prime}t.
Definition 3.3.

Let (δ,δ)2(\delta,\delta^{\prime})\in\mathbb{R}^{2}. For a bundle EE which is periodic on W[,0]{W}[-\infty,0] and cylindrical on [0,)×Y[0,\infty)\times Y, we define the norm Lk,(δ,δ)2(E)\|-\|_{L^{2}_{k,(\delta,\delta^{\prime})}(E)} by

fLk,(δ,δ)2(E):=eτδ,δfLk2(E)\|f\|_{L^{2}_{k,(\delta,\delta^{\prime})}(E)}:=\|e^{\tau_{\delta,\delta^{\prime}}}f\|_{L^{2}_{k}(E)}

and define Lk,(δ,δ)2(E)L^{2}_{k,(\delta,\delta^{\prime})}(E) to be the completion of compactly supported sections with respect to Lk,(δ,δ)2(E)\|-\|_{L^{2}_{k,(\delta,\delta^{\prime})}(E)}.

Note that the formal adjoint with respect to L(δ,δ)2L^{2}_{(\delta,\delta^{\prime})} of dd is given as

d(δ,δ)(w)=eτδ,δd(eτδ,δw).d^{*_{(\delta,\delta^{\prime})}}(w)=e^{-\tau_{\delta,\delta^{\prime}}}d^{*}(e^{\tau_{\delta,\delta^{\prime}}}w).

We also consider the ‘sliced’ Atiyah–Hitchin–Singer operator with APS-boundary condition:

(22) d++p0r:Lk,δ2(iΛW[,0]1)CCLk1,δ2(iΛW[,0]+)V0(Y;).\displaystyle\begin{split}d^{+}+{p}^{0}_{-\infty}\circ{r}:L^{2}_{k,\delta}(i\Lambda_{W[-\infty,0]}^{1})_{CC}\to L^{2}_{k-1,\delta}(i\Lambda_{W[-\infty,0]}^{+})\oplus{V}^{0}_{-\infty}(Y;\mathbb{R}).\end{split}

We calculate the kernel and the cokernel of (22) :

Theorem 3.4.

There exists δ1>0\delta_{1}>0 such that for any δ(0,δ1)\delta\in(0,\delta_{1}), the operator (22) is an isomorphism.

We take the constant δ1\delta_{1} to be smaller than δ0\delta_{0} given in Remark 2.6. The rest of this subsection is devoted to prove Theorem 3.4.

To prove Theorem 3.4, it is sufficient to prove the operator (17) is invertible for a sufficiently small δ>0\delta>0. First we shall calculate the kernel and the cokernel of

(23) dδ+d++p^0r^:Lk,δ2(iΛW[,0]1)Lk1,δ2(iΛW[,0]0ΛW[,0]+)V^0(Y;).\displaystyle\begin{split}d^{*_{\delta}}+d^{+}+\widehat{p}^{0}_{-\infty}\circ\widehat{r}:L^{2}_{k,\delta}(i\Lambda_{W[-\infty,0]}^{1})\\ \to L^{2}_{k-1,\delta}(i\Lambda_{W[-\infty,0]}^{0}\oplus\Lambda_{W[-\infty,0]}^{+})\oplus\widehat{V}^{0}_{-\infty}(Y;\mathbb{R}).\end{split}

The following lemma can be proved by considering the similar discussion given in [APSI].

Lemma 3.5.

We have the following identifications:

{Ker(dδ+d++p^0r^)={aLk,(δ,0)2(iΛW^[,0]1)|d(δ,0)a=0,d+a=0},Coker(dδ+d++p^0r^)={(0,b)Lk1,(δ,0)2(iΛW^[,0]0ΛW^[,0]+)|d(δ,0)b=0}.\begin{cases}\mathop{\mathrm{Ker}}\nolimits(d^{*_{\delta}}+d^{+}+\widehat{p}^{0}_{-\infty}\circ\widehat{r})=\Set{a\in L^{2}_{k,(\delta,0)}(i\Lambda_{\hat{W}[-\infty,0]}^{1})}{d^{*_{(\delta,0)}}a=0,d^{+}a=0},\\ \operatorname{Coker}(d^{*_{\delta}}+d^{+}+\widehat{p}^{0}_{-\infty}\circ\widehat{r})\\ =\Set{(0,b)\in L^{2}_{k-1,(\delta,0)}(i\Lambda_{\hat{W}[-\infty,0]}^{0}\oplus\Lambda_{\hat{W}[-\infty,0]}^{+})}{d^{*_{(\delta,0)}}b=0}.\end{cases}
Proof.

By the same discussion in [APSI, Proposition 3.11], a solution under the spectral boundary condition can be identified with an L2L^{2}-solution on a cylindrical end manifold. Thus one has an isomorphism

Ker(dδ+d++p^0r^){aLk,(δ,0)2(iΛW^[,0]1)|d(δ,0)a=0,d+a=0},\mathop{\mathrm{Ker}}\nolimits(d^{*_{\delta}}+d^{+}+\widehat{p}^{0}_{-\infty}\circ\widehat{r})\cong\Set{a\in L^{2}_{k,(\delta,0)}(i\Lambda_{\hat{W}[-\infty,0]}^{1})}{d^{*_{(\delta,0)}}a=0,d^{+}a=0},

and the cokernel can be written by using extended Lk2L^{2}_{k}-solutions:

Coker(dδ+d++p^0r^)\displaystyle\operatorname{Coker}(d^{*_{\delta}}+d^{+}+\widehat{p}^{0}_{-\infty}\circ\widehat{r})
\displaystyle\cong {(a,b)|d(δ,0)b=0,da=0,(a,b)|W[,0]Lk1,δ2(iΛW[,0]0ΛW[,0]+),(ac,b)Lk12(iΛ[0,)×Y0Λ[0,)×Y+),c}\displaystyle\Set{(a,b)}{\begin{matrix}&d^{*_{(\delta,0)}}b=0,\ da=0,\\ &(a,b)|_{W[-\infty,0]}\in L^{2}_{k-1,\delta}(i\Lambda_{W[-\infty,0]}^{0}\oplus\Lambda_{W[-\infty,0]}^{+}),\\ &(a-c,b)\in L^{2}_{k-1}(i\Lambda_{[0,\infty)\times Y}^{0}\oplus\Lambda_{[0,\infty)\times Y}^{+}),\ \exists c\in\mathbb{R}\end{matrix}}
\displaystyle\subset Lk1,loc2(iΛW^[,0]0ΛW^[,0]+).\displaystyle L^{2}_{k-1,\text{loc}}(i\Lambda_{\hat{W}[-\infty,0]}^{0}\oplus\Lambda_{\hat{W}[-\infty,0]}^{+}).

Here we used H(Y;)H(S3;)H^{*}(Y;\mathbb{R})\cong H^{*}(S^{3};\mathbb{R}). On the other hand, da=0da=0 implies aa is a constant and cc should be zero. This gives the conclusion. ∎

Lemma 3.6.

For δ\delta sufficiently small, the space

{aLk,(δ,0)2(iΛW^[,0]1)|d(δ,0)a=0,da=0}\Set{a\in L^{2}_{k,(\delta,0)}(i\Lambda_{\hat{W}[-\infty,0]}^{1})}{d^{*_{(\delta,0)}}a=0,da=0}

can be identified with the middle cohomology of

Lk+1,(δ,δ)2(iΛW^[,0]0)𝑑Lk,(δ,δ)2(iΛW^[,0]1)𝑑Lk1,(δ,δ)2(iΛW^[,0]2).L^{2}_{k+1,(\delta,\delta)}(i\Lambda_{\hat{W}[-\infty,0]}^{0})\xrightarrow{d}L^{2}_{k,(\delta,\delta)}(i\Lambda_{\hat{W}[-\infty,0]}^{1})\xrightarrow{d}L^{2}_{k-1,(\delta,\delta)}(i\Lambda_{\hat{W}[-\infty,0]}^{2}).
Proof.

By the exponential decay result, one can see the correspondence

feτδ,δτδ,0ff\mapsto e^{\tau_{\delta,\delta}-\tau_{\delta,0}}f

gives an identification

{aLk,(δ,0)2(iΛW^[,0]1)|eτδ,0d(eτδ,0b)=0,da=0}\Set{a\in L^{2}_{k,(\delta,0)}(i\Lambda_{\hat{W}[-\infty,0]}^{1})}{e^{-\tau_{\delta,0}}d^{*}(e^{\tau_{\delta,0}}b)=0,da=0}
{aLk,(δ,δ)2(iΛW^[,0]1)|eτδ,δd(eτδ,δb)=0,da=0}.\to\Set{a\in L^{2}_{k,(\delta,\delta)}(i\Lambda_{\hat{W}[-\infty,0]}^{1})}{e^{-\tau_{\delta,\delta}}d^{*}(e^{\tau_{\delta,\delta}}b)=0,da=0}.

On the other hand, for an appropriate δ\delta, the complex

(24) Lk,(δ,δ)2(iΛW^[,0]0)𝑑Lk,(δ,δ)2(iΛW^[,0]1)𝑑Lk,(δ,δ)2(iΛW^[,0]2)\displaystyle L^{2}_{k,(\delta,\delta)}(i\Lambda_{\hat{W}[-\infty,0]}^{0})\xrightarrow{d}L^{2}_{k,(\delta,\delta)}(i\Lambda_{\hat{W}[-\infty,0]}^{1})\xrightarrow{d}L^{2}_{k,(\delta,\delta)}(i\Lambda_{\hat{W}[-\infty,0]}^{2})

is Fredholm, and we can identify the kernel of

Lk,(δ,δ)2(iΛW^[,0]1)d+d(δ,δ)Lk1,(δ,δ)2(iΛW^[,0]0iΛW^[,0]2)L^{2}_{k,(\delta,\delta)}(i\Lambda_{\hat{W}[-\infty,0]}^{1})\xrightarrow{d+d^{*_{(\delta,\delta)}}}L^{2}_{k-1,(\delta,\delta)}(i\Lambda_{\hat{W}[-\infty,0]}^{0}\oplus i\Lambda_{\hat{W}[-\infty,0]}^{2})

with the middle cohomology of (24). ∎

The exponential decay result enables us to prove the following correspondence:

Lemma 3.7.

For δ>0\delta>0 sufficiently small, the space

{bLk1,(δ,0)2(iΛW^[,0]+)|d(δ,0)b=0}\Set{b\in L^{2}_{k-1,(\delta,0)}(i\Lambda_{\hat{W}[-\infty,0]}^{+})}{d^{*_{(\delta,0)}}b=0}

can be identified with

{bLk1,(δ,δ)2(iΛW^[,0]+)|d(δ,δ)b=0}.\Set{b\in L^{2}_{k-1,(\delta,\delta)}(i\Lambda_{\hat{W}[-\infty,0]}^{+})}{d^{*_{(\delta,\delta)}}b=0}.
Proof.

The proof is the same as in that of Lemma 3.6. ∎

Lemma 3.8.

We have that

(25) {aLk,(δ,0)2(iΛW^[,0]1)|d(δ,0)a=0,da=0}={0},\displaystyle\Set{a\in L^{2}_{k,(\delta,0)}(i\Lambda_{\hat{W}[-\infty,0]}^{1})}{d^{*_{(\delta,0)}}a=0,da=0}=\{0\},

and one can construct an injective homomorphism

{(0,b)Lk1,(δ,0)2(iΛW^[,0]0ΛW^[,0]+)|d(δ,0)b=0}\displaystyle\Set{(0,b)\in L^{2}_{k-1,(\delta,0)}(i\Lambda_{\hat{W}[-\infty,0]}^{0}\oplus\Lambda_{\hat{W}[-\infty,0]}^{+})}{d^{*_{(\delta,0)}}b=0}
H2(W[n,0][0,n]×Y,(W[n,0][0,n]×Y);)\displaystyle\to H^{2}(W[-n,0]\cup[0,n]\times Y,\partial(W[-n,0]\cup[0,n]\times Y);\mathbb{R})

for nn sufficiently large.

