Positive scalar curvature and homology cobordism invariants
Abstract.
We determine the local equivalence class of the Seiberg–Witten Floer stable homotopy type of a spin rational homology 3-sphere embedded into a spin rational homology with a positive scalar curvature metric so that generates the third homology. The main tool of the proof is a relative Bauer–Furuta-type invariant on a periodic-end -manifold. As a consequence, we give obstructions to positive scalar curvature metrics on spin rational homology , 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 -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 -sphere equipped with a spinc structure, and recovers the monopole Floer homology defined by Kronheimer and Mrowka [KM07] for this class of -manifolds [LM18]. Therefore, in principle, the Seiberg–Witten Floer stable homotopy type contains all Floer-theoretic information from Seiberg–Witten theory for rational homology -spheres.
In this paper, we will consider a spin rational homology -sphere embedded into a spin -manifold with the rational homology of so that the fundamental class of generates . The main theorem of this paper states that, if admits a metric with positive scalar curvature (PSC), we can determine the Seiberg–Witten Floer stable homotopy type of such , denoted by , up to the local equivalence relation explained below. This result gives a strong obstruction to PSC metrics of spin rational homology , and this is the authors’ original motivation for this study. To the same class of -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 following [Ma16], Manolescu’s invariants [Ma14, Ma16], and Stoffregen’s invariants [Sto171]. These invariants have different applications, for example: The Frøyshov invariant was used to extend Donaldson’s diagonalization theorem [Do83] to negative-definite -manifolds with boundary [Fr96, Fr10, Ma03]. Manolescu used the invariant to disprove the triangulation conjecture [Ma16], and used to extend Furuta’s 10/8-inequality [Fu01] to spin -manifolds with boundary [Ma14]. Stoffregen’s invariants should correspond, respectively, to in involutive Heegaard Floer homology [HM17], using -equivariant ordinary cohomology.
These invariants are spin rational homology cobordism invariants, and obtained from described above. However, these invariants factor through a weaker invariant than , the local equivalence class of , defined by Stoffregen [Sto20]. The local equivalence is an equivalence relation on a certain class of spaces including for rational homology -spheres , and this is an abstraction of a relation between and for and which are spin rational homology cobordant to each other. To summarize this situation, let us denote by the -dimensional homology cobordism group, and denote by the -dimensional spin rational homology cobordism group. Namely, an element of is the equivalence class of a spin rational homology -sphere, and the equivalence relation is given by a spin rational homology cobordism. Stoffregen [Sto20] introduced the local equivalence group , which consists of the local equivalence classes of certain spaces modeled on . Then one has group homomorphisms
For a spin rational homology -sphere , the local equivalence class is valued in , and the above numerical invariants factor through , such as :
1.2. Main theorem
As described, the local equivalence class of is, so far at least, a candidate of the ‘universal’ Seiberg–Witten theoretic homology cobordism invariant of : it contains information of all known homology cobordism invariants obtained from Seiberg–Witten theory. The main theorem of this paper determines when is embedded into a spin rational homology admitting a PSC metric so that generates :
Theorem 1.1.
Let be an oriented spin rational homology , and be an oriented spin rational homology -sphere. Suppose that is a cross-section of , i.e. is embedded into so that it represents a fixed generator of , and that is isomorphic to . Assume that admits a PSC metric. Then the local equivalence class of is given by
(1) |
In particular, for an arbitrary spin rational homology cobordism invariant which factors through , the invariant of coincides with the invariant of the right-hand side of (1).
Here is the Casson-type invariant defined by the Mrowka–Ruberman–Saveliev [MRS11] for an integral homology , which was later generalized for a rational homology by J. Lin–Ruberman–Saveliev [LRS17]. Recall that an element of is expressed as the class of a triple , where is a space of type SWF [Ma16, Definition 2.7], and , .
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 , and compare this with known obstructions on PSC metrics for the same class of -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 -spheres admit PSC metrics: only connected sums of spherical -manifolds. Rational homology is a class of -manifold that may be seen to be closed to rational homology -sphere, but it is not easy to rule out the existence of PSC metrics on such -manifolds. In dimension , the Seiberg–Witten invariant is known as a powerful obstruction to PSC metric, but it cannot be used to rational homology , since the Seiberg–Witten invariant is not well-defined for such -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 in [Lin19], and later this result was generalized by himself with Ruberman and Saveliev to any rational homology in [LRS17]. J. Lin’s obstruction is described as follows: under the same assumption with Theorem 1.1, one has the equality
(2) |
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 be an oriented spin rational homology , and be an oriented spin rational homology -sphere. Suppose that is a cross-section of . Assume that admits a PSC metric. Then we have
(3) |
Proof.
