By Lemma 1.2, the extension
$$\begin{aligned} 0\longrightarrow \Theta _{2n+1}\longrightarrow \Gamma ^n_{g,1}\longrightarrow \Gamma ^n_{g,1/2}\longrightarrow 0 \end{aligned}$$
discussed in Sect. 1.3 is central and is as such classified by a class in \(\mathrm {H}^2(\Gamma ^n_{g,1/2};\Theta _{2n+1})\) with \(\Gamma ^n_{g,1/2}\) acting trivially on \(\Theta _{2n+1}\). In this section, we identify this extension class in terms of the algebraic description of \(\Gamma ^n_{g,1/2}\) provided in the previous section, leading to a proof of our main result Theorem B. Our approach is partially based on ideas of Galatius–Randal-Williams [21, Sect. 7], who determined the extension for \(n\equiv 5\ (\mathrm {mod}\ 8)\) and \(g\ge 5\) up to automorphisms of \(\Theta _{2n+1}\).
We begin with an elementary recollection on the relation between Pontryagin classes and obstructions to extending trivialisations of vector bundles, mainly to fix notation.
Obstructions and Pontryagin classes
Let \(k\ge 1\) and \(\xi :X\rightarrow \tau _{>4k-1}\mathrm {BO}\) be a map to the \((4k-1)\)st connected cover of \(\mathrm {BO}\) with a lift \(\bar{\xi }:A\rightarrow \tau _{>4k}\mathrm {BO}\) over a subspace \(A\subset X\) along the canonical map \(\tau _{>4k}\mathrm {BO}\rightarrow \tau _{>4k-1}\mathrm {BO}\). Such data has a relative Pontryagin class \(p_k(\xi ,\bar{\xi })\in \mathrm {H}^{4k}(X,A;\mathbf {Z})\) given as the pullback along the map \((\xi ,\bar{\xi }):(X,A)\rightarrow (\tau _{>4k-1}\mathrm {BO},\tau _{>4k}\mathrm {BO})\) of the unique lift to \(\mathrm {H}^{4k}(\tau _{>4k-1}\mathrm {BO},\tau _{>4k}\mathrm {BO};\mathbf {Z})\) of the pullback \(p_k\in \mathrm {H}^{4k}(\tau _{>4k-1}\mathrm {BO};\mathbf {Z})\) of the usual Pontryagin class \(p_k\in \mathrm {H}^{4k}(\mathrm {BSO};\mathbf {Z})\). The class \(p_k(\xi ,\bar{\xi })\) is related to the primary obstruction \(\chi (\xi ,\bar{\xi })\in \mathrm {H}^{4k}(X,A;\pi _{4k-1}\mathrm {SO})\) to solving the lifting problem
by the equality
$$\begin{aligned} p_k(\xi ,\bar{\xi })=\pm a_k(2k-1)!\cdot \chi (\xi ,\bar{\xi }), \end{aligned}$$
up to the choice of a generator \(\pi _{4k-1}\mathrm {SO}\cong \mathbf {Z}\) (cf. [44, Lem. 2]). We suppress the lift \(\bar{\xi }\) from the notation whenever there is no source of confusing. For us, \(X=M\) will usually be a compact oriented 8k-manifold and \(A=\partial M\) its boundary, in which case we can evaluate \(\chi ^2(\xi ,\bar{\xi })\in \mathrm {H}^{8k}(M,\partial M;\mathbf {Z})\) against the relative fundamental class \([M,\partial M]\) to obtain a number \(\chi ^2(\xi ,\bar{\xi })\in \mathbf {Z}\). The following two sources of manifolds are relevant for us.
Example 3.1
Fix an integer \(n\equiv 3\ (\mathrm {mod}\ 4)\).
-
(i)
For a compact oriented n-connected \((2n+2)\)-manifold whose boundary is a homotopy sphere, there is a (up to homotopy) unique lift \(M\rightarrow \tau _{>n}\mathrm {BO}\) of the stable oriented normal bundle. On the boundary \(\partial M\), this lifts uniquely further to \(\tau _{>n+1}\mathrm {BO}\), so we obtain a canonical class \(\chi (M)\in \mathrm {H}^{n+1}(M,\partial M;\mathbf {Z})\) and a characteristic number \(\chi ^2(M)\in \mathbf {Z}\).
-
(ii)
Consider a \((W_{g,1},D^{2n-1})\)-bundle \(\pi :E\rightarrow B\), i.e. a smooth \(W_{g,1}\)-bundle with a trivialised \(D^{2n-1}\)-subbundle of its \(\partial W_{g,1}\)-bundle \(\partial \pi :\partial E\rightarrow B\) of boundaries. The standard framing of \(D^{2n-1}\) induces a trivialisation of stable vertical tangent bundle \(T_\pi E:E\rightarrow \mathrm {BSO}\) over the subbundle \(B\times D^{2n-1}\subset \partial E\), which extends uniquely to a \(\tau _{>n+1}\mathrm {BO}\)-structure on \(T_\pi E|_{\partial E}\) by obstruction theory. Using that \(W_g\) is n-parallelisable, another application of obstruction theory shows that the induced \(\tau _{>n}\mathrm {BO}\)-structure on \(T_\pi E|_{\partial E}\) extends uniquely to a \(\tau _{>n}\mathrm {BO}\)-structure on \(T_\pi E\), so the above discussion provides a class \(\chi (T_\pi E)\in \mathrm {H}^{n+1}(E,\partial E;\mathbf {Z})\), and, assuming B is an oriented closed surface, a number \(\chi ^2(T_\pi E)\in \mathbf {Z}\).
Highly connected almost closed manifolds
As a consequence of Theorem 3.12, we shall see that \((W_{g,1},D^{2n-1})\)-bundles over surfaces are closely connected to n-connected almost closed \((2n+2)\)-manifolds. These manifolds were classified by Wall [55], which we now recall for \(n\ge 3\) in a form tailored to later applications, partly following [34, Sect. 2].