Proof.

This is essentially the same argument given in [T87, Proof of Proposition 5.1]. To prove the first assertion, we construct an injective homomorphism

{aLk,(δ,0)2(iΛW^[,0]1)|d(δ,0)a=0,da=0}\displaystyle\{a\in L^{2}_{k,(\delta,0)}(i\Lambda_{\hat{W}[-\infty,0]}^{1})\ |\ d^{*_{(\delta,0)}}a=0,da=0\}
H1(W[2,0][0,2]×Y,(W[2,0][0,2]×Y);),\displaystyle\to H^{1}(W[-2,0]\cup[0,2]\times Y,\partial(W[-2,0]\cup[0,2]\times Y);\mathbb{R}),

and show that this map factors through {0}\{0\}. The proof uses the condition H1(Wn;)=0H^{1}(W_{n};\mathbb{R})=0. This map is defined by choosing a bump function β:W^[,0]\beta:\hat{W}[-\infty,0]\to\mathbb{R} such that

  • β|W[1,0][0,1]×Y=0\beta|_{W[-1,0]\cup[0,1]\times Y}=0 and

  • β|W^[,0]intW[2,0][0,2]×Y=1\beta|_{\hat{W}[-\infty,0]\setminus\operatorname{int}W[-2,0]\cup[0,2]\times Y}=1.

For a given wKerdLk,(δ,δ)2(iΛW^[,0]1)w\in\mathop{\mathrm{Ker}}\nolimits d\subset L^{2}_{k,(\delta,\delta)}(i\Lambda_{\hat{W}[-\infty,0]}^{1}), one can assume ww is smooth. Since H1(W[,1];)=0H^{1}(W[-\infty,-1];\mathbb{R})=0 and H1(Y;)=0H^{1}(Y;\mathbb{R})=0, one can choose smooth functions ff_{-} on W[,1]W[-\infty,-1] and f+f_{+} on [1,)×Y[1,\infty)\times Y such that

w|W[,1]=df and w|[1,)×Y=df+.w|_{W[-\infty,-1]}=df_{-}\text{ and }w|_{[1,\infty)\times Y}=df_{+}.

Since w|W[,1]=dfLk,δ2w|_{W[-\infty,-1]}=df_{-}\in L^{2}_{k,\delta} and w|[1,)×Y=df+Lk,02w|_{[1,\infty)\times Y}=df_{+}\in L^{2}_{k,0} , up to adding constants, one can assume

fLk+1,δ2(W[,1]) and f+Lk+1,δ2([1,)×Y).f_{-}\in L^{2}_{k+1,\delta}(W[-\infty,-1])\text{ and }f_{+}\in L^{2}_{k+1,\delta}([1,\infty)\times Y).

Define

ϕ([w]):=wd(βf++βf).\phi([w]):=w-d(\beta f_{+}+\beta f_{-}).

One can see that ϕ\phi induces a homomorphism

ϕ:(KerdLk,(δ,δ)2(iΛW^[,0]1))/d(Lk,(δ,δ)2(iΛW^[,0]0))\phi:(\mathop{\mathrm{Ker}}\nolimits d\subset L^{2}_{k,(\delta,\delta)}(i\Lambda_{\hat{W}[-\infty,0]}^{1}))/d(L^{2}_{k,(\delta,\delta)}(i\Lambda_{\hat{W}[-\infty,0]}^{0}))\to
H1(W[2,0][0,2]×Y,W[2,0][0,2]×Y;).H^{1}(W[-2,0]\cup[0,2]\times Y,\partial W[-2,0]\cup[0,2]\times Y;\mathbb{R}).

The same argument given in [T87, (5.8)-(5.10)] shows that ϕ\phi is an injection. Under our assumption, the cohomology

H1(W[2,0][0,2]×Y,W[2,0][0,2]×Y;)H^{1}(W[-2,0]\cup[0,2]\times Y,\partial W[-2,0]\cup[0,2]\times Y;\mathbb{R})

is generated by

[dg0]H1(W[2,0][0,2]×Y,W[2,0][0,2]×Y;)[dg_{0}]\in H^{1}(W[-2,0]\cup[0,2]\times Y,\partial W[-2,0]\cup[0,2]\times Y;\mathbb{R})

which is constant near (W[2,0][0,2]×Y)\partial(W[-2,0]\cup[0,2]\times Y) and g0|Y3g0|{2}×Yg_{0}|_{Y^{-}_{-3}}\neq g_{0}|_{\{2\}\times Y}, where Wi=YiYi+\partial W_{i}=Y^{-}_{i}\cup Y^{+}_{i} and Y=Yi-Y=Y^{-}_{i} as oriented manifolds.

Next we show that Imϕ={0}\mathop{\mathrm{Im}}\nolimits\phi=\{0\}. Suppose that [dg0]=ϕ([w])[dg_{0}]=\phi([w]). Then we have

ϕ([w])=wdf=dg0+dg\phi([w])=w-df=dg_{0}+dg^{\prime}

for some gΩ0(W[2,0][0,2]×Y,W[2,0][0,2]×Y)g^{\prime}\in\Omega^{0}(W[-2,0]\cup[0,2]\times Y,\partial W[-2,0]\cup[0,2]\times Y) and for some fLk+1,(δ,δ)2(W^[,0])f\in L^{2}_{k+1,(\delta,\delta)}(\hat{W}[-\infty,0]). Up to image dd, we can assume d(δ,δ)w=0d^{*_{(\delta,\delta)}}w=0. Thus we have

d(δ,δ)df=d(δ,δ)dg0+d(δ,δ)dg,-d^{*_{(\delta,\delta)}}df=d^{*_{(\delta,\delta)}}dg_{0}+d^{*_{(\delta,\delta)}}dg^{\prime},

here we consider g0g_{0} and gg^{\prime} as constant extensions on the ends. This implies

f=g0+g+constant.-f=g_{0}+g^{\prime}+\text{constant}.

Since ff goes to 0 on the ends, this gives a contradiction.

Combining this with the injectivity of ϕ\phi, we have that the domain of ϕ\phi is {0}\{0\}. Here Lemma 3.6 implies that the domain of ϕ\phi is isomorphic to the left-hand side of (25), and now the first assertion of the Lemma follows.

On the second assertion, a homomorphism

ϕn:\displaystyle\phi_{n}: {(0,b)Lk1,(δ,0)2(iΛW^[,0]0ΛW^[,0]+)|d(δ,0)b=0}\displaystyle\{(0,b)\in L^{2}_{k-1,(\delta,0)}(i\Lambda_{\hat{W}[-\infty,0]}^{0}\oplus\Lambda_{\hat{W}[-\infty,0]}^{+})\ |\ d^{*_{(\delta,0)}}b=0\}
H2(W[n,0][0,n]×Y,(W[n,0][0,n]×Y);)\displaystyle\to H^{2}(W[-n,0]\cup[0,n]\times Y,\partial(W[-n,0]\cup[0,n]\times Y);\mathbb{R})

is given as follows. First note that the domain of ϕn\phi_{n} is identified with the corresponding functional space for the weight (δ,δ)(\delta,\delta) because of Lemma 3.7. Take a sequence of bump functions βn:W^[,0]\beta_{n}:\hat{W}[-\infty,0]\to\mathbb{R} satisfying

βn|W[n+1,0][0,n1]×Y=1,βn|W[,n][n,]×Y=0, and |dβn|C0<C.\beta_{n}|_{W[-n+1,0]\cup[0,n-1]\times Y}=1\ ,\beta_{n}|_{W[-\infty,-n]\cup[n,\infty]\times Y}=0,\text{ and }|d\beta_{n}|_{C^{0}}<C.

For a given w{bLk1,(δ,δ)2(iΛW^[,0]+)|d(δ,δ)b=0}w\in\{b\in L^{2}_{k-1,(\delta,\delta)}(i\Lambda_{\hat{W}[-\infty,0]}^{+})\ |\ d^{*_{(\delta,\delta)}}b=0\}, one can see

deτ^δw=0.de^{\hat{\tau}\delta}w=0.

Since H2(W[n2,n+2];)=0H^{2}(W[-n-2,-n+2];\mathbb{R})=0 and H2([n2,n+2]×Y;)=0H^{2}([n-2,n+2]\times Y;\mathbb{R})=0, one can choose γ\gamma_{-} and γ+\gamma_{+} such that

eτ^δw|W[n2,n+2]=dγn and eτ^δw|[n2,n+2]×Y=dγ+n.e^{\hat{\tau}\delta}w|_{W[-n-2,-n+2]}=d\gamma^{n}_{-}\text{ and }e^{\hat{\tau}\delta}w|_{[n-2,n+2]\times Y}=d\gamma^{n}_{+}.

Define

ϕn(w):={eτ^δw on W[n2,0][0,n+1]×Yd(βnγn+βnγ+n) on W[n2,n+2][n2,n+2]×Y0 otherwise .\phi_{n}(w):=\begin{cases}e^{\hat{\tau}\delta}w\text{ on }W[-n-2,0]\cup[0,n+1]\times Y\\ d(\beta_{n}\gamma^{n}_{-}+\beta_{n}\gamma^{n}_{+})\text{ on }W[-n-2,-n+2]\cup[n-2,n+2]\times Y\\ 0\text{ otherwise }.\end{cases}

The proof of the injectivity of ϕn\phi_{n} is the same as the proof of [T87, Lemma 5.4]. ∎

Proof of Theorem 3.4.

Note that Proposition 3.2 gives isomorphisms of the kernels and cokernels between the operators (22) and (23), since the operator dealt with in Proposition 3.2 is the direct sum of a real operator and a complex operator. Using this and Lemma 3.5, to show the Theorem, it suffices to see that

(26) {aLk,(δ,0)2(iΛW^[,0]1)|d(δ,0)a=0,d+a=0}={0}\displaystyle\Set{a\in L^{2}_{k,(\delta,0)}(i\Lambda_{\hat{W}[-\infty,0]}^{1})}{d^{*_{(\delta,0)}}a=0,d^{+}a=0}=\{0\}

and

(27) {(0,b)Lk1,(δ,0)2(iΛW^[,0]0ΛW^[,0]+)|d(δ,0)b=0}={0}.\displaystyle\Set{(0,b)\in L^{2}_{k-1,(\delta,0)}(i\Lambda_{\hat{W}[-\infty,0]}^{0}\oplus\Lambda_{\hat{W}[-\infty,0]}^{+})}{d^{*_{(\delta,0)}}b=0}=\{0\}.

By integration by parts, one has

{aLk,(δ,0)2(iΛW^[,0]1)|d(δ,0)a=0,d+a=0}\displaystyle\Set{a\in L^{2}_{k,(\delta,0)}(i\Lambda_{\hat{W}[-\infty,0]}^{1})}{d^{*_{(\delta,0)}}a=0,d^{+}a=0}
=\displaystyle= {aLk,(δ,0)2(iΛW^[,0]1)|d(δ,0)a=0,da=0}.\displaystyle\Set{a\in L^{2}_{k,(\delta,0)}(i\Lambda_{\hat{W}[-\infty,0]}^{1})}{d^{*_{(\delta,0)}}a=0,da=0}.

The vanishing (26) follows from this and the first assertion of Lemma 3.8.

Our assumption implies H2(W[n,0],W[n,0])=0H^{2}(W[-n,0],\partial W[-n,0])=0, and the vanishing (27) follows from the second assertion of Lemma 3.8. ∎

In the proof of Theorem 1.1, we also consider the ‘right-periodic’ manifold W[0,]W[0,\infty]. Fix a Riemann metric gW[0,]g_{W[0,\infty]} on W[0,]W[0,\infty] such that

  • gW[0,]|W[1,]g_{W[0,\infty]}|_{W[1,\infty]} is periodic and PSC, and

  • gW[0,]g_{W[0,\infty]} is product metric near W[0,]=Y\partial W[0,\infty]=-Y.