By the definition of [Ma16, Ma14, Sto171], it is easy to see that the values of these invariants for the right-hand side of (1) are given by . Therefore the Corollary directly follows from Theorem 1.1. ∎
Note that, by Corollary 1.3, we can replace in the right-hand side of (1) with various invariants of .
An obvious consequence of Corollary 1.3 is:
Corollary 1.4.
Let be an oriented homology 3-sphere. Suppose that at least two of do not coincide with each other. Then, for any homology cobordism from to itself, the homology obtained from by gluing the boundary components does not admit a PSC metric.
Here we drop the unique spin structure from our notation for (integral) homology -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 be as in Theorem 1.1. Take a compact smooth spin -manifold bounded by . Suppose that is a cross-section of . Assume that admits a PSC metric. Then we have
(4) |
Proof.
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 -theory, whereas Manolescu’s inequality is based on -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 , to consider finite-dimensional approximations of the Seiberg–Witten equations on a periodic-end -manifold. More precisely, we shall construct a relative Bauer–Furuta-type invariant over a half-periodic-end -manifold
along the spirit of Furuta [Fu01], Bauer–Furuta [BF04], and Manolescu [Ma03]. Here is the -manifold defined by cutting open along , 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 . 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 with such a periodic PSC metric on the end looks like a homology cobordism from to from Seiberg–Witten theoretic point of view. The relative Bauer–Furuta invariant over gives a local map from to . The quantity emerges from the spin Dirac index over , discussed in Subsection 3.3.
Similarly, by considering the relative Bauer–Furuta invariant over
we get a local map from to , and we can conclude that is locally equivalent to .
1.5. Examples
In Section 8 we shall give examples of concrete 3-manifolds 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 having a cross-section with does not admit a PSC metric. Here 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 be a Seifert homology 3-sphere such that
where is the Neumann–Siebenmann invariant for graph homology 3-spheres, introduced in [N80, Si80]. Let be an oriented homology 3-sphere which is homology cobordant to . Then, for any homology cobordism from to itself, the 4-manifold obtained from by gluing the boundary components does not admit a PSC metric.
-
(ii)
Let be negative Seifert homology 3-spheres of projective type. Suppose that . Set . Suppose that at least two of following four integers are distinct:
Let be an oriented homology 3-sphere which is homology cobordant to . Then, for any homology cobordism from to itself, the 4-manifold obtained from 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 . In Section 2 we give several notations related to infinite cyclic covering spaces of a -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 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 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 -manifolds into -manifolds admitting PSC metrics. In Section 8 we provide several families of examples of homology ’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 -manifolds. Let be an oriented spin rational homology , i.e. a spin -manifold whose rational homology is isomorphic to that of . Fix a Riemannian metric on and a generator of , denoted by . Note that is isomorphic to , and hence to . Let be an oriented rational homology -sphere, and assume that is embedded into so that . We call such a cross-section of . Let be the rational homology cobordism from to itself obtained by cutting open along . The manifold is equipped with an orientation and a spin structure induced by those of . For with , we define the periodic -manifold
where is a copy of for each . This -manifold is also equipped with an orientation and a spin structure as well as . The element of corresponding to via the Poincaré duality gives the isomorphism class of an infinite cyclic covering
(5) |
and an identification
(6) |
Via the identification (6), let us think of as a map from to . Define the map as the restriction of . We call an object defined on , such as connection, metric, bundle, and differential operator, a periodic object if the restriction of the object to can be identified with the pull-back of an object on under . Considering the pull-back under , the Riemannian metric on induces a Riemannian metric, denoted by , on . Let be the positive/negative spinor bundles respectively over with respect to the metric and the spin structure above. Fixing a trivialization of the determinant line bundle of the spin structure on , we obtain the canonical reference connection on corresponding to the trivial connection.