A compact manifold M is almost closed if its boundary is a homotopy sphere. We write \(A_{d}^{\tau _{>n}}\) for the abelian group of almost closed oriented n-connected d-manifolds up to oriented n-connected bordism restricting to an h-cobordism on the boundary. Recall that \(\Omega _{d}^{\tau _{>n}}\) denotes the bordism group of closed d-manifolds M equipped with a \(\tau _{>n}\mathrm {BO}\)-structure on their stable normal bundle \(M\rightarrow \mathrm {BO}\), i.e. a lift \(M\rightarrow \tau _{>n}\mathrm {BO}\) to the n-connected cover. By classical surgery, the group \(\Omega _{d}^{\tau _{>n}}\) is canonically isomorphic to the bordism group of closed oriented n-connected d-manifolds up to n-connected bordism as long as \(d\ge 2n+1\), so we will use both descriptions interchangeably. There is an exact sequence
$$\begin{aligned} \Theta _{2n+2}\longrightarrow \Omega _{2n+2}^{\tau _{>n}}\longrightarrow A_{2n+2}^{\tau _{>n}}\overset{\partial }{\longrightarrow }\Theta _{2n+1}\longrightarrow \Omega _{2n+1}^{\tau _{>n}}\longrightarrow 0 \end{aligned}$$
(3.1)
due to Wall [57, p. 293] in which the two outer morphisms are the obvious ones, noting that homotopy d-spheres n-connected for \(n<d\). The morphisms \(\Omega _{2n+2}^{\tau _{>n}}\rightarrow A_{2n+2}^{\tau _{>n}}\) and \(\partial :A_{2n+2}^{\tau _{>n}}\rightarrow \Theta _{2n+1}\) are given by cutting out an embedded disc and by assigning to an almost closed manifold its boundary, respectively. By surgery theory, the subgroup
$$\begin{aligned} \mathrm {bA}_{2n+2}\,{:}{=}\,\mathrm {im}(A_{2n+2}^{\tau _{>n}}\overset{\partial }{\longrightarrow }\Theta _{2n+1}) \end{aligned}$$
of homotopy \((2n+1)\)-spheres bounding n-connected manifolds contains the cyclic subgroup \(\mathrm {bP}_{2n+2}\subset \Theta _{2n+2}\) of homotopy \((2n+1)\)-spheres bounding parallelisable manifolds, so the right end of (3.1) receives canonical a map from Kervaire–Milnor’s exact sequence [30],
which in particular induces a morphism \(\mathrm {coker}(J)_{2n+1}\rightarrow \Omega ^{\tau _{>n}}_{2n+1}\), concretely given by representing a class in \(\mathrm {coker}(J)_{2n+1}\) by a stably framed manifold and restricting its stable framing to a \(\tau _{>n}\mathrm {BO}\)-structure.
Wall’s classification
For our purposes, Wall’s computation [55, 57] of \(A_{2n+2}^{\tau _{>n}}\) is for \(n\ge 3\) odd is most conveniently stated in terms of two particular almost closed n-connected \((2n+2)\)-manifolds, namely
-
(i)
Milnor’s \(E_8\)-plumbing P, arising from plumbing together 8 copies of the disc bundle of the tangent bundle of the standard \((n+1)\)-sphere such that the intersection form of P agrees with the \(E_8\)-form (see e.g.[8, Ch. V.2]), and
-
(ii)
the manifold Q, obtained from plumbing together two copies of a linear \(D^{n+1}\)-bundle over the \((n+1)\)-sphere representing a generator of \(S\pi _n\mathrm {SO}(n)\).
The following can be derived from Wall’s work (see e.g. [34, Thm 2.1]).
Theorem 3.2
(Wall) For \(n\ge 3\) odd, the bordism group \(A_{2n+2}^{\tau _{>n}}\) satisfies
$$\begin{aligned} A_{2n+2}^{\tau _{>n}}\cong {\left\{ \begin{array}{ll} \mathbf {Z}\oplus \mathbf {Z}/2&{}\quad \text{ if } n\equiv 1\ (\mathrm {mod}\ 8)\\ \mathbf {Z}\oplus \mathbf {Z}&{}\quad \text{ if } n\equiv 3\ (\mathrm {mod}\ 4)\\ \mathbf {Z}&{}\quad \text{ if } n\equiv 5\ (\mathrm {mod}\ 8). \end{array}\right. } \end{aligned}$$
The first summand is generated by P in all cases but \(n=3,7\) where it is generated by \(\mathbf {H}P^2\) and \(\mathbf {O}P^2\). The second summand for \(n\equiv 5\ (\mathrm {mod}\ 8)\) is generated by Q.
From a consultation of Table 1, one sees that the group \(S\pi _n\mathrm {SO}(n)\) vanishes for \(n\equiv 5\ (\mathrm {mod}\ 8)\), so \(Q\in A_{2n+2}^{\tau _{>n}}\) is trivial in this case, which shows that the subgroup \(\mathrm {bA}_{2n+2}\) is for all \(n\ge 3\) odd generated by the boundaries
$$\begin{aligned} \Sigma _P\,{:}{=}\,\partial P\in \mathrm {bA}_{2n+2}\quad \text {and}\quad \Sigma _Q\,{:}{=}\,\partial Q\in \mathrm {bA}_{2n+3}. \end{aligned}$$
In the cases \(n\equiv 1\ (\mathrm {mod}\ 8)\) in which Q defines a \(\mathbf {Z}/2\)-summand, its boundary \(\Sigma _Q\) is trivial by a result of Schultz [50, Cor. 3.2, Thm 3.4 iii)]. For \(n\equiv 3\ (\mathrm {mod}\ 4)\) on the other hand, it is nontrivial by a calculation of Kosinski [31, p. 238–239].
Theorem 3.3
(Kosinski, Schultz) The homotopy sphere \(\Sigma _Q\in \Theta _{2n+1}\) is trivial for \(n\equiv 1\ (\mathrm {mod}\ 4)\) and nontrivial for \(n\equiv 3\ (\mathrm {mod}\ 4)\)
In the exceptional dimensions \(n=3,7\), the homotopy sphere \(\Sigma _Q\) agrees with the inverse of the Milnor sphere \(\Sigma _P\), as explained in [34, Cor. 2.8 ii)].
Lemma 3.4
\(\Sigma _Q=-\Sigma _P\) for \(n=3,7\).
For \(n\ge 3\) odd, the Milnor sphere \(\Sigma _P\in \Theta _{2n+1}\) is well-known to be nontrival and to generate the cyclic subgroup \(\mathrm {bP}_{2n+2}\subset \Theta _{2n+1}\) whose order can be expressed in terms of numerators of divided Bernoulli numbers (see e.g. [41, Lem. 3.5 (2), Cor. 3.20]), so Theorem 3.3 has the following corollary.
Corollary 3.5
For \(n\ge 3\) odd, the subgroup \(\mathrm {bA}_{2n+1}\) is nontrivial. It is generated by \(\Sigma _P\) for \(n\equiv 1\ (\mathrm {mod}\ 4)\), by \(\Sigma _Q\) for \(n=3,7\), and by \(\Sigma _P\) and \(\Sigma _Q\) for \(n\equiv 3\ (\mathrm {mod}\ 4)\).
Combining the previous results with the diagram (3.2), we obtain the following result, which we already mentioned in the introduction.