Let us consider the following operators:

  • the Atiyah–Hitchin–Singer operator with APS-boundary condition:

    (28) d++p0r:Lk,δ2(iΛW[0,]1)CCLk1,δ2(iΛW[0,]+)V0(Y;),\displaystyle\begin{split}d^{+}+{p}^{0}_{-\infty}\circ{r}:L^{2}_{k,\delta}(i\Lambda_{W[0,\infty]}^{1})_{CC}\to L^{2}_{k-1,\delta}(i\Lambda_{W[0,\infty]}^{+})\oplus{V}^{0}_{-\infty}(-Y;\mathbb{R}),\end{split}
  • the linearlization of the Seiberg–Witten equation

    (29) LW[0,](p0r):𝒰k,δLk1,δ2(iΛ+S)V0(Y).\displaystyle L_{W[0,\infty]}\oplus(p^{0}_{-\infty}\circ r):\mathcal{U}_{k,\delta}\to L^{2}_{k-1,\delta}(i\Lambda^{+}\oplus S^{-})\oplus V^{0}_{-\infty}(-Y).
Theorem 3.9.

There exists δ1>0\delta^{\prime}_{1}>0 such that for any δ(0,δ1)\delta\in(0,\delta^{\prime}_{1}), the followings are true:

  • (i)

    the operator (29) is Fredholm, and

  • (ii)

    the operator (28) is isomorphism.

Proof.

The proof of (i) is the same as that of Proposition 3.1. The proof of (ii) is also essentially the same as that of Theorem 3.4. ∎

3.2. Global slice theorem

In this subsection we prove the global slice theorem in our situation. We follow the method given in [IT20]. In [IT20], for 4-manifolds with conical end, a global slice theorem is given and the essentially same method can be applied to our situation.

The following proposition is a key lemma to prove the global slice theorem:

Proposition 3.10.

There exists a small positive number δ2\delta_{2} such that for any positive real number 0<δδ20<\delta\leq\delta_{2},

(30) Lk,δ2(iΛW[,0]1)=Lk,δ2(iΛW[,0]1)CCdLk+1,δ2(iΛW[,0]0).\displaystyle L^{2}_{k,\delta}(i\Lambda^{1}_{W[-\infty,0]})=L^{2}_{k,\delta}(i\Lambda^{1}_{W[-\infty,0]})_{CC}\oplus dL^{2}_{k+1,\delta}(i\Lambda^{0}_{W[-\infty,0]}).

This proposition corresponds to [IT20, Proposition 3.5].

Proof.

The proof is essentially same as the proof of [IT20, Proposition 3.5]. We first prove

(31) Lk,δ2(iΛW[,0]1)CCdLk+1,δ2(iΛW[,0]0)={0}.\displaystyle L^{2}_{k,\delta}(i\Lambda^{1}_{W[-\infty,0]})_{CC}\cap dL^{2}_{k+1,\delta}(i\Lambda^{0}_{W[-\infty,0]})=\{0\}.

However, the proof of (31) is the same as the proof of (21) in [IT20, Proposition 3.5], and we omit this.

Next, we will see

Lk,δ2(iΛW[,0]1)=Lk,δ2(iΛW[,0]1)CC+dLk+1,δ2(iΛW[,0]0).L^{2}_{k,\delta}(i\Lambda^{1}_{W[-\infty,0]})=L^{2}_{k,\delta}(i\Lambda^{1}_{W[-\infty,0]})_{CC}+dL^{2}_{k+1,\delta}(i\Lambda^{0}_{W[-\infty,0]}).

We need to prove that, for any αLk,δ2(iΛW[,0]1)\alpha\in L^{2}_{k,\delta}(i\Lambda^{1}_{W[-\infty,0]}), there exists ξLk+1,δ2(iΛW[,0]0)\xi\in L^{2}_{k+1,\delta}(i\Lambda^{0}_{W[-\infty,0]}) such that αdξLk,δ2(iΛW[,0]1)CC\alpha-d\xi\in L^{2}_{k,\delta}(i\Lambda^{1}_{W[-\infty,0]})_{CC}, i.e.

dδdξ=dδα\displaystyle d^{*_{\delta}}d\xi=d^{*_{\delta}}\alpha
d𝐭dξ=d𝐭α\displaystyle d^{*}{\bf t}d\xi=d^{*}{\bf t}\alpha

hold. These equations are equivalent to

Δδξ=dδα\displaystyle\Delta_{\delta}\xi=d^{*_{\delta}}\alpha
𝐭ξ=GYd𝐭α,\displaystyle{\bf t}\xi=G_{Y}d^{*}{\bf t}\alpha,

where GYG_{Y} is the Green operator on YY. Therefore we need to prove surjectivity of the map

Δδ(W[,0],):Lk+1,δ2(iΛW[,0]0)Lk1,δ2(iΛW[,0]0)Lk+122(iΛY0),\Delta_{\delta}(W[-\infty,0],\partial):L^{2}_{k+1,\delta}(i\Lambda^{0}_{W[-\infty,0]})\to L^{2}_{k-1,\delta}(i\Lambda^{0}_{W[-\infty,0]})\oplus L^{2}_{k+\frac{1}{2}}(i\Lambda^{0}_{Y}),

defined by

Δδ(W[,0],)ξ=(Δδξ,𝐭ξ).\Delta_{\delta}(W[-\infty,0],\partial)\xi=(\Delta_{\delta}\xi,{\bf t}\xi).

In order to prove this, we use the excision principle and reduce the surjectivity of Δα(W[,0],)\Delta_{\alpha}(W[-\infty,0],\partial) to calculations of indexes for several Laplacian operators. The calculation of indicies of Laplacian operators are also given in [IT20, Proposition 3.5, page 18]. We can confirm the surjectivity of Δδ(W[,0],)\Delta_{\delta}(W[-\infty,0],\partial) and obtain the conclusion. ∎

Proposition 3.10 implies the following global slice theorem:

Lemma 3.11.

Let δ2\delta_{2} be the constant given in Proposition 3.10. Then, for δ(0,δ2)\delta\in(0,\delta_{2}), there is a 𝒢k+1,δ(W[,0])\mathcal{G}_{k+1,\delta}(W[-\infty,0])-equivariant diffeomorphism

𝒰k,δ(W[,0])Lk,δ2(iΛW[,0]1)CC×𝒢k+1,δ(W[,0]).\mathcal{U}_{k,\delta}(W[-\infty,0])\cong L^{2}_{k,\delta}(i\Lambda^{1}_{W[-\infty,0]})_{CC}\times\mathcal{G}_{k+1,\delta}(W[-\infty,0]).

The proof is the essentially same as in the case of closed 4-manifolds.

3.3. Dirac index on W[,0]W[-\infty,0]

In this Subsection, we shall calculate the spin Dirac index indDW[,0]+\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{W[-\infty,0]} on the half-periodic 44-manifold W[,0]W[-\infty,0]:

Proposition 3.12.

Assuming that a PSC metric is equipped with XX, we have

(32) indDW[,0]+=λSW(X,𝔰)+n(Y,𝔱,g),\displaystyle\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{W[-\infty,0]}=\lambda_{SW}(X,\mathfrak{s})+n(Y,\mathfrak{t},g),

where indDW[,0]+\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{W[-\infty,0]} means the index of the Dirac operator under the APS-boundary condition and n(Y,𝔱,g)n(Y,\mathfrak{t},g) is given in (10).

Before proving Proposition 3.12, we note a few lemmas:

Lemma 3.13.

Let M1,M2M_{1},M_{2} be compact spin 44-manifolds with common boundary YY with orientation M1=Y=M2\partial M_{1}=Y=-\partial M_{2}. Equip M1,M2M_{1},M_{2} with metrics so that the metrics are the product metric

dt2+prgYdt^{2}+\text{pr}^{*}g_{Y}

near the boundary for a Riemann metric gYg_{Y} on YY, where tt is a collar coordinate of the product neighborhood and pr means the projection from the collar neighborhoods of YY to YY. Then we have

indDM1++indDM2++dimKer=indDM1YM2+,\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{1}}+\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{2}}+\dim_{\mathbb{C}}\mathop{\mathrm{Ker}}\nolimits\cancel{\partial}=\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{1}\cup_{Y}M_{2}},

where \cancel{\partial} is the 33-dimensional Dirac operator on YY and indDMi+\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{i}} denotes the index of the Dirac operator under the APS-boundary condition.

Proof.

This can be checked by the Atiyah–Singer–Patodi index theorem [APSI] immediately, but we give a bit more direct proof to make clear the following Lemma 3.14.

We follow an argument given in Donaldson’s book [Do02], mainly [Do02, Subsubsection 3.3.1]. For α\alpha\in\mathbb{R} which is not a spectrum of \cancel{\partial}, denote by indDM1,α+\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{1},\alpha} the Fredholm index defined using the weighted Sobolev norm described as

fLk,α2=eαtfLk2\|f\|_{L^{2}_{k,\alpha}}=\|e^{\alpha t}f\|_{L^{2}_{k}}

at the end of M1=M1[0,)×YM_{1}=M_{1}\cup[0,\infty)\times Y. Take α>0\alpha>0 so that |α||\alpha| is smaller than the absolute value of the smallest non-zero eigenvalue of \cancel{\partial}. Then we obtain

indDM1YM2+=indDM1,α++indDM2,α+\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{1}\cup_{Y}M_{2}}=\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{1},\alpha}+\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{2},-\alpha}

by the gluing formula, Equation (3.2) of [Do02]. Hence it suffices to show that

(33) indDM1,α++indDM2,α+=indDM1++indDM2++dimKer.\displaystyle\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{1},\alpha}+\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{2},-\alpha}=\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{1}}+\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{2}}+\dim_{\mathbb{C}}\mathop{\mathrm{Ker}}\nolimits\cancel{\partial}.

By the definition of the APS-boundary condition, we have

indDM1,α+=indDM1+.\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{1},\alpha}=\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{1}}.

On the other hand, we have that

indDM2,α+=indDM2++dimKer\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{2},-\alpha}=\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{2}}+\dim_{\mathbb{C}}\mathop{\mathrm{Ker}}\nolimits\cancel{\partial}

by [Do02, Proposition 3.10], which is shown considering a certain ordinary equation [Do02, Lemma 3.11] corresponding to the cylinder (,)×Y(-\infty,\infty)\times Y appearing the neck stretching of M1YM2M_{1}\cup_{Y}M_{2}. Now we have checked (33) and this completes the proof. ∎

The proof of Lemma 3.13 involves only near the neck of M1YM2M_{1}\cup_{Y}M_{2}. Even if we replace M1M_{1} with a manifold with an additional end, we obtain a similar result as far as we work in Fredholm setting. This makes clear the following Lemma:

Lemma 3.14.

Let MM be a compact spin manifold bounded by YY with the orientation M=Y\partial M=-Y. Equip MM with a metric so that the metrics are product metrics near the boundary. Then we have

(34) indDW[,0]++indDM++dimKer=indDM+,\displaystyle\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{W[-\infty,0]}+\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M}+\dim_{\mathbb{C}}\mathop{\mathrm{Ker}}\nolimits\cancel{\partial}=\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{\infty}},

where

M=YWYWYM.M_{\infty}=\cdots\cup_{Y}W\cup_{Y}W\cup_{Y}M.

Now we are ready to prove Proposition 3.12.

Proof of Proposition 3.12.

Take a compact spin bound MM of YY with the orientation M=Y\partial M=-Y. Take a metric on MM so that the metrics are product metrics near the boundary.