To consider the weighted Sobolev norms on , fix a function
with , where is the deck transform determined by . Note that defined a cohomology class which is equal to corresponding to via the Poincaré duality.
Definition 2.1.
Let be a periodic vector bundle on with a periodic inner product. For a fixed and , we define the weighted Sobolev norm by
for a smooth comactly supported section of . Here we used a periodic metric and a periodic connection on to define the -norm. Let denote the -completion of compactly supported smooth sections of .
Note that the equivalence class of norms does not depend on the choices of a periodic metric and a periodic connection on .
2.2. Fredholm theory on
In this Subsection we review the Fredholm property of periodic elliptic operators on the infinite cyclic covering developed by C. Taubes [T87]. He showed that a periodic elliptic operator is Fredholm under some condition with respect to -norms for generic . For the details, see [T87], or [KT20, Subsection 2.1].
Let be a periodic elliptic complex on , i.e. the complex
(7) |
consisting of first order periodic linear differential operators between periodic vector bundles on with exact symbol sequence. Here, for a vector bundle , the notation denotes the set of compactly supported smooth sections of . As well as Definition 2.1, define the weighted Sobolev norm on by
using a periodic connection and a periodic metric. The complex (7) gives rise to the complex of bounded operators
(8) |
for each and .
Note that, since the operators in (8) are periodic differential operators, there exist differential operators on such that is given as the pull-back .
Definition 2.2.
For , define the complex by
where the operators are defined by
Theorem 2.3 ([T87, Lemmas 4.3 and 4.5]).
Suppose that there exists where the complex is acyclic. Then there exists a discrete subset in with no accumulation points such that the complex (8) is an acyclic complex for all in .
Definition 2.4 ([RS07]).
We call an admissible metric on if the kernel of
is zero, where the map is a smooth classifying map of (5).
The admissibility condition does not depend on the choice of classifying map . One can show that every PSC metric on 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 with respect to the pull-back of an admissible metric on .
-
•
The Atiyah–Hitchin–Singer complex
-
•
The de Rham complex
Remark 2.6.
Since the subset of given in Theorem 2.3 has no accumulation points, we can take a sufficiently small so that for any the operators in Lemma 2.5 are Fredholm. Henceforth we fix the notation .
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 be an oriented rational homology -sphere with a Riemannian metric . Let be a spinc structure on , and be the spinor bundle of . We fix a flat spinc reference connection of the determinant line bundle of .
Definition 2.7.
For an integer , we define the configuration space by
The Chern–Simons–Dirac functional is deined by
where is the Dirac operator with respect to the connection .
The gauge group and a subgroup of are defined by
and
The gauge group naturally acts on and the functional is invariant under the action. The global slice of the action of on is given by
on which we still have the remaining -action. We often drop and/or from our notation to denote . The -equivariant formal gradient flow on of with respect to the Coulomb projection of the -metric can be written as the sum of the linear term
and some quadratic term, denoted by .
For , we define as the direct sum of eigenspaces of , regarded as an unbounded operator on , whose eigenvalues belong to . Here we think of as a subspace of . We denote by
the -projection of onto . We often abbreviate as . Since is the sum of a real operator and a complex operator, decomposes into a real vector space and a complex vector space, denoted by
(9) |
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 of sufficiently large radius centered at the origin is an isolating neighborhood of the invariant part of the ball. Precisely, the flow on considered here is a flow obtained from by cutting off outside a larger ball (see [Ma03, page 907]). We denote by the -equivariant Conley index of the invariant part. The Seiberg–Witten Floer homotopy type is defined as the triple , which is symbolically denoted by
The triple is regarded as an object a certain suspension category . In general an object of is given as a triple , where is a pointed topological -space, , and . The quantity is defined to be
(10) |
where is a compact 4-manifold satisfying and 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 comes from a spin structure. In this case, the formal gradient flow of admits a larger symmetry of the group defined by
This group acts on for any non-negative integer as follows: the -action on spinors given as the restriction of the natural -action on spinor bundles, and the -action on is given via the non-trivial homomorphism . We denote by the real -dimensional representation of , and by the space of quaternions, on which naturally acts. Thus we have decompositions
and
(11) |
Considering -equivariant Conley index instead, we obtain a stable homotopy type of a pointed -space
which lies in a suspension category . An object of is given as a triple , where is a pointed topological -space, , and .
