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

Simply-connected manifolds with large homotopy stable classes

Anthony Conway Massachusetts Institute of Technology, Cambridge MA 02139 anthonyyconway@gmail.com Diarmuid Crowley School of Mathematics and Statistics, University of Melbourne, Australia dcrowley@unimelb.edu.au Mark Powell Department of Mathematical Sciences, Durham University, United Kingdom mark.a.powell@durham.ac.uk  and  Joerg Sixt sixtj@yahoo.de
Abstract.

For every k2k\geq 2 and n2n\geq 2 we construct nn pairwise homotopically inequivalent simply-connected, closed 4k4k-dimensional manifolds, all of which are stably diffeomorphic to one another. Each of these manifolds has hyperbolic intersection form and is stably parallelisable. In dimension 44, we exhibit an analogous phenomenon for spinc structures on S2×S2S^{2}\times S^{2}.

For m1m\geq 1, we also provide similar (4m1)(4m{-}1)-connected 8m8m-dimensional examples, where the number of homotopy types in a stable diffeomorphism class is related to the order of the image of the stable JJ-homomorphism π4m1(SO)π4m1s\pi_{4m-1}(SO)\to\pi^{s}_{4m-1}.

Key words and phrases:
Stable diffeomorphism, homotopy equivalence, 4k4k-manifold
1991 Mathematics Subject Classification:
Primary 57R65, 57R67.

1. Introduction

Let qq be a positive integer and let Wg:=#g(Sq×Sq)W_{g}:=\#_{g}(S^{q}\times S^{q}) be the gg-fold connected sum of the manifold Sq×SqS^{q}\times S^{q} with itself. Two compact, connected smooth 2q2q-manifolds M0M_{0} and M1M_{1} with the same Euler characteristic are stably diffeomorphic, written M0stM1M_{0}\cong_{\text{st}}M_{1}, if there exists a non-negative integer gg and a diffeomorphism

M0#WgM1#Wg.M_{0}\#W_{g}\to M_{1}\#W_{g}.

Note that Sq×SqS^{q}\times S^{q} admits an orientation-reversing diffeomorphism. Hence the same is true of WgW_{g} and it follows that when the MiM_{i} are orientable the diffeomorphism type of the connected sum does not depend on orientations.

A paradigm of modified surgery, as developed by Kreck [Kre99], is that one first seeks to classify 2q2q-manifolds up to stable diffeomorphism, and then for each M0M_{0} one tries to understand its stable class:

𝒮st(M0):={M1M1stM0}/diffeomorphism.\mathcal{S}^{\rm st}(M_{0}):=\{M_{1}\mid M_{1}\cong_{\text{st}}M_{0}\}/\text{diffeomorphism.}

The efficacy of this method was first demonstrated by Hambleton and Kreck, who applied it to 44-manifolds with finite fundamental group in a series of papers [HK88a, HK88b, HK93b, HK93a].

On the other hand, the Browder-Novikov-Sullivan-Wall surgery exact sequence [Wal99] aims instead to classify manifolds within a fixed homotopy class. In general there is no obvious relationship between homotopy equivalence and stable diffeomorphism, although in some cases there are implications e.g. [Dav05]. To enable a comparison between the two approaches, we define the homotopy stable class of M0M_{0} to be

𝒮hst(M0)={M1M1stM0}/homotopy equivalence.\mathcal{S}^{\rm st}_{\rm h}(M_{0})=\{M_{1}\mid M_{1}\cong_{\text{st}}M_{0}\}/\text{homotopy equivalence}.

Our aim is to investigate the cardinality of 𝒮hst(M0)\mathcal{S}^{\rm st}_{\rm h}(M_{0}), and in particular we shall exhibit new examples of simply-connected manifolds with arbitrarily large homotopy stable class.

Throughout this article we will consider closed, connected, simply-connected, smooth manifolds. In order to define the intersection form and related invariants we orient all manifolds. When necessary, to achieve unoriented results, we will later factor out by the effect of the choice of orientation.

When the dimension is 4k+24k{+}2, Kreck showed that the stable class of such manifolds is trivial [Kre99, Theorem D]. We therefore focus on dimensions 4k4k with k>1k>1 (dimension 4 will be discussed separately below). Kreck also showed that for every such simply-connected manifold M4kM^{4k}, the stable class of M4k#W1M^{4k}\#W_{1} is trivial. But as pointed out by Kreck and Schafer [KS84, I], for k>1k\!>\!1 examples of closed, simply-connected (2k1)(2k{-}1)-connected 4k4k-manifolds MM with arbitrarily large homotopy stable class have been implicit in the literature since Wall’s classification of these manifolds up to the action of the group of homotopy spheres [Wal62]. These examples are distinguished by their intersection form

λM:H2k(M;)×H2k(M;),\lambda_{M}\colon H_{2k}(M;\mathbb{Z})\times H_{2k}(M;\mathbb{Z})\to\mathbb{Z},

which must be definite (in order to have inequivalent forms) and in order to realise the forms by closed, almost-parallelisable manifolds they must have signature divisible by 8|bP4k|8|bP_{4k}|, where |bP4k||bP_{4k}| is the order of the group of homotopy (4k1)(4k{-}1)-spheres which bound parallelisable manifolds [MK60, Corollary on p. 457].

In this paper we consider examples where the intersection form is isomorphic to the standard hyperbolic form

H+()=(2,(0110))H^{+}(\mathbb{Z})=\left(\mathbb{Z}^{2},\left(\begin{array}[]{cc}0&1\\ 1&0\end{array}\right)\right)

and where there is an additional invariant, a homomorphism f:2f\colon\mathbb{Z}^{2}\to\mathbb{Z}. The pair (H+(),f)(H^{+}(\mathbb{Z}),f) is an example of an extended symmetric form; see Definition  3.5. The isometries of the rank two hyperbolic form are highly restricted: they are generated by switching the two basis vectors and multiplying both basis vectors by 1-1. As such the unordered pair

{a,b}:={f(1,0),f(0,1)}/(±1),\{a,b\}:=\{f(1,0),f(0,1)\}/(\pm 1),

considered up to multiplication of both integers by 1-1, gives an invariant of the isometry class of the extended symmetric form (H+(),f)(H^{+}(\mathbb{Z}),f). On the other hand, in the Witt class, or stable equivalence class, only the divisibility d:=gcd(a,b)d:=\gcd(a,b) and the product A=abA=ab are invariants. Since a fixed number AA can often be factorised in many ways as a product of coprime integers a,ba,b, if we can define a suitable ff, this simple algebra has the chance to detect large stable classes. In the proof of our first main theorem, we will define such an ff using the cohomology ring of the manifolds we construct.

Theorem 1.1.

Fix positive integers nn and k2k\geq 2. There are infinitely many stable diffeomorphism classes of closed, smooth, simply-connected 4k4k-manifolds {[Mi]st}i=1\{[M_{i}]_{\rm st}\}_{i=1}^{\infty}, such that |𝒮hst(Mi)|n|\mathcal{S}^{\rm st}_{h}(M_{i})|\geq n. Moreover 𝒮hst(Mi)\mathcal{S}^{\rm st}_{h}(M_{i}) contains a subset {Mij}j=1n\{M^{j}_{i}\}_{j=1}^{n} of cardinality nn, where Mi1=MiM^{1}_{i}=M_{i}, and each MijM^{j}_{i} is stably parallelisable and has hyperbolic intersection form.

Here stably parallelisable means that the tangent bundle becomes trivial after taking the Whitney sum with a trivial bundle of sufficiently high rank. More than one notion of stabilisation appears in this article, one for manifolds and one for vector bundles.

Kreck and Schafer [KS84] constructed examples of 4k4k-manifolds MM with nontrivial finite fundamental groups, such that the homotopy stable class of MM contains distinct elements with hyperbolic intersection forms. However as far as we know our construction gives the first simply-connected examples and the first for which the homotopy stable class has been shown to have arbitrary cardinality. In a companion paper [CCPS21], we will investigate the homotopy stable class in more detail, also for manifolds with nontrivial fundamental group, and we shall relate the homotopy stable class to computations of the \ell-monoid from [CS11].

The manifolds we construct in order to prove Theorem  1.1 are shown to be homotopically inequivalent using their cohomology rings. An alternative construction to obtain nontrivial homotopy stable class instead uses Pontryagin classes to define the homomorphism ff in an extended symmetric form. This was alluded to in [KS84], but not carried through. Section 3 proves a theorem which implies the following result.

Theorem 1.2.

For every m1m\geq 1 there exists a pair of closed, smooth, (4m1)(4m{-}1)-connected 8m8m-manifolds M1M_{1} and M2M_{2} with hyperbolic intersection forms, that are stably diffeomorphic but not homotopy equivalent.

Compared with the manifolds from Theorem 1.1 (for even mm, in the notation of that theorem), the manifolds M1M_{1} and M2M_{2} from Theorem 1.2 are not stably parallelisable, but on the other hand since they are (4m1)(4m{-}1)-connected and have the same intersection pairing, their cohomology rings are isomorphic. In particular, once again the intersection form does not help.

To show that the manifolds in Theorem 1.2 are not homotopy equivalent, we use Wall’s homotopy classification of (4m1)(4m{-}1)-connected 8m8m-manifolds [Wal62, Lemma 8], which makes use of an extended symmetric form (H+(),f:2/jm)(H^{+}(\mathbb{Z}),f\colon\mathbb{Z}^{2}\to\mathbb{Z}/j_{m}), where jmj_{m} is the order of the image of the stable JJ-homomorphism J:π4m1(SO)π4m1sJ\colon\pi_{4m-1}(SO)\to\pi^{s}_{4m-1}; see Section 3.

Remark 1.3.

The limiting factor preventing us from exhibiting arbitrarily large homotopy stable classes in Theorem 1.2 is that our lower bound depends only on the number of primes dividing jmj_{m}. This grows with mm, but in a fixed dimension cannot be made arbitrarily large. On the other hand, if we instead count diffeomorphism classes, then we show in Theorem 3.3 (2) that the stable class can be arbitrarily large for (4m1)(4m{-}1)-connected 8m8m-manifolds with hyperbolic intersection forms.

Dimension 4

Dimension 4 was absent from the above discussion. This is because closed, smooth, simply-connected 4-manifolds MM and NN are stably diffeomorphic if and only if they are homotopy equivalent. Here is an outline of why this holds. First, two such 4-manifolds are stably diffeomorphic if and only if there are orientations such that they have the same signatures, Euler characteristics, and w2w_{2}-types i.e. σ(M)=σ(N)\sigma(M)=\sigma(N), χ(M)=χ(N)\chi(M)=\chi(N), and their intersection forms have the same parity (even or odd). Thus homotopy equivalence implies stable diffeomorphism. For the other direction, σ(M)=σ(N)\sigma(M)=\sigma(N) and χ(M)=χ(N)\chi(M)=\chi(N) implies that the intersection forms are either both definite or both indefinite. In the definite case, the intersection forms must be diagonal by Donaldson’s theorem [Don83], and so the intersections forms are isometric and therefore the manifolds are homotopy equivalent [Whi49, Mil58a]. In the indefinite case, the intersection form is determined up to isometry by its rank, parity, and signature, and so again MM and NN are homotopy equivalent. Thus the assumption that k2k\geq 2 was essential in Theorem 1.1.

One way in which an analogous phenomenon does occur in dimension 4 is by considering spinc structures. Seiberg-Witten invariants of 4-manifolds and Heegaard-Floer cobordism maps are indexed by spinc structures. The first Chern class c1c_{1} of the spinc structure then defines the map ff in the extended symmetric forms. We illustrate this in Section 4, using the 4-manifold S2×S2S^{2}\times S^{2}.

Theorem 1.4.

Let CC\in\mathbb{Z} with |C|16|C|\geq 16 and 8C8\mid C. Define P(C)P(C) to be the number of distinct primes dividing C/8C/8. There are n:=2P(C)1n:=2^{P(C)-1} stably equivalent spinc structures 𝔰1,,𝔰n\mathfrak{s}_{1},\dots,\mathfrak{s}_{n} on S2×S2S^{2}\times S^{2} with c1(𝔰i)2=CH4(S2×S2)c_{1}(\mathfrak{s}_{i})^{2}=C\in H^{4}(S^{2}\times S^{2})\cong\mathbb{Z}, that are all pairwise inequivalent.

Organisation

Section 2 proves Theorem 1.1, Section 3 proves Theorem 1.2, and Section 4 proves Theorem 1.4.

Conventions

Throughout this paper all manifolds are compact, simply-connected, and smooth. As mentioned above we will also equip our manifolds with an orientation. For the remainder of this paper all (co)homology groups have integral coefficients. We write 0:={0}\mathbb{N}_{0}:=\mathbb{N}\cup\{0\}.

Acknowledgements

We would like to thank Manuel Krannich for advice about the homotopy sphere ΣQ\Sigma_{Q}, Jens Reinhold for comments on an earlier draft of this paper, and Csaba Nagy for pointing out a mistake in a previous version of the proof of Theorem 4.11.

MP is grateful to the Max Planck Institute for Mathematics in Bonn, where he was a visitor while this paper was written. MP was partially supported by EPSRC New Investigator grant EP/T028335/1 and EPSRC New Horizons grant EP/V04821X/1.

2. Simply-connected 4k4k-manifolds with arbitrarily large stable class

We prove Theorem 1.1 by stating and proving Proposition 2.2 below. In the proposition, we construct a collection of 4k4k-manifolds Na,bN_{a,b}, for each unordered pair of positive integers {a,b}\{a,b\} such that (2k)!(2k)! divides 2ab2ab. If {a,b}{a,b}\{a,b\}\neq\{a^{\prime},b^{\prime}\}, then Na,bN_{a,b} and Na,bN_{a^{\prime},b^{\prime}} are not homotopy equivalent. On the other hand ab=abab=a^{\prime}b^{\prime} if and only if Na,bN_{a,b} and Na,bN_{a^{\prime},b^{\prime}} are stably diffeomorphic. Moreover every manifold Na,bN_{a,b} is closed, simply-connected, has hyperbolic intersection form, and is stably parallelisable. Thus the proposition immediately implies Theorem 1.1.

First we have a lemma. In order to rule out orientation-reversing homotopy equivalences, we shall appeal to the following observation.

Lemma 2.1.

Let NN and NN^{\prime} be closed, oriented 4k4k-manifolds. Suppose that a class zz freely generates H2(N)H^{2}(N) and satisfies that z2k=nz^{2k}=n for some nonzero n=H4k(N)n\in\mathbb{Z}=H^{4k}(N), and similarly for (N,z)(N^{\prime},z^{\prime}). Then any homotopy equivalence f:NNf\colon N\to N^{\prime} must be orientation preserving.

Proof.

Assume that ff is of degree ε=±1\varepsilon=\pm 1. Since ff is a homotopy equivalence, NN and NN^{\prime} have isomorphic cohomology rings. In particular H2(N)H^{2}(N^{\prime})\cong\mathbb{Z} is generated by z=(f)1(z)z^{\prime}=(f^{*})^{-1}(z). Since z2k=n{z^{\prime}}^{2k}=n in H4k(N)H^{4k}(N^{\prime})\cong\mathbb{Z}, and f(z2k)=z2kf^{*}({z^{\prime}}^{2k})=z^{2k}, properties of the cap and cup products show that

n=f(z2k[N])=f(f(z2k)[N])=z2kf([N])=z2kε[N]=εn.n=f_{*}(z^{2k}\cap[N])=f_{*}(f^{*}({z^{\prime}}^{2k})\cap[N])={z^{\prime}}^{2k}\cap f_{*}([N])={z^{\prime}}^{2k}\cap\varepsilon[N^{\prime}]=\varepsilon n.

Since n0n\neq 0, this implies that ff must be orientation-preserving. ∎

Now we proceed with the construction of the promised manifolds.

Proposition 2.2.

Fix k>1k>1. Given an unordered pair {a,b}\{a,b\} of positive coprime integers such that (2k)!(2k)! divides 2ab2ab, there exists a closed, oriented, 4k4k-manifold Na,b4kN^{4k}_{a,b} with the following properties.

  1. (i)

    The manifold Na,bN_{a,b} is simply-connected and stably parallelisable.

  2. (ii)

    The ring H(Na,b)H^{*}(N_{a,b}) has generators w,x,y,zw,x,y,z and 11 of degrees 2k+22k{+}22k2k, 2k2k22 and 0 respectively, with zk=ax+by,x2=0=y2,2abw=zk+1,xz=bw,yz=awz^{k}=ax+by,x^{2}=0=y^{2},2abw=z^{k+1},xz=bw,yz=aw and xyxy generates H4k(Na,b)H^{4k}(N_{a,b}).

In particular, the intersection form of Na,bN_{a,b} is hyperbolic and z2k=2abxyz^{2k}=2abxy is 2ab2ab times a fundamental class of Na,bN_{a,b}. If {a,b}{a,b}\{a,b\}\neq\{a^{\prime},b^{\prime}\} then Na,bN_{a,b} and Na,bN_{a^{\prime},b^{\prime}} have non-isomorphic integral cohomology rings and so are not homotopy equivalent. Moreover ab=abab=a^{\prime}b^{\prime} if and only if Na,bN_{a,b} and Na,bN_{a^{\prime},b^{\prime}} are stably diffeomorphic.

Proof.

Note that if we have a manifold Na,bN_{a,b} and if we choose a stable normal framing on Na,bN_{a,b}, then the pair (Na,b,z)(N_{a,b},z) corresponds to a (normally) framed manifold over P\mathbb{C}P^{\infty} using the identification H2(Na,b)[Na,b,P]H^{2}(N_{a,b})\cong[N_{a,b},\mathbb{C}P^{\infty}]. This motivates the method we shall use, constructing Na,bN_{a,b} by framed surgery on stably normally framed manifolds over P\mathbb{C}P^{\infty}. It will then follow automatically that the manifolds we obtain are stably parallelisable, since a manifold with trivial stable normal bundle has trivial stable tangent bundle too.