Now we shall check

(35) indDM+=σ(M)8+λSW(X,𝔰).\displaystyle\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{\infty}}=-\frac{\sigma(M)}{8}+\lambda_{SW}(X,\mathfrak{s}).

Note the sign: this comes from the orientation of MM with M=Y\partial M=-Y. Indeed, by [Lin19, Lemma 2.21], it follows from the existence of PSC metric on XX that

(36) λSW(X,𝔰)=λSW(X,𝔰).\displaystyle-\lambda_{SW}(-X,\mathfrak{s})=\lambda_{SW}(X,\mathfrak{s}).

(Precisely, XX is supposed to be an integral homology S1×S3S^{1}\times S^{3} in [Lin19], but the proof of [Lin19, Lemma 2.21] is valid also for rational homology S1×S3S^{1}\times S^{3}’s without any changes.) On the other hand, for a PSC metric gg on XX, we have

(37) λSW(X,𝔰)=w(X,g,0)=indDM+σ(M)8.\displaystyle\lambda_{SW}(-X,\mathfrak{s})=-w(-X,g,0)=-\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M_{\infty}}-\frac{\sigma(M)}{8}.

Equation (35) is deduced from (36) and (37).

On the other hand, we also have

(38) indDM+=σ(M)8n(Y,𝔱,g)dimKer.\displaystyle\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M}=-\frac{\sigma(M)}{8}-n(Y,\mathfrak{t},g)-\dim_{\mathbb{C}}\mathop{\mathrm{Ker}}\nolimits\cancel{\partial}.

Indeed, it follows that

(39) indDM++indDM++dimKer=0\displaystyle\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{-M}+\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M}+\dim_{\mathbb{C}}\mathop{\mathrm{Ker}}\nolimits\cancel{\partial}=0

because of Lemma 3.13 and

indDMYM+=σ(M)8+σ(M)8=0.\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{-M\cup_{Y}M}=\frac{\sigma(-M)}{8}+\frac{\sigma(M)}{8}=0.

By the definition of n(Y,𝔱,g)n(Y,\mathfrak{t},g), we have

(40) n(Y,𝔱,g)=indDM++σ(M)8.\displaystyle n(Y,\mathfrak{t},g)=\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{-M}+\frac{\sigma(-M)}{8}.

Equation (38) is deduced from (39), (40).

Combining Lemma 3.14 with (35) and (38), we obtain the desired equality (32). ∎

4. The boundedness result

In this section, we prove a certain boundedness result in order to construct Bauer–Furuta type invariant. We mainly follow the methods given in [Ma03, Kha15]. The situation is similar to that in [IT20], which gives a Bauer-Furuta invariant for 4-manifolds with conical end.

Our main result in this section is:

Theorem 4.1.

There exists δ3>0\delta_{3}>0 and a constant R>0R>0 such that the following conclusion holds. Let δ\delta be an element in (0,δ3](0,\delta_{3}]. Suppose that a pair (x,y)(x,y) of

x𝒰k,δ(W[,0])x\in\mathcal{U}_{k,\delta}({W[-\infty,0]})

and y:[0,)V(Y)y:[0,\infty)\to V(Y) satisfy the following conditions:

  • (i)

    the element x+(A0,Φ0)x+(A_{0},\Phi_{0}) is a solution to the equation (16) on W[,0]W[-\infty,0],

  • (ii)

    the element yy is a solution to the Seiberg–Witten equations on 0×Y\mathbb{R}^{\geq 0}\times Y,

  • (iii)

    yy is temporal gauge, i.e. db(t)=0d^{*}b(t)=0 for each tt, where y(t)=(b(t),ψ(t))y(t)=(b(t),\psi(t)), and yy is of finite type,

  • (iv)

    x|Y=y(0)x|_{Y}=y(0), and

  • (v)

    |limtCSD(y(t))|<|\lim_{t\to\infty}CSD(y(t))|<\infty.

Then we have the following universal bounds:

xLk,δ2<R and y(t)Lk122<R(t0).\|x\|_{L^{2}_{k,\delta}}<R\text{ and }\|y(t)\|_{L^{2}_{k-\frac{1}{2}}}<R\ (\forall t\geq 0).

In order to prove Theorem 4.1, we use several corresponding notions used in [Ma03].

Definition 4.2.

We consider a Riemannian manifold

W^[,0]=W[,0](0×Y)\hat{W}[-\infty,0]=W[-\infty,0]\cup(\mathbb{R}^{\geq 0}\times Y)

obtained by gluing the half-cylinder (0×Y,dt2+gY)(\mathbb{R}^{\geq 0}\times Y,dt^{2}+g_{Y}) and W[,0]W[-\infty,0] along their boundary. A solution (A,Φ)(A,\Phi) to the Seiberg–Witten equations on W^[,0]\hat{W}[-\infty,0] is called W[,0]{W}[-\infty,0]-trajectories. If a W[,0]W[-\infty,0]-trajectory (A,Φ)(A,\Phi) satisfies

supt0|CSD(A|{t}×Y)|< and ΦC0(0×Y)<,\sup_{t\in\mathbb{R}^{\geq 0}}|CSD(A|_{\{t\}\times Y})|<\infty\text{ and }\|\Phi\|_{C^{0}(\mathbb{R}^{\geq 0}\times Y)}<\infty,

then (A,Φ)(A,\Phi) is called a finite type W[,0]W[-\infty,0]-trajectory.

Let us note the following boundedness result:

Theorem 4.3.

Let CC be a positive real number and

(A,Φ)(A0,0)+Lk,δ2(iΛW[,0][0,1]×Y1)Lk,δ2(SW[,0][0,1]×Y+)(A,\Phi)\in(A_{0},0)+L^{2}_{k,\delta}(i\Lambda_{W[-\infty,0]\cup[0,1]\times Y}^{1})\oplus L^{2}_{k,\delta}(S^{+}_{W[-\infty,0]\cup[0,1]\times Y})

be a solution to (A,Φ)=0\mathcal{F}(A,\Phi)=0 such that

top(A,Φ)C\mathcal{E}^{top}(A,\Phi)\leq C

and

(A,Φ)𝒰k,δ(W[,0]).(A,\Phi)\in\mathcal{U}_{k,\delta}({W}[-\infty,0]).

Then, there exists δ3\delta_{3} such that for any δ(0,δ3)\delta\in(0,\delta_{3}), the inequality

(A,Φ)(A0,0)Lk,δ2(W[,0])D(C)\|(A,\Phi)-(A_{0},0)\|_{L^{2}_{k,\delta}(W[-\infty,0])}\leq D(C)

holds, where D(C)D(C) is a constant depending only on DD.

Proof.

We compare gauge transformations constructed by J. Lin [Lin19, Subsection 4.2] with the global slice obtained in Proposition 3.10. The proof of [Lin19, Lemma 4.10] implies that there exists a constant δ3\delta^{\prime}_{3} and a gauge transformation gg^{\prime} on W[,0]W[-\infty,0] such that for any δ(0,δ3)\delta\in(0,\delta_{3}^{\prime}),

(g)(A,Φ)(A0,0)Lk,δ2(W[,0])D(C).\|(g^{\prime})^{*}(A,\Phi)-(A_{0},0)\|_{L^{2}_{k,\delta}(W[-\infty,0])}\leq D(C).

Define

δ3:=min{δ3,δ2}.\delta_{3}:=\min\{\delta_{3}^{\prime},\delta_{2}\}.

On the other hand, by Lemma 3.11, the map obtained by giving a slice

𝒰k,δ(W[,0])Lk,δ2(iΛW[,0]1)CC×𝒢k+1,δ(W[,0])\mathcal{U}_{k,\delta}(W[-\infty,0])\xrightarrow{\cong}L^{2}_{k,\delta}(i\Lambda^{1}_{W[-\infty,0]})_{CC}\times\mathcal{G}_{k+1,\delta}(W[-\infty,0])

is continuous. This implies there is a gauge transformation gg such that

g(A,Φ)𝒰k,δ(W[,0])g^{*}(A,\Phi)\in\mathcal{U}_{k,\delta}({W}[-\infty,0])

and

g(A,Φ)(A0,0)Lk,δ2(W[,0])C(g)(A,Φ)(A0,0)Lk,δ2(W[,0])CD(C).\|g^{*}(A,\Phi)-(A_{0},0)\|_{L^{2}_{k,\delta}(W[-\infty,0])}\leq C^{\prime}\|(g^{\prime})^{*}(A,\Phi)-(A_{0},0)\|_{L^{2}_{k,\delta}(W[-\infty,0])}\leq C^{\prime}D(C).

This gives the desired result. ∎

The topological energy top\mathcal{E}^{\text{top}} and the analytic energy top\mathcal{E}^{\text{top}} for configurations on W^[,0]\hat{W}[-\infty,0] are defined along the book by Kronheimer–Mrowka [KM07, Definition 4.5.4]. Note that, for a configuration (A,Φ)(A,\Phi) converging to (A0,0)(A_{0},0) on the periodic end, the boundary terms in the topological energy corresponding to the end vanishes, while the boundary terms corresponding to the cylindrical end may survive. If such a configuration (A,Φ)(A,\Phi) is a W[,0]{W}[-\infty,0]-trajectory and is asymptotic to 𝔠\mathfrak{c} on the cylindrical end, we have that

(41) top(A,Φ)=CXCSD(𝔠),\displaystyle\mathcal{E}^{\text{top}}(A,\Phi)=C_{X}-CSD(\mathfrak{c}),

where CXC_{X} depends only on XX and the fixed metric and spin structure on XX. Moreover, we have that top(A,Φ)=an(A,Φ)\mathcal{E}^{\text{top}}(A,\Phi)=\mathcal{E}^{\text{an}}(A,\Phi) as well as for a configuration over a compact 44-manifold.

Proof of Theorem 4.1.

Let δ3\delta_{3} be the constant given in Theorem 4.3. Suppose that

(x,y)𝒰k,δ(Map([0,),Lk122(iΛY1)Lk122(SY+))(x,y)\in\mathcal{U}_{k,\delta}\oplus(\operatorname{Map}([0,\infty),L^{2}_{k-\frac{1}{2}}(i\Lambda^{1}_{Y})\oplus L^{2}_{k-\frac{1}{2}}(S^{+}_{Y}))

satisfies the assumption of Theorem 4.1. First, we state a pasting lemma:

Lemma 4.4.

The pair (x,y)(x,y) gives rise to a finite type W[,0]W[-\infty,0]-trajectory (A,Φ)(A,\Phi).

Proof.

This is essentially the same as the proof of [Khan15, Corollary 4.3]. ∎

It follows from Lemma 4.4 that we have a solution (A,Φ)(A,\Phi) to the Seiberg–Witten equations on W^[,0]\hat{W}[-\infty,0] whose topological energy is finite.

Recall that the set of critical points of CSDCSD modulo gauge is compact. Since we consider a spin structure now, CSDCSD is gauge invariant. Therefore the set of critical values of CSDCSD is compact.

Since we have assumed that |limtCSD(y(t))|<|\lim_{t\to\infty}CSD(y(t))|<\infty, we have that

|CSD(y(t))CSD(y(t+1))|0|CSD(y(t))-CSD(y(t+1))|\to 0

as tt\to\infty, and therefore there exists a critical point of CSDCSD to which (A,Φ)(A,\Phi) is Lk122L^{2}_{k-\frac{1}{2}}-asymptotic as tt\to\infty. This combined with (41) implies that top(A,Φ)\mathcal{E}^{\text{top}}(A,\Phi) is uniformly bounded, and hence so is an(A,Φ)\mathcal{E}^{\text{an}}(A,\Phi).