Let us recall the definition of local equivalence.
Definition 2.8 ([Sto20]).
For two objects and in , a local map is a -equivariant map
for some and such that the -invariant part is a -homotopy equivalence. Two objects and are locally equivalent if there exist local maps and .
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
In this subsection we describe the Seiberg–Witten equations on , mainly to fix notations. We use the double Cloumb gauge condition introduced in [Kha15].
Definition 2.9.
Let be a positive integer with and a positive real number. We first define the configuration space by
The gauge group is given by
(12) |
Here denotes the trivial bundle over with fiber . The action of on is given by
The double Coulomb slice introduced in [Kha15] is defined by
where
Here denotes the restriction of -forms as differential forms and is the formal adjoint of with respect to .
We will prove that gives a global slice with respect to the action of on . Note that, on , the ‘full gauge group’
also acts. Thus we have an additional -symmetry on coming from the limits with respect to the end of gauge transformations.
Based on the Sobolev embedding , we can naturally define the group structure on by pointwise multiplication.
On , one can define the Seiberg–Witten map
(13) |
by
(14) |
When we write , we often decompose the Seiberg–Witten map as the sum of the linear part
(15) |
the quadratic part
We regard also as an operator with domain by the restriction. The quadratic part is a compact operator by [Lin19, Proposition 2.13] for a positive . The differential equation
(16) |
is called the Seiberg–Witten equation for . The linearlization of is given by .
3. Linear analysis on
Fix a Riemann metric on such that
-
•
is periodic and PSC, and
-
•
is product metric near .
3.1. Fredholm theory on
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 , the spinor bundles are written as and . In this section, we use the following completions:
We prove the Fredholm properties of the following two types of operators on :
-
•
the Atiyah–Hitchin–Singer operator with APS-boundary condition:
(17) where
-
(i)
the space is the -completion of the negative eigenspaces of the operator
-
(ii)
the map is the restriction,
-
(iii)
the operator
is the -projection to .
-
(i)
-
•
the Dirac operator with APS-boundary condition:
(18) where
-
(i)
the space is the -completion of the negative eigenspaces of the operator
-
(ii)
the map is the restriction,
-
(iii)
the operator
is the -projection to .
-
(i)
We first prove the following proposition:
Proposition 3.1.
The following facts hold:
-
(i)
For any , the operator (18) is Fredholm.
-
(ii)
Let be a positive real number given in Remark 2.6. For any , 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
is an isomorphism for any , and we get a continuous inverse . By patching a local parametrix of (18) near the boundary and , we obtain a parametrix of (18). This implies the conclusion.
-
•
Next, we prove (ii). By Lemma 2.5,
is an acyclic complex for , where is a discrete subset of given in Theorem 2.3. This implies that
is an isomorphism for . Since does not have accumulation points, there exists a small positive real number such that
Then the remaining part is the same as the proof of (i).
∎
Set
and consider the operators
over . Here is defined in (15), and is defined by
It follows from Proposition 3.1 that the operator is Fredholm for all .
Proposition 3.2.
Let be the positive real number given in Remark 2.6. For any , we obtain
where denotes the algebraic cokernel. In particular, is Fredholm and the index of coincides with that of .
Proof.
The proof is essentially the same as the proof in [Khan15]. First, by the choice of , Proposition 3.1 implies that is Fredholm. Set
and let
be the -orthogonal projection, and consider an operator
(19) |
as an intermediary between the two operators in the statement of the Proposition.
We first show that
(20) |
Set
and let
be the operator defined by
We denote the -completion of by the same notation. Then we have
and
Let be the span of non-positive eigenvectors of . As shown in [Khan15], the projection is an isomorphism, and hence so is
Thus we obtain the following commutative diagram between functional spaces over :
From this diagram we obtain the isomorphisms (20). Moreover, as noted, it follows from Proposition 3.1 that the operator is Fredholm. Therefore this diagram implies that is also Fredholm.
The remaining task is to show that
(21) |
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 , we can obtain (21):
∎
We consider a Riemannian manifold
obtained by gluing the half-cylinder with along their boundary. We will compare formal adjoints for several weights, and would like to introduce a family of weight functions
such that
Definition 3.3.