We start with S2S^{2} together with the unique framing of its stable normal bundle corresponding to a choice of orientation, and consider the corresponding dual orientation class αH2(S2)\alpha\in H^{2}(S^{2}). Take the 2k2k-fold product of S2S^{2} with itself,

X0:=S2××S2,X_{0}:=S^{2}\times\cdots\times S^{2},

and define β0H2(X0)\beta_{0}\in H^{2}(X_{0}) to be the class that restricts to α\alpha in each S2S^{2} factor. This means that under the inclusion

ιj:{}××S2×{}S2××S2\iota_{j}\colon\{*\}\times\cdots\times S^{2}\times\cdots\{*\}\to S^{2}\times\cdots\times S^{2}

in the  jjth factor, ιj(β0)=α\iota_{j}^{*}(\beta_{0})=\alpha. Equivalently, let pi:S2××S2S2p_{i}\colon S^{2}\times\cdots\times S^{2}\to S^{2} be the iith projection. Then β0=i=12kpi(α)\beta_{0}=\sum_{i=1}^{2k}p_{i}^{*}(\alpha). An elementary calculation shows that

β02k=(2k)![X0]H4k(X0).\beta_{0}^{2k}=(2k)![X_{0}]^{*}\in H^{4k}(X_{0}).

Here we write [X0]H4k(X0)[X_{0}]^{*}\in H^{4k}(X_{0}) for the dual of the fundamental class [X0]H4k(X0)[X_{0}]\in H_{4k}(X_{0}). To make this calculation, use β0=i=12kpi(α)\beta_{0}=\sum_{i=1}^{2k}p_{i}^{*}(\alpha) and note that:

  1. (i)

    pi(α)pj(α)=pj(α)pi(α)p_{i}^{*}(\alpha)\cup p_{j}^{*}(\alpha)=p_{j}^{*}(\alpha)\cup p_{i}^{*}(\alpha) for iji\neq j,

  2. (ii)

    pi(α)pi(α)=pi(αα)=pi(0)=0p_{i}^{*}(\alpha)\cup p_{i}^{*}(\alpha)=p_{i}^{*}(\alpha\cup\alpha)=p_{i}^{*}(0)=0, and

  3. (iii)

    p1(α)p2k(α)=[X0]p_{1}^{*}(\alpha)\cup\cdots\cup p_{2k}^{*}(\alpha)=[X_{0}]^{*}.

By assumption there is a positive integer jj such that 2ab=j(2k)!2ab=j(2k)!. Take X1:=#jX0X_{1}:=\#^{j}X_{0} to be the framed jj-fold connected sum of X0X_{0} and β1H2(X1)\beta_{1}\in H^{2}(X_{1}) to be the class that restricts to β0\beta_{0} in each summand. That is, H2(X1)jH2(X0)H^{2}(X_{1})\cong\bigoplus^{j}H^{2}(X_{0}) and β1=(β0,,β0)\beta_{1}=(\beta_{0},\dots,\beta_{0}). Then

β12k=jβ02k=j(2k)![X1]=2ab[X1]H4k(X1).\beta_{1}^{2k}=j\beta_{0}^{2k}=j(2k)![X_{1}]^{*}=2ab[X_{1}]^{*}\in H^{4k}(X_{1}).

The element β1H2(X1)\beta_{1}\in H^{2}(X_{1}) and the normal framing on X1X_{1} defines a normal map

(β1,β¯1):X1P,(\beta_{1},\overline{\beta}_{1})\colon X_{1}\to\mathbb{C}P^{\infty},

where we take the trivial bundle over P\mathbb{C}P^{\infty}. By surgery below the middle dimension, the normal map (β1,β¯1)(\beta_{1},\overline{\beta}_{1}) is normally bordant to a 2k2k-connected map (β2,β¯2):X2P(\beta_{2},\overline{\beta}_{2})\colon X_{2}\to\mathbb{C}P^{\infty}. Since X0X_{0} has signature zero, the same holds for X1X_{1} and X2X_{2}. Since the stable normal bundle of X2X_{2} is framed, so is the stable tangent bundle. Therefore the stable tangent bundle has trivial 2k2k-th Wu class vanishes and so the intersection form on X2X_{2} is even. Let zH2(P)z_{\infty}\in H^{2}(\mathbb{C}P^{\infty}) be the generator restricting to αH2(P1)=H2(S2)\alpha\in H^{2}(\mathbb{C}P^{1})=H^{2}(S^{2}) via the inclusion P1P\mathbb{C}P^{1}\to\mathbb{C}P^{\infty}, and consider the Poincaré dual of β2(zk)\beta_{2}^{*}(z_{\infty}^{k}),

u:=PD(β2(zk))H2k(X2).u:=\mathrm{PD}(\beta_{2}^{*}(z_{\infty}^{k}))\in H_{2k}(X_{2}).

Since β2:X2P\beta_{2}\colon X_{2}\to\mathbb{C}P^{\infty} is 2k2k-connected, H2k(X2)H2k(P)H_{2k}(X_{2})\to H_{2k}(\mathbb{C}P^{\infty})\cong\mathbb{Z} is onto and therefore splits since \mathbb{Z} is free. Since all homology groups are torsion-free, the dual map can be identified with the map β2:H2k(P)H2k(X2)\beta_{2}^{*}\colon H^{2k}(\mathbb{C}P^{\infty})\to H^{2k}(X_{2}) on cohomology. The splitting for β2\beta_{2} dualises to a splitting for β2\beta_{2}^{*}, so the image of a generator β2(zk)\beta_{2}^{*}(z^{k}_{\infty}) generates a summand. Applying Poincaré duality we see that uH2k(X2)u\in H_{2k}(X_{2}) is a primitive element; i.e.  uu generates a summand of H2k(X2)H_{2k}(X_{2}).

We take connected sum with an additional copy of S2k×S2kS^{2k}\times S^{2k} with null-bordant framing and trivial map to P\mathbb{C}P^{\infty} to obtain

X3:=X2#(S2k×S2k)X_{3}:=X_{2}\#(S^{2k}\times S^{2k})

and a normal map (β3,β¯3):X3P(\beta_{3},\overline{\beta}_{3})\colon X_{3}\to\mathbb{C}P^{\infty}. Note that up until this point we have only used the product abab, rather than the data of the pair {a,b}\{a,b\}. This will change for the upcoming construction of X4=Na,bX_{4}=N_{a,b}.

The intersection form λX3\lambda_{X_{3}} of X3X_{3} has an orthogonal decomposition corresponding to the connected sum decomposition of X3X_{3}:

(H2k(X3),λX3)=(H2k(X2),λX2)H+(),(H_{2k}(X_{3}),\lambda_{X_{3}})=(H_{2k}(X_{2}),\lambda_{X_{2}})\oplus H^{+}(\mathbb{Z}),

where H+()H^{+}(\mathbb{Z}) is the standard symmetric hyperbolic form. Let {e,f}\{e,f\} be a standard basis for H+()H^{+}(\mathbb{Z}). Since aa and bb are coprime, we may and shall choose integers c,dc,d such that adbc=1ad-bc=1. We also write u=PD(β3(zk))u=\mathrm{PD}(\beta_{3}^{*}(z_{\infty}^{k})). Here note that uu is essentially the same element as the element uH2(X2)u\in H_{2}(X_{2}) that we defined above thinking of H2(X2)H_{2}(X_{2}) as a subgroup of H2(X3)H_{2}(X_{3}). Keeping this in mind, we have that

λX3(u,u)=β3(zk)β3(zk),[X3]=β3(z2k),[X3]=z2k,(β3)[X3]=2ab,\lambda_{X_{3}}(u,u)=\langle{\beta_{3}^{*}(z^{k}_{\infty})\cup\beta_{3}^{*}(z^{k}_{\infty}),[X_{3}]}\rangle=\langle{\beta_{3}^{*}(z_{\infty}^{2k}),[X_{3}]}\rangle=\langle{z^{2k}_{\infty},(\beta_{3})_{*}[X_{3}]}\rangle=2ab,

since z2kz_{\infty}^{2k} generates H4k(P)H^{4k}(\mathbb{C}P^{\infty}) and since (β3)(\beta_{3})_{*} sends [X3][X_{3}] to the same multiple of the generator of H4k(P)H_{4k}(\mathbb{C}P^{\infty}) as (β1)(\beta_{1})_{*} sends [X1][X_{1}] to. Since uH2k(X2)H2k(X3)u\in H_{2k}(X_{2})\subseteq H_{2k}(X_{3}) is primitive and since λX2\lambda_{X_{2}} is nonsingular, there is an element v′′H2k(X2)H2k(X3)v^{\prime\prime}\in H_{2k}(X_{2})\subseteq H_{2k}(X_{3}) such that

λX3(u,v′′)=λX2(u,v′′)=1.\lambda_{X_{3}}(u,v^{\prime\prime})=\lambda_{X_{2}}(u,v^{\prime\prime})=1.

Now set v:=(ad+bc)v′′v^{\prime}:=(ad+bc)v^{\prime\prime} as well as

v:=v+e+2cdλX3(v,v)2f.v:=v^{\prime}+e+\frac{2cd-\lambda_{X_{3}}(v^{\prime},v^{\prime})}{2}f.

Since uH2(X2)u\in H_{2}(X_{2}) and e,fH+()e,f\in H^{+}(\mathbb{Z}), we observe that λX3(u,e)=λX3(u,f)=0\lambda_{X_{3}}(u,e)=\lambda_{X_{3}}(u,f)=0. As a consequence, the elements u,vu,v span a subspace Hu,vH2k(X3)H_{u,v}\subseteq H_{2k}(X_{3}) where λX3\lambda_{X_{3}} restricted to Hu,vH_{u,v} has matrix

A=(2abad+bcad+bc2cd),A=\left(\begin{array}[]{cc}2ab&ad+bc\\ ad+bc&2cd\end{array}\right),

which has determinant 4abcd(ad+bc)2=(adbc)2=14abcd-(ad+bc)^{2}=-(ad-bc)^{2}=-1. Hence Hu,vH_{u,v} is an orthogonal summand of (H2k(X3),λX3)(H_{2k}(X_{3}),\lambda_{X_{3}}) and a calculation shows that Hu,vH_{u,v} is hyperbolic with standard basis {e1,f1}\{e_{1},f_{1}\} where u=ae1+bf1u=ae_{1}+bf_{1} and v=ce1+df1v=ce_{1}+df_{1}. To see this, let P:=(abcd)P:=\left(\begin{smallmatrix}a&b\\ c&d\end{smallmatrix}\right) and note that P(0110)PT=AP\left(\begin{smallmatrix}0&1\\ 1&0\end{smallmatrix}\right)P^{T}=A.

The orthogonal complement of Hu,vH_{u,v}, namely Hu,vH_{u,v}^{\perp}, has signature equal to the signature of X3X_{3}, which is zero and hence since the intersection form is even, Hu,vH_{u,v}^{\perp} is stably hyperbolic.

We assert that Hu,vH_{u,v}^{\perp} maps trivially to H2k(P)H_{2k}(\mathbb{C}P^{\infty}) under β3\beta_{3*}. To see this, first note that H2k(P)H^{2k}(\mathbb{C}P^{\infty})\cong\mathbb{Z}, generated by zkz^{k}_{\infty}. We have an isomorphism

zk:H2k(P)H0().z^{k}_{\infty}\cap-\colon H_{2k}(\mathbb{C}P^{\infty})\xrightarrow{\cong}H_{0}(\mathbb{CP}^{\infty})\cong\mathbb{Z}.

Recall that now u=PD(β3(zk))Hu,vu=\mathrm{PD}(\beta_{3}^{*}(z_{\infty}^{k}))\in H_{u,v} and let xHu,vx\in H_{u,v}^{\perp}. Then

0=λX3(u,x)=PD1(u)x=β3(zk)x=zk(β3)(x).0=\lambda_{X_{3}}(u,x)=\mathrm{PD}^{-1}(u)\cap x=\beta_{3}^{*}(z_{\infty}^{k})\cap x=z_{\infty}^{k}\cap(\beta_{3})_{*}(x).

Since zkz^{k}_{\infty}\cap- is an isomorphism, this implies that (β3)(x)=0(\beta_{3})_{*}(x)=0, which proves the assertion.

Now, since β3:X3P\beta_{3}\colon X_{3}\to\mathbb{C}P^{\infty} is  2k2k-connected and since Hu,vH_{u,v}^{\perp} maps trivially to H2k(P)H_{2k}(\mathbb{C}P^{\infty}), the Hurewicz theorem and the linked long exact sequences

\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π2k+1(P,X3)\textstyle{\pi_{2k+1}(\mathbb{C}P^{\infty},X_{3})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}π2k(X3)\textstyle{\pi_{2k}(X_{3})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π2k(P)\textstyle{\pi_{2k}(\mathbb{C}P^{\infty})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H2k+1(P,X3)\textstyle{H_{2k+1}(\mathbb{C}P^{\infty},X_{3})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H2k(X3)\textstyle{H_{2k}(X_{3})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(β3)\scriptstyle{(\beta_{3})_{*}}H2k(P)\textstyle{H_{2k}(\mathbb{C}P^{\infty})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}

show that every element of Hu,vH_{u,v}^{\perp} is represented by a map from a 2k2k-sphere in π2k(X3)\pi_{2k}(X_{3}). Hence standard surgery arguments allow us to perform framed surgery on (β3,β¯3):X3P(\beta_{3},\overline{\beta}_{3})\colon X_{3}\to\mathbb{C}P^{\infty} to kill Hu,vH_{u,v}^{\perp}. We obtain a normal map (β4,β¯4):X4P(\beta_{4},\overline{\beta}_{4})\colon X_{4}\to\mathbb{C}P^{\infty}, with intersection form isomorphic to (Hu,v,λX3|Hu,v)(H_{u,v},\lambda_{X_{3}}|_{H_{u,v}}). The manifold

Na,b:=X4N_{a,b}:=X_{4}

is the required manifold, as we verify next. For the rest of the proof we shall write N:=Na,bN:=N_{a,b} for brevity. We use the orientation corresponding to the fundamental class [Na,b][N_{a,b}] induced from tracking [X0][X_{0}] through the construction.

We have already noted at the beginning of the proof that the construction via normally framed surgery implies that Na,bN_{a,b} is stably parallelisable. As the map β4:Na,bP\beta_{4}\colon N_{a,b}\to\mathbb{C}P^{\infty} is 2k2k-connected and since there is an isomorphism θ:H2k(Na,b)Hu,v2\theta\colon H_{2k}(N_{a,b})\to H_{u,v}\cong\mathbb{Z}^{2}, the manifold Na,bN_{a,b} is simply-connected and has the correct integral (co)homology groups. To verify that Na,bN_{a,b} has the required cohomology ring we set

z:=β4(z),x:=PD1(θ1(e1)),y:=PD1(θ1(f1)).z:=\beta_{4}^{*}(z_{\infty}),\quad x:=\mathrm{PD}^{-1}(\theta^{-1}(e_{1})),\quad y:=\mathrm{PD}^{-1}(\theta^{-1}(f_{1})).

Since u=ae1+bf1u=ae_{1}+bf_{1}, it follows that zk=ax+byz^{k}=ax+by. Since θ1(e1),θ1(f1)\theta^{-1}(e_{1}),\theta^{-1}(f_{1}) form a standard hyperbolic basis for (H2k(Na,b),λNa,b)(H_{2k}(N_{a,b}),\lambda_{N_{a,b}}), it follows that xyxy generates H4k(Na,b)H^{4k}(N_{a,b}) and z2k[Na,b]>0z^{2k}\cap[N_{a,b}]>0. Finally, since zk1z^{k-1} generates H2k2(Na,b)H^{2k-2}(N_{a,b})\cong\mathbb{Z}, there is a generator wH2k+2(Na,b)w\in H^{2k+2}(N_{a,b}) such that zk1w=xyz^{k-1}w=xy. The remaining properties of H(Na,b)H^{*}(N_{a,b}) follow from Poincaré duality.

Finally, let zkH2k(Na,b)\langle{z^{k}}\rangle\subseteq H^{2k}(N_{a,b}) be the subgroup generated by zkz^{k} and consider the isomorphism class of the pair (H2k(Na,b),zk)(H^{2k}(N_{a,b}),\langle{z^{k}}\rangle). This pair, modulo the action of the self-equivalences of Na,bN_{a,b} on H2k(Na,b)H^{2k}(N_{a,b}), is a homotopy invariant of Na,bN_{a,b}. Since z2k0z^{2k}\neq 0, and since z2k[Na,b]>0z^{2k}\cap[N_{a,b}]>0, every self-homotopy equivalence of Na,bN_{a,b} is orientation preserving by Lemma 2.1.

Thus zk\langle{z^{k}}\rangle modulo the action of Aut(H+())\mathrm{Aut}(H^{+}(\mathbb{Z})) is a homotopy invariant. We claim that the pair {a,b}\{a,b\} is an invariant of this action. To see this, from the form of the matrix AA above, it is easy to see that the automorphisms of the hyperbolic form are

±Id and ±(0110).\pm\operatorname{Id}\text{ and }\pm\left(\begin{array}[]{cc}0&1\\ 1&0\end{array}\right).

So automorphisms can change the sign of both aa and bb simultaneously, and they can switch aa and bb. Then since we always take a,b>0a,b>0, the unordered pair of positive integers {a,b}\{a,b\} is an invariant of the homotopy type. Hence if there is a homotopy equivalence Na,bNa,bN_{a,b}\to N_{a^{\prime},b^{\prime}}, then we have {a,b}={a,b}\{a,b\}=\{a^{\prime},b^{\prime}\}.