We claim that the analytic energy of (A,Φ)(A,\Phi) restricted to W[,1]W[-\infty,-1] is also uniformly bounded. To see this, let us decompose W^[,1]\hat{W}[-\infty,-1] into three parts: the periodic part W[,1]W[-\infty,-1], the cylindrical part 0×Y\mathbb{R}^{\geq 0}\times Y, and the ‘joint’ between the periodic part and the cylindrical part. We have seen that the analytic energy of (A,Φ)(A,\Phi) on W^[,1]\hat{W}[-\infty,-1] is uniformly bounded, and this energy is the sum of the energies on these three parts. Therefore, to prove that the analytic energy of (A,Φ)(A,\Phi) restricted to W[,1]W[-\infty,-1] is also uniformly bounded, it suffices to show that all of the energies on these three parts are bounded from below. But this is obvious to recalling the definition of the analytic energy. (See the proof of [Lin19, Lemma 4.8].)

This uniform boundedness enables us to apply Theorem 4.3, and thus we obtain the boundedness of xLk,δ2<R\|x\|_{L^{2}_{k,\delta}}<R:

(A,Φ)(A0,0)Lk,δ2(W[,0])R\|(A,\Phi)-(A_{0},0)\|_{L^{2}_{k,\delta}(W[-\infty,0])}\leq R

for any δ[0,δ3)\delta\in[0,\delta_{3}). The remaining boundedness result y(t)Lk122<R\|y(t)\|_{L^{2}_{k-\frac{1}{2}}}<R follows from the same argument for XX-trajectories, where XX is a compact 44-manifold bounded by YY. See [Khan15, Corollary 4.3] for example. ∎

5. Relative Bauer–Furuta type invariant

In this section, we construct a relative Bauer–Furuta type invariant for 4-manifolds with periodic end and boundary W[,0]W[-\infty,0]. We mainly follow the methods given by Manolescu [Ma03] and Khandhawit [Kha15].

We consider a finite-dimensional approximation of the map

W[,0]:𝒰k,δ𝒱k1,δV(Y).\mathcal{F}_{W[-\infty,0]}:\mathcal{U}_{k,\delta}\to\mathcal{V}_{k-1,\delta}\oplus V(Y).

We fix a weight δ(0,)\delta\in(0,\infty) satisfying

δmin{δ0,δ1,δ2,δ3}\delta\leq\min\{\delta_{0},\delta_{1},\delta_{2},\delta_{3}\}

in the rest of this paper, where δi\delta_{i} are the constants appeared in Remark 2.6, Theorem 3.4, Proposition 3.10, and Theorem 4.1. Take sequences of subspaces

𝒱1𝒱2𝒱k1,δ and Vλ1λ1Vλ2λ2V(Y)\mathcal{V}_{1}\subset\mathcal{V}_{2}\subset\cdots\subset\mathcal{V}_{k-1,\delta}\text{ and }V^{\lambda_{1}}_{-\lambda_{1}}\subset V^{\lambda_{2}}_{-\lambda_{2}}\subset\cdots\subset V(Y)

such that

  • (i)

    (ImLW[,0]+pλnλnr)𝒱k1,δV(Y)𝒱nVλnλn(Y)(\mathop{\mathrm{Im}}\nolimits L_{W[-\infty,0]}+p^{\lambda_{n}}_{-\lambda_{n}}\circ r)^{\perp_{\mathcal{V}_{k-1,\delta}\oplus V(Y)}}\subset\mathcal{V}_{n}\oplus V^{\lambda_{n}}_{-\lambda_{n}}(Y)

  • (ii)

    the L2L^{2}-projection Pn:𝒱k1,αV(Y)𝒱nVλnλn(Y)P_{n}:\mathcal{V}_{k-1,\alpha}\oplus V(Y)\to\mathcal{V}_{n}\oplus V^{\lambda_{n}}_{-\lambda_{n}}(Y) satisfies

    limnPn(v)=v\lim_{n\to\infty}P_{n}(v)=v

    for any v𝒱k1,δV(Y)v\in\mathcal{V}_{k-1,\delta}\oplus V(Y).

Then we define a sequence of subspaces

𝒰n:=(LW[,0]+pλnλnr)1(𝒱nVλnλn).\mathcal{U}_{n}:=(L_{W[-\infty,0]}+p^{\lambda_{n}}_{-\lambda_{n}}\circ r)^{-1}(\mathcal{V}_{n}\oplus V^{\lambda_{n}}_{-\lambda_{n}}).

This gives a family of the approximated Seiberg–Witten maps

{n:=Pn(LW[,0]+CW[,0],pλnλnr):𝒰n𝒱nVλnλn(Y)}.\{\mathcal{F}_{n}:=P_{n}(L_{W[-\infty,0]}+C_{W[-\infty,0]},p^{\lambda_{n}}_{-\lambda_{n}}\circ r)\colon\mathcal{U}_{n}\to\mathcal{V}_{n}\oplus V^{\lambda_{n}}_{-\lambda_{n}}(Y)\}.

The following proposition gives us a well-defined continuous map between spheres.

Proposition 5.1.

For a large nn and a large positive real number RR, there exists an index pair (Nn,Ln)(N_{n},L_{n}) of Vλnλn(Y)V^{\lambda_{n}}_{-\lambda_{n}}(Y) and a sequence {εn}\{\varepsilon_{n}\} of positive numbers such that

(42) B(𝒰n;R)/S(𝒰n;R)(𝒱n/B(𝒱n,εn)c)(Nn/Ln)\displaystyle B(\mathcal{U}_{n};R)/S(\mathcal{U}_{n};R)\to(\mathcal{V}_{n}/B(\mathcal{V}_{n},\varepsilon_{n})^{c})\wedge(N_{n}/L_{n})

is well-defined, where B(V;R)B(V;R) is the closed ball in VV with radius RR and S(V;R)S(V;R) is the sphere in VV with radius RR.

For the proof of Proposition 5.1, we use the following proposition.

Proposition 5.2.

Let {xn}\{x_{n}\} be a bounded sequence in 𝒰k,δ\mathcal{U}_{k,\delta} such that

(LW[,0](xn),pλnr(xn))𝒱n×Vλnλn(L_{W[-\infty,0]}(x_{n}),p^{\lambda_{n}}_{-\infty}\circ r(x_{n}))\in\mathcal{V}_{n}\times V^{\lambda_{n}}_{-\lambda_{n}}

and

Pn(LW[,0]+CW[,0])xn0.P_{n}(L_{W[-\infty,0]}+C_{W[-\infty,0]})x_{n}\to 0.

Let yn:[0,)Vλnλny_{n}:[0,\infty)\to V^{\lambda_{n}}_{-\lambda_{n}} be a uniformly bounded sequence of trajectories such that

yn(0)=pλnr(xn).y_{n}(0)=p^{\lambda_{n}}_{-\infty}\circ r(x_{n}).

Then, after taking a subsequence, {xn}\{x_{n}\} converges to a solution x𝒰k,δx\in\mathcal{U}_{k,\delta} (in the topology of 𝒰k,δ\mathcal{U}_{k,\delta}) and {yn(t)}\{y_{n}(t)\} converges to y(t)(t[0,))y(t)(\forall t\in[0,\infty)) in Lk122L^{2}_{k-\frac{1}{2}} which is a solution of the Seiberg–Witten equations on 0×Y\mathbb{R}^{\geq 0}\times Y.

Proof.

The proof is similar to the proof of [Kha15, Proposition 3]. By the same argument, one sees the following result: for any compact set I(0,)I\subset(0,\infty), after taking a subsequence, yn(t)y_{n}(t) uniformly converges to y(t)y(t) in Lk122L^{2}_{k-\frac{1}{2}}, where y(t)y(t) is the weak limit.

For the sequence {xn}\{x_{n}\}, we need to ensure:

  • after taking a subsequence, p0yn(0)p0r(x)p^{0}_{-\infty}y_{n}(0)\to p^{0}_{-\infty}r(x) in Lk122L^{2}_{k-\frac{1}{2}}, where xx is the weak limit and

  • after taking a subsequence, the sequence {xn}\{x_{n}\} converges to xx in Lk,δ2(X)L^{2}_{k,\delta}(X).

The proof of the second statement is the only difference between our construction and the usual Bauer–Furuta invariant. Here we again follow the method given in [IT20]. To obtain the convergence of {xn}\{x_{n}\}, we will use the following inequality obtained by the Fredholm property of LW[,0]L_{W[-\infty,0]}: there exists a constant C>0C>0 such that, for any x𝒰k,δx\in\mathcal{U}_{k,\delta},

xLk,δ2C(LW[,0](x)Lk1,δ2+p0r(x)Lk122+xL2).\|x\|_{L^{2}_{k,\delta}}\leq C(\|L_{W[-\infty,0]}(x)\|_{L^{2}_{k-1,\delta}}+\|p^{0}_{-\infty}r(x)\|_{L^{2}_{k-\frac{1}{2}}}+\|x\|_{L^{2}}).

Then, by the same discussion given in the proof of [IT20, Lemma 3.18], we complete the proof. ∎

Proof of Proposition 5.1.

We combine Proposition 5.2, Theorem 4.1 and the proof of [Kha15, Proposition 4.5] and complete the proof. ∎

By Proposition 5.1, we obtain a family of the continuous maps (42). By the definition of Fredholm index, we have

ind(LW[,0]pλnr)=dim𝒰ndim𝒱ndimVλnλn.\mathop{\mathrm{ind}}\nolimits_{\mathbb{R}}(L_{W[-\infty,0]}\oplus p^{\lambda_{n}}_{-\infty}\circ r)=\dim_{\mathbb{R}}\mathcal{U}_{n}-\dim_{\mathbb{R}}\mathcal{V}_{n}-\dim_{\mathbb{R}}V^{\lambda_{n}}_{-\lambda_{n}}.

We obtain a map stably written by

Ψ:(~mn)+(~mn)+ΣVλn0(Nn/Ln),\Psi:(\widetilde{\mathbb{R}}^{m}\oplus\mathbb{H}^{n})^{+}\to(\widetilde{\mathbb{R}}^{m^{\prime}}\oplus\mathbb{H}^{n^{\prime}})^{+}\wedge\Sigma^{-V^{0}_{-\lambda_{n}}}(N_{n}/L_{n}),

here we fixed trivializations of vector spaces.

Remark 5.3.

Our construction gives an invariant of 4-manifolds with periodic end admitting periodic PSC metric on the end. This can be regarded as relative Bauer–Furuta invariant corresponding to [Ve14].

6. The proof of Theorem 1.1

In this Section, we prove Theorem 1.1. Recalling the definition of local equivalence [Sto20], what we have to do is to construct a certain type of map called local map from SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}) to [(S0,0,λSW(X,𝔰)/2)]\left[\left(S^{0},0,-\lambda_{SW}(X,\mathfrak{s})/2\right)\right], and also a local map from [(S0,0,λSW(X,𝔰)/2)]\left[\left(S^{0},0,-\lambda_{SW}(X,\mathfrak{s})/2\right)\right] to SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}).

We shall consider the relative Bauer–Furuta invariant on the ‘left-periodic’ manifold W[,0]W[-\infty,0] and that on ‘right-periodic’ manifold W[0,]W[0,\infty]. These two relative Bauer–Furuta invariants give the desired two local maps.

Proof of Theorem 1.1.