Let . For a bundle which is periodic on and cylindrical on , we define the norm by
and define to be the completion of compactly supported sections with respect to .
Note that the formal adjoint with respect to of is given as
We also consider the ‘sliced’ Atiyah–Hitchin–Singer operator with APS-boundary condition:
(22) |
We calculate the kernel and the cokernel of (22) :
Theorem 3.4.
There exists such that for any , the operator (22) is an isomorphism.
We take the constant to be smaller than 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 . First we shall calculate the kernel and the cokernel of
(23) |
The following lemma can be proved by considering the similar discussion given in [APSI].
Lemma 3.5.
We have the following identifications:
Proof.
By the same discussion in [APSI, Proposition 3.11], a solution under the spectral boundary condition can be identified with an -solution on a cylindrical end manifold. Thus one has an isomorphism
and the cokernel can be written by using extended -solutions:
Here we used . On the other hand, implies is a constant and should be zero. This gives the conclusion. ∎
Lemma 3.6.
For sufficiently small, the space
can be identified with the middle cohomology of
Proof.
By the exponential decay result, one can see the correspondence
gives an identification
On the other hand, for an appropriate , the complex
(24) |
is Fredholm, and we can identify the kernel of
with the middle cohomology of (24). ∎
The exponential decay result enables us to prove the following correspondence:
Lemma 3.7.
For sufficiently small, the space
can be identified with
Proof.
The proof is the same as in that of Lemma 3.6. ∎
Lemma 3.8.
We have that
(25) |
and one can construct an injective homomorphism
for 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
and show that this map factors through . The proof uses the condition . This map is defined by choosing a bump function such that
-
•
and
-
•
.
For a given , one can assume is smooth. Since and , one can choose smooth functions on and on such that
Since and , up to adding constants, one can assume
Define
One can see that induces a homomorphism
The same argument given in [T87, (5.8)-(5.10)] shows that is an injection. Under our assumption, the cohomology
is generated by
which is constant near and , where and as oriented manifolds.
Next we show that . Suppose that . Then we have
for some and for some . Up to image , we can assume . Thus we have
here we consider and as constant extensions on the ends. This implies
Since goes to on the ends, this gives a contradiction.
Combining this with the injectivity of , we have that the domain of is . Here Lemma 3.6 implies that the domain of is isomorphic to the left-hand side of (25), and now the first assertion of the Lemma follows.
On the second assertion, a homomorphism
is given as follows. First note that the domain of is identified with the corresponding functional space for the weight because of Lemma 3.7. Take a sequence of bump functions satisfying
For a given , one can see
Since and , one can choose and such that
Define
The proof of the injectivity of 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) |
and
(27) |
By integration by parts, one has
The vanishing (26) follows from this and the first assertion of Lemma 3.8.
In the proof of Theorem 1.1, we also consider the ‘right-periodic’ manifold . Fix a Riemann metric on such that
-
•
is periodic and PSC, and
-
•
is product metric near .
Let us consider the following operators:
-
•
the Atiyah–Hitchin–Singer operator with APS-boundary condition:
(28) -
•
the linearlization of the Seiberg–Witten equation
(29)
Theorem 3.9.
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 such that for any positive real number ,
(30) |
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) |
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
We need to prove that, for any , there exists such that , i.e.
hold. These equations are equivalent to
where is the Green operator on . Therefore we need to prove surjectivity of the map
defined by
In order to prove this, we use the excision principle and reduce the surjectivity of 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 and obtain the conclusion. ∎
Proposition 3.10 implies the following global slice theorem:
Lemma 3.11.
Let be the constant given in Proposition 3.10. Then, for , there is a -equivariant diffeomorphism
The proof is the essentially same as in the case of closed 4-manifolds.
3.3. Dirac index on
In this Subsection, we shall calculate the spin Dirac index on the half-periodic -manifold :
Proposition 3.12.
Assuming that a PSC metric is equipped with , we have
(32) |
where means the index of the Dirac operator under the APS-boundary condition and is given in (10).
Before proving Proposition 3.12, we note a few lemmas:
Lemma 3.13.