Now we address the final statement of the proposition, which concerns stable diffeomorphism. Observe that H2(Na,b)H2(Na,b#S2k×S2k)\mathbb{Z}\cong H^{2}(N_{a,b})\cong H^{2}(N_{a,b}\#S^{2k}\times S^{2k}), and that the image of zz, which we call zstH2(Na,b#S2k×S2k)z_{st}\in H^{2}(N_{a,b}\#S^{2k}\times S^{2k}), satisfies the equality zst2k=2ab[Na,b#S2k×S2k]z_{st}^{2k}=2ab[N_{a,b}\#S^{2k}\times S^{2k}]. Since this property of zstz_{st} and the fundamental class are preserved under diffeomorphism, it follows that if Na,bN_{a,b} and Na,bN_{a^{\prime},b^{\prime}} are stably diffeomorphic, then ab=abab=a^{\prime}b^{\prime}.

On the other hand, for a fixed product ab=abab=a^{\prime}b^{\prime}, the manifolds Na,bN_{a,b} and Na,bN_{a^{\prime},b^{\prime}} are obtained from the 4k4k-manifold X3X_{3} by surgering away a stably hyperbolic form Hu,vH_{u,v}^{\perp}. Recall that uu and vv depend on a,ba,b, so in particular we may need to stabilise a different number of times for Hu,vH_{u,v}^{\perp} versus Hu,vH_{u^{\prime},v^{\prime}}^{\perp} to make them hyperbolic. Let h(u,v)h(u,v) and h(u,v)h(u^{\prime},v^{\prime}) be the number of stabilisations required, and let h:=max{h(u,v),h(u,v)}h:=\max\{h(u,v),h(u^{\prime},v^{\prime})\}. Then for some gg we have

Na,b#WgX3#WhNa,b#Wg,N_{a,b}\#W_{g}\cong X_{3}\#W_{h}\cong N_{a^{\prime},b^{\prime}}\#W_{g},

as desired. So indeed ab=abab=a^{\prime}b^{\prime} if and only if Na,bstNa,bN_{a,b}\cong_{\text{st}}N_{a^{\prime},b^{\prime}}. ∎

3. (4m1)(4m{-}1)-connected 8m8m-manifolds with nontrivial homotopy stable class

In this section, for every m1m\geq 1 we construct (4m1)(4m{-}1)-connected 8m8m-manifolds with hyperbolic intersection form and with nontrivial homotopy stable class. Specifically, we describe certain 8m8m-manifolds Ma,bM_{a,b}, for positive integers aa and bb, and we will give bounds from above and below on the size of the homotopy stable class of Ma,bM_{a,b} in terms of aa, bb, and mm. In particular, for each mm there are infinitely many choices of a,ba,b such that |𝒮hst(Ma,b)|>1|\mathcal{S}^{\rm st}_{\rm h}(M_{a,b})|>1.

In contrast to the manifolds in the previous section, the homotopically inequivalent manifolds constructed here have isomorphic integral cohomology rings, but are not stably parallelisable. We will detect that our manifolds are not homotopy equivalent using a refinement of the mmth Pontryagin class.

This section is organised as follows. In Section 3.1 we recall some facts about exotic spheres and the JJ homomorphism, which we will need for the statement and the proof of Theorem 3.3. We state this theorem in Section 3.2. In Section 3.3 we recall Wall’s classification of (4m1)(4m{-}1)-connected 8m8m-manifolds up to the action of the group of homotopy 8m8m-spheres, then in Section 3.4 we determine the stable classification of such manifolds, again up to the action of the homotopy spheres. Next, in Section 3.5 we construct the manifolds Ma,bM_{a,b} appearing in Theorem 3.3 and we prove this theorem in Section 3.6.

3.1. Exotic spheres and the JJ-homomorphism

Let Θn\Theta_{n} denote the group of hh-cobordism classes of homotopy nn-spheres, that is closed, connected, oriented nn-manifolds that are homotopy equivalent to SnS^{n}, with the group operation given by connected sum. By [KM63] these are finite abelian groups. We will briefly recall some of what is known about them, focussing on dimensions n=8mn=8m and n=8m1n=8m{-}1, for m1m\geq 1.

Recall that bPn+1ΘnbP_{n+1}\subseteq\Theta_{n} is the subgroup of hh-cobordism classes of homotopy nn-spheres which bound parallelisable (n+1)(n{+}1)-manifolds. Kervaire and Milnor showed that this is a finite cyclic group, and for n+1=4>4n{+}1=4\ell>4 the order of bPn+1bP_{n+1} is given by a formula in terms of Bernoulli numbers and the image of the JJ-homomorphism [KM63]. Following results of Adams [Ada66] and Quillen [Qui71] on the JJ-homomorphism, this formula led to the computation of |bP4||bP_{4\ell}|; we will give more details shortly. The group bP4bP_{4\ell} is generated by the boundary of Milnor’s E8E_{8} plumbing [Bro72, V], a 44\ell-manifold obtained from plumbing disc bundles according to the E8E_{8} lattice.

Let

Jn:πn(SO)πnsJ_{n}\colon\pi_{n}(SO)\to\pi^{s}_{n}

be the stable JJ-homomorphism [Whi42, §3], where πns\pi^{s}_{n} is the stable nn-stem. Kervaire and Milnor [KM63] showed that Θ8mcokerJ8m\Theta_{8m}\cong\mathop{\rm coker}\nolimits J_{8m} and that there is a short exact sequence

0bP8mΘ8m1cokerJ8m10.0\to bP_{8m}\to\Theta_{8m-1}\to\mathop{\rm coker}\nolimits J_{8m-1}\to 0.

Later Brumfiel [Bru68] defined a splitting Θ8m1bP8m\Theta_{8m-1}\to bP_{8m} and so proved that

Θ8m1bP8mcokerJ8m1.\Theta_{8m-1}\cong bP_{8m}\oplus\mathop{\rm coker}\nolimits J_{8m-1}.

Consider a (4m1)(4m{-}1)-connected 8m8m-manifold WW with boundary WΘ8m1\partial W\in\Theta_{8m-1}. Extending work of Stolz [Sto85] and Burklund, Hahn and Senger [BHS19], Burklund and Senger [BS20, Theorem 1.2] proved that [W]bP8m[\partial W]\in bP_{8m}, except possibly when m=3m=3, when they also show that 2[W]bP242[\partial W]\in bP_{24}. For our purposes later in this section, we also assume that WW has signature 0 and this ensures that W\partial W is a multiple of the homotopy sphere denoted ΣQ\Sigma_{Q} by Krannich and Reinhold [KR20, §2] (see just below Lemma 3.9 for the definition of ΣQ\Sigma_{Q}.)

Definition 3.1.

Let 𝔟𝔭m\mathfrak{bp}_{m} be the order of ΣQ\Sigma_{Q} in Θ8m1\Theta_{8m-1}.

Remark 3.2.

The precise value of 𝔟𝔭m\mathfrak{bp}_{m} can be calculated, assuming knowledge of the relevant Bernoulli numbers, from [KR20, Lemma 2.7]. In particular, 𝔟𝔭m|bP8m|\mathfrak{bp}_{m}\mid|bP_{8m}|. This is clear when m3m\neq 3, since ΣQbP8m\Sigma_{Q}\in bP_{8m}. It follows from a direct calculation when m=3m=3, given that the projection of ΣQ\Sigma_{Q} to bP24bP_{24} has order divisible by 22.

We now recall some facts about the JJ-homomorphism for context and later use. We start with the stable JJ-homomorphism J4m1:π4m1(SO)π4m1sJ_{4m-1}\colon\pi_{4m-1}(SO)\to\pi^{s}_{4m-1} and write

jm:=|Im(J4m1)|.j_{m}:=|\operatorname{Im}(J_{4m-1})|.

For example

j1=24,j2=240,andj3=504.j_{1}=24,\quad j_{2}=240,\quad\text{and}\quad j_{3}=504.

Later we will use the fact that 4jm4\mid j_{m}, for m=1,2m=1,2, as we see here. Since the stable homotopy groups of spheres are finite, so is jmj_{m}. Since π4m1(SO)\pi_{4m-1}(SO)\cong\mathbb{Z}, in fact Im(J4m1)/jm\operatorname{Im}(J_{4m-1})\cong\mathbb{Z}/j_{m}. By [Ada66] (see e.g. [Lüc02, Theorem 6.26]), jmj_{m} can be computed using the denominator of the rational number Bm/4mB_{m}/4m, where BmB_{m}\in\mathbb{Q} is the mmth Bernoulli number, defined by the generating function

etet1=1t2+n=1(1)n+1Bn(2n)!t2n.\frac{e^{t}}{e^{t}-1}=1-\frac{t}{2}+\sum_{n=1}^{\infty}\frac{(-1)^{n+1}B_{n}}{(2n)!}t^{2n}.

By [KM63, Section 7], |bP8m|/(24m2(24m11))|bP_{8m}|/(2^{4m-2}(2^{4m-1}-1)) equals the numerator of the rational number 2B2m/m2B_{2m}/m, from which one can compute |bP8m||bP_{8m}|.

Next we consider the unstable JJ-homomorphism, J4m1,4m:π4m1(SO4m)π8m1(S4m)J_{4m-1,4m}\colon\pi_{4m-1}(SO_{4m})\to\pi_{8m-1}(S^{4m}), which, along with the stable JJ-homomorphism, the Euler class e\mathrm{e} and the Hopf-invariant HH, fits into the following commutative diagram with exact rows:

(*) 0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π4m1(SO4m)\textstyle{\pi_{4m-1}(SO_{4m})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}eS\scriptstyle{\mathrm{e}\oplus S}J4m1,4m\scriptstyle{J_{4m-1,4m}}π4m1(SO)\textstyle{\mathbb{Z}\oplus\pi_{4m-1}(SO)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}IdJ4m1\scriptstyle{\operatorname{Id}\oplus J_{4m-1}}/2\textstyle{\mathbb{Z}/2\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}=\scriptstyle{=}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π8m1(S4m)\textstyle{\pi_{8m-1}(S^{4m})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}HS\scriptstyle{H\oplus S}π4m1s\textstyle{\mathbb{Z}\oplus\pi^{s}_{4m-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}/2\textstyle{\mathbb{Z}/2\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

The commutativity of left hand square in (*3.1) is equivalent to the classical statements that e=HJ4m1,4m\mathrm{e}=H\circ J_{4m-1,4m} and that the JJ-homomorphism commutes with stabilisation [JW54, 1.2 & 1.3]. That eS\mathrm{e}\oplus S is injective with index 22 is reviewed in [Wal62, p. 171]. That the same statements hold for HSH\oplus S follows from Toda’s calculations in the exceptional cases m=1,2m=1,2 [Tod62, V, (iii) & (vii)] and from Adam’s solution of the Hopf invariant 11 problem for m>2m>2 [Ada60]. For m>2m>2, both e(π4m1(SO4m))\mathrm{e}(\pi_{4m-1}(SO_{4m}))\subseteq\mathbb{Z} and H(π8m1(S4m))H(\pi_{8m-1}(S^{4m}))\subseteq\mathbb{Z} are index two subgroups and stabilisation is a split surjection, [BM58, Ada60]. In particular this means that for m>2m>2 the Euler class ee is always even for rank 4m4m oriented vector bundles over S4mS^{4m}. When m=1,2m=1,2, the maps e\mathrm{e} and HH are both onto and e=HJ4m1,4mS\mathrm{e}=H\circ J_{4m-1,4m}\equiv S mod 22 [Wal62, p. 171] and HSH\equiv S mod 22 by Toda’s computations mentioned above. These computations show that for m=1m=1, HS:π7(S4)/12/24H\oplus S\colon\pi_{7}(S^{4})\cong\mathbb{Z}\oplus\mathbb{Z}/12\to\mathbb{Z}\oplus\mathbb{Z}/24 sends (x,y)(x,x+2y)(x,y)\mapsto(x,x+2y). For m=2m=2, the map HS:π15(S8)/120/240H\oplus S\colon\pi_{15}(S^{8})\cong\mathbb{Z}\oplus\mathbb{Z}/120\to\mathbb{Z}\oplus\mathbb{Z}/240 is also given by (x,y)(x,x+2y)(x,y)\mapsto(x,x+2y). It follows that HSH\equiv S mod 22 as asserted.

3.2. Estimating 𝒮hst(M)\mathcal{S}^{\rm st}_{\rm h}(M)

In this section we give upper and lower bounds for the homotopy stable class of certain (4m1)(4m{-}1)-connected 8m8m-manifolds. To state these bounds we require a certain amount of notation.

Let mm be a positive integer and let {a,b}\{a,b\} be a pair of positive integers. Since the dimensions 88 and 1616 are exceptional, we introduce the factor

cm:={2m=1 or 2,1m>2,c_{m}:=\begin{cases}2&m=\text{$1$ or $2$,}\\ 1&m>2,\end{cases}

to handle the exceptional dimensions. We define

d:=gcd(a,b)cmd:=\gcd(a,b)c_{m}

and write

acm=da and bcm=dbac_{m}=da^{\prime}\text{ and }bc_{m}=db^{\prime}

for some coprime a,ba^{\prime},b^{\prime}. Set

A:=ab=abcm2/d2.A:=a^{\prime}b^{\prime}=abc_{m}^{2}/d^{2}.

For a positive integer nn we let 𝒫n\mathcal{P}_{n} be the set of prime factors of nn:

𝒫n:={p:p prime, pn}.\mathcal{P}_{n}:=\{p\in\mathbb{N}\,:\,p\text{ prime, }p\mid n\}.

We set j¯m=jm/gcd(jm,d)\overline{j}_{m}=j_{m}/\gcd(j_{m},d) and consider the sets 𝒫A\mathcal{P}_{\!A}, 𝒫j¯m\mathcal{P}_{\overline{j}_{m}} and their intersection

𝒫A,m:=𝒫A𝒫j¯m,\mathcal{P}_{\!A,m}:=\mathcal{P}_{\!A}\cap\mathcal{P}_{\overline{j}_{m}},

the set of primes dividing both j¯m\overline{j}_{m} and AA. We define the non-negative integers

qA:=|𝒫A|1 and qA,m:=|𝒫A,m|1.q_{A}:=|\mathcal{P}_{\!A}|-1\text{ and }q_{A,m}:=|\mathcal{P}_{\!A,m}|-1.

Now we can state the main theorem of this section. Its proof will occupy the remainder of the section.

Theorem 3.3.

Let mm be a positive integer and let {a,b}\{a,b\} be a pair of positive integers such that 𝔟𝔭mab\mathfrak{bp}_{m}\mid ab. If d=gcd(a,b)d=\gcd(a,b) and j¯m=jm/gcd(jm,d)\overline{j}_{m}=j_{m}/\gcd(j_{m},d), then the closed, (4m1)(4m{-}1)-connected 8m8m-manifolds Ma,bM_{a,b} constructed in Section 3.5 satisfy the following:

  1. (1)

    Ma,bM_{a,b} has hyperbolic intersection form,

  2. (2)

    |𝒮st(Ma,b)/Θ8m|=2qA|\mathcal{S}^{\rm st}(M_{a,b})/\Theta_{8m}|=2^{q_{A}}, and

  3. (3)

    2qA,m|𝒮hst(Ma,b)|j¯m2+2j¯m+442^{q_{A,m}}\leq|\mathcal{S}^{\rm st}_{\rm h}(M_{a,b})|\leq\Big{\lfloor}\frac{\overline{j}_{m}^{2}+2\overline{j}_{m}+4}{4}\Big{\rfloor}.

Adam’s work on jmj_{m} [Ada66], a theorem of von Staudt and Clausen (see [IR90, Theorem 3, p. 233]) on the denominator of BmB_{m}, and a result of von Staudt on the numerator of BmB_{m} (see [Mil58b, Lemma 2]) combine to show that

𝒫jm={p prime:(p1)2m}.\mathcal{P}_{j_{m}}=\{p\text{ prime}:(p-1)\mid 2m\}.

Since 22 and 33 certainly lie in the latter set, |𝒫jm|2|\mathcal{P}_{j_{m}}|\geq 2. Now define

qm:=|𝒫jm|11.q_{m}:=|\mathcal{P}_{j_{m}}|-1\geq 1.

By choosing aa and bb with some care, we obtain the following corollary, which implies Theorem 1.2.

Corollary 3.4.

Let mm be a positive integer and let {a,b}\{a,b\} be a pair of positive, coprime integers such that 𝔟𝔭mab\mathfrak{bp}_{m}\mid ab and jm/cmA=abcm2j_{m}/c_{m}\mid A=abc_{m}^{2}. Then the closed, (4m1)(4m{-}1)-connected 8m8m-manifolds Ma,bM_{a,b} constructed in Section 3.5 have hyperbolic intersection form and satisfy that 22qm|𝒮hst(Ma,b)|2\leq 2^{q_{m}}\leq|\mathcal{S}^{\rm st}_{\rm h}(M_{a,b})|.

In particular, any coprime, positive a,ba,b such that 𝔟𝔭mjm/cm\mathfrak{bp}_{m}\cdot j_{m}/c_{m} divides A=abcm2A=abc_{m}^{2} satisfies the hypotheses of the corollary. Note that changing AA does not alter the lower bound, which is purely a function of mm.

Proof.