In Section 5, under the assumption of the existence of PSC metric on XX, we constructed a Pin(2)\operatorname{Pin}(2)-equivariant continuous map of the form

(43) f:(~m0n0)+(~m1n1)+Iλλ\displaystyle f:(\tilde{\mathbb{R}}^{m_{0}}\oplus\mathbb{H}^{n_{0}})^{+}\to(\tilde{\mathbb{R}}^{m_{1}}\oplus\mathbb{H}^{n_{1}})^{+}\wedge I_{-\lambda}^{\lambda}

as the relative Bauer–Furuta invariant over W[,0]W[-\infty,0]. One sees that fS1f^{S^{1}} induces a Pin(2)\operatorname{Pin}(2)-homotopy equivalence by Theorem 3.4. The numbers m0m1,n0n1m_{0}-m_{1},n_{0}-n_{1} are given by

(44) m0m1=dimVλ0(),2(n0n1)=indDW[,0]++dimVλ0()=λSW(X,𝔰)+n(Y,𝔱,g)+dimVλ0().\displaystyle\begin{split}m_{0}-m_{1}&=\dim V^{0}_{-\lambda}(\mathbb{R}),\\ 2(n_{0}-n_{1})&=\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{W[-\infty,0]}+\dim_{\mathbb{C}}V^{0}_{-\lambda}(\mathbb{H})\\ &=\lambda_{SW}(X,\mathfrak{s})+n(Y,\mathfrak{t},g)+\dim_{\mathbb{C}}V^{0}_{-\lambda}(\mathbb{H}).\end{split}

For the notations Vλ0()V^{0}_{-\lambda}(\mathbb{R}) and Vλ0()V^{0}_{-\lambda}(\mathbb{H}), see (11). Here we have used Proposition 3.12 to get the second equality of (44) and Theorem 3.4 to get the first equality.

Equations (43) and (44) mean that the map ff gives a local map from [(S0,0,λSW(X,𝔰)/2)]\left[\left(S^{0},0,-\lambda_{SW}(X,\mathfrak{s})/2\right)\right] to SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}).

Next, instead of the ‘left-periodic’ manifold W[,0]W[-\infty,0], we consider the ‘right-periodic’ manifold

W[0,]=WYWYWY.W[0,\infty]=W\cup_{Y}W\cup_{Y}W\cup_{Y}\cdots.

Repeating analysis in Section 5 for W[0,]W[0,\infty] instead of W[,0]W[-\infty,0], we obtain a Pin(2)\operatorname{Pin}(2)-map of the form

(45) f:(~m0n0)+(~m1n1)+I¯λλ\displaystyle f^{\prime}:(\tilde{\mathbb{R}}^{m^{\prime}_{0}}\oplus\mathbb{H}^{n^{\prime}_{0}})^{+}\to(\tilde{\mathbb{R}}^{m^{\prime}_{1}}\oplus\mathbb{H}^{n^{\prime}_{1}})^{+}\wedge\bar{I}_{-\lambda}^{\lambda}

as the relative Bauer–Furuta invariant over W[0,]W[0,\infty]. Here I¯λλ\bar{I}_{-\lambda}^{\lambda} denotes the Conley index for Y-Y. As well as ff above, (f)S1(f^{\prime})^{S^{1}} induces a Pin(2)\operatorname{Pin}(2)-homotopy equivalence by Theorem 3.9. For μ0λ\mu\leq 0\leq\lambda, as in [Ma16, Proof of Proposition 3.8], let us denote by V¯μλ\bar{V}^{\lambda}_{\mu} the vector space VμλV^{\lambda}_{\mu} defined for Y-Y. Note that, for μ<0<λ\mu<0<\lambda, we have an identification V¯μλVλμ\bar{V}^{\lambda}_{\mu}\cong V^{-\mu}_{-\lambda}, and in particular V¯λλVλλ\bar{V}^{\lambda}_{-\lambda}\cong V^{\lambda}_{-\lambda}. Under this notation, m0m1,n0n1m^{\prime}_{0}-m^{\prime}_{1},n^{\prime}_{0}-n^{\prime}_{1} are given by

(46) m0m1=dimV¯λ0(),2(n0n1)=indDW[0,]++dimV¯λ0().\displaystyle\begin{split}m^{\prime}_{0}-m^{\prime}_{1}&=\dim\bar{V}^{0}_{-\lambda}(\mathbb{R}),\\ 2(n^{\prime}_{0}-n^{\prime}_{1})&=\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{W[0,\infty]}+\dim_{\mathbb{C}}\bar{V}^{0}_{-\lambda}(\mathbb{H}).\end{split}

By an argument using a duality map as in [Ma16, page 168], we obtain a Pin(2)\operatorname{Pin}(2)-map

f′′:(~m0n0)+Iλλ(~m1n1)+(Vλλ)+\displaystyle f^{\prime\prime}:(\tilde{\mathbb{R}}^{m^{\prime}_{0}}\oplus\mathbb{H}^{n^{\prime}_{0}})^{+}\wedge I_{-\lambda}^{\lambda}\to(\tilde{\mathbb{R}}^{m^{\prime}_{1}}\oplus\mathbb{H}^{n^{\prime}_{1}})^{+}\wedge(V^{\lambda}_{-\lambda})^{+}

from (45). The vector space VλλV^{\lambda}_{-\lambda} can be decomposed so that Vλλ()Vλλ()V^{\lambda}_{-\lambda}(\mathbb{R})\oplus V^{\lambda}_{-\lambda}(\mathbb{H}). Set

(47) m1′′=m1+dimVλλ(),n1′′=n1+dimVλλ().\displaystyle\begin{split}&m_{1}^{\prime\prime}=m_{1}^{\prime}+\dim V^{\lambda}_{-\lambda}(\mathbb{R}),\\ &n_{1}^{\prime\prime}=n_{1}^{\prime}+\dim_{\mathbb{H}}V^{\lambda}_{-\lambda}(\mathbb{H}).\end{split}

Then the domain and codomain of f′′f^{\prime\prime} are given by

(48) f′′:(~m0n0)+Iλλ(~m1′′n1′′)+.\displaystyle f^{\prime\prime}:(\tilde{\mathbb{R}}^{m^{\prime}_{0}}\oplus\mathbb{H}^{n^{\prime}_{0}})^{+}\wedge I_{-\lambda}^{\lambda}\to(\tilde{\mathbb{R}}^{m^{\prime\prime}_{1}}\oplus\mathbb{H}^{n^{\prime\prime}_{1}})^{+}.

We shall show that f′′f^{\prime\prime} gives a local map from SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}) to [(S0,0,λSW(X,𝔰)/2)]\left[\left(S^{0},0,-\lambda_{SW}(X,\mathfrak{s})/2\right)\right]. The restriction (f′′)S1(f^{\prime\prime})^{S^{1}} is a Pin(2)Pin(2)-homotopy equivalence since ff^{\prime} is so. One may assume λ\lambda was taken to avoid the eigenvalues of the linearization ll of the flow equations. Then we have

(49) V¯λ0()=V0λ(),V¯λ0()=V0λ()ker.\displaystyle\begin{split}&\bar{V}_{-\lambda}^{0}(\mathbb{R})=V_{0}^{\lambda}(\mathbb{R}),\\ &\bar{V}_{-\lambda}^{0}(\mathbb{H})=V_{0}^{\lambda}(\mathbb{H})\oplus\ker\cancel{\partial}.\end{split}

Here, to obtain the first equality, we have used Ker(d:kerdΩ1(Y))=0\mathop{\mathrm{Ker}}\nolimits(\ast d:\ker d^{\ast}\to\Omega^{1}(Y))=0 deduced from the assumption that b1(Y)=0b_{1}(Y)=0. Using (LABEL:eq:_zero_eig_proof_of_main), we have

(50) Vλλ()Vλ0()Vλ0()Vλ0()V¯λ0()\displaystyle\begin{split}V^{\lambda}_{-\lambda}(\mathbb{R})&\cong V^{0}_{-\lambda}(\mathbb{R})\oplus V^{0}_{-\lambda}(\mathbb{R})\\ &\cong V^{0}_{-\lambda}(\mathbb{R})\oplus\bar{V}_{-\lambda}^{0}(\mathbb{R})\end{split}

and

(51) Vλλ()kerVλ0()Vλ0()kerVλ0()V¯λ0().\displaystyle\begin{split}V^{\lambda}_{-\lambda}(\mathbb{H})\oplus\ker\cancel{\partial}&\cong V^{0}_{-\lambda}(\mathbb{H})\oplus V^{0}_{-\lambda}(\mathbb{H})\oplus\ker\cancel{\partial}\\ &\cong V^{0}_{-\lambda}(\mathbb{H})\oplus\bar{V}_{-\lambda}^{0}(\mathbb{H}).\end{split}

Combining (46) with (LABEL:eq:_prf_main_double_pr), (50) and (51), we obtain

(52) m0m1′′=m0m1dimV¯λλ()=dimV¯λ0()dimVλλ()=dimVλ0(),\displaystyle\begin{split}m_{0}^{\prime}-m_{1}^{\prime\prime}&=m_{0}^{\prime}-m_{1}^{\prime}-\dim\bar{V}^{\lambda}_{-\lambda}(\mathbb{R})\\ &=\dim\bar{V}^{0}_{-\lambda}(\mathbb{R})-\dim V^{\lambda}_{-\lambda}(\mathbb{R})=-\dim V^{0}_{-\lambda}(\mathbb{R}),\end{split}
(53) n0n1′′=n0n1dimV¯λλ()=dimV¯λ0()dimVλλ()=indDW[0,]++dimkerdimVλ0().\displaystyle\begin{split}n_{0}^{\prime}-n_{1}^{\prime\prime}&=n_{0}^{\prime}-n_{1}^{\prime}-\dim\bar{V}^{\lambda}_{-\lambda}(\mathbb{H})\\ &=\dim\bar{V}^{0}_{-\lambda}(\mathbb{H})-\dim V^{\lambda}_{-\lambda}(\mathbb{H})\\ &=\mathop{\mathrm{ind}}\nolimits_{\mathbb{H}}D^{+}_{W[0,\infty]}+\dim\ker\cancel{\partial}-\dim_{\mathbb{H}}V^{0}_{-\lambda}(\mathbb{H}).\end{split}

Let us calculate indDW[0,]+\mathop{\mathrm{ind}}\nolimits_{\mathbb{H}}D^{+}_{W[0,\infty]} in the last equality. Let MM^{\prime} be an oriented compact smooth 44-manifold with boundary M=Y\partial M^{\prime}=Y. Set

M=MYWYWY.M^{\prime}_{\infty}=M^{\prime}\cup_{Y}W\cup_{Y}W\cup_{Y}\cdots.

Then, as well as Lemma 3.14, we obtain

(54) indDW[0,]++indDM++dimKer=indDM+,\displaystyle\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{W[0,\infty]}+\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M^{\prime}}+\dim_{\mathbb{C}}\mathop{\mathrm{Ker}}\nolimits\cancel{\partial}=\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M^{\prime}_{\infty}},

On the other hand, for a PSC metric gg on XX, we have

(55) λSW(X,𝔰)=w(X,g,0)=indDM+σ(M)8.\displaystyle\lambda_{SW}(X,\mathfrak{s})=-w(X,g,0)=-\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M^{\prime}_{\infty}}-\frac{\sigma(M^{\prime})}{8}.

Recalling the definition of n(Y,𝔱,g)n(Y,\mathfrak{t},g), we have

(56) n(Y,𝔱,g)=indDM++σ(M)8.\displaystyle n(Y,\mathfrak{t},g)=\mathop{\mathrm{ind}}\nolimits_{\mathbb{C}}D^{+}_{M^{\prime}}+\frac{\sigma(M^{\prime})}{8}.

Combining (54) with (55) and (56), we have

(57) indDW[0,]+=12(λSW(X,𝔰)+n(Y,𝔱,g))dimker.\displaystyle\mathop{\mathrm{ind}}\nolimits_{\mathbb{H}}D^{+}_{W[0,\infty]}=-\frac{1}{2}(\lambda_{SW}(X,\mathfrak{s})+n(Y,\mathfrak{t},g))-\dim_{\mathbb{H}}\ker\cancel{\partial}.