Corollary 3.6
The natural morphism \(\mathrm {coker}(J)_{2n+1}\rightarrow \Omega ^{\tau _{>n}}_{2n+2}\) is an isomorphism for \(n\equiv 1\ (\mathrm {mod}\ 4)\) and for \(n=3,7\). For \(n\equiv 3\ (\mathrm {mod}\ 4)\), it is an epimorphism whose kernel is generated by the class \([\Sigma _Q]\in \mathrm {coker}(J)_{2n+1}\).
\([\Sigma _Q]\in \mathrm {coker}(J)_{2n+1}\) is conjecturally trivial [21, Conj. A] for all n odd. Until recently (see Remark 3.8), this was only known for \(n=3,7\) and \(n\equiv 1\ (\mathrm {mod}\ 4)\) by the results above.
Conjecture 3.7
(Galatius–Randal-Williams) \([\Sigma _Q]=0\) for all \(n\ge 3\) odd.
Remark 3.8
As mentioned in the introduction, after the completion of this work, Burklund–Hahn–Senger [11] and Burklund–Senger [12] showed that \([\Sigma _Q]\) vanishes in \(\mathrm {coker}(J)_{2n+1}\) for n odd if and only if \(n\ne 11\), confirming Conjecture 3.7 for \(n\ne 11\) and disproving it for \(n=11\). This has as a consequence that, for \(n\ne 11\) odd, the subgroup \(\mathrm {bA}_{2n+1}\) is generated by \(\Sigma _P\) even for \(n\equiv 3\ (\mathrm {mod}\ 4)\) and that the morphism \(\mathrm {coker}(J)_{2n+1}\rightarrow \Omega _{2n+2}^{\tau _{>n}}\) discussed in Corollary 3.6 is an isomorphism.
Invariants
It follows from Theorems 3.2 and 3.3 that the boundary of an n-connected almost closed \((2n+2)\)-manifold M is determined by at most two integral bordism invariants of M. Concretely, we consider the signature \(\mathrm {sgn}:A_{2n+2}^{\tau _{>n}}\rightarrow \mathbf {Z}\) and for \(n\equiv 3\ (\mathrm {mod}\ 4)\) the characteristic number \(\chi ^2:A_{2n+2}^{\tau _{>n}}\rightarrow \mathbf {Z}\), explained in Example 3.1. As discussed for example in [34, Sect. 2.1], these functionals evaluate to
$$\begin{aligned} \mathrm {sgn}(P)=8 \quad \quad \mathrm {sgn}(Q)=0\quad \chi ^2(P)=0\quad \chi ^2(Q)={\left\{ \begin{array}{ll}8&{}\quad \text {for }n=3,7\\ 2&{}\quad \text {otherwise}\end{array}\right. }, \end{aligned}$$
(3.3)
and on the closed manifolds \(\mathbf {H}P^2\) and \(\mathbf {O}P^2\) to
$$\begin{aligned} \mathrm {sgn}(\mathbf {H}P^2)=\mathrm {sgn}(\mathbf {O}P^2)=1\quad \quad \chi ^2(\mathbf {H}P^2)=\chi ^2(\mathbf {O}P^2)=1, \end{aligned}$$
(3.4)
which results in the following formula for boundary spheres of highly connected manifolds when combined with the discussion above.
Proposition 3.9
For \(n\ge 3\) odd, the boundary \(\partial M\in \Theta _{2n+1}\) of an almost closed oriented n-connected \((2n+2)\)-manifold M satisfies
$$\begin{aligned} \partial M= {\left\{ \begin{array}{ll} \mathrm {sgn}(M)/8\cdot \Sigma _P&{}\quad \text{ if } n\equiv 1\ (\mathrm {mod}\ 4)\\ \mathrm {sgn}(M)/8\cdot \Sigma _P+\chi ^2(M)/2\cdot \Sigma _Q&{}\quad \text{ if } n\equiv 3\ (\mathrm {mod}\ 4)\text { and }n\ne 3,7\\ \big (\chi ^2(M)-\mathrm {sgn}(M)\big )/8\cdot \Sigma _Q&{}\quad \text{ if } n=3,7. \end{array}\right. } \end{aligned}$$
The minimal signature
As in the introduction, we denote by \(\sigma _n'\) the minimal positive signature of a smooth closed n-connected \((2n+2)\)-manifold. This satisfies \(\sigma _n'=1\) for \(n=1,3,7\) as witnessed by \(\mathbf {C}P^2\), \(\mathbf {H}P^2\), and \(\mathbf {O}P^2\), and in all other cases, it can be expressed in terms of the subgroup \(\mathrm {bA}_{2n+2}\subset \Theta _{2n+1}\) as follows.
Lemma 3.10
For \(n\ge 3\) odd, the quotient \(\mathrm {bA}_{2n+2}/\langle \Sigma _Q \rangle \) is a cyclic group generated by the class of \(\Sigma _P\). It is trivial if \(n=3,7\) and of order \(\sigma _n'/8\) otherwise.
Proof
For \(n=3,7\), the claim is a consequence of Theorem 3.2 and Lemma 3.4. In the case \(n\ne 3,7\), it follows from taking vertical cokernels in the commutative diagram
with exact rows, obtained from a combination of Theorem 3.2 with (3.1) and (3.3). \(\square \)
Remark 3.11
In [34, Prop. 2.15], the minimal signature \(\sigma _n'\) was expressed in terms of Bernoulli numbers and the order of \([\Sigma _Q]\in \mathrm {coker}(J)_{2n+1}\), from which one can conclude that for \(n\ne 1,3,7\), the signature of such manifolds is divisible by \(2^{n+3}\) if \((n+1)/2\) is odd and by \(2^{n-2\nu _2(n+1)}\) otherwise, where \(\nu _2(-)\) denotes the 2-adic valuation (see [34, Cor. 2.18]).
Bundles over surfaces and almost closed manifolds
In order to identify the cohomology class in \(\mathrm {H}^2(\Gamma _{g,1/2}^n;\Theta _{2n+1})\) that classifies the central extension
$$\begin{aligned} 0\longrightarrow \Theta _{2n+1}\longrightarrow \Gamma ^n_{g,1}\longrightarrow \Gamma ^n_{g,1/2}\longrightarrow 0, \end{aligned}$$
(3.5)
we first determine how it evaluates against homology classes, i.e. identify its image
$$\begin{aligned} d_2:\mathrm {H}_2(\Gamma ^n_{g,1/2};\mathbf {Z})\longrightarrow \Theta _{2n+1} \end{aligned}$$
under the map h participating in the universal coefficient theorem
$$\begin{aligned}&0\rightarrow \mathrm {Ext}(\mathrm {H}_1(\Gamma ^n_{g,1/2};\mathbf {Z}),\Theta _{2n+1})\rightarrow \mathrm {H}^2(\Gamma ^n_{g,1/2};\Theta _{2n+1})\\&\quad \overset{h}{\longrightarrow }\mathrm {Hom}(\mathrm {H}_2(\Gamma ^n_{g,1/2};\mathbf {Z});\Theta _{2n+1})\rightarrow 0, \end{aligned}$$
followed by resolving the remaining ambiguity originating from the Ext-term. Indicated by our choice of notation, the morphism \(d_2\) can be viewed alternatively as the first possibly nontrivial differential in the \(E_2\)-page of the Serre spectral sequence of the extension (3.5) (cf. [26, Thm 4]). Before identifying this differential, we remind the reader of two standard facts we shall make frequent use of.