Since aa and bb are coprime, d=cmd=c_{m}, j¯m=jm/cm\overline{j}_{m}=j_{m}/c_{m} and 𝒫j¯m=𝒫jm\mathcal{P}_{\overline{j}_{m}}=\mathcal{P}_{j_{m}} (using 4jm4\mid j_{m} for m=1,2m=1,2). Since jm/cm=j¯mAj_{m}/c_{m}=\overline{j}_{m}\mid A we see that 𝒫j¯m𝒫A\mathcal{P}_{\overline{j}_{m}}\subseteq\mathcal{P}_{\!A} and therefore 𝒫A,m=𝒫j¯m=𝒫jm\mathcal{P}_{\!A,m}=\mathcal{P}_{\overline{j}_{m}}=\mathcal{P}_{j_{m}}, so that qA,m=qmq_{A,m}=q_{m}. Since qm1q_{m}\geq 1, the corollary follows from the lower bound in Theorem 3.3 (3). ∎

3.3. The almost-diffeomorphism classification of (4m1)(4m{-}1)-connected 8m8m-manifolds

In this section we recall the relevant part of Wall’s classification results for closed, (4m1)(4m{-}1)-connected 8m8m-manifolds. Recall that two closed manifolds M0M_{0} and M1M_{1} are almost diffeomorphic if there is a homotopy sphere Σ\Sigma and a diffeomorphism f:M0#ΣM1f\colon M_{0}\#\Sigma\to M_{1}.

Let MM be a closed, (4m1)(4m{-}1)-connected 8m8m-manifold, and equip MM with an orientation. The intersection form of MM is a symmetric bilinear form

λM:H4m(M)×H4m(M).\lambda_{M}\colon H_{4m}(M)\times H_{4m}(M)\to\mathbb{Z}.

The obstruction class of MM is the homomorphism

SαM:H4m(M)π4m1(SO)S\alpha_{M}\colon H_{4m}(M)\to\pi_{4m-1}(SO)\cong\mathbb{Z}

defined by representing a homology class xx by a smoothly embedded sphere Sx4mMS^{4m}_{x}\hookrightarrow M, whose existence is ensured by Hurewicz theorem and [Hae61, Theorem 1(a)], and then taking the homotopy class of the clutching map of the stable normal bundle of Sx4mS^{4m}_{x}. The map SαMS\alpha_{M} is the stabilisation of a map αM\alpha_{M} defined by taking the normal bundle of Sx4mS^{4m}_{x}. This will be important in the proof of Theorem 3.7 below. As shown by Wall [Wal62, p. 171 & Lemma 2], if m=1,2m=1,2 then the existence of rank 4m4m vector bundles over S4mS^{4m} with odd Euler class implies that the obstruction class is characteristic for the intersection form; i.e. if m=1m=1 or 22 then for all xH4m(M)x\in H_{4m}(M)

({\dagger}) λM(x,x)SαM(x)mod2.\lambda_{M}(x,x)\equiv S\alpha_{M}(x)~{}\mathrm{mod}~{}2.

For m>2m>2, by Wall [Wal62, p. 171], there is no relation between SαMS\alpha_{M} and λM\lambda_{M}. As also shown in [Wal62, p. 171 & Lemma 2], since e=HJ4m1,4m\mathrm{e}=H\circ J_{4m-1,4m} and since for m>2m>2 we have that HJ4m1,4mH\circ J_{4m-1,4m} is even, the Euler number is always even and therefore λM(x,x)0mod2\lambda_{M}(x,x)\equiv 0\mod{2} for all xH4m(M)x\in H_{4m}(M).

For the homotopy classification, we consider the stable JJ-homomorphism

J4m1:π4m1(SO)/jmπ4m1s.J_{4m-1}\colon\pi_{4m-1}(SO)\to\mathbb{Z}/j_{m}\subseteq\pi^{s}_{4m-1}.

The homotopy obstruction class of MM, SαMhS\alpha^{\mathrm{h}}_{M}, is the composition of SαMS\alpha_{M} with J4m1J_{4m-1},

SαMh:=J4m1SαM:H4m(M)/jm.S\alpha^{\mathrm{h}}_{M}:=J_{4m-1}\circ S\alpha_{M}\colon H_{4m}(M)\to\mathbb{Z}/j_{m}.

Since j1j_{1} and j2j_{2} are divisible by 22 the congruence of ({\dagger}3.3) implies that if m=1,2m=1,2 then

(1) λM(x,x)SαMh(x)mod2.\lambda_{M}(x,x)\equiv S\alpha^{\mathrm{h}}_{M}(x)~{}\mathrm{mod}~{}2.

We now define the invariants we use to classify (4m1)(4m{-}1)-connected 8m8m-manifolds up to almost diffeomorphism and homotopy equivalence.

Definition 3.5 (Extended symmetric form).

Fix a homomorphism v:G/2v\colon G\to\mathbb{Z}/2 from an abelian group GG to /2\mathbb{Z}/2. An extended symmetric form over vv consists of a triple (H,λ,p)(H,\lambda,p) where:

  1. (1)

    HH is a finitely generated free \mathbb{Z}-module;

  2. (2)

    λ:H×H\lambda\colon H\times H\to\mathbb{Z} is a symmetric, bilinear form; and

  3. (3)

    f:HGf\colon H\to G is a homomorphism such that λ(x,x)vf(x)\lambda(x,x)\equiv v\circ f(x) mod 22.

Two extended symmetric forms (H,λ,f)(H,\lambda,f) and (H,λ,f)(H^{\prime},\lambda^{\prime},f^{\prime}) are equivalent if there is an isometry h:(H,λ)(H,λ)h\colon(H,\lambda)\to(H^{\prime},\lambda^{\prime}) such that fh=f:HGf^{\prime}\circ h=f\colon H\to G.

In our applications to 8m8m-manifolds, the group GG will either be the infinite cyclic group π4m1(SO)\pi_{4m-1}(SO)\cong\mathbb{Z} or the finite cyclic group Im(J4m1)/jm\mathrm{Im}(J_{4m-1})\cong\mathbb{Z}/j_{m}. Due to the existence of rank 4m4m bundles over S4mS^{4m} with odd Euler number when m=1,2m=1,2, and the non-existence of such bundles for m3m\geq 3, we set vv to be nonzero for m=1,2m=1,2 (recall 2j12\mid j_{1} and 2j22\mid j_{2}) and zero for m>2m>2. Hence for m>2m>2, (3) is just the requirement that λM\lambda_{M} be even. With these conventions on vv, the following assignments define extended symmetric forms.

Definition 3.6 (The extended symmetric forms of MM).

Let MM be an oriented (4m1)(4m{-}1)-connected 8m8m-manifold.

  1. (1)

    The smooth extended symmetric form of MM is the triple

    (H4m(M),λM,SαM)\bigl{(}H_{4m}(M),\lambda_{M},S\alpha_{M}\bigr{)}

    with GG\cong\mathbb{Z}.

  2. (2)

    The homotopy extended symmetric form of MM is the triple

    (H4m(M),λM,SαMh)\bigl{(}H_{4m}(M),\lambda_{M},S\alpha^{\mathrm{h}}_{M}\bigr{)}

    with G/jmG\cong\mathbb{Z}/j_{m}.

The following result is a direct consequence of classification results of Wall [Wal62, p. 170 & Lemma 8].

Theorem 3.7 (Wall).

Let M1M_{1} and M2M_{2} be closed, oriented, (4m1)(4m{-}1)-connected 8m8m-manifolds. The manifolds M1M_{1} and M2M_{2} are:

  1. (1)

    almost diffeomorphic, via an orientation-preserving diffeomorphism, if and only if their smooth extended symmetric forms are equivalent;

  2. (2)

    homotopy equivalent, via a degree one homotopy equivalence, if and only if their homotopy extended symmetric forms are equivalent.

When applying these classifications, we will later have to factor out by the effect of the orientation choice on the extended symmetric forms.

Proof.

We start with the almost diffeomorphism classification (1). As mentioned above, the homomorphism SαMS\alpha_{M} is the stabilisation of a certain quadratic form, the extended quadratic form of MM, which is the map

αM:H4m(M)π4m1(SO4m),\alpha_{M}\colon H_{4m}(M)\to\pi_{4m-1}(SO_{4m}),

defined by representing a homology class by a smoothly embedded sphere S4mMS^{4m}\hookrightarrow M, and then taking the classifying map in π4m1(SO4m)\pi_{4m-1}(SO_{4m}) of the normal bundle of the embedded sphere. For all x,yH4m(M)x,y\in H_{4m}(M), [Wal62, Lemma 2] (and the fact that e=HJ4m1,4m\mathrm{e}=H\circ J_{4m-1,4m}) proves that αM\alpha_{M} relates to the intersection form of MM by the equations

λM(x,x)=e(αM(x))andαM(x+y)=αM(x)+αM(y)+λ(x,y)τ.\lambda_{M}(x,x)=\mathrm{e}(\alpha_{M}(x))\quad\text{and}\quad\alpha_{M}(x+y)=\alpha_{M}(x)+\alpha_{M}(y)+\lambda(x,y)\tau.

Here the map e:π4m1(SO4m)\mathrm{e}\colon\pi_{4m-1}(SO_{4m})\to\mathbb{Z} is the Euler number of the corresponding bundle and τπ4m1(SO4m)\tau\in\pi_{4m-1}(SO_{4m}) is the clutching function of the tangent bundle of S4mS^{4m}. Wall also proved [Wal62, p. 170] that the triple (H4m(M),λM,αM)(H_{4m}(M),\lambda_{M},\alpha_{M}) is a complete almost diffeomorphism invariant of MM. In fact, Wall stated his classification in terms of almost closed manifolds: compact manifolds with boundary a homotopy sphere. But this also yields the almost diffeomorphism classification, as follows. If the extended symmetric forms of two closed (4m1)(4m{-}1)-connected 8m8m-manifolds are equivalent then by the almost closed classification the manifolds are diffeomorphic after removing a ball D8mD^{8m} from each. Gluing the balls back in compatibly with the diffeomorphism might change one of the manifolds by connected sum with a homotopy sphere, but nonetheless the two closed manifolds are almost diffeomorphic. On the other hand almost diffeomorphic manifolds are diffeomorphic after removing a ball from each, and then by the classification the extended symmetric forms are equivalent.

As mentioned above, SαM:=SαMS\alpha_{M}:=S\circ\alpha_{M}, where S:π4m1(SO4m)π4m1(SO)S\colon\pi_{4m-1}(SO_{4m})\to\pi_{4m-1}(SO) is the stabilisation homomorphism. The homotopy exact sequence of the fibration SO4mSO4m+1S4mSO_{4m}\to SO_{4m+1}\to S^{4m} shows that the kernel of SS is generated by τ\tau [Lev85, Lemma 1.3 and Theorem 1.4] and since

eS:π4m1(SO4m)π4m1(SO)\mathrm{e}\oplus S\colon\pi_{4m-1}(SO_{4m})\to\mathbb{Z}\oplus\pi_{4m-1}(SO)

is injective by (*3.1) it follows that the pair (λM(x,x),SαM(x))=(e(αM(x)),S(αM(x)))π4m1(SO)(\lambda_{M}(x,x),S\alpha_{M}(x))=\bigl{(}\mathrm{e}(\alpha_{M}(x)),S(\alpha_{M}(x))\bigr{)}\in\mathbb{Z}\oplus\pi_{4m-1}(SO)\cong\mathbb{Z}\oplus\mathbb{Z} determines αM(x)\alpha_{M}(x) for all xH4m(M)x\in H_{4m}(M). The theorem now follows from Wall’s almost diffeomorphism classification.

The proof of the homotopy classification is similar. By Wall [Wal62, Lemma 8], the triple (H4m(M),λM,αMh:=J4m1,4mαM)(H_{4m}(M),\lambda_{M},\alpha_{M}^{h}:=J_{4m-1,4m}\circ\alpha_{M}) is a complete homotopy invariant of the manifolds under consideration. Since e=HJ\mathrm{e}=HJ and

HS:π8m1(S4m)π4m1sH\oplus S\colon\pi_{8m-1}(S^{4m})\to\mathbb{Z}\oplus\pi^{s}_{4m-1}

is injective by (*3.1), it follows that the pair (λM(x,x),SαMh(x))=(H(αMh(x)),S(αMh(x)))π4m1s(\lambda_{M}(x,x),S\alpha^{\mathrm{h}}_{M}(x))=\bigl{(}H(\alpha^{\mathrm{h}}_{M}(x)),S(\alpha^{\mathrm{h}}_{M}(x))\bigr{)}\in\mathbb{Z}\oplus\pi^{s}_{4m-1} determines αMh(x)\alpha^{\mathrm{h}}_{M}(x) for all xH4m(M)x\in H_{4m}(M). The theorem now follows from Wall’s homotopy classification. ∎

3.4. Stable almost-diffeomorphism classification of (4m1)(4m{-}1)-connected 8m8m-manifolds

In this section we give the stable classification of closed (4m1)(4m{-}1)-connected 8m8m-manifolds up to connected sum with homotopy 8m8m-spheres. Define the non-negative integer dMd_{M} by the equation

SαM(H4m(M))=dM.S\alpha_{M}(H_{4m}(M))=d_{M}\mathbb{Z}.

Equivalently, dMd_{M} is the divisibility of SαMH4m(M)S\alpha_{M}\in H^{4m}(M), where, since MM is (4m1)(4m{-}1)-connected, we may regard SαMS\alpha_{M} as an element of the group H4m(M)H^{4m}(M) via the inverse of the evaluation map ev:H4m(M)Hom(H4m(M),)\operatorname{ev}\colon H^{4m}(M)\to\operatorname{Hom}(H_{4m}(M),\mathbb{Z}), which is an isomorphism. In particular, it makes sense to consider the class (SαM)2H8m(M)(S\alpha_{M})^{2}\in H^{8m}(M)\cong\mathbb{Z}.

Theorem 3.8.

Two closed, oriented, (4m1)(4m{-}1)-connected 8m8m-manifolds MM and NN with the same Euler characteristic are almost stably diffeomorphic, via an orientation-preserving diffeomorphism, if and only if the following hold:

  1. (1)

    dM=dNd_{M}=d_{N},

  2. (2)

    σ(M)=σ(N)\sigma(M)=\sigma(N),

  3. (3)

    (SαM)2,[M]=(SαN)2,[N]\langle(S\alpha_{M})^{2},[M]\rangle=\langle(S\alpha_{N})^{2},[N]\rangle.

Proof.

First, we note that dMd_{M}, the signature, and (SαM)2(S\alpha_{M})^{2} are invariants of orientation preserving almost stable diffeomorphisms, so one implication holds.

For the other implication we assume that MM and NN are such that dM=dNd_{M}=d_{N}, σ(M)=σ(N)\sigma(M)=\sigma(N), and (SαM)2=(SαN)2(S\alpha_{M})^{2}=(S\alpha_{N})^{2} and we show that MM and NN are stably diffeomorphic. The normal (4m1)(4m{-}1)-type of MM and NN is determined by d=dM=dNd=d_{M}=d_{N} and is described as follows. Let dd be a non-negative integer. Let BO4m1BOBO\langle{4m{-}1}\rangle\to BO be the (4m1)(4m{-}1)-connected cover of BOBO and let pH4m(BO4m1)p\in H^{4m}(BO\langle{4m{-}1}\rangle)\cong\mathbb{Z} be a generator. We regard ρd(p)\rho_{d}(p), the mod dd reduction of pp, as a map ρd(p):BO4m1K(/d,4m)\rho_{d}(p)\colon BO\langle{4m{-}1}\rangle\to K(\mathbb{Z}/d,4m) and define BO4m1,dMBO\langle{4m{-}1,d_{M}}\rangle to be the homotopy fibre of ρd(p)\rho_{d}(p). The normal (4m1)(4m{-}1)-type of MM and NN is represented by the fibration given by the composition

BO4m1,dBO4m1BO.BO\langle{4m{-}1,d}\rangle\to BO\langle{4m{-}1}\rangle\to BO.

For brevity, use (Bd,ηd)(B_{d},\eta_{d}) to denote the fibration ηd:BO4m1,dBO\eta_{d}\colon BO\langle{4m{-}1,d}\rangle\to BO. We assert that MM and NN admit unique normal (4m1)(4m{-}1)-smoothings ν¯M:MBd\overline{\nu}_{M}\colon M\to B_{d} and ν¯N:NBd\overline{\nu}_{N}\colon N\to B_{d}. We prove the assertion for MM, as the proof for NN is identical. Since MM is 4m4m-connected, its stable normal bundle νM:MBO\nu_{M}\colon M\to BO lifts (up to homotopy) uniquely to ν4m:MBO4m\nu_{4m}\colon M\to BO\langle 4m\rangle. In order to lift ν4m\nu_{4m} to BdB_{d}, we consider the long exact sequence (of pointed sets) of the fibration

0=H4m1(M;/d)[M,Bd][M,BO4m]H4m(M;/d),0=H^{4m-1}(M;\mathbb{Z}/d)\to[M,B_{d}]\to[M,BO\langle 4m\rangle]\to H^{4m}(M;\mathbb{Z}/d)\to\cdots,

where on the left, we used [M,ΩK(/d,4m)]=[M,K(/d,4m1)]=H4m1(M;/d)=0[M,\Omega K(\mathbb{Z}/d,4m)]=[M,K(\mathbb{Z}/d,4m-1)]=H^{4m-1}(M;\mathbb{Z}/d)=0, because MM is 4m4m-connected. The assertion is now proved by noting that ν4m[M,BO4m]\nu_{4m}\in[M,BO\langle 4m\rangle] maps to SαMH4m(M;/d)S\alpha_{M}\in H^{4m}(M;\mathbb{Z}/d), which is zero by definition of the divisibility dMd_{M}.

By [Kre99, Theorem 2], MM and NN are orientation preserving stably diffeomorphic if

[M,ν¯M]=[N,ν¯N]Ω8m(Bd,ηd).[M,\overline{\nu}_{M}]=[N,\overline{\nu}_{N}]\in\Omega_{8m}(B_{d},\eta_{d}).