It follows from (53) and (57) that

(58) n0n1′′=12(λSW(X,𝔰)+n(Y,𝔱,g))dimVλ0().\displaystyle\begin{split}n_{0}^{\prime}-n_{1}^{\prime\prime}=-\frac{1}{2}(\lambda_{SW}(X,\mathfrak{s})+n(Y,\mathfrak{t},g))-\dim_{\mathbb{H}}V^{0}_{-\lambda}(\mathbb{H}).\end{split}

Now we deduce from (48), (52), and (58) that f′′f^{\prime\prime} gives a local map from SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}) to [(S0,0,λSW(X,𝔰)/2)]\left[\left(S^{0},0,-\lambda_{SW}(X,\mathfrak{s})/2\right)\right]. ∎

7. Obstruction to embeddings of 3-manifolds into 4-manifolds with PSC metric

Theorem 1.1 gives an obstruction to embedding of 3-manifolds into 4-manifolds with PSC metric under a homological assumption. By a standard surgery argument enables us to prove the following generalization of Theorem 1.1.

Theorem 7.1.

Let (X,𝔰)(X,\mathfrak{s}) be an oriented spin closed connected 4-manifold with b2(X)=0b_{2}(X)=0 and YY a smooth oriented closed codimension-1 submanifold of XX. Suppose b1(Y)=0b_{1}(Y)=0 and XX admits a PSC metric. Then the local equivalence class of SWF(Y,𝔱)\mathrm{SWF}(Y,\mathfrak{t}) is given by

(59) [SWF(Y,𝔱)]=[(S0,0,δ(Y,𝔱)2)],\displaystyle[\mathrm{SWF}(Y,\mathfrak{t})]=\left[\left(S^{0},0,-\frac{\delta(Y,\mathfrak{t})}{2}\right)\right],

where 𝔱:=𝔰|Y\mathfrak{t}:=\mathfrak{s}|_{Y}.

This theorem can be seen as a Seiberg–Witten analogue of the result proven by Yang–Mills gauge theory [T19, Theorem 1.9]. Using the Heegaard Floer correction term, Levine–Ruberman [LR19] gave an obstruction of codimension-1 smooth embeddings into homology S1×S3S^{1}\times S^{3}’s. For the obstructions to codimension-1 smooth embeddings into indefinite spin 4-maniolds, see [PMK17].

Proof.

We argue the case that [Y]0[Y]\neq 0 and that [Y]=0[Y]=0 individually. First, let us assume [Y]0[Y]\neq 0. In this case, the cobordism W0:=XY¯W_{0}:=\overline{X\setminus Y} from YY to itself is connected. When b2(W0)=0b_{2}(W_{0})=0, one can see XX is a rational homology S1×S3S^{1}\times S^{3} and [Y][Y] generates H3(X)H_{3}(X). Thus, by Theorem 1.1, one has

[SWF(Y,𝔱)]=[(S0,0,λSW(X,𝔰)2)]=[(S0,0,δ(Y,𝔱)2)].[\mathrm{SWF}(Y,\mathfrak{t})]=\left[\left(S^{0},0,-\frac{\lambda_{SW}(X,\mathfrak{s})}{2}\right)\right]=\left[\left(S^{0},0,-\frac{\delta(Y,\mathfrak{t})}{2}\right)\right].

When b2(W0)>1b_{2}(W_{0})>1, we take disjoint simple closed curves l1,,lb2(W0)l_{1},\cdots,l_{b_{2}(W_{0})} in XX which generate H2(W0;)H_{2}(W_{0};\mathbb{Z}). We extend l1,,lb2(W0)l_{1},\cdots,l_{b_{2}(W_{0})} to disjoint smooth embeddings from S1×D3S^{1}\times D^{3}’s into W0W_{0} and denote them by the same notations. We consider the manifold

W0(l1,,lb2(W0))W_{0}({l_{1},\cdots,l_{b_{2}(W_{0})}})

obtained by the surgery of W0W_{0} along l1lb2(W0)l_{1}\cup\cdots\cup l_{b_{2}(W_{0})}. One can see W0(l1,,lb2(W0))W_{0}({l_{1},\cdots,l_{b_{2}(W_{0})}}) also admits a spin structure. We write the glued manifold along the boundary of W0(l1,,lb2(W0))W_{0}({l_{1},\cdots,l_{b_{2}(W_{0})}}) by X(l1,,lb2(W0))X({l_{1},\cdots,l_{b_{2}(W_{0})}}).

Since we are considering codimension-3 surgeries, [GL80, Theorem A] implies that X(l1,,lb2(W0))X({l_{1},\cdots,l_{b_{2}(W_{0})}}) also admits a PSC metric. The manifold X(l1,,lb2(W0))X({l_{1},\cdots,l_{b_{2}(W_{0})}}) is a spin rational homology S1×S3S^{1}\times S^{3}. By construction, YY is smoothly embedded into X(l1,,lb2(W0))X({l_{1},\cdots,l_{b_{2}(W_{0})}}) such that

0[Y]H3(X(l1,,lb2(W0));).0\neq[Y]\in H_{3}(X({l_{1},\cdots,l_{b_{2}(W_{0})}});\mathbb{Z})\cong\mathbb{Z}.

An easy observation shows that [Y][Y] generates H3(X(l1,,lb2(W0));)H_{3}(X({l_{1},\cdots,l_{b_{2}(W_{0})}});\mathbb{Z}). Thus one can use Theorem 1.1 and see

[SWF(Y,𝔱)]=[(S0,0,λSW(X(l1,,lb2(W0)))2)]=[(S0,0,δ(Y,𝔱)2)].[\mathrm{SWF}(Y,\mathfrak{t})]=\left[\left(S^{0},0,-\frac{\lambda_{SW}(X({l_{1},\cdots,l_{b_{2}(W_{0})}}))}{2}\right)\right]=\left[\left(S^{0},0,-\frac{\delta(Y,\mathfrak{t})}{2}\right)\right].

Next, we consider the case [Y]=0[Y]=0. In this case, our cobordism W0W_{0} should have two connected components: W0+W0W_{0}^{+}\cup W_{0}^{-}. Suppose W0+=Y\partial W_{0}^{+}=Y and W0=Y\partial W_{0}^{-}=-Y. By 1-handle surgery, one can assume that W0+W_{0}^{+} and W0W_{0}^{-} are spin rational homology D4D^{4}’s. Thus the relative Bauer–Furuta invariants BFW0+BF_{W_{0}^{+}} and BFW0BF_{W_{0}^{-}} gives rise to the local equivalence between SWF(Y,𝔱)SWF(Y,\mathfrak{t}) and [(S0,0,0)]\left[\left(S^{0},0,0\right)\right]. This completes the proof. ∎

Corollary 7.2.

Let YY be an integral homology 33-sphere. Suppose that at least two of α(Y),β(Y),γ(Y),δ(Y),δ¯(Y),δ¯(Y),κ(Y)\alpha(Y),\beta(Y),\gamma(Y),\delta(Y),\overline{\delta}(Y),\underline{\delta}(Y),\kappa(Y) do not coincide with each other. Then YY does not admit any smooth embedding into a spin closed 4-manifold with a PSC metric satisfying b2(X)=0b_{2}(X)=0.

Remark 7.3.

Freedman’s result ([F82]) implies that all homology 3-spheres have a locally flat embedding into S4S^{4}, and Corollary 7.2 is false for locally flat topological embeddings.

8. Examples

In this section we use Corollary 1.4 to obtain a concrete family of 4-manifolds which does not admit PSC metrics. In order to use Corollary 1.4, we need to calculate the homology cobordism invariants α\alpha, β\beta, γ\gamma, δ\delta. The following remark gives a method to calculate δ\delta for a large class of 3-manifolds:

Remark 8.1.

In [LRS18, Remark 1.1], it is mentioned that Heegaad Floer correction term d(Y,𝔰)d(Y,\mathfrak{s}) and the monopole Frøyshov invariant h(Y,𝔰)h(Y,\mathfrak{s}) satisfy

d(Y,𝔰)=2h(Y,𝔰),d(Y,\mathfrak{s})=-2h(Y,\mathfrak{s}),

for any spinc rational homology 3-sphere (Y,𝔰)(Y,\mathfrak{s}). Moreover, it is proved in [LM18] that

h(Y,𝔰)=δ(Y,𝔰).\displaystyle-h(Y,\mathfrak{s})=\delta(Y,\mathfrak{s}).

Therefore one can use calculations of correction terms in Heegaard Floer theory ([OS03, BN13, Tw13, KS19]) in order to calculate δ(Y,𝔰)\delta(Y,\mathfrak{s}).

For the invariants α\alpha, β\beta and γ\gamma, we mainly use Stoffregen’s computation results [Sto20] for Seifert homology 3-spheres and connected sums of them.

Before considering to the connected sum, we start with a single Seifert homology 3-sphere. The following result is proved by Stoffregen [Sto20]. Recall that a Seifert rational homology 3-sphere YY is called negative if the underlying orbifold line bundle of YY is of negative degree (see [Sto20, Section 5]).

Theorem 8.2 ([Sto20]).

The following results hold.

  • (i)

    Let YY be a Seifert homology 3-sphere with negative fibration. Then

    β(Y)=γ(Y)=μ¯(Y), and \beta(Y)=\gamma(Y)=-\overline{\mu}(Y),\ \text{ and }
    α(Y)={d(Y)/2=δ(Y) if d(Y)/2μ¯(Y)mod2d(Y)/2+1=δ(Y)+1 otherwise \alpha(Y)=\begin{cases}d(Y)/2=\delta(Y)\text{ if }d(Y)/2\equiv-\overline{\mu}(Y)\operatorname{mod}2\\ d(Y)/2+1=\delta(Y)+1\text{ otherwise }\end{cases}

    hold.

  • (ii)

    Let YY be a Seifert homology 3-sphere with positive fibration. Then

    α(Y)=β(Y)=μ¯(Y), and \alpha(Y)=\beta(Y)=-\overline{\mu}(Y),\ \text{ and }
    γ(Y)={d(Y)/2=δ(Y) if d(Y)/2μ¯(Y)mod2d(Y)/21=δ(Y)1 otherwise \gamma(Y)=\begin{cases}d(Y)/2=\delta(Y)\text{ if }d(Y)/2\equiv-\overline{\mu}(Y)\operatorname{mod}2\\ d(Y)/2-1=\delta(Y)-1\text{ otherwise }\end{cases}

    hold.

Combining Corollary 1.3 with Theorem 8.2, we obtain:

Theorem 8.3.

Let YY^{\prime} be a Seifert homology 3-sphere such that

μ¯(Y)δ(Y),-\overline{\mu}(Y^{\prime})\neq\delta(Y^{\prime}),

where μ¯\overline{\mu} is the Neumann–Siebenmann invariant for graph homology 3-spheres introduced in [N80, Si80]. Let YY be an oriented homology 3-sphere which is homology cobordant to YY^{\prime}. Then, for any homology cobordism WW from YY to itself, the 4-manifold obtained from WW by gluing the boundary components does not admit a PSC metric.

The invariant μ¯\overline{\mu} has a concrete recursion formula for Σ(a1,,an)\Sigma(a_{1},\cdots,a_{n}). See [Sa02, (2.8), (2.9) in Subsection 2.4.2]. Although (2.8) and (2.9) in [Sa02, Subsection 2.4.2] are formulae for the Rochlin invariant, it is pointed out in [Sa02, page 197] that the same formula holds also for the invariant μ¯\overline{\mu}. We also note another way to compute μ¯\overline{\mu} based on the ww-invariant. For the definition of ww-invariant, see [Fuk09, Definition 2.2]. In [Fuk09], the ww-invariants of several types of Seifert homology 3-spheres are computed, and the following relation is given in [Sa07, FFU01, Fuk00]: for any Seifert homology 3-sphere of type Σ(2,q,r)\Sigma(2,q,r),

w(Σ(2,q,r),X(2,q,r),𝔰)=μ¯(Σ(2,q,r)).w(\Sigma(2,q,r),X(2,q,r),\mathfrak{s})=-\overline{\mu}(\Sigma(2,q,r)).