-
(i)
The canonical map of spectra \(\mathbf {MSO}\rightarrow \mathbf {HZ}\) is 4-connected, so pushing forward fundamental classes induces an isomorphism \(\Omega ^{\mathrm {SO}}_*(X)\rightarrow \mathrm {H}_*(X;\mathbf {Z})\) for \(*\le 3\) and any space X, and
-
(ii)
the 1-truncation of a connected space X (in particular the natural map \(\mathrm {B}G\rightarrow \mathrm {B}\pi _0G\) for a topological group G) induces a surjection \(\mathrm {H}_2(X;\mathbf {Z}) \twoheadrightarrow \mathrm {H}_2(K(\pi _1X,1);\mathbf {Z})\), whose kernel agrees with the image of the Hurewicz homomorphism \(\pi _2X\rightarrow \mathrm {H}_2(X;\mathbf {Z})\).
The key geometric ingredient to identify the differential \(d_2\) is the following result.
Theorem 3.12
Let \(n\ge 3\) be odd and \(\pi :E\rightarrow S\) a \((W_{g,1},D^{2n-1})\)-bundle over an oriented closed surface S. There exists a class \(E'\in A_{2n+2}^{\tau _{>n}}\) such that
-
(i)
its boundary \(\partial E'\in \Theta _{2n+1}\) is the image of the class \([\pi ]\in \mathrm {H}^2(\mathrm {BDiff}_{\partial /2}(W_{g,1});\mathbf {Z})\) under the composition
-
(ii)
it satisfies \(\mathrm {sgn}(E)=\mathrm {sgn}(E')\), and \(\chi ^2(T_\pi E)=\chi ^2(E')\) if \(n\equiv 3\ (\mathrm {mod}\ 4).\)
Proof
By the isotopy extension theorem, the restriction map to the moving part of the boundary \(\mathrm {Diff}_{\partial /2}(W_{g,1})\rightarrow \mathrm {Diff}_\partial (D^{2n-1})\) is a fibration. As its image is contained in the component of the identity (see the proof of Lemma 1.2), this fibration induces the upper row in a map of fibrations
whose bottom row is induced by the extension (3.5). The two right vertical maps are induced by taking components and the left vertical map is the induced map on homotopy fibres. The latter agrees with the delooping of the Gromoll map \(\Omega \mathrm {Diff}_\partial (D^{2n-1})\rightarrow \mathrm {Diff}_\partial (D^{2n})\) followed by taking components, which one checks by looping the fibre sequences and using that
$$\begin{aligned} \Omega \mathrm {Diff}^{\mathrm {id}}_\partial (D^{2n-1})\longrightarrow \mathrm {Diff}_{\partial }(W_{g,1}) \end{aligned}$$
is given by “twisting” a collar \( [0,1]\times S^{2n-1}\subset W_{g,1}\), meaning that it sends a smooth loop \(\gamma \in \Omega \mathrm {Diff}^{\mathrm {id}}_\partial (D^{2n-1})\subset \Omega \mathrm {Diff}^{\mathrm {id}}_\partial (S^{2n-1})\) to the diffeomorphism that is the identity outside the collar and is given by \((t,x)\mapsto (t,\gamma (t)\cdot x)\) on the collar. Now consider the commutative square
obtained from delooping (3.6) once to the right and using \(\mathrm {H}_2(B^2\Theta _{2n+1})\cong \Theta _{2n+1}\). By transgression, the bottom arrow agrees with the differential \(d_2\) in the statement. Combining this with the Hurewicz theorem, the square (3.7) gives a factorisation
$$\begin{aligned} \mathrm {H}_2(\mathrm {BDiff}_{\partial /2}(W_{g,1}))\rightarrow \mathrm {H}_2(\mathrm {BDiff}_{\partial }^{\mathrm {id}}(D^{2n-1}))\cong \pi _2(\mathrm {BDiff}_{\partial }^{\mathrm {id}}(D^{2n-1}))\rightarrow \Theta _{2n+1} \end{aligned}$$
of the map in the first part of the statement, which thus has the following geometric description: a smooth \((W_{g,1},D^{2n-1})\)-bundle \(\pi :E\rightarrow S\) represents a class \([\pi ]\in \mathrm {H}^2(\mathrm {BDiff}_{\partial /2}(W_{g,1});\mathbf {Z})\) and its image under the first map in the composition is the class \([\pi _+]\in \mathrm {H}_2(\mathrm {BDiff}^{\mathrm {id}}_{\partial }(D^{2n-1});\mathbf {Z})\) of its \((\partial W_{g,1},D^{2n-1})\)-bundle \(\pi _+:E_+\rightarrow S\) of boundaries, which in turn maps under the inverse of the Hurewicz homomorphism to a \((\partial W_{g,1},D^{2n-1})\)-bundle \(\pi _-:E_-\rightarrow S^2\) over the 2-sphere that is bordant, as a bundle, to \(\pi _-\). That is, there exists a \((\partial W_{g,1},D^{2n-1})\)-bundle \(\bar{\pi }:\bar{E}\rightarrow K\) over an oriented bordism K between S and \(S^2\) that restricts to \(\pi _+\) over S and to \(\pi _-\) over \(S^2\). We claim that the image of the \((\partial W_{g,1},D^{2n-1})\)-bundle \(\pi _-\) under the final map in the composition is the homotopy sphere \(\Sigma _\pi \in \Theta _{2n+1}\) obtained by doing surgery on the total space \(E_-\) along the trivialised subbundle \(D^{2n-1}\times S^2\subset E_-\). This is most easily seen by thinking of a class in \(\pi _k\mathrm {BDiff}_\partial (D^d)\) as a smooth bundle \(D^d\rightarrow P\rightarrow D^k\) together with a trivialisation \(\varphi :D^d\times \partial D^k\cong P|_{\partial D^k}\) and a trivialised \(\partial D^k\)-subbundle \(\psi :\partial D^d\times D^k\hookrightarrow P\) such that \(\varphi \) and \(\psi \) agree on \(\partial D^d\times \partial D^k\). From this point of view, the morphism \(\pi _k\mathrm {BDiff}_\partial (D^d)\rightarrow \pi _{k-1}\mathrm {BDiff}_\partial (D^{d}\times D^1)\cong \pi _{k-1}\mathrm {BDiff}_\partial (D^{d+1})\) induced by the Gromoll map is given by sending such a bundle \(p:P\rightarrow D^k\cong D^{k-1}\times D^1\) to the \((D^d\times D^1)\)-bundle \((\mathrm {pr}_1\circ p):P\rightarrow D^{k-1}\), and the isomorphism \(\pi _1\mathrm {BDiff}_\partial (D^{d})\cong \Theta _{d+1}\) is given by assigning to a disc bundle \(D^d\rightarrow P\rightarrow D^1\) the manifold \(P\cup _{\partial D^d\times D^1\cup D^d\times \partial D^1} D^d\times D^1\). Deccomposing the sphere into half-discs \(S^n=D^n_+\cup D^n_-\), we see from this description that the composition \(\pi _k\mathrm {BDiff}_\partial (D^d)\rightarrow \Theta _{d+k}\) of the iterated Gromoll map with the isomorphism \(\pi _1\mathrm {BDiff}_\partial (D^{d+k-1})\cong \Theta _{d+k}\) maps a class represented by an \(S^{d}\)-bundle \(S^d\rightarrow Q\rightarrow S^k\) with a trivialisation \(\varphi :D_+^k\times S^d\cong Q|_{D_+^k}\) and a trivialised subbundle \(\psi :S^k\times D_+^d\hookrightarrow Q\) that agree on \(D^k_+\times D^d_+\) to the homotopy sphere