Since homotopy 8m8m-spheres have a unique (Bd,ηd)(B_{d},\eta_{d})-structure, there is a well-defined homomorphism Θ8mΩ8m(Bd,ηd)\Theta_{8m}\to\Omega_{8m}(B_{d},\eta_{d}). Now the arguments in Wall’s computation of the Grothendieck groups of almost closed (4m1)(4m{-}1)-connected 8m8m-manifolds [Wal62, Theorem 2] show that there is an exact sequence

(Ω\Omega) Θ8mΩ8m(Bd,ηd)(σ,(Sα)2)2,\Theta_{8m}\to\Omega_{8m}(B_{d},\eta_{d})\xrightarrow{(\sigma,\,(S\alpha)^{2})}\mathbb{Z}^{2},

where σ([M,ν¯M])=σM\sigma([M,\overline{\nu}_{M}])=\sigma_{M} and Sα2([M,ν¯M])=(SαM)2([M])S\alpha^{2}([M,\overline{\nu}_{M}])=(S\alpha_{M})^{2}([M]). It follows that there is a homotopy 8m8m-sphere Σ\Sigma such that [M#Σ,ν¯M#Σ]=[N,ν¯N]Ω8m(Bd,η)[M\#\Sigma,\overline{\nu}_{M\#\Sigma}]=[N,\overline{\nu}_{N}]\in\Omega_{8m}(B_{d},\eta). Hence M#ΣM\#\Sigma and NN are stably diffeomorphic and so MM and NN are almost stably diffeomorphic. ∎

3.5. Construction of the manifolds Ma,bM_{a,b}

In this section we construct the manifolds Ma,bM_{a,b} appearing in Theorem 3.3. Let aa and bb be positive integers such that 𝔟𝔭mab\mathfrak{bp}_{m}\mid ab. We will build simply-connected, closed 8m8m-manifolds Ma,bM_{a,b} with the cohomology ring of S4m×S4mS^{4m}\times S^{4m} by attaching handles to an 8m8m-ball. We attach two 4m4m-handles hxh_{x} and hyh_{y}, diffeomorphic to D4m×D4mD^{4m}\times D^{4m}, to D8mD^{8m} using attaching maps ϕx,ϕy:S4m1×{0}S8m1\phi_{x},\phi_{y}\colon S^{4m-1}\times\{0\}\to S^{8m-1} with linking number 11. Note that for m1m\geq 1, 2-component links S4m1S4m1S8m1S^{4m-1}\sqcup S^{4m-1}\hookrightarrow S^{8m-1} are classified up to smooth isotopy by the linking number, an integer [Hae62, Theorem in Section 5]. There is more data needed for the attaching maps, which for each 4m4m-handle corresponds to a choice of framing for the attaching sphere S4m1S8m1S^{4m{-}1}\subseteq S^{8m-1}. The framings that induce a given orientation are in one to one correspondence with homotopy classes of maps [S4m1,SO4m][S^{4m{-}1},SO_{4m}] where the class of the constant map corresponds to the framing which extends over an embedded 4m4m-disc D4mS8m1D^{4m}\subseteq S^{8m-1}. Recall from (*3.1) that π4m1(SO4m)\pi_{4m{-}1}(SO_{4m})\cong\mathbb{Z}\oplus\mathbb{Z}, detected by eS\mathrm{e}\oplus S (although this map is not an isomorphism). We are attaching 4m4m-handles hxh_{x} hyh_{y}; let xx and yy denote the corresponding classes in (4m)(4m)th homology and let ξx,ξyπ4m1(SO4m)\xi_{x},\xi_{y}\in\pi_{4m-1}(SO_{4m}) be the framings for the attaching maps.

Since we want λ(x,x)=0\lambda(x,x)=0, we require that e(ξx)=0\mathrm{e}(\xi_{x})=0 but we are otherwise free to choose ξx\xi_{x}. Recall that cm=2c_{m}=2 if m=1,2m=1,2 and cm=1c_{m}=1 if m>2m\!>\!2, fix an isomorphism π4m1(SO)=\pi_{4m-1}(SO)=\mathbb{Z} and choose ξx\xi_{x} such that S(ξx)=acmS(\xi_{x})=ac_{m}. By the discussion following (*3.1), we can find such a ξx\xi_{x} for any choice of aa. Similarly, we attach the handle hyh_{y} with e(ξy)=0\mathrm{e}(\xi_{y})=0 and S(ξy)=bcmS(\xi_{y})=bc_{m}. Again, we can find such a ξy\xi_{y} for any bb. After attaching the pair of 4m4m-handles, we write W:=Wa,bW:=W_{a,b} for the resulting compact 8m8m-manifold with boundary. Note that there is a homotopy equivalence WS4mS4mW\simeq S^{4m}\vee S^{4m}. As above let xx and yy be generators of 2H4m(W)\mathbb{Z}^{2}\cong H_{4m}(W) and let {x,y}\{x^{*},y^{*}\} be the dual basis for 2H4m(W)=H4m(W)\mathbb{Z}^{2}\cong H^{4m}(W)=H_{4m}(W)^{*}. The manifold W=Wa,bW=W_{a,b} has smooth extended symmetric form given by

(H4m(W),λW,SαW)=(2,(0110),(acmbcm):2),\bigl{(}H_{4m}(W),\lambda_{W},S\alpha_{W}\bigr{)}=\left(\mathbb{Z}^{2},\left(\begin{array}[]{cc}0&1\\ 1&0\end{array}\right),\left(\begin{array}[]{c}ac_{m}\\ bc_{m}\end{array}\right)\colon\mathbb{Z}^{2}\to\mathbb{Z}\right),

where the notation for SαWS\alpha_{W} means that SαW(x)=acmS\alpha_{W}(x)=ac_{m} and SαW(y)=bcmS\alpha_{W}(y)=bc_{m}.

Alternatively, the construction thus far can be achieved by taking the two D4mD^{4m}-bundles over S4mS^{4m} determined by ξx\xi_{x} and ξy\xi_{y}, and plumbing them together once.

The boundary of Wa,bW_{a,b} is a homotopy (8m1)(8m{-1})-sphere, which we denote by Σa,b\Sigma_{a,b}. In particular, W1,1\partial W_{1,1} is by definition the homotopy sphere ΣQ\Sigma_{Q} from [KR20, §2]. More generally, Wa,b\partial W_{a,b} is given as follows.

Lemma 3.9.

[Wa,b]=[abΣQ]Θ8m1[\partial W_{a,b}]=[ab\Sigma_{Q}]\in\Theta_{8m-1}.

Proof.

Recall from [Wal67, §17] the group A8m4mA^{\langle{4m}\rangle}_{8m} of bordism classes of (4m1)(4m{-}1)-connected 8m8m-manifolds with boundary a homotopy sphere, where the bordisms are required to be hh-cobordisms on the boundary. Addition is via boundary connected sum. Taking the boundary defines a homomorphism A8m4mΘ8m1A^{\langle{4m}\rangle}_{8m}\to\Theta_{8m-1}, and the characteristic numbers σ\sigma and (Sα)2(S\alpha)^{2} of (Ω\Omega) are also well-defined on A8m4mA^{\langle{4m}\rangle}_{8m}. Indeed, Wall [Wal62, Theorems 2 & 3] proved that σ(Sα)2:A8m4m2\sigma\oplus(S\alpha)^{2}\colon A^{\langle{4m}\rangle}_{8m}\to\mathbb{Z}^{2} is an injective homomorphism. Since Wa,bW_{a,b} satisfies σ(Wa,b)=0\sigma(W_{a,b})=0 and (SαWa,b)2=2abcm2(S\alpha_{W_{a,b}})^{2}=2abc_{m}^{2}, we have that (SαW1,1)2=2cm2(S\alpha_{W_{1,1}})^{2}=2c_{m}^{2} and so

(SαWa,b)2=2abcm2=ab(SαW1,1)2=(SαabW1,1)2,(S\alpha_{W_{a,b}})^{2}=2abc_{m}^{2}=ab(S\alpha_{W_{1,1}})^{2}=(S\alpha_{\natural^{ab}W_{1,1}})^{2},

where the last equality used that (Sα)2(S\alpha)^{2} is a homomorphism. Since σ(Sα)2\sigma\oplus(S\alpha)^{2} is injective, Wa,b=abW1,1A8m4mW_{a,b}=\natural^{ab}W_{1,1}\in A^{\langle{4m}\rangle}_{8m}. So Wa,b\partial W_{a,b} and (abW1,1)=abΣQ\partial(\natural^{ab}W_{1,1})=ab\Sigma_{Q} are hh-cobordant and therefore diffeomorphic. ∎

From Lemma 3.9 and our assumption that 𝔟𝔭mab\mathfrak{bp}_{m}\mid ab, it follows that [Σa,b]=0bP8m[\Sigma_{a,b}]=0\in bP_{8m}, so that there is a choice of diffeomorphism f:Σa,bS8m1f\colon\Sigma_{a,b}\to S^{8m-1}. We write Ma,b,f=Wa,bfD8mM_{a,b,f}=W_{a,b}\cup_{f}D^{8m} for the closure of Wa,bW_{a,b} built using a diffeomorphism f:Σa,bS8m1f\colon\Sigma_{a,b}\to S^{8m-1}. We will also use Ma,bM_{a,b} to ambiguously denote any Ma,b,fM_{a,b,f}. For any other choice of diffeomorphism ff^{\prime}, Ma,b,fM_{a,b,f} and Ma,b,fM_{a,b,f^{\prime}} are almost diffeomorphic.

Let us record the values of the key invariants on Ma,bM_{a,b}. The stable almost diffeomorphism invariants of Ma,bM_{a,b} are dMa,b=gcd(a,b)cmd_{M_{a,b}}=\mathrm{gcd}(a,b)c_{m}, σ(Ma,b)=0\sigma(M_{a,b})=0, and (SαMa,b)2=2abcm2(S\alpha_{M_{a,b}})^{2}=2abc_{m}^{2}. The extended symmetric form of M:=Ma,bM:=M_{a,b} is the same as that of WW:

(H4m(M),λM,SαM)=(2,(0110),(acmbcm):2).\bigl{(}H_{4m}(M),\lambda_{M},S\alpha_{M}\bigr{)}=\left(\mathbb{Z}^{2},\left(\begin{array}[]{cc}0&1\\ 1&0\end{array}\right),\left(\begin{array}[]{c}ac_{m}\\ bc_{m}\end{array}\right)\colon\mathbb{Z}^{2}\to\mathbb{Z}\right).

This completes the construction of the manifolds Ma,bM_{a,b}.

3.6. The proof of Theorem 3.3

Now that we have constructed the (4m1)(4m{-}1)-connected 8m8m-manifolds Ma,bM_{a,b}, we are ready to prove Theorem 3.3.

Proof of Theorem 3.3.

Let aa and bb be positive integers such that 𝔟𝔭mab\mathfrak{bp}_{m}\mid ab. By construction the oriented manifolds Ma,bM_{a,b} have hyperbolic intersection form, so Theorem 3.3 (1) is immediate.

As before write d:=gcd(a,b)cmd:=\gcd(a,b)c_{m} and define A:=abcm2/d2A:=abc_{m}^{2}/d^{2}. Let p1,,pqA+1p_{1},\dots,p_{q_{A}+1} be the prime-power factors of AA, which are powers of pairwise distinct primes. Then there are 2qA2^{q_{A}} ways to express AA as a product yiziy_{i}z_{i} of coprime positive integers, counting unordered pairs {yi,zi}\{y_{i},z_{i}\}. We consider the 8m8m-manifolds

{Mi:=Mdyi,dzi}i=12qA.\{M_{i}:=M_{dy_{i},dz_{i}}\}_{i=1}^{2^{q_{A}}}.

For each ii, dMi=dd_{M_{i}}=d, σ(Mi)=0\sigma(M_{i})=0, and (SαMi)2,[Mi]=2dyidzi=2d2A=2abcm2\langle(S\alpha_{M_{i}})^{2},[M_{i}]\rangle=2dy_{i}dz_{i}=2d^{2}A=2abc_{m}^{2}. Therefore the manifolds MiM_{i} are pairwise almost stably diffeomorphic by Theorem 3.8. A priori they could not all lie in 𝒮st(Ma,b)\mathcal{S}^{\rm st}(M_{a,b}), but the ambiguity of whether they are actually stably diffeomorphic can be removed by more carefully choosing the diffeomorphisms fi:ΣiS8m1f_{i}\colon\Sigma_{i}\to S^{8m-1} used to glue on D8mD^{8m} in the construction of the MiM_{i}. By changing the choice of the identification fif_{i} we can change MiM_{i} by connected sum with an exotic sphere of our choice. The MiM_{i} were only determined up to this choice in our construction, so let us assume we have made this consistently so that Mi𝒮st(Ma,b)M_{i}\in\mathcal{S}^{\rm st}(M_{a,b}) for every i=1,,2qAi=1,\dots,2^{q_{A}}. In 𝒮st(Ma,b)/Θ8m\mathcal{S}^{\rm st}(M_{a,b})/\Theta_{8m} this choice of the fif_{i} is in any case irrelevant.

When we discuss extended symmetric forms on 2\mathbb{Z}^{2}, we will always mean with respect to a particular choice of basis. For MiM_{i}, with its fixed choice of fundamental class [Mi][M_{i}], we shall use a basis with respect to which the intersection form is represented by H+()=(0110)H^{+}(\mathbb{Z})=\left(\begin{smallmatrix}0&1\\ 1&0\end{smallmatrix}\right). We have constructed the MiM_{i} so that with respect to such a basis SαMi:2S\alpha_{M_{i}}\colon\mathbb{Z}^{2}\to\mathbb{Z} is represented by (acmbcm)\left(\begin{smallmatrix}ac_{m}\\ bc_{m}\end{smallmatrix}\right) with ab>0ab>0 and cm{1,2}c_{m}\in\{1,2\}.

The smooth extended symmetric forms of the MiM_{i} are pairwise distinct, since isometries of the rank two hyperbolic intersection form can only change the sign and permute the basis elements. The map SαMi:2S\alpha_{M_{i}}\colon\mathbb{Z}^{2}\to\mathbb{Z} is given by (dyi,dzi)(dy_{i},dz_{i}). Since the unordered pairs {dyi,dzi}\{dy_{i},dz_{i}\} are pairwise distinct, by the almost diffeomorphism classification of Theorem 3.7 (1), the MiM_{i} are pairwise distinct up to orientation-preserving almost diffeomorphism. We will be able to deduce that |𝒮st(Ma,b)/Θ8m|2qA|\mathcal{S}^{\rm st}(M_{a,b})/\Theta_{8m}|\geq 2^{q_{A}} once we have factored out by the effect of the choice of orientation of the MiM_{i}. In other words we must show that there are also no orientation-reversing almost diffeomorphisms from MiM_{i} to MjM_{j}, for iji\neq j, or equivalently that there is no orientation-preserving diffeomorphism MiMjM_{i}\cong-M_{j}.

Changing the orientation of MjM_{j} changes the smooth extended symmetric form (with respect to the same basis for H4m(M)H_{4m}(M)) by altering the sign of the intersection form, but does not affect SαMjS\alpha_{M_{j}}. To see this, note that while changing the orientation of MjM_{j} changes the induced orientation of the fibres of the normal bundle of an embedded sphere xx, SαMj(x)π4m1(SO)S\alpha_{M_{j}}(x)\in\pi_{4m-1}(SO) is the clutching map of this normal bundle, and this is unaffected by the orientation of the fibres.

The isometries from the rank 2 hyperbolic form H+()=(0110)H^{+}(\mathbb{Z})=\left(\begin{smallmatrix}0&1\\ 1&0\end{smallmatrix}\right) to its negative H+()-H^{+}(\mathbb{Z}) consist of the self-isometries of the hyperbolic form, namely ±Id\pm\operatorname{Id} and (0110)\left(\begin{smallmatrix}0&1\\ 1&0\end{smallmatrix}\right), composed with either (1001)\left(\begin{smallmatrix}1&0\\ 0&-1\end{smallmatrix}\right) or (1001)\left(\begin{smallmatrix}-1&0\\ 0&1\end{smallmatrix}\right). Thus an orientation-reversing almost diffeomorphism could identify the smooth extended symmetric form characterised by (H+(),±{v,w})(H^{+}(\mathbb{Z}),\pm\{v,w\}) with one of the extended symmetric forms (H+(),±{v,w})(-H^{+}(\mathbb{Z}),\pm\{-v,w\}) or (H+(),±{v,w})(-H^{+}(\mathbb{Z}),\pm\{v,-w\}). But for both MiM_{i} and Mi-M_{i}, the corresponding pair of integers is ±{v,w}=±{dyi,dzi}\pm\{v,w\}=\pm\{dy_{i},dz_{i}\}, where both elements have the same sign. So our manifolds {Mi}\{M_{i}\} are indeed distinct up to almost diffeomorphism. This proves that |𝒮st(Ma,b)/Θ8m|2qA|\mathcal{S}^{\rm st}(M_{a,b})/\Theta_{8m}|\geq 2^{q_{A}}.