Let be compact spin -manifolds with common boundary with orientation . Equip with metrics so that the metrics are the product metric
near the boundary for a Riemann metric on , where is a collar coordinate of the product neighborhood and pr means the projection from the collar neighborhoods of to . Then we have
where is the -dimensional Dirac operator on and 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 which is not a spectrum of , denote by the Fredholm index defined using the weighted Sobolev norm described as
at the end of . Take so that is smaller than the absolute value of the smallest non-zero eigenvalue of . Then we obtain
by the gluing formula, Equation (3.2) of [Do02]. Hence it suffices to show that
(33) |
By the definition of the APS-boundary condition, we have
On the other hand, we have that
by [Do02, Proposition 3.10], which is shown considering a certain ordinary equation [Do02, Lemma 3.11] corresponding to the cylinder appearing the neck stretching of . Now we have checked (33) and this completes the proof. ∎
The proof of Lemma 3.13 involves only near the neck of . Even if we replace 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 be a compact spin manifold bounded by with the orientation . Equip with a metric so that the metrics are product metrics near the boundary. Then we have
(34) |
where
Now we are ready to prove Proposition 3.12.
Proof of Proposition 3.12.
Take a compact spin bound of with the orientation . Take a metric on so that the metrics are product metrics near the boundary.
Now we shall check
(35) |
Note the sign: this comes from the orientation of with . Indeed, by [Lin19, Lemma 2.21], it follows from the existence of PSC metric on that
(36) |
(Precisely, is supposed to be an integral homology in [Lin19], but the proof of [Lin19, Lemma 2.21] is valid also for rational homology ’s without any changes.) On the other hand, for a PSC metric on , we have
(37) |
On the other hand, we also have
(38) |
Indeed, it follows that
(39) |
because of Lemma 3.13 and
By the definition of , we have
(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 and a constant such that the following conclusion holds. Let be an element in . Suppose that a pair of
and satisfy the following conditions:
-
(i)
the element is a solution to the equation (16) on ,
-
(ii)
the element is a solution to the Seiberg–Witten equations on ,
-
(iii)
is temporal gauge, i.e. for each , where , and is of finite type,
-
(iv)
, and
-
(v)
.
Then we have the following universal bounds:
In order to prove Theorem 4.1, we use several corresponding notions used in [Ma03].
Definition 4.2.
We consider a Riemannian manifold
obtained by gluing the half-cylinder and along their boundary. A solution to the Seiberg–Witten equations on is called -trajectories. If a -trajectory satisfies
then is called a finite type -trajectory.
Let us note the following boundedness result:
Theorem 4.3.
Let be a positive real number and
be a solution to such that
and
Then, there exists such that for any , the inequality
holds, where is a constant depending only on .
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 and a gauge transformation on such that for any ,
Define
On the other hand, by Lemma 3.11, the map obtained by giving a slice
is continuous. This implies there is a gauge transformation such that
and
This gives the desired result. ∎
The topological energy and the analytic energy for configurations on are defined along the book by Kronheimer–Mrowka [KM07, Definition 4.5.4]. Note that, for a configuration converging to 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 is a -trajectory and is asymptotic to on the cylindrical end, we have that
(41) |
where depends only on and the fixed metric and spin structure on . Moreover, we have that as well as for a configuration over a compact -manifold.
Proof of Theorem 4.1.
Let be the constant given in Theorem 4.3. Suppose that
satisfies the assumption of Theorem 4.1. First, we state a pasting lemma:
Lemma 4.4.
The pair gives rise to a finite type -trajectory .
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 to the Seiberg–Witten equations on whose topological energy is finite.
Recall that the set of critical points of modulo gauge is compact. Since we consider a spin structure now, is gauge invariant. Therefore the set of critical values of is compact.
Since we have assumed that , we have that
as , and therefore there exists a critical point of to which is -asymptotic as . This combined with (41) implies that is uniformly bounded, and hence so is .
We claim that the analytic energy of restricted to is also uniformly bounded. To see this, let us decompose into three parts: the periodic part , the cylindrical part , and the ‘joint’ between the periodic part and the cylindrical part. We have seen that the analytic energy of on is uniformly bounded, and this energy is the sum of the energies on these three parts. Therefore, to prove that the analytic energy of restricted to 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 :
for any . The remaining boundedness result follows from the same argument for -trajectories, where is a compact -manifold bounded by . 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 . We mainly follow the methods given by Manolescu [Ma03] and Khandhawit [Kha15].