$$\begin{aligned}&\Big (Q\backslash \mathrm {int}\big (\varphi (D_+^k\times S^d)\cup \psi (S^k\times D_+^d)\big )\Big )\cup (D^k\times D^d)\\&\quad \cong \Big (Q\backslash \mathrm {int}\big (\psi (S^k\times D^d_+)\big )\Big )\cup _{S^k\times \partial D^d}(D^{k+1}\times \partial D^d) \end{aligned}$$
obtained by doing surgery along the trivialised \(D^d\)-subbundle, where \(D^k\times D^d\) is glued to \(Q\backslash \mathrm {int}\big (\varphi (D_+^k\times S^d)\cup _{\partial (D^k\times S^d)} \psi (S^k\times D_+^d)\big )\) along the embedding
$$\begin{aligned}&\partial (D^k\times D^d)=\partial D^k\times D^d\cup D^k\times \partial D^d\\&\quad \xrightarrow {\varphi |_{\partial D^k_+\times D^d_-}\cup \psi |_{D^k_-\times \partial D^d_+}}\partial \Big (Q\backslash \mathrm {int}\big (\varphi (D_+^k\times S^d)\cup \psi (S^k\times D_+^d)\Big ). \end{aligned}$$
This in particular implies the claim we made above in the case \(k=2\) and \(d=2n-1\).
As a consequence of this description of the morphism in consideration, the image \(\Sigma _\pi \in \Theta _{2n+1}\) of the class \([\pi ]\) comes equipped with a nullbordism, namely \(N\,{:}{=}\,E\cup _{E_+}\bar{E}\cup _{E_-}W,\) where W is the trace of the performed surgery. Omitting the trivialised \(D^{2n-1}\)-subbundles, the situation can be summarised schematically as follows
A choice of a stable framing of K induces stable framings on S and \(S^2\) and thus a stable isomorphism \(TE\cong T_\pi E\oplus \pi ^*TS\cong _sT_\pi E\) using which the canonical \(\tau _{>n}\mathrm {BO}\)-structure on \(T_\pi E\) and the \(\tau _{>n+1}\mathrm {BO}\)-one on \(T_\pi E|_{E_+}\) (see Example 3.1) induce a \(\tau _{>n}\mathrm {BO}\)-structure on TE and a \(\tau _{>n+1}\mathrm {BO}\)-structure on \(TE|_{E_+}\cong _sTE_+\). With these choices, we have \(\chi (T_\pi E)=\chi (TE,TE_+)\). By construction, the restriction of this \(\tau _{>n+1}\mathrm {BO}\)-structure to \(TE_+|_{S\times D^{2n-1}}\cong _s TS\) agrees with the \(\tau _{>n+1}\mathrm {BO}\)-structure on TS obtained from the stable framing of K, so we obtain a \(\tau _{>n+1}\mathrm {BO}\)-structure on \(T\bar{E}|_{E_+\cup K\times D^{2n-1}}\), which by obstruction theory extends to one on \(T(\bar{E}\cup _{E_-}W)\): the relative Serre spectral sequence shows that \(H^*(\bar{E},E_+\cup K\times D^{2n-1})\) vanishes for \(*\le 2n-2\) and thus that \(H^{i+1}(\bar{E}\cup _{E_-}W,E_+\cup K\times D^{2n-1};\pi _i(\tau _{\le n}\mathrm {SO}))\cong H^{i+1}(\bar{E}\cup _{E_-}W,\bar{E};\pi _i(\tau _{\le n}\mathrm {SO}))\cong H^{i+1}(W,E_-;\pi _i(\tau _{\le n}SO))\cong H^{i+1}(D^3,S^2;\pi _i(\tau _{\le n}SO))=0\) for \(i\le 2n-3\), using \(W\simeq E_{-}\cup _{S^2}D^3\) and \(\pi _2SO=0\). The restriction of this \(\tau _{>n+1}\mathrm {BO}\)-structure on \(T\bar{E}|_{E_+\cup K\times D^{2n-1}}\) to a \(\tau _{>n}\mathrm {BO}\) and the canonical \(\tau _{>n}\mathrm {BO}\)-structure on \(TE\cong T_\pi E\) (see Example 3.1) assemble to a \(\tau _{>n}\mathrm {BO}\)-structure on N. By construction, the canonical restriction map (using excision)
$$\begin{aligned} H^*(E,E_+;\mathbf {Z})\cong H^*(N,\bar{E}\cup _{E_-}W; \mathbf {Z})\longrightarrow H^*(N,\Sigma _\pi ; \mathbf {Z}) \end{aligned}$$
sends \(\chi (T_\pi E)=\chi (TE,TE_+)\) to \(\chi (TN,T\Sigma _\pi )\), so we conclude \(\chi ^2(T_\pi E)=\chi ^2(TN,T\Sigma _\pi )\). To finish the proof, note that the \(\tau _{>n}\mathrm {BO}\)-structure on TN allows us to do surgery away from the boundary on N to obtain an n-connected manifold \(E'\), which gives a class in \(A_{2n+2}^{\tau _{>n}}\) as aimed for: \(\partial E'=\Sigma _\pi \) holds by construction, \(\chi ^2(E')=\chi ^2(TN,T\Sigma _\pi )=\chi ^2(T_\pi E)\) by the bordism invariance of Pontryagin numbers (see Example 3.1), and \(\mathrm {sgn}(E')=\mathrm {sgn}(N)=\mathrm {sgn}(E)\) by the additivity and bordism invariance of the signature.\(\square \)
Combining the previous result with Proposition 3.9, we conclude that
sends a homology class \([\pi ]\) represented by a bundle \(\pi :E\rightarrow S\) to a certain linear combination of \(\Sigma _P\) and \(\Sigma _Q\) whose coefficients involve the invariants \(\mathrm {sgn}(E)\) and \(\chi ^2(T_\pi E)\). In the following two subsections, we shall see that these functionals
$$\begin{aligned} \mathrm {sgn}:\mathrm {H}_2(\mathrm {BDiff}_{\partial /2}(W_{g,1});\mathbf {Z})\longrightarrow \mathbf {Z}\quad \text {and}\quad \chi ^2:\mathrm {H}_2(\mathrm {BDiff}_{\partial /2}(W_{g,1});\mathbf {Z})\longrightarrow \mathbf {Z}\end{aligned}$$
factor through the composition
and have a more algebraic description in terms of \(H(g)\otimes \pi _n\mathrm {SO})\rtimes G_g\). This uses
$$\begin{aligned} (s_F,p):\Gamma ^n_{g,1/2}\longrightarrow (H(g)\otimes \pi _n\mathrm {SO})\rtimes G_g, \end{aligned}$$
induced by acting on a stable framing of \(W_{g,1}\) (agreeing with the usual framing on \(D^{2n-1}\)) as explained in Sect. 2.