Next we prove that |𝒮st(Ma,b)/Θ8m|2qA|\mathcal{S}^{\rm st}(M_{a,b})/\Theta_{8m}|\leq 2^{q_{A}}. Any closed 8m8m-manifold MM that is almost stably diffeomorphic to Ma,bM_{a,b} is also necessarily (4m1)(4m{-}1)-connected, the divisibility of SαMS\alpha_{M} is dd, and the intersection form is rank 2, indefinite, and even, and therefore either hyperbolic or H+()-H^{+}(\mathbb{Z}). If MM and Ma,bM_{a,b} are almost stably diffeomorphic then there is an orientation on MM such that MM and Ma,bM_{a,b} are almost stably diffeomorphic via an orientation-preserving stable diffeomorphism. Use this orientation, and choose a basis for H4m(M)H_{4m}(M) with respect to which the intersection form of MM is H+()H^{+}(\mathbb{Z}). Observe that the manifolds MiM_{i} cover all possibilities for SαMS\alpha_{M} while keeping (SαM)2(S\alpha_{M})^{2} a fixed multiple of the dual fundamental class. (If SαM=(acmbcm)S\alpha_{M}=\left(\begin{smallmatrix}-ac_{m}\\ bc_{m}\end{smallmatrix}\right), for example, then (SαM)2=2abcm2<0(S\alpha_{M})^{2}=-2abc_{m}^{2}<0, whereas (SαMa,b)2=2abcm2>0(S\alpha_{M_{a,b}})^{2}=2abc_{m}^{2}>0. This would contradict that Ma,bM_{a,b} and MM are orientation-preserving almost stably diffeomorphic.) It follows by Theorem 3.7 (1) that every such MM is almost-stably diffeomorphic to one of the MiM_{i}, and therefore |𝒮st(Ma,b)/Θ8m|2qA|\mathcal{S}^{\rm st}(M_{a,b})/\Theta_{8m}|\leq 2^{q_{A}} as desired. This completes the proof of Theorem 3.3 (2).

To prove (3), we need to estimate the size of the homotopy stable class of Ma,bM_{a,b} from above and below. We begin with the upper bound. As above, every closed 8m8m-manifold MM stably diffeomorphic to Ma,bM_{a,b} has an orientation such that MM has hyperbolic intersection form and dM=dd_{M}=d. The possibilities for SαMhS\alpha_{M}^{h}, up to equivalence of extended symmetric forms, are therefore given by an unordered pair of elements of /jm\mathbb{Z}/j_{m}, both of which are divisible by dd. Such an element of /jm\mathbb{Z}/j_{m} lies in the subgroup generated by gcd(jm,d)\gcd(j_{m},d), and so there are j¯m=jm/gcd(jm,d)\overline{j}_{m}=j_{m}/\gcd(j_{m},d) possibilities. We assert that there are

j¯m(j¯m+1)2\displaystyle\frac{\overline{j}_{m}(\overline{j}_{m}+1)}{2}

such pairs. To see this, there are (j¯m2)=j¯m(j¯m1)2{\overline{j}_{m}\choose 2}=\mbox{\large$\frac{\overline{j}_{m}(\overline{j}_{m}-1)}{2}$} choices with distinct elements (x,y)(x,y), and j¯m\overline{j}_{m} choices of the form (x,x)(x,x). Then j¯m(j¯m1)2+j¯m=j¯m(j¯m+1)2\mbox{\footnotesize$\displaystyle\frac{\overline{j}_{m}(\overline{j}_{m}-1)}{2}$}+\overline{j}_{m}=\mbox{\large$\frac{\overline{j}_{m}(\overline{j}_{m}+1)}{2}$}, which is the count asserted. Next, we also factor out by the action of /2\mathbb{Z}/2 on our set of unordered pairs which multiplies both numbers by 1-1. In the case that j¯m\overline{j}_{m} is even, there are j¯m2+1\mbox{\large$\frac{\overline{j}_{m}}{2}$}{+}1 fixed points of this action of the form (x,x)(x,-x), and also (0,j¯m2)(0,\mbox{\large$\frac{\overline{j}_{m}}{2}$}) is a fixed point. Thus there are precisely j¯m2+2\mbox{\large$\frac{\overline{j}_{m}}{2}$}+2 fixed points of a /2\mathbb{Z}/2 action on a set with j¯m(j¯m+1)2\frac{\overline{j}_{m}(\overline{j}_{m}+1)}{2} elements. A short calculation then shows that there are

j¯m2+2j¯m+44\displaystyle\frac{\overline{j}_{m}^{2}+2\overline{j}_{m}+4}{4}

orbits. A similar calculation for j¯m\overline{j}_{m} odd gives

(j¯m+1)24=j¯m2+2j¯m+44\mbox{\small$\displaystyle\frac{(\overline{j}_{m}+1)^{2}}{4}$}=\Big{\lfloor}\mbox{\small$\displaystyle\frac{\overline{j}_{m}^{2}+2\overline{j}_{m}+4}{4}$}\Big{\rfloor}

orbits. The right hand side is equal for both parities of j¯m\overline{j}_{m}, and gives our desired upper bound. Note that this upper bound does not take into account the requirement for the product abab to be constant within a stable diffeomorphism class.

It remains to prove that 2qA,m|𝒮hst(Ma,b)|2^{q_{A,m}}\leq|\mathcal{S}^{\rm st}_{\rm h}(M_{a,b})|. As above let p1,,pqA+1p_{1},\dots,p_{q_{A}+1} be the prime-power factors of AA, which are powers of pairwise distinct primes. By reordering if necessary, assume that p1,,pqA,m+1p_{1},\dots,p_{q_{A,m}+1} are the prime-powers of the form pp^{\ell} where pj¯mp\mid\overline{j}_{m}. (It could be that the highest exponent of pp that divides j¯m\overline{j}_{m} is less than the highest exponent of pp that divides AA.) Recall that d=gcd(a,b)cmd=\gcd(a,b)c_{m} and write

d:=dι=qA,m+2qA+1pι and A:=ι=1qA,m+1pι.d^{\prime}:=d\cdot\prod_{\iota=q_{A,m}+2}^{q_{A}+1}p_{\iota}\text{ and }A^{\prime}:=\prod_{\iota=1}^{q_{A,m}+1}p_{\iota}.

Note that dA=dAd^{\prime}A^{\prime}=dA. There are 2qA,m2^{q_{A,m}} essentially distinct ways to express AA^{\prime} as a product viwiv_{i}w_{i} of coprime positive integers, counting unordered pairs {vi,wi}\{v_{i},w_{i}\}. We consider the 8m8m-manifolds

{Mi:=Mdvi,dwi}i=12qA,m.\{M_{i}:=M_{dv_{i},d^{\prime}w_{i}}\}_{i=1}^{2^{q_{A,m}}}.

For each ii, dMi=dd_{M_{i}}=d, σ(Mi)=0\sigma(M_{i})=0, and (SαMi)2,[Mi]=2dvidwi=2ddA=2d2A=2abcm2\langle(S\alpha_{M_{i}})^{2},[M_{i}]\rangle=2dv_{i}d^{\prime}w_{i}=2dd^{\prime}A^{\prime}=2d^{2}A=2abc_{m}^{2}. Therefore the MiM_{i} are pairwise almost stably diffeomorphic by Theorem 3.8, so up to homotopy equivalence they are all stably diffeomorphic. As above, the ambiguity of whether they are actually stably diffeomorphic can be removed by more carefully choosing the diffeomorphisms fi:ΣS8m1f_{i}\colon\Sigma\to S^{8m-1} used to glue on D8mD^{8m} in the construction of the MiM_{i}. Let us assume once again that we have made this choice consistently so that Mi𝒮hst(Ma,b)M_{i}\in\mathcal{S}^{\rm st}_{\rm h}(M_{a,b}) for every i=1,,2qA,mi=1,\dots,2^{q_{A,m}}.

Next we show that the MiM_{i} are distinct up to homotopy equivalence. For this, by Theorem 3.7 we need to distinguish their homotopy extended symmetric forms, by showing that the maps SαMih:2/jmS\alpha_{M_{i}}^{h}\colon\mathbb{Z}^{2}\to\mathbb{Z}/j_{m} are pairwise distinct, up to precomposing with an isometry of the hyperbolic form, or to allow for the possibility of an orientation-reversing homotopy equivalence, up to an isometry between the hyperbolic form and its negative. This means we have to show that the unordered pair of elements of /jm\mathbb{Z}/j_{m} determining SαMihS\alpha_{M_{i}}^{h} and SαMjhS\alpha_{M_{j}}^{h} are distinct up to changing signs.

Let MiM_{i} and MjM_{j} be two of our manifolds, for iji\neq j. We will show that they are not homotopy equivalent. First, gcd(d,jm)\gcd(d,j_{m}) divides dd, so divides dvidv_{i} and dwid^{\prime}w_{i}. As above write j¯m:=jm/gcd(d,jm)\overline{j}_{m}:=j_{m}/\gcd(d,j_{m}). The map SαMih:2/jmS\alpha_{M_{i}}^{h}\colon\mathbb{Z}^{2}\to\mathbb{Z}/j_{m} factors as

SαMih:2/j¯m/jmS\alpha_{M_{i}}^{h}\colon\mathbb{Z}^{2}\to\mathbb{Z}/\overline{j}_{m}\to\mathbb{Z}/j_{m}

for all ii, where /j¯m/jm\mathbb{Z}/\overline{j}_{m}\to\mathbb{Z}/j_{m} is the standard inclusion sending 1gcd(d,jm)1\mapsto\gcd(d,j_{m}). Define

d¯:=dgcd(d,jm) and d¯:=dgcd(d,jm)=d¯ι=qA,m+2qA+1pι.\overline{d}:=\mbox{\footnotesize$\displaystyle\frac{d}{\gcd(d,j_{m})}$}\text{ and }\overline{d}^{\prime}:=\mbox{\footnotesize$\displaystyle\frac{d^{\prime}}{\gcd(d,j_{m})}$}=\overline{d}\cdot\prod_{\iota=q_{A,m}+2}^{q_{A}+1}p_{\iota}.

We obtain

Sα¯Mih=(d¯vid¯wi):2/j¯m.S\overline{\alpha}^{h}_{M_{i}}=\begin{pmatrix}\overline{d}v_{i}\\ \overline{d}^{\prime}w_{i}\end{pmatrix}\colon\mathbb{Z}^{2}\to\mathbb{Z}/\overline{j}_{m}.

It suffices to prove that for iji\neq j the resulting pairs {d¯vi,d¯wi}\{\overline{d}v_{i},\overline{d}^{\prime}w_{i}\} and {d¯vj,d¯wj}\{\overline{d}v_{j},\overline{d}^{\prime}w_{j}\} are distinct, up to signs and switching the orders. Note that gcd(d¯,j¯m)=1=gcd(d¯,j¯m)\gcd(\overline{d},\overline{j}_{m})=1=\gcd(\overline{d}^{\prime},\overline{j}_{m}).

Let pp be a prime dividing AA^{\prime}. Up to possibly changing the orders of viv_{i} and wiw_{i}, and of vjv_{j} and wjw_{j}, assume that pp divides viv_{i} and vjv_{j}. If so, pp does not divide wiw_{i} and wjw_{j}, since gcd(vi,wi)=1=gcd(vj,wj)\gcd(v_{i},w_{i})=1=\gcd(v_{j},w_{j}).

Now let qpq\neq p be a prime dividing AA^{\prime} such that either:

  1. (i)

    qq divides wjw_{j} but qq does not divide wiw_{i}; or

  2. (ii)

    qq divides wiw_{i} but qq does not divide wjw_{j}.

Without loss of generality suppose that (i) holds. Then also qq divides viv_{i} but qq does not divide vjv_{j}, since both pairs (vi,wi)(v_{i},w_{i}) and (vj,wj)(v_{j},w_{j}) are coprime. There exists such a qq, unless qA,m=0q_{A,m}=0, in which case 2qA,m=12^{q_{A,m}}=1 and we have nothing to prove anyway. So we can assume that qA,mq_{A,m} is positive and that such a qq exists. The idea is that the primes pp and qq are chosen so that they divide the same element of the unordered pair associated with the homotopy extended symmetric form for MiM_{i}, but divide different elements of the unordered pair for MjM_{j}. It is this distinction we want to detect.

We consider the images of the four elements d¯vi\overline{d}v_{i}, d¯wi\overline{d}^{\prime}w_{i}, d¯vj\overline{d}v_{j}, and d¯wj\overline{d}^{\prime}w_{j} of /j¯m\mathbb{Z}/\overline{j}_{m} under the canonical surjections

ρp:/j¯m/p and ρq:/j¯m/q.\rho_{p}\colon\mathbb{Z}/\overline{j}_{m}\to\mathbb{Z}/p\text{ and }\rho_{q}\colon\mathbb{Z}/\overline{j}_{m}\to\mathbb{Z}/q.

Since pp and qq divide j¯m\overline{j}_{m} and gcd(d¯,j¯m)=1=gcd(d¯,j¯m)\gcd(\overline{d},\overline{j}_{m})=1=\gcd(\overline{d}^{\prime},\overline{j}_{m}), we know that pp and qq do not divide d¯\overline{d} and do not divide d¯\overline{d}^{\prime}. Therefore for the /p\mathbb{Z}/p reductions we have

ρp(d¯vi)=0,ρp(d¯wi)0,ρp(d¯vj)=0, and ρp(d¯wj)0,\rho_{p}(\overline{d}v_{i})=0,\;\rho_{p}(\overline{d}^{\prime}w_{i})\neq 0,\;\rho_{p}(\overline{d}v_{j})=0,\text{ and }\rho_{p}(\overline{d}^{\prime}w_{j})\neq 0,

while for the /q\mathbb{Z}/q reductions we have:

ρq(d¯vi)=0,ρq(d¯wi)0,ρq(d¯vj)0, and ρq(d¯wj)=0.\rho_{q}(\overline{d}v_{i})=0,\;\rho_{q}(\overline{d}^{\prime}w_{i})\neq 0,\;\rho_{q}(\overline{d}v_{j})\neq 0,\text{ and }\rho_{q}(\overline{d}^{\prime}w_{j})=0.

We indicate one of these calculations briefly, that ρp(d¯wi)0\rho_{p}(\overline{d}^{\prime}w_{i})\neq 0, to give the idea. If d¯wi\overline{d}^{\prime}w_{i} were 0 modulo pp then for some a,ba,b\in\mathbb{Z} we would have ap=d¯wi+bj¯map=\overline{d}^{\prime}w_{i}+b\overline{j}_{m}\in\mathbb{Z}. But pj¯mp\mid\overline{j}_{m} and so pp divides d¯wi\overline{d}^{\prime}w_{i}, which is a contradiction.

Note that switching the sign of an element in /j¯m\mathbb{Z}/\overline{j}_{m} preserves whether or not its image under ρp\rho_{p} or ρq\rho_{q} is zero. Let us summarise the calculations above. For {d¯vi,d¯wi}\{\overline{d}v_{i},\overline{d}^{\prime}w_{i}\}, one element is zero under the reductions modulo pp and qq, while the other element is nonzero under both reductions. On the other hand, for the pair {d¯vj,d¯wj}\{\overline{d}v_{j},\overline{d}^{\prime}w_{j}\} we have shown that precisely one element is zero under each of the modulo pp and modulo qq reductions. Switching the orders of the elements and switching signs preserves these descriptions, and therefore MiM_{i} and  MjM_{j} are not homotopy equivalent. It follows that |𝒮hst(Ma,b)||\mathcal{S}^{\rm st}_{\rm h}(M_{a,b})| is at least 2qA,m2^{q_{A,m}}, as desired. ∎

4. spinc structures on 4-manifolds

As explained in the introduction, the homotopy stable class is trivial for every closed, simply-connected 4-manifold. However a parallel phenomenon occurs when one considers equivalence classes of spinc structures on the tangent bundle. In this section we illustrate this on S2×S2S^{2}\times S^{2}.

For all n2n\geq 2, the group Spinn\operatorname{Spin}_{n} is the connected double cover of SOnSO_{n} and the group /2\mathbb{Z}/2 acts by deck transformations. The group /2\mathbb{Z}/2 acts on U(1)S1U(1)\cong S^{1} by complex conjugation. We quotient out by the diagonal action on the product to obtain:

Spinnc:=U(1)×/2Spinn\operatorname{Spin}^{c}_{n}:=U(1)\times_{\mathbb{Z}/2}\operatorname{Spin}_{n}

There are well-defined maps

U(1)\textstyle{U(1)}Spinnc\textstyle{\operatorname{Spin}^{c}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pr2\scriptstyle{\operatorname{pr}_{2}}pr1\scriptstyle{\operatorname{pr}_{1}}SOn\textstyle{SO_{n}}

obtained as the composition of the double cover SpinncU(1)×SOn\operatorname{Spin}^{c}_{n}\to U(1)\times SO_{n} with the projections.

There are natural inclusions SpinncSpinn+1c\operatorname{Spin}^{c}_{n}\hookrightarrow\operatorname{Spin}^{c}_{n+1} and the stable spinc group is defined by Spinc:=colimnSpinnc\operatorname{Spin}^{c}:=\operatorname{colim}_{n\to\infty}\operatorname{Spin}^{c}_{n}. There are also stable projections

U(1)\textstyle{U(1)}Spinc\textstyle{\operatorname{Spin}^{c}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pr2\scriptstyle{\operatorname{pr}_{2}}pr1\scriptstyle{\operatorname{pr}_{1}}SO,\textstyle{SO,}

where SOSO is the stable special orthogonal group. We will use the same notation pr1\operatorname{pr}_{1}, pr2\operatorname{pr}_{2} for the induced maps on classifying spaces.

Definition 4.1.

Let MM be a closed, oriented nn-manifold. A spinc structure on MM is a lift

BSpinc\textstyle{B\operatorname{Spin}^{c}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pr2\scriptstyle{\operatorname{pr}_{2}}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔰\scriptstyle{\mathfrak{s}}τM\scriptstyle{\tau_{M}}BSO\textstyle{BSO}

of the stable tangent bundle’s classifying map to BSpincB\operatorname{Spin}^{c}.

For more background on spinc structures on 4-manifolds, we refer to e.g. [GS99, Section 2.4.1] and [Sco05, Sections 10.2 & 10.7].