We consider a finite-dimensional approximation of the map
We fix a weight satisfying
in the rest of this paper, where are the constants appeared in Remark 2.6, Theorem 3.4, Proposition 3.10, and Theorem 4.1. Take sequences of subspaces
such that
-
(i)
-
(ii)
the -projection satisfies
for any .
Then we define a sequence of subspaces
This gives a family of the approximated Seiberg–Witten maps
The following proposition gives us a well-defined continuous map between spheres.
Proposition 5.1.
For a large and a large positive real number , there exists an index pair of and a sequence of positive numbers such that
(42) |
is well-defined, where is the closed ball in with radius and is the sphere in with radius .
For the proof of Proposition 5.1, we use the following proposition.
Proposition 5.2.
Let be a bounded sequence in such that
and
Let be a uniformly bounded sequence of trajectories such that
Then, after taking a subsequence, converges to a solution (in the topology of ) and converges to in which is a solution of the Seiberg–Witten equations on .
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 , after taking a subsequence, uniformly converges to in , where is the weak limit.
For the sequence , we need to ensure:
-
•
after taking a subsequence, in , where is the weak limit and
-
•
after taking a subsequence, the sequence converges to in .
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 , we will use the following inequality obtained by the Fredholm property of : there exists a constant such that, for any ,
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
We obtain a map stably written by
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 to , and also a local map from to .
We shall consider the relative Bauer–Furuta invariant on the ‘left-periodic’ manifold and that on ‘right-periodic’ manifold . 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 , we constructed a -equivariant continuous map of the form
(43) |
as the relative Bauer–Furuta invariant over . One sees that induces a -homotopy equivalence by Theorem 3.4. The numbers are given by
(44) |
For the notations and , 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.
Next, instead of the ‘left-periodic’ manifold , we consider the ‘right-periodic’ manifold
Repeating analysis in Section 5 for instead of , we obtain a -map of the form
(45) |
as the relative Bauer–Furuta invariant over . Here denotes the Conley index for . As well as above, induces a -homotopy equivalence by Theorem 3.9. For , as in [Ma16, Proof of Proposition 3.8], let us denote by the vector space defined for . Note that, for , we have an identification , and in particular . Under this notation, are given by
(46) |
By an argument using a duality map as in [Ma16, page 168], we obtain a -map
from (45). The vector space can be decomposed so that . Set
(47) |
Then the domain and codomain of are given by
(48) |
We shall show that gives a local map from to . The restriction is a -homotopy equivalence since is so. One may assume was taken to avoid the eigenvalues of the linearization of the flow equations. Then we have
(49) |
Here, to obtain the first equality, we have used deduced from the assumption that . Using (LABEL:eq:_zero_eig_proof_of_main), we have
(50) |
and
(51) |
Combining (46) with (LABEL:eq:_prf_main_double_pr), (50) and (51), we obtain
(52) |
(53) |
Let us calculate in the last equality. Let be an oriented compact smooth -manifold with boundary . Set
Then, as well as Lemma 3.14, we obtain
(54) |
On the other hand, for a PSC metric on , we have
(55) |
Recalling the definition of , we have
(56) |
Combining (54) with (55) and (56), we have
(57) |
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 be an oriented spin closed connected 4-manifold with and a smooth oriented closed codimension-1 submanifold of . Suppose and admits a PSC metric. Then the local equivalence class of is given by
(59) |
where .
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 ’s. For the obstructions to codimension-1 smooth embeddings into indefinite spin 4-maniolds, see [PMK17].
Proof.
We argue the case that and that individually. First, let us assume . In this case, the cobordism from to itself is connected. When , one can see is a rational homology and generates . Thus, by Theorem 1.1, one has
When , we take disjoint simple closed curves in which generate . We extend to disjoint smooth embeddings from ’s into and denote them by the same notations. We consider the manifold
obtained by the surgery of along . One can see also admits a spin structure. We write the glued manifold along the boundary of by .