Signatures of bundles of symplectic lattices
The standard action of the symplectic group \(\mathrm {Sp}_{2g}(\mathbf {Z})\) on \(\mathbf {Z}^{2g}\) gives rise to a local system \(\mathcal {H}(g)\) over \(\mathrm {BSp}_{2g}(\mathbf {Z})\) and the usual symplectic form on \(\mathbf {Z}^{2g}\) gives a morphism \(\lambda :\mathcal {H}(g)\otimes \mathcal {H}(g)\rightarrow \mathbf {Z}\) of local systems to the constant system. To an oriented closed surface S with a map \(f:S\rightarrow \mathrm {BSp}_{2g}(\mathbf {Z})\), we can associate a bilinear form
$$\begin{aligned} \langle -,-\rangle _f:\mathrm {H}^1(S; f^*\mathcal {H}(g))\otimes \mathrm {H}^1(S; f^*\mathcal {H}(g))\rightarrow \mathbf {Z}, \end{aligned}$$
defined as the composition
$$\begin{aligned} \mathrm {H}^1(S; f^*\mathcal {H}(g))\otimes \mathrm {H}^1(S; f^*\mathcal {H}(g))\xrightarrow {\smile } \mathrm {H}^2(S; f^*\mathcal {H}(g)\otimes \mathcal {H}(g))\overset{\lambda }{\longrightarrow }\mathrm {H}^2(S;\mathbf {Z})\xrightarrow {[S]}\mathbf {Z}. \end{aligned}$$
As both the cup-product and the symplectic pairing \(\lambda \) are antisymmetric, the form \(\langle -,-\rangle _f\) is symmetric. The usual argument for the bordism invariance of the signature shows that its signature \(\mathrm {sgn}(\langle -,-\rangle _f)\) depends only on the bordism class \([f]\in \Omega ^{\mathrm {SO}}_{2}(\mathrm {BSp}_{2g}(\mathbf {Z}))\cong \mathrm {H}_2(\mathrm {Sp}_{2g}(\mathbf {Z});\mathbf {Z})\) and thus induces a morphism
$$\begin{aligned} \mathrm {sgn}:\mathrm {H}_2(\mathrm {Sp}_{2g}(\mathbf {Z});\mathbf {Z})\longrightarrow \mathbf {Z}, \end{aligned}$$
(3.8)
which is compatible with the usual inclusion \(\mathrm {Sp}_{2g}(\mathbf {Z})\subset \mathrm {Sp}_{2g+2}(\mathbf {Z})\).
Remark 3.13
As \(\mathrm {Sp}_{2g}(\mathbf {Z})\) is perfect for \(g\ge 3\) (see Lemma A.1), the morphism (3.8) determines a unique cohomology class \(\mathrm {sgn}\in \mathrm {H}^2(\mathrm {Sp}_{2g}(\mathbf {Z});\mathbf {Z})\). There is a well-known purely algebraically defined cocycle representative of this class due to Meyer [43], known as the Meyer cocycle.
The morphism (3.8) measures signatures of total spaces of smooth bundles over surfaces (even of fibrations of Poincaré complexes). More precisely, for a compact oriented \((4k+2)\)-manifold M, the action of its group of diffeomorphisms on the middle cohomology induces a morphism
for \(2g=\mathrm {rk}(\mathrm {H}^{2k+1}(M))\) and the resulting composition
can be shown to map a homology class represented by a smooth bundle over a surface to the signature of its total space. This fact can either be proved along the lines of [14] or extracted from [42] and it has in particular the following consequence.
Lemma 3.14
For n odd, the composition
$$\begin{aligned} \mathrm {H}_2(\mathrm {BDiff}_{\partial /2}(W_{g,1});\mathbf {Z})\longrightarrow \mathrm {H}_2(\mathrm {Sp}_{2g}(\mathbf {Z});\mathbf {Z})\xrightarrow {\mathrm {sgn}}\mathbf {Z}\end{aligned}$$
sends the class of an \((W_{g,1},D^{2n-1})\)-bundle \(\pi :E\rightarrow S\) to the signature \(\mathrm {sgn}(E)\) of its total space.
We proceed by computing the image of the signature morphism (3.8) and of its pullback to the theta-subgroup \(\mathrm {Sp}_{2g}^q(\mathbf {Z})\subset \mathrm {Sp}_{2g}(\mathbf {Z})\) as defined in Sect. 1.2.
Lemma 3.15
The signature morphism satisfies
$$\begin{aligned} \mathrm {im}\left( \mathrm {H}_2(\mathrm {Sp}_{2g}(\mathbf {Z});\mathbf {Z})\xrightarrow {\mathrm {sgn}}\mathbf {Z}\right)= & {} {\left\{ \begin{array}{ll}0&{}\quad \text{ if } g=1\\ 4\cdot \mathbf {Z}&{}\quad \text{ if } g\ge 2\\ \end{array}\right. }\\ \mathrm {im}\left( \mathrm {H}_2(\mathrm {Sp}^q_{2g}(\mathbf {Z});\mathbf {Z})\xrightarrow {\mathrm {sgn}}\mathbf {Z}\right)= & {} {\left\{ \begin{array}{ll}0&{}\quad \text{ if } g=1\\ 8\cdot \mathbf {Z}&{}\quad \text{ if } g\ge 2.\\ \end{array}\right. } \end{aligned}$$
Proof
The signatures realised by classes in \(\mathrm {H}_2(\mathrm {Sp}_{2g}(\mathbf {Z});\mathbf {Z})\) are well-known (see e.g. [3, Lem 6.5, Thm 6.6 (vi)]). To prove that the signature of classes in \(\mathrm {H}_2(\mathrm {Sp}_{2g}^q(\mathbf {Z}))\) is divisible by 8, recall from Sects. 1.2 and 1.3 that for n odd the morphism \(\mathrm {Diff}_{\partial /2}(W_{g,1})\rightarrow \mathrm {Sp}_{2g}(\mathbf {Z})\) lands in the subgroup \(\mathrm {Sp}_{2g}^q(\mathbf {Z})\subset \mathrm {Sp}_{2g}(\mathbf {Z})\) as long as \(n\ne 1,3,7\), so we have a composition
$$\begin{aligned} \mathrm {H}_2(\mathrm {BDiff}_{\partial /2}(W_{g,1});\mathbf {Z})\longrightarrow \mathrm {H}_2(\Gamma ^n_{g,1/2};\mathbf {Z})\longrightarrow \mathrm {H}_2(\mathrm {Sp}_{2g}^q(\mathbf {Z});\mathbf {Z})\xrightarrow {\mathrm {sgn}}\mathbf {Z}, \end{aligned}$$
which maps the class of a bundle \(\pi :E\rightarrow S\) by Lemma 3.14 to \(\mathrm {sgn}(E)\). The latter agrees by Theorem 3.12 with the signature of an almost closed n-connected \((2n+2)\)-manifold, so it is divisible by 8 as the intersection form of such manifolds is unimodular and even (see e.g. [55]). This proves the claimed divisibility, since the first two morphisms in the composition are surjective, the first one because of the second reminder at the beginning of Sect. 3.3 and the second one by Corollary 2.4. As the signature morphism vanishes on \(\mathrm {H}_2(\mathrm {Sp}_{2}(\mathbf {Z});\mathbf {Z})\) by the first part, it certainly vanishes on \(\mathrm {H}_{2g}(\mathrm {Sp}^{q}_{2}(\mathbf {Z});\mathbf {Z})\). Consequently, by the compatibility of the signature with the inclusion \(\mathrm {Sp}_{2g}(\mathbf {Z})\subset \mathrm {Sp}_{2g+2}(\mathbf {Z})\), the remaining claim follows from constructing a class in \(\mathrm {H}_2(\mathrm {Sp}_{2g}^q(\mathbf {Z});\mathbf {Z})\) of signature 8 for \(g=2\). Using \(\mathrm {H}_2(\mathrm {Sp}_{4}(\mathbf {Z});\mathbf {Z})\cong \mathbf {Z}\oplus \mathbf {Z}/2\) (see e.g. [3, Lem. A.1(iii)]) and the first part of the claim, the existence of such a class is equivalent to the image of \(\mathrm {H}_2(\mathrm {Sp}^q_4(\mathbf {Z});\mathbf {Z})\) in the torsion free quotient \(\mathrm {H}_2(\mathrm {Sp}_{4}(\mathbf {Z});\mathbf {Z})_{\mathrm {free}}\cong \mathbf {Z}\) containing 2. That it contains 10 is ensured by transfer, since the index of \(\mathrm {Sp}_{4}^q(\mathbf {Z})\subset \mathrm {Sp}_{4}(\mathbf {Z})\) is 10 (see Sect. 1.2). As \(\mathrm {H}_1(\mathrm {Sp}_4(\mathbf {Z});\mathbf {Z})\) and \(\mathrm {H}_1(\mathrm {Sp}^q_4(\mathbf {Z});\mathbf {Z})\) are 2-torsion by Lemma A.1, it therefore suffices to show that \(\mathrm {H}_2(\mathrm {Sp}^q_4(\mathbf {Z});\mathbf {F}_5)\rightarrow \mathrm {H}_2(\mathrm {Sp}_{4}(\mathbf {Z});\mathbf {F}_5)\) is nontrivial, for which we consider the level 2 congruence subgroup \(\mathrm {Sp}_{4}(\mathbf {Z},2)\subset \mathrm {Sp}_4(\mathbf {Z})\), i.e. the kernel of the reduction map \(\mathrm {Sp}_4(\mathbf {Z})\rightarrow \mathrm {Sp}_4(\mathbf {Z}/2)\), which is surjective (see e.g. [45, Thm 1] for an elementary proof). From the explicit description of \(\mathrm {Sp}_{2g}^q(\mathbf {Z})\) presented in Sect. 1.2, one sees that it contains the congruence subgroup \(\mathrm {Sp}_{4}(\mathbf {Z},2)\). As a result, it is enough to prove that \(\mathrm {H}_2(\mathrm {Sp}_4(\mathbf {Z},2);\mathbf {F}_5)\rightarrow \mathrm {H}_2(\mathrm {Sp}_{4}(\mathbf {Z});\mathbf {F}_5)\) is nontrivial, which follows from an application of the Serre spectral sequence of the extension
$$\begin{aligned} 0\longrightarrow \mathrm {Sp}_4(\mathbf {Z},2)\longrightarrow \mathrm {Sp}_4(\mathbf {Z})\longrightarrow \mathrm {Sp}_4(\mathbf {Z}/2)\longrightarrow 0, \end{aligned}$$
using that \(\mathrm {H}_1(\mathrm {Sp}_{4}(\mathbf {Z},2);\mathbf {F}_5)\) vanishes by a result of Sato [49, Cor. 10.2] and that the groups \(\mathrm {H}_*(\mathrm {Sp}_4(\mathbf {Z}/2);\mathbf {F}_5)\) are trivial in low degrees, because of the exceptional isomorphism between \(\mathrm {Sp}_4(\mathbf {Z}/2)\) and the symmetric group in 6 letters as explained for instance in [46, p. 37]. \(\square \)
Remark 3.16
-
(i)
There are at least two other proofs for the divisibility of the signature of classes in \(\mathrm {H}_2(\mathrm {Sp}_{2g}^q(\mathbf {Z});\mathbf {Z})\) by 8. One can be extracted from the proof of [21, Lem. 7.5 i)] and another one is given in [4, Thm 12.1]. The proof in [21] shows actually something stronger, namely that the form \(\langle -,-\rangle _f\) associated to a class \([f]\in \mathrm {H}_2(\mathrm {Sp}^q_{2g}(\mathbf {Z});\mathbf {Z})\) is always even. We shall give a different proof of this fact as part of the second part of Lemma 3.19 below.
-
(ii)
For \(g\ge 4\), the existence of a class in \(\mathrm {H}_2(\mathrm {Sp}_{2g}^q(\mathbf {Z});\mathbf {Z})\) of signature 8 was shown as part of the proof of [21, Thm 7.7], using that the image of
$$\begin{aligned} \mathrm {H}_2(\mathrm {Sp}_{2g}(\mathbf {Z},2);\mathbf {Z})\longrightarrow \mathrm {H}_2(\mathrm {Sp}_{2g}(\mathbf {Z});\mathbf {Z})\cong \mathbf {Z}\end{aligned}$$
for \(g\ge 4\) is known to be \(2\cdot \mathbf {Z}\) by a result of Putman [47, Thm F]. However, this argument breaks for small values of g, in which case the image of \(\mathrm {sgn}:\mathrm {H}_2(\mathrm {Sp}_{2g}^q(\mathbf {Z});\mathbf {Z})\rightarrow \mathbf {Z}\) was not known before, at least to the knowledge of the author.
By Lemma 3.15, the signatures of classes in \(\mathrm {H}_2(\mathrm {Sp}_{2g}^q(\mathbf {Z});\mathbf {Z})\) are divisible by 8, so we obtain a morphism of the form
$$\begin{aligned} \mathrm {sgn}/8:\mathrm {H}_2(\mathrm {Sp}_{2g}^{q}(\mathbf {Z});\mathbf {Z})\longrightarrow \mathbf {Z}. \end{aligned}$$
To lift this morphism to a cohomology class in \(\mathrm {H}^2(\mathrm {Sp}_{2g}^q(\mathbf {Z});\mathbf {Z})\), we consider
$$\begin{aligned} a:\mathbf {Z}/4\longrightarrow \mathrm {Sp}_2^q(\mathbf {Z})\subset \mathrm {Sp}_{2g}^q(\mathbf {Z}), \end{aligned}$$
(3.9)
induced by the matrix \(\left( {\begin{matrix}0 &{} -1 \\ 1 &{} 0 \end{matrix}}\right) \in \mathrm {Sp}_2^q(\mathbf {Z})\). By Lemma A.1, this is an isomorphism on abelianisations for \(g\ge 3\) and thus induces a splitting of the universal coefficient theorem
$$\begin{aligned} a^*\oplus h:\mathrm {H}^2(\mathrm {Sp}_{2g}^q(\mathbf {Z});\mathbf {Z})\overset{\cong }{\longrightarrow } \mathrm {H}^2(\mathbf {Z}/4;\mathbf {Z})\oplus \mathrm {Hom}(\mathrm {H}_2(\mathrm {Sp}^q_{2g}(\mathbf {Z});\mathbf {Z}),\mathbf {Z}). \end{aligned}$$
(3.10)
This splitting is compatible with the inclusion \(\mathrm {Sp}_{2g}^q(\mathbf {Z})\subset \mathrm {Sp}^q_{2g+2}(\mathbf {Z})\), so we can define a lift of the divided signature \(\mathrm {sgn}/8\) to a class \(\mathrm {H}^2(\mathrm {Sp}_{2g}^q(\mathbf {Z});\mathbf {Z})\) as follows.
Definition 3.17
-
(i)
Define the class
$$\begin{aligned} \textstyle {\frac{\mathrm {sgn}}{8}\in \mathrm {H}^2(\mathrm {Sp}_{2g}^q(\mathbf {Z});\mathbf {Z})} \end{aligned}$$
for \(g\gg 0\) via the splitting (3.10) by declaring its image in the first summand to be trivial and to be \(\mathrm {sgn}/8\) in the second. For small g, the class \(\frac{\mathrm {sgn}}{8}\in \mathrm {H}^2(\mathrm {Sp}^q_{2g}(\mathbf {Z});\mathbf {Z})\) is defined as the pullback of \(\frac{\mathrm {sgn}}{8}\in \mathrm {H}^2(\mathrm {Sp}^q_{2g+2h}(\mathbf {Z});\mathbf {Z})\) for \(h\gg 0\).
-
(ii)
Define the class
$$\begin{aligned} \textstyle {\frac{\mathrm {sgn}}{8}\in \mathrm {H}^2(\Gamma ^n_{g,1/2};\mathbf {Z})}\quad \text {for } n\ne 1,3,7\text { odd} \end{aligned}$$
as the pullback of the same-named class along the map \(\Gamma ^n_{g,1/2}\rightarrow G_g\cong \mathrm {Sp}_{2g}^q(\mathbf {Z})\) induced by the action on the middle cohomology.
Framing obstructions
To describe the invariant
$$\begin{aligned} \chi ^2:\mathrm {H}_2(\mathrm {BDiff}_{\partial /2}(W_{g,1});\mathbf {Z})\longrightarrow \mathbf {Z}\end{aligned}$$
explained in Sect. 3.1 more algebraically, note that a map \(f:S\rightarrow \mathrm {B}(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}(\mathbf {Z}))\) from an oriented closed connected surface S to \(\mathrm {B}(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}(\mathbf {Z}))\) induces a 1-cocycle
$$\begin{aligned} \pi _1(S;*)\longrightarrow \mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}(\mathbf {Z})\longrightarrow \mathbf {Z}^{2g} \end{aligned}$$
and hence a class \([f]\in \mathrm {H}^1(S; f^*\mathcal {H}(g))\). The composition \(S\rightarrow \mathrm {B}(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}(\mathbf {Z}))\rightarrow \mathrm {BSp}_{2g}(\mathbf {Z})\) defines a bilinear form \(\langle -,-\rangle _f\) on \(\mathrm {H}^1(S; f^*\mathcal {H}(g))\) as explained in Sect. 3.4, and hence a number \(\langle [f],[f]\rangle _f\in \mathbf {Z}\). Varying f, this gives a morphism
$$\begin{aligned} \chi ^2:\mathrm {H}_2(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}(\mathbf {Z});\mathbf {Z})\longrightarrow \mathbf {Z}, \end{aligned}$$
which is compatible with natural inclusion \(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}(\mathbf {Z})\subset \mathbf {Z}^{2g+2}\rtimes \mathrm {Sp}_{2g+2}(\mathbf {Z})\) and takes for \(n\equiv 3\ (\mathrm {mod}\ 4)\) part in a composition
$$\begin{aligned} \mathrm {H}_2(\Gamma