Lemma 4.2 ([GS99, Prop. 2.4.16]).

Every oriented 44-manifold admits a spinc structure.

Proof.

In [GS99], spinc structures on 44-manifolds are defined by using BSpin4cB\operatorname{Spin}^{c}_{4} in place of BSpincB\operatorname{Spin}^{c}, and [GS99, Prop. 2.4.16] proves the existence of a lift of the classifying map of the (unstable) tangent bundle to BSpin4cB\operatorname{Spin}^{c}_{4}. Composing with the maps to the colimit, this implies the existence of a spinc-structure in the sense of Definition 4.1. ∎

Definition 4.3 (Equivalence of spinc structures).

Let MM be a closed, oriented 44-manifold.

  1. (1)

    Two spinc structures 𝔰1\mathfrak{s}_{1} and 𝔰2\mathfrak{s}_{2} on MM are equivalent if there is an orientation-preserving diffeomorphism f:MMf\colon M\to M such that 𝔰1,𝔰2f:MBSpinc\mathfrak{s}_{1},\mathfrak{s}_{2}\circ f\colon M\to B\operatorname{Spin}^{c} are homotopic over BSOBSO; i.e. there is homotopy KK and a commutative diagram

    BSpinc\textstyle{B\operatorname{Spin}^{c}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pr2\scriptstyle{\operatorname{pr}_{2}}M×I\textstyle{M\times I\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K\scriptstyle{K}τM×I\scriptstyle{\tau_{M\times I}}BSO\textstyle{BSO}

    where KK restricts to 𝔰1\mathfrak{s}_{1} on M×{0}M\times\{0\} and 𝔰2f\mathfrak{s}_{2}\circ f on M×{1}M\times\{1\}.

  2. (2)

    Two spinc structures 𝔰1\mathfrak{s}_{1} and 𝔰2\mathfrak{s}_{2} are homotopic if they are equivalent as in the previous item, with f=IdMf=\operatorname{Id}_{M}.

Recall that the projection onto the first component gives a compatible collection of maps pr1:BSpinncBU(1)\operatorname{pr}_{1}\colon B\operatorname{Spin}^{c}_{n}\to BU(1), nn\in\mathbb{N}. Therefore, passing to the colimit and keeping the same notation, we obtain a map pr1:BSpincBU(1)\operatorname{pr}_{1}\colon B\operatorname{Spin}^{c}\to BU(1).

Definition 4.4.

Via the map pr1:BSpincBU(1)\operatorname{pr}_{1}\colon B\operatorname{Spin}^{c}\to BU(1), a spinc structure 𝔰:MBSpinc\mathfrak{s}\colon M\to B\operatorname{Spin}^{c} on a 4-manifold MM determines a line bundle 𝔰\mathcal{L}_{\mathfrak{s}}. The first Chern class of 𝔰\mathfrak{s} is defined by

c1(𝔰):=c1(𝔰)H2(M).c_{1}(\mathfrak{s}):=c_{1}(\mathcal{L}_{\mathfrak{s}})\in H^{2}(M).

Noting that BU(1)BU(1) is a K(,2)K(\mathbb{Z},2), c1(𝔰)c_{1}(\mathfrak{s}) corresponds to pr1𝔰:MBU(1)\operatorname{pr}_{1}\circ\mathfrak{s}\colon M\to BU(1) under the isomorphism H2(M)[M,BU(1)]H^{2}(M)\cong[M,BU(1)]. The map pr1\operatorname{pr}_{1} can be interpreted as a determinant, and 𝔰\mathcal{L}_{\mathfrak{s}} is called the determinant line bundle of 𝔰\mathfrak{s}. The next lemma follows from [GS99, Proposition 2.4.16].

Lemma 4.5.

Let MM be a closed, oriented 4-manifold.

  1. (i)

    For every spinc structure 𝔰\mathfrak{s} on MM, reduction modulo two is such that:

    H2(M)\displaystyle H^{2}(M) H2(M;/2)\displaystyle\to H^{2}(M;\mathbb{Z}/2)
    c1(𝔰)\displaystyle c_{1}(\mathfrak{s}) w2(M),\displaystyle\mapsto w_{2}(M),

    where w2(M)w_{2}(M) is the second Stiefel-Whitney class.

  2. (ii)

    There is a transitive action of H2(M)H^{2}(M) on the set of homotopy classes of spinc structures on MM, such that for xH2(M)x\in H^{2}(M) we have

    c1(x𝔰)=c1(𝔰)+2xH2(M).c_{1}(x\cdot\mathfrak{s})=c_{1}(\mathfrak{s})+2x\in H^{2}(M).
  3. (iii)

    If H1(M)H_{1}(M) is 2-torsion free, then this action is free.

Proof.

As mentioned during the proof of Lemma 4.2, in [GS99] spinc structures are defined by using BSpin4cB\operatorname{Spin}^{c}_{4} in place of BSpincB\operatorname{Spin}^{c} and therefore the Chern class of a spinc-structure is defined using pr1:BSpin4cBU(1)\operatorname{pr}_{1}\colon B\operatorname{Spin}^{c}_{4}\to BU(1). However, since the map BSpin4cBU(1)B\operatorname{Spin}^{c}_{4}\to BU(1) factors through BSpincB\operatorname{Spin}^{c}, both definitions of the Chern class coincide and so the lemma follows from [GS99, Proposition 2.4.16]. ∎

As a consequence every characteristic cohomology class yH2(M)y\in H^{2}(M) can be realised as the first Chern class of some spinc structure on MM, and if H1(M)H_{1}(M) is 2-torsion free then this spinc structure is uniquely determined by yy. Here recall that yy being characteristic means that xx,[M]xy,[M]mod2\langle x\cup x,[M]\rangle\equiv\langle x\cup y,[M]\rangle\mod{2} for every xH2(M)x\in H^{2}(M).

The next lemma is immediate from the fact that the Chern class is an invariant of a spinc structure, and is natural.

Lemma 4.6.

If two spinc structures 𝔰1\mathfrak{s}_{1} and 𝔰2\mathfrak{s}_{2} on a closed, oriented 4-manifold MM are equivalent, then there is an isometry of the intersection form on H2(M)H^{2}(M) sending c1(𝔰1)c_{1}(\mathfrak{s}_{1}) to c1(𝔰2)c_{1}(\mathfrak{s}_{2}). ∎

To define stable equivalence of spinc structures, fix once and for all the preferred spinc structure 𝔰g\mathfrak{s}_{g} on Wg:=#gS2×S2W_{g}:=\#^{g}S^{2}\times S^{2}, to be the spinc structure with c1(𝔰g)=0H2(Wg)c_{1}(\mathfrak{s}_{g})=0\in H^{2}(W_{g}). Such a spinc structure exists by Lemma 4.5.

Definition 4.7 (Stable equivalence of spinc structures).

Two spinc structures 𝔰1\mathfrak{s}_{1} and 𝔰2\mathfrak{s}_{2} on a closed, oriented 4-manifold MM are stably equivalent if there exists g0g\in\mathbb{N}_{0} such that the induced spinc structures on M#WgM\#W_{g}, extending 𝔰1\mathfrak{s}_{1} and 𝔰2\mathfrak{s}_{2} using the fixed spinc structure 𝔰g\mathfrak{s}_{g} on WgW_{g}, are equivalent.

The stable classification of spinc structures 𝔰\mathfrak{s} on simply-connected 44-manifolds MM is analogous to the almost stable classification of (4m1)(4m{-}1)-connected 8m8m-manifolds from Theorem 3.8.

We will want to apply Kreck’s stable diffeomorphism theorem [Kre99, Theorem C], with appropriate 11-smoothings. In particular, a 1-smoothing has to be 2-connected. While we will work with simply-connected 4-manifolds, so π1(M)=0=π1(BSpinc)\pi_{1}(M)=0=\pi_{1}(B\operatorname{Spin}^{c}), the map MBSpincM\to B\operatorname{Spin}^{c} classifying a spinc structure need not be surjective on π2\pi_{2}. To mitigate this we make the following definition.

Given (M,𝔰)(M,\mathfrak{s}) we define the divisibility d(𝔰)0d(\mathfrak{s})\in\mathbb{N}_{0} of c1(𝔰)c_{1}(\mathfrak{s}) by the equation

c1(𝔰)(H2(M))=d(𝔰).c_{1}(\mathfrak{s})(H_{2}(M))=d(\mathfrak{s})\mathbb{Z}.

Let BSpinc(d)B\operatorname{Spin}^{c}(d) be the homotopy fibre of the mod dd spinc first Chern class, so that there is a fibre sequence

BSpinc(d)𝜋BSpincK(/d,2).B\operatorname{Spin}^{c}(d)\xrightarrow{\pi}B\operatorname{Spin}^{c}\to K(\mathbb{Z}/d,2).

By construction, π:BSpinc(d)BSpinc\pi\colon B\operatorname{Spin}^{c}(d)\to B\operatorname{Spin}^{c} is a fibration, and the universal stable bundle over BSpincB\operatorname{Spin}^{c} pulls back to a stable bundle over BSpinc(d)B\operatorname{Spin}^{c}(d).

Definition 4.8.

Let MM be a closed, oriented nn-manifold. A spin(d)c{}^{c}(d) structure on MM is a lift

BSpinc(d)\textstyle{B\operatorname{Spin}^{c}(d)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pr2π\scriptstyle{\operatorname{pr}_{2}\circ\pi}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔰(d)\scriptstyle{\mathfrak{s}(d)}τM\scriptstyle{\tau_{M}}BSO\textstyle{BSO}

of the stable tangent bundle’s classifying map to BSpinc(d)B\operatorname{Spin}^{c}(d). We denote a manifold with a BSpinc(d)B\operatorname{Spin}^{c}(d)-structure by (M,𝔰(d))(M,\mathfrak{s}(d)) and the corresponding bordism groups by ΩSpinc(d)\Omega_{*}^{\operatorname{Spin}^{c}(d)}.

Lemma 4.9.

The following assertions hold:

  1. (1)

    π2(BSpinc(d))=\pi_{2}(B\operatorname{Spin}^{c}(d))=\mathbb{Z} for d0d\neq 0 and π2(BSpinc(d))=0\pi_{2}(B\operatorname{Spin}^{c}(d))=0 for d=0d=0;

  2. (2)

    if (M,𝔰)(M,\mathfrak{s}) is a Spinc\operatorname{Spin}^{c}-manifold, then MM is a Spinc(d)\operatorname{Spin}^{c}(d)-manifold for d:=d(𝔰)d:=d(\mathfrak{s}) and the map 𝔰(d):MBSpinc(d)\mathfrak{s}(d)\colon M\to B\operatorname{Spin}^{c}(d) is 22-connected.

Proof.

We have π2(BSpinc)\pi_{2}(B\operatorname{Spin}^{c})\cong\mathbb{Z}, and so the long exact sequence of a fibration in homotopy groups yields

0=π3(K(/d,2))/d=π2(K(/d,2))0.0=\pi_{3}(K(\mathbb{Z}/d,2))\to\mathbb{Z}\to\mathbb{Z}\twoheadrightarrow\mathbb{Z}/d=\pi_{2}(K(\mathbb{Z}/d,2))\xrightarrow{}0.

Since also π1(BSpinc)=0\pi_{1}(B\operatorname{Spin}^{c})=0 we have that BSpinc(d)B\operatorname{Spin}^{c}(d) is 11-connected and π2(BSpinc(d))\pi_{2}(B\operatorname{Spin}^{c}(d))\cong\mathbb{Z} for d0d\neq 0. A similar calculation shows that π2(BSpinc(0))=π2(BSpin)=0.\pi_{2}(B\operatorname{Spin}^{c}(0))=\pi_{2}(B\operatorname{Spin})=0. In fact it then follows from Whitehead’s theorem that the map BSpinBSpinc(0)B\operatorname{Spin}\to B\operatorname{Spin}^{c}(0), obtained from factoring the canonical map BSpinBSpincB\operatorname{Spin}\to B\operatorname{Spin}^{c} through BSpinc(0)B\operatorname{Spin}^{c}(0), is a homotopy equivalence. This concludes the proof of the first assertion.

We now assume that (M,𝔰)(M,\mathfrak{s}) is a Spinc\operatorname{Spin}^{c}-manifold and prove the second assertion. The first point follows by observing that in the exact sequence

[M,BSpinc(d)][M,BSpinc][M,K(/d,2)]=H2(M;/d),[M,B\operatorname{Spin}^{c}(d)]\to[M,B\operatorname{Spin}^{c}]\to[M,K(\mathbb{Z}/d,2)]=H^{2}(M;\mathbb{Z}/d),

the spinc structure 𝔰[M,BSpinc]\mathfrak{s}\in[M,B\operatorname{Spin}^{c}] is mapped to zero, by definition of d(𝔰)d(\mathfrak{s}). It only remains to show that 𝔰(d)\mathfrak{s}(d) is 22-connected. Since this is clear for d=0d=0, we assume that d0d\neq 0. As we know that π2(BSpinc)=\pi_{2}(B\operatorname{Spin}^{c})=\mathbb{Z} and π2(BSpinc(d))=\pi_{2}(B\operatorname{Spin}^{c}(d))=\mathbb{Z}, the long exact sequence of the fibration and the definition of d=d(𝔰)d=d(\mathfrak{s}) imply that Im(𝔰)=d=π2(BSpinc)\operatorname{Im}(\mathfrak{s}_{*})=d\mathbb{Z}\subseteq\mathbb{Z}=\pi_{2}(B\operatorname{Spin}^{c}) and therefore 𝔰(d)\mathfrak{s}(d) is surjective on π2\pi_{2}, as required. ∎

Our aim is now to construct an injective map ΩSpinc(d)\Omega_{*}^{\operatorname{Spin}^{c}(d)}\to\mathbb{Z}\oplus\mathbb{Z}. The first component of this map will be the signature, while the second will arise as a characteristic number obtained from 𝔰(d):MBSpinc(d)\mathfrak{s}(d)\colon M\to B\operatorname{Spin}^{c}(d) by pulling back a universal class c1/dH2(BSpinc(d))c_{1}/d\in H^{2}(B\operatorname{Spin}^{c}(d)) that we now define. For d0d\neq 0, Lemma 4.9 implies that H2(BSpinc(d))H^{2}(B\operatorname{Spin}^{c}(d)) is an infinite cyclic group. It is generated by a class c1/dH2(BSpinc(d))c_{1}/d\in H^{2}(B\operatorname{Spin}^{c}(d)) such that the pullback π(c1)\pi^{*}(c_{1}) of the spinc first Chern class, satisfies

(ϖ\varpi) d(c1/d)=π(c1)H2(BSpinc(d)).d\big{(}c_{1}/d\big{)}=\pi^{*}(c_{1})\in H^{2}(B\operatorname{Spin}^{c}(d)).

For d=0d=0, H2(BSpinc(0))=H2(BSpin)=0H^{2}(B\operatorname{Spin}^{c}(0))=H^{2}(B\operatorname{Spin})=0, and we set c1/d=0c_{1}/d=0. As is conventional for characteristic classes, given a Spinc(d)\operatorname{Spin}^{c}(d)-structure 𝔰(d):MBSpinc(d)\mathfrak{s}(d)\colon M\to B\operatorname{Spin}^{c}(d) we write c1/d(𝔰(d)):=𝔰(d)(c1/d)H2(M)c_{1}/d(\mathfrak{s}(d)):=\mathfrak{s}(d)^{*}(c_{1}/d)\in H^{2}(M).

Lemma 4.10.

There is an injective homomorphism

Θ:Ω4Spinc(d)\displaystyle\Theta\colon\Omega_{4}^{\operatorname{Spin}^{c}(d)} \displaystyle\to\mathbb{Z}\oplus\mathbb{Z}
[N,𝔰(d)]\displaystyle[N,\mathfrak{s}(d)] (σ(N),(c1/d(𝔰(d)))2,[N]).\displaystyle\mapsto\big{(}\sigma(N),\langle(c_{1}/d(\mathfrak{s}(d)))^{2},[N]\rangle\big{)}.
Proof.

The given map is a homomorphism, and is a bordism invariant because the signature is bordism invariant, and because c12c_{1}^{2} is a characteristic number and therefore so is (c1/d)2(c_{1}/d)^{2}.