Since we are considering codimension-3 surgeries, [GL80, Theorem A] implies that also admits a PSC metric. The manifold is a spin rational homology . By construction, is smoothly embedded into such that
An easy observation shows that generates . Thus one can use Theorem 1.1 and see
Next, we consider the case . In this case, our cobordism should have two connected components: . Suppose and . By 1-handle surgery, one can assume that and are spin rational homology ’s. Thus the relative Bauer–Furuta invariants and gives rise to the local equivalence between and . This completes the proof. ∎
Corollary 7.2.
Let be an integral homology -sphere. Suppose that at least two of do not coincide with each other. Then does not admit any smooth embedding into a spin closed 4-manifold with a PSC metric satisfying .
Remark 7.3.
Freedman’s result ([F82]) implies that all homology 3-spheres have a locally flat embedding into , 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 , , , . The following remark gives a method to calculate for a large class of 3-manifolds:
Remark 8.1.
In [LRS18, Remark 1.1], it is mentioned that Heegaad Floer correction term and the monopole Frøyshov invariant satisfy
for any spinc rational homology 3-sphere . Moreover, it is proved in [LM18] that
Therefore one can use calculations of correction terms in Heegaard Floer theory ([OS03, BN13, Tw13, KS19]) in order to calculate .
For the invariants , and , 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 is called negative if the underlying orbifold line bundle of is of negative degree (see [Sto20, Section 5]).
Theorem 8.2 ([Sto20]).
The following results hold.
-
(i)
Let be a Seifert homology 3-sphere with negative fibration. Then
hold.
-
(ii)
Let be a Seifert homology 3-sphere with positive fibration. Then
hold.
Combining Corollary 1.3 with Theorem 8.2, we obtain:
Theorem 8.3.
Let be a Seifert homology 3-sphere such that
where is the Neumann–Siebenmann invariant for graph homology 3-spheres introduced in [N80, Si80]. Let be an oriented homology 3-sphere which is homology cobordant to . Then, for any homology cobordism from to itself, the 4-manifold obtained from by gluing the boundary components does not admit a PSC metric.
The invariant has a concrete recursion formula for . 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 . We also note another way to compute based on the -invariant. For the definition of -invariant, see [Fuk09, Definition 2.2]. In [Fuk09], the -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 ,
Here is a certain spin 4-orbifold. For the unique way to construct , see the sentences after [Fuk09, Theorem 3.1].
Also, in [Sto20], there are direct computations of , and . Using them, we can prove:
Corollary 8.4.
Suppose a homology 3-sphere is homology cobordant to one of Seifert homology 3-spheres with types:
Then, for any homology cobordism from to itself, the 4-manifold obtained from by gluing the boundary components does not admit a PSC metric.
Proof.
We just combine computation results [Fuk09, Ma16, Sto20] of and and Corollary 1.4. ∎
Remark 8.5.
We remark that for homology ’s obtained as mapping tori, enlargeable obstruction [GL80] can be used to obstruct PSC metrics. A large class of homology which are not obtained as mapping tori are introduced in [KT20, Subsection 4.4.1]. Also, a review of several known obstructions for homology ’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 , and 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 with a spin structure projective if its Heegaard Floer homology is of the form
for some , , , , and some index set , where
-
•
, where is the field of two elements,
-
•
, and
-
•
whose grading is shifted by .
There are many examples of projective Seifert homology 3-spheres [Ne07, BN13, Tw13]. It is confirmed in [Ne07, BN13, Tw13] that is projective for a relatively prime pair and positive integer .
Theorem 8.6 ([Sto20]).
Let be negative Seifert homology 3-spheres of projective type. Suppose . Set . Then
-
•
-
•
-
•
Combining Corollary 1.3 with Theorem 8.6, we can prove:
Theorem 8.7.
Let be negative Seifert homology 3-spheres of projective type. Suppose . Set . Suppose that at least two of the following four integers are distinct:
Let be an oriented homology 3-sphere which is homology cobordant to . Then, for any homology cobordism from to itself, the 4-manifold obtained from 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 is homology cobordant to one of homology 3-spheres:
-
•
,
-
•
, and
-
•
for some , where means the connected sum of -copies of . Then, for any homology cobordism from to itself, the 4-manifold obtained from by gluing the boundary components does not admit a PSC metric.
Proof.
As it is calculated in [Sto20], one has
-
•
,
-
•
, and
-
•
.
Since are projective, it follows from from Theorem 8.6 that
-
•
,
-
•
, and
-
•
, .
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 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.