Here (X(2,q,r),𝔰)(X(2,q,r),\mathfrak{s}) is a certain spin 4-orbifold. For the unique way to construct X(2,q,r)X(2,q,r), see the sentences after [Fuk09, Theorem 3.1].

Also, in [Sto20], there are direct computations of α\alpha, β\beta and γ\gamma. Using them, we can prove:

Corollary 8.4.

Suppose a homology 3-sphere YY is homology cobordant to one of Seifert homology 3-spheres with types:

(2,3,12k1),(2,3,12n+7),(2,5,20k+11),(2,5,20k1),(2,5,20k3),\displaystyle(2,3,12k-1),(2,3,12n+7),(2,5,20k+11),(2,5,20k-1),(2,5,20k-3),
(2,5,20k+13),(2,7,28k1),(2,7,28k+15),(2,7,28k3),(2,7,14k+3),\displaystyle(2,5,20k+13),(2,7,28k-1),(2,7,28k+15),(2,7,28k-3),(2,7,14k+3),
and (2,7,14k5).\displaystyle\text{ and }(2,7,14k-5).

Then, for any homology cobordism WW from YY to itself, the 4-manifold obtained from WW by gluing the boundary components does not admit a PSC metric.

Proof.

We just combine computation results [Fuk09, Ma16, Sto20] of α,β,γ\alpha,\beta,\gamma and μ¯\overline{\mu} and Corollary 1.4. ∎

Remark 8.5.

We remark that for homology S1×S3S^{1}\times S^{3}’s obtained as mapping tori, enlargeable obstruction [GL80] can be used to obstruct PSC metrics. A large class of homology S1×S3S^{1}\times S^{3} which are not obtained as mapping tori are introduced in [KT20, Subsection 4.4.1]. Also, a review of several known obstructions for homology S1×S3S^{1}\times S^{3}’s is given in [KT20, Subsection 4.4].

Next, we consider the connected sums of Seifert homology 3-spheres. In order to obtain a certain connected sum formula of invariants α\alpha, β\beta and γ\gamma for Seifert homology 3-spheres, Stoffregen considered a class of Seifert homology 3-spheres, called projective type. We call a negative Seifert rational homology 3-sphere YY with a spin structure 𝔰\mathfrak{s} projective if its Heegaard Floer homology is of the form

HF+(Y,𝔰)𝒯d+𝒯2n+1+(m)iI𝒯ai+(mi)2HF^{+}(Y,\mathfrak{s})\cong\mathcal{T}^{+}_{d}\oplus\mathcal{T}_{-2n+1}^{+}(m)\oplus\bigoplus_{i\in I}\mathcal{T}_{a_{i}}^{+}(m_{i})^{\oplus 2}

for some nn, mm, dd, aia_{i}, mim_{i} and some index set II, where

  • 𝒯+:=𝔽[U,U1]/𝔽[U]\mathcal{T}^{+}:=\mathbb{F}[U,U^{-1}]/\mathbb{F}[U], where 𝔽\mathbb{F} is the field of two elements,

  • 𝒯+(i):=𝔽[Ui+1,Ui+2]/𝔽[U]\mathcal{T}^{+}(i):=\mathbb{F}[U^{-i+1},U^{-i+2}]/\mathbb{F}[U], and

  • 𝒯d+(n):=𝒯+(n)\mathcal{T}^{+}_{d}(n):=\mathcal{T}^{+}(n) whose grading is shifted by d-d.

There are many examples of projective Seifert homology 3-spheres [Ne07, BN13, Tw13]. It is confirmed in [Ne07, BN13, Tw13] that Σ(p,q,pqk±1)\Sigma(p,q,pqk\pm 1) is projective for a relatively prime pair (p,q)(p,q) and positive integer kk.

Theorem 8.6 ([Sto20]).

Let Y1,,YnY_{1},\cdots,Y_{n} be negative Seifert homology 3-spheres of projective type. Suppose δ(Y1)δ(Yn)\delta(Y_{1})\leq\cdots\leq\delta(Y_{n}). Set δ~i:=δ(Yi)+μ¯(Yi)\widetilde{\delta}_{i}:=\delta(Y_{i})+\overline{\mu}(Y_{i}). Then

  • α(Y1##Yn)=2i=1nδ~i+12i=1nμ¯(Yi)\alpha(Y_{1}\#\cdots\#Y_{n})=2\lfloor\frac{\sum_{i=1}^{n}\widetilde{\delta}_{i}+1}{2}\rfloor-\sum_{i=1}^{n}\overline{\mu}(Y_{i})

  • β(Y1##Yn)=2i=1n1δ~i+12i=1nμ¯(Yi)\beta(Y_{1}\#\cdots\#Y_{n})=2\lfloor\frac{\sum_{i=1}^{n-1}\widetilde{\delta}_{i}+1}{2}\rfloor-\sum_{i=1}^{n}\overline{\mu}(Y_{i})

  • γ(Y1##Yn)=2i=1n2δ~i+12i=1nμ¯(Yi)\gamma(Y_{1}\#\cdots\#Y_{n})=2\lfloor\frac{\sum_{i=1}^{n-2}\widetilde{\delta}_{i}+1}{2}\rfloor-\sum_{i=1}^{n}\overline{\mu}(Y_{i})

Combining Corollary 1.3 with Theorem 8.6, we can prove:

Theorem 8.7.

Let Y1,,YnY_{1},\cdots,Y_{n} be negative Seifert homology 3-spheres of projective type. Suppose δ(Y1)δ(Yn)\delta(Y_{1})\leq\cdots\leq\delta(Y_{n}). Set δ~i:=δ(Yi)+μ¯(Yi)\widetilde{\delta}_{i}:=\delta(Y_{i})+\overline{\mu}(Y_{i}). Suppose that at least two of the following four integers are distinct:

i=1nδ(Yi),2i=1nδ~i+12i=1nμ¯(Yi),\displaystyle\sum_{i=1}^{n}{\delta}(Y_{i}),\quad 2\lfloor\frac{\sum_{i=1}^{n}\widetilde{\delta}_{i}+1}{2}\rfloor-\sum_{i=1}^{n}\overline{\mu}(Y_{i}),
2i=1n1δ~i+12i=1nμ¯(Yi),2i=1n2δ~i+12i=1nμ¯(Yi).\displaystyle 2\lfloor\frac{\sum_{i=1}^{n-1}\widetilde{\delta}_{i}+1}{2}\rfloor-\sum_{i=1}^{n}\overline{\mu}(Y_{i}),\quad 2\lfloor\frac{\sum_{i=1}^{n-2}\widetilde{\delta}_{i}+1}{2}\rfloor-\sum_{i=1}^{n}\overline{\mu}(Y_{i}).

Let YY be an oriented homology 3-sphere which is homology cobordant to Y1##YnY_{1}\#\cdots\#Y_{n}. Then, for any homology cobordism WW from YY to itself, the 4-manifold obtained from WW by gluing the boundary components does not admit a PSC metric.

For a concrete family, one can see the following non-existence of PSC metrics for connected sums:

Corollary 8.8.

Suppose a homology 3-sphere YY is homology cobordant to one of homology 3-spheres:

  • #jΣ(2,3,12n1)\#_{j}\Sigma(2,3,12n-1),

  • #jΣ(2,5,20n1)\#_{j}\Sigma(2,5,20n-1), and

  • #jΣ(2,7,28n1)\#_{j}\Sigma(2,7,28n-1)

for some j>0j\in\mathbb{Z}_{>0}, where #jY\#_{j}Y means the connected sum of jj-copies of YY. Then, for any homology cobordism WW from YY to itself, the 4-manifold obtained from WW by gluing the boundary components does not admit a PSC metric.

Proof.

As it is calculated in [Sto20], one has

  • α(Σ(2,3,12n1))=2,β(Σ(2,3,12n1))=0,γ(Σ(2,3,12n1))=0,\alpha(\Sigma(2,3,12n-1))=2,\beta(\Sigma(2,3,12n-1))=0,\gamma(\Sigma(2,3,12n-1))=0, μ¯(Σ(2,3,12n1))=1,δ(Σ(2,3,12n1))=1\overline{\mu}(\Sigma(2,3,12n-1))=1,\delta(\Sigma(2,3,12n-1))=1,

  • α(Σ(2,5,20n1))=2,β(Σ(2,5,20n1))=0,γ(Σ(2,5,20n1))=0,\alpha(\Sigma(2,5,20n-1))=2,\beta(\Sigma(2,5,20n-1))=0,\gamma(\Sigma(2,5,20n-1))=0, μ¯(Σ(2,5,20n1))=1,δ(Σ(2,5,20n1))=1\overline{\mu}(\Sigma(2,5,20n-1))=1,\delta(\Sigma(2,5,20n-1))=1, and

  • α(Σ(2,7,28n1))=2,β(Σ(2,7,28n1))=0,γ(Σ(2,7,28n1))=0,\alpha(\Sigma(2,7,28n-1))=2,\beta(\Sigma(2,7,28n-1))=0,\gamma(\Sigma(2,7,28n-1))=0, μ¯(Σ(2,7,28n1))=1,δ(Σ(2,7,28n1))=2\overline{\mu}(\Sigma(2,7,28n-1))=1,\delta(\Sigma(2,7,28n-1))=2.

Since Σ(p,q,pq±1)\Sigma(p,q,pq\pm 1) are projective, it follows from from Theorem 8.6 that

  • α(Σ(2,3,12n1))=2j+12,β(Σ(2,3,12n1))=2j2,\alpha(\Sigma(2,3,12n-1))=2\lfloor\frac{j+1}{2}\rfloor,\beta(\Sigma(2,3,12n-1))=2\lfloor\frac{j}{2}\rfloor,

    γ(Σ(2,3,12n1))=2j12,δ(#jΣ(2,3,12n1))=j\gamma(\Sigma(2,3,12n-1))=2\lfloor\frac{j-1}{2}\rfloor,\delta(\#_{j}\Sigma(2,3,12n-1))=j,

  • α(Σ(2,5,20n1))=2j+12,β(Σ(2,5,20n1))=2j2,\alpha(\Sigma(2,5,20n-1))=2\lfloor\frac{j+1}{2}\rfloor,\beta(\Sigma(2,5,20n-1))=2\lfloor\frac{j}{2}\rfloor,

    γ(Σ(2,5,20n1))=2j12,δ(#jΣ(2,5,20n1))=j\gamma(\Sigma(2,5,20n-1))=2\lfloor\frac{j-1}{2}\rfloor,\delta(\#_{j}\Sigma(2,5,20n-1))=j, and

  • α(Σ(2,7,28n1))=22j+12,β(Σ(2,7,28n1))=2j\alpha(\Sigma(2,7,28n-1))=2\lfloor\frac{2j+1}{2}\rfloor,\beta(\Sigma(2,7,28n-1))=2j, γ(Σ(2,7,28n1))=22j12,\gamma(\Sigma(2,7,28n-1))=2\lfloor\frac{2j-1}{2}\rfloor, δ(#jΣ(2,7,28n1))=j\delta(\#_{j}\Sigma(2,7,28n-1))=j.

Therefore, in these cases, the assumptions of Theorem 8.7 are satisfied, and Theorem 8.7 implies the desired conclusion. ∎

Remark 8.9.

We expect that the connected Seiberg–Witten Floer homology SWFHconn(Y,𝔰)SWFH_{\rm conn}(Y,\mathfrak{s}) introduced in [Sto20] can be used to obstruct PSC metrics. Also, the equivariant KO-theoretic homology cobordism invariants introduced in [Lin15] should give another obstruction.

References