It remains to prove injectivity of Θ\Theta. Let (M,𝔰(d))(M,\mathfrak{s}(d)) be a spin(d)c{}^{c}(d)-manifold with vanishing signature and (c1/d(𝔰(d)))2=0(c_{1}/d(\mathfrak{s}(d)))^{2}=0. Since π1(BSpinc(d))=0\pi_{1}(B\operatorname{Spin}^{c}(d))=0, after preliminary surgeries over BSpinc(d)B\operatorname{Spin}^{c}(d) we may assume that MM is simply-connected. Since σ(M)=0\sigma(M)=0, the homeomorphism classification of smooth simply-connected 44-manifolds [Fre82] means that we can assume that MM is homeomorphic to one of the following model manifolds:

MTOP{Wgd even,Xgd odd,M\cong_{\rm TOP}\begin{cases}W_{g}&\text{$d$ even,}\\ X_{g}&\text{$d$ odd,}\end{cases}

where Xg:=#gS2×~S2X_{g}:=\#^{g}S^{2}\widetilde{\times}S^{2}. In other words MM is a possibly exotic WgW_{g} or XgX_{g}. Now, exotic pairs of simply-connected 4-manifolds are hh-cobordant [Wal64, Theorem 2], and the spin(d)c{}^{c}(d)-structure on MM propagates along an hh-cobordism to a spin(d)c{}^{c}(d) structure on either WgW_{g} or XgX_{g}, as appropriate. Hence we may assume that (M,𝔰(d))(M,\mathfrak{s}(d)) is diffeomorphic to either (Wg,𝔰g(d))(W_{g},\mathfrak{s}^{\prime}_{g}(d)) or (Xg,𝔰g′′(d))(X_{g},\mathfrak{s}^{\prime\prime}_{g}(d)) for some spin(d)c{}^{c}(d)-structures 𝔰g(d)\mathfrak{s}^{\prime}_{g}(d) or 𝔰g′′(d)\mathfrak{s}^{\prime\prime}_{g}(d). Now, MM has a standard coboundary NN, N=M\partial N=M, where

N{Ygd even,Zgd odd.N\cong\begin{cases}Y_{g}&\text{$d$ even,}\\ Z_{g}&\text{$d$ odd.}\end{cases}

Here Yg:=gD3×S2Y_{g}:=\natural^{g}D^{3}\times S^{2} and Zg:=gD3×~S2Z_{g}:=\natural^{g}D^{3}\widetilde{\times}S^{2}, where D3×~S2S2D^{3}\widetilde{\times}S^{2}\to S^{2} is the nontrivial bundle. By assumption (c1/d(𝔰(d)))2=0(c_{1}/d(\mathfrak{s}(d)))^{2}=0 and it follows that c1/d(𝔰(d))Lc_{1}/d(\mathfrak{s}(d))\in L, for some lagrangian LH2(M)L\subseteq H^{2}(M). Now, the automorphisms of the intersection form act transitively on the set of lagrangians (see for example [Wal64, pp. 144-5]), and Wall [Wal64, p. 144] also showed that every isometry of the intersection form of H2(M)H^{2}(M) is realised by a diffeomorphism. Hence we may assume that c1/d(𝔰(d))H2(M)c_{1}/d(\mathfrak{s}(d))\in H^{2}(M) lies in the standard lagrangian of H2(M)H^{2}(M), and so is the restriction to the boundary of cc for some cH2(N)c\in H^{2}(N). Since H2(N)H2(M)H^{2}(N)\to H^{2}(M) is onto a summand, it follows that NN admits a spin(d)c{}^{c}(d)-structure 𝔰N(d)\mathfrak{s}_{N}(d) that restricts to 𝔰(d)\mathfrak{s}(d). Hence (N,𝔰N(d))(N,\mathfrak{s}_{N}(d)) is a spin(d)c{}^{c}(d) null-bordism of (M,𝔰)(M,\mathfrak{s}), and so Θ\Theta is indeed injective. ∎

Next, using Lemma 4.10 we deduce the stable classification of spinc structures on simply-connected 4-manifolds.

The fibration sequence defining BSpinc(d)B\operatorname{Spin}^{c}(d) gives rise to an exact sequence

[M,ΩK(/d,2)][M,BSpinc(d)][M,BSpinc][M,K(/d,2)].[M,\Omega K(\mathbb{Z}/d,2)]\to[M,B\operatorname{Spin}^{c}(d)]\to[M,B\operatorname{Spin}^{c}]\to[M,K(\mathbb{Z}/d,2)].

If MM is simply-connected then [M,ΩK(/d,2)][M,K(/d,1)]H1(M;/d)=0[M,\Omega K(\mathbb{Z}/d,2)]\cong[M,K(\mathbb{Z}/d,1)]\cong H^{1}(M;\mathbb{Z}/d)=0, so if a lift of Spinc\operatorname{Spin}^{c} structure 𝔰\mathfrak{s} to a spin(d)c{}^{c}(d) structure 𝔰(d)\mathfrak{s}(d) exists, then it is essentially unique.

A spin(d)c{}^{c}(d) structure 𝔰(d)\mathfrak{s}(d) on MM induces a spin(d)c{}^{c}(d) structure on M#WgM\#W_{g}, for any gg: as in Definition 4.7 we extend the associated Spinc(d)\operatorname{Spin}^{c}(d) structure on MM by the spin(d)c{}^{c}(d) structure on WgW_{g} with c1=0c_{1}=0. By the previous paragraph, since WgW_{g} is simply connected, there is an essentially unique such Spinc(d)\operatorname{Spin}^{c}(d) structure on WgW_{g}. Then a lift to a spin(d)c{}^{c}(d) structure on MM determines such a lift on M#WgM\#W_{g}. We can therefore define stable equivalence of spin(d)c{}^{c}(d) structures. The definition is identical to the definition for spin(d)c{}^{c}(d) structure, just replacing Spinc\operatorname{Spin}^{c} with Spinc(d)\operatorname{Spin}^{c}(d) throughout Definition 4.3 (1) and Definition 4.7.

Theorem 4.11.

Let MM be a closed, oriented, simply-connected 4-manifold. Two spinc structures 𝔰1\mathfrak{s}_{1} and 𝔰2\mathfrak{s}_{2} on MM are stably equivalent if and only if d(𝔰1)=d(𝔰2)0d(\mathfrak{s}_{1})=d(\mathfrak{s}_{2})\in\mathbb{N}_{0} and c1(𝔰1)2=c1(𝔰2)2H4(M)c_{1}(\mathfrak{s}_{1})^{2}=c_{1}(\mathfrak{s}_{2})^{2}\in H^{4}(M).

Proof.

For the forward direction, the square of the Chern class and its divisibility are preserved by stable equivalence because we fixed the spinc structure on WgW_{g} to be the structure with trivial first Chern class, and because equivalence of spinc structures preserve Chern numbers and the divisibility.

For the reverse direction, by Lemma 4.10, for a fixed 4-manifold MM, two spinc structures 𝔰1\mathfrak{s}_{1} and 𝔰2\mathfrak{s}_{2} on MM with d(𝔰1)=d(𝔰2)=dd(\mathfrak{s}_{1})=d(\mathfrak{s}_{2})=d determine BSpinc(d)B\operatorname{Spin}^{c}(d)-structures 𝔰1(d)\mathfrak{s}_{1}(d) and 𝔰2(d)\mathfrak{s}_{2}(d), and therefore elements of Ω4Spinc(d)\Omega_{4}^{\operatorname{Spin}^{c}(d)}. Since c1(𝔰1)2=c1(𝔰2)2c_{1}(\mathfrak{s}_{1})^{2}=c_{1}(\mathfrak{s}_{2})^{2}, it follows that (c1/d(𝔰1(d)))2=(c1/d(𝔰2(d)))2(c_{1}/d(\mathfrak{s}_{1}(d)))^{2}=(c_{1}/d(\mathfrak{s}_{2}(d)))^{2}: for d=0d=0 this is automatic; for d0d\neq 0 apply π\pi^{*} to c1(𝔰1)2=c1(𝔰2)2c_{1}(\mathfrak{s}_{1})^{2}=c_{1}(\mathfrak{s}_{2})^{2} and use (ϖ\varpi). Therefore, since σ(M)\sigma(M) is independent of tangential structures, by Lemma 4.10 (M,𝔰1(d))(M,\mathfrak{s}_{1}(d)) and (M,𝔰1(d))(M,\mathfrak{s}_{1}(d)) are bordant over BSpinc(d)B\operatorname{Spin}^{c}(d). Then we apply Kreck’s stable diffeomorphism theorem [Kre99, Theorem C], which in the current situation implies that Spinc(d)\operatorname{Spin}^{c}(d) structures on MM are stably equivalent if they are bordant. Here we use that the maps MBSpinc(d)M\to B\operatorname{Spin}^{c}(d) are 1-smoothings by Lemma 4.9. Via π:BSpinc(d)BSpinc\pi\colon B\operatorname{Spin}^{c}(d)\to B\operatorname{Spin}^{c}, a stable equivalence of Spinc(d)\operatorname{Spin}^{c}(d) structures determines a stable equivalence of Spinc\operatorname{Spin}^{c} structures. ∎

Now we are ready to prove the main result of this section. Let CC\in\mathbb{Z} be such that |C|16|C|\geq 16 and 8C8\mid C. Define P(C):=|𝒫C/8|P(C):=|\mathcal{P}_{C/8}|, namely the number of distinct primes dividing C/8C/8. We consider S2×S2S^{2}\times S^{2} with a fixed orientation. This determines an identification H4(S2×S2)=H^{4}(S^{2}\times S^{2})=\mathbb{Z}.

Theorem 4.12.

For every CC\in\mathbb{Z} with |C|16|C|\geq 16 and 8C8\mid C, there are n:=2P(C)1n:=2^{P(C)-1} stably equivalent spinc structures 𝔰1,,𝔰n\mathfrak{s}_{1},\dots,\mathfrak{s}_{n} on S2×S2S^{2}\times S^{2} with c1(𝔰i)2=CH4(S2×S2)=c_{1}(\mathfrak{s}_{i})^{2}=C\in H^{4}(S^{2}\times S^{2})=\mathbb{Z}, that are all pairwise inequivalent.

Proof.

Let M:=S2×S2M:=S^{2}\times S^{2}. Let x,yH2(M)2x,y\in H^{2}(M)\cong\mathbb{Z}^{2} be generators dual to [pt×S2][\operatorname{pt}\times S^{2}] and [S2×pt][S^{2}\times\operatorname{pt}] respectively. So xy=1H4(M)xy=1\in H^{4}(M) while x2=y2=0x^{2}=y^{2}=0. Henceforth we identify H4(M)H^{4}(M)\cong\mathbb{Z}. Let Q:=C/8Q:=C/8. There are P(C)P(C) prime powers dividing QQ. Up to switching the order and multiplying both by 1-1, there are 2P(C)12^{P(C)-1} ways to write QQ as a product of coprime integers Q=q1q2Q=q_{1}q_{2}. For each such factorisation, let 𝔰i\mathfrak{s}_{i} be a spinc structure with

c1(𝔰i)=2q1x+2q2y.c_{1}(\mathfrak{s}_{i})=2q_{1}x+2q_{2}y.

Such spinc structures exist by Lemma 4.5: every characteristic element of H2(M)H^{2}(M) can be realised as the first Chern class of some spinc structure. Note that

c1(𝔰i)2=8q1q2=8Q=C=H4(S2×S2)c_{1}(\mathfrak{s}_{i})^{2}=8q_{1}q_{2}=8Q=C\in\mathbb{Z}=H^{4}(S^{2}\times S^{2})

and d(𝔰i)=2d(\mathfrak{s}_{i})=2 for every ii. Thus by Theorem 4.11, all the 𝔰i\mathfrak{s}_{i} are stably equivalent to one another. But as we saw in the proof of Proposition 2.2 there is no isometry of the intersection pairing of MM that sends (2q1,2q2)(2q_{1},2q_{2}) to (2q1,2q2)(2q_{1}^{\prime},2q_{2}^{\prime}) in H2(M)2H^{2}(M)\cong\mathbb{Z}^{2} for distinct unordered pairs {q1,q2}\{q_{1},q_{2}\} and {q1,q2}\{q_{1}^{\prime},q_{2}^{\prime}\}. By Lemma 4.6 it follows that the {𝔰i}\{\mathfrak{s}_{i}\} are pairwise inequivalent spinc structures. ∎

References

  • [Ada60] J. Frank Adams. On the non-existence of elements of Hopf invariant one. Ann. of Math. (2), 72:20–104, 1960.
  • [Ada66] J. Frank Adams. On the groups J(X)J(X). IV. Topology, 5:21–71, 1966.
  • [BHS19] Robert Burklund, Jeremy Hahn, and Andrew Senger. On the boundaries of highly connected, almost closed manifolds. Preprint, available at arXiv:1910.14116, 2019.
  • [BM58] Raoul Bott and John Milnor. On the parallelizability of the spheres. Bull. Amer. Math. Soc., 64:87–89, 1958.
  • [Bro72] William Browder. Surgery on simply-connected manifolds. Springer-Verlag, New York, 1972. Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 65.
  • [Bru68] Gregory Brumfiel. On the homotopy groups of BPL and PL/O. Ann. Math. (2), 88:291–311, 1968.
  • [BS20] Robert Burklund and Andrew Senger. On the high-dimensional geography problem. Preprint, available at arXiv:2007.05127, 2020.
  • [CCPS21] Anthony Conway, Diarmuid Crowley, Mark Powell, and Joerg Sixt. Stably diffeomorphic manifolds and modified surgery obstructions. Preprint; available at arXiv:2109.05632, 2021.
  • [CS11] Diarmuid Crowley and Joerg Sixt. Stably diffeomorphic manifolds and l2q+1([π])l_{2q+1}(\mathbb{Z}[\pi]). Forum Math., 23(3):483–538, 2011.
  • [Dav05] James Davis. The Borel/Novikov conjectures and stable diffeomorphisms of 4-manifolds. In Geometry and topology of manifolds, volume 47 of Fields Inst. Commun., pages 63–76. Amer. Math. Soc., Providence, RI, 2005.
  • [Don83] Simon K. Donaldson. An application of gauge theory to four-dimensional topology. J. Differential Geom., 18(2):279–315, 1983.
  • [Fre82] Michael Freedman. The topology of four-dimensional manifolds. J. Differential Geom., 17(3):357–453, 1982.
  • [GS99] Robert Gompf and András Stipsicz. 44-manifolds and Kirby calculus, volume 20 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 1999.
  • [Hae61] André Haefliger. Differentiable imbeddings. Bull. Amer. Math. Soc., 67:109–112, 1961.
  • [Hae62] André Haefliger. Differentiable links. Topology, 1:241–244, 1962.
  • [HK88a] Ian Hambleton and Matthias Kreck. On the classification of topological 44-manifolds with finite fundamental group. Math. Ann., 280(1):85–104, 1988.
  • [HK88b] Ian Hambleton and Matthias Kreck. Smooth structures on algebraic surfaces with cyclic fundamental group. Invent. Math., 91(1):53–59, 1988.
  • [HK93a] Ian Hambleton and Matthias Kreck. Cancellation, elliptic surfaces and the topology of certain four-manifolds. J. Reine Angew. Math., 444:79–100, 1993.
  • [HK93b] Ian Hambleton and Matthias Kreck. Cancellation of hyperbolic forms and topological four-manifolds. J. Reine Angew. Math., 443:21–47, 1993.
  • [IR90] Kenneth Ireland and Michael Rosen. A classical introduction to modern number theory, volume 84 of Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1990.
  • [JW54] I. M. James and J. H. C. Whitehead. The homotopy theory of sphere bundles over spheres. I. Proc. London Math. Soc. (3), 4:196–218, 1954.
  • [KM63] Michel A. Kervaire and John W. Milnor. Groups of homotopy spheres. I. Ann. of Math. (2), 77:504–537, 1963.
  • [KR20] Manuel Krannich and Jens Reinhold. Characteristic numbers of manifold bundles over surfaces with highly connected fibers. Journal of the London Mathematical Society, 102(2):879–904, 2020.
  • [Kre99] Matthias Kreck. Surgery and duality. Ann. of Math. (2), 149(3):707–754, 1999.
  • [KS84] Matthias Kreck and James A. Schafer. Classification and stable classification of manifolds: some examples. Comment. Math. Helv., 59:12–38, 1984.
  • [Lev85] J. P. Levine. Lectures on groups of homotopy spheres. In Algebraic and geometric topology (New Brunswick, N.J., 1983), volume 1126 of Lecture Notes in Math., pages 62–95. Springer, Berlin, 1985.
  • [Lüc02] Wolfgang Lück. A basic introduction to surgery theory. In Topology of high-dimensional manifolds, No. 1, 2 (Trieste, 2001), volume 9 of ICTP Lect. Notes, pages 1–224. Abdus Salam Int. Cent. Theoret. Phys., Trieste, 2002.
  • [Mil58a] John Milnor. On simply connected 44-manifolds. In Symposium internacional de topología algebraica International symposi um on algebraic topology, pages 122–128. Universidad Nacional Autónoma de México and UNESCO, Mexico City, 1958.
  • [Mil58b] John Milnor. On the Whitehead homomorphism JJ. Bull. Amer. Math. Soc., 64:79–82, 1958.
  • [MK60] John W. Milnor and M. Kervaire. Bernoulli numbers, homotopy groups, and a theorem of Rohlin. In Proc. Internat. Congress Math. 1958, pages 454–458. Cambridge Univ. Press, New York, 1960.
  • [Qui71] Daniel Quillen. The Adams conjecture. Topology, 10:67–80, 1971.
  • [Sco05] Alexandru Scorpan. The wild world of 4-manifolds. American Mathematical Society, Providence, RI, 2005.
  • [Sto85] Stephan Stolz. Hochzusammenhängende Mannigfaltigkeiten und ihre Ränder, volume 1116. Springer, Cham, 1985.
  • [Tod62] Hirosi Toda. Composition methods in homotopy groups of spheres. Annals of Mathematics Studies, No. 49. Princeton University Press, Princeton, N.J., 1962.
  • [Wal62] C. T. C. Wall. Classification of (n1)(n{-}1)-connected 2n2n-manifolds. Ann. of Math. (2), 75:163–189, 1962.
  • [Wal64] C. T. C. Wall. On simply-connected 4-manifolds. J. Lond. Math. Soc., 39:141–149, 1964.
  • [Wal67] C. T. C. Wall. Classification problems in differential topology. VI. Classification of (s1)(s{-}1)-connected (2s+1)(2s{+}1)-manifolds. Topology, 6:273–296, 1967.
  • [Wal99] Charles Terence Clegg Wall. Surgery on compact manifolds. American Mathematical Society, Providence, RI, second edition, 1999. Edited and with a foreword by A. A. Ranicki.
  • [Whi42] George W. Whitehead. On the homotopy groups of spheres and rotation groups. Ann. of Math. (2), 43:634–640, 1942.
  • [Whi49] J. H. C. Whitehead. On simply connected, 44-dimensional polyhedra. Comment. Math. Helv., 22:48–92, 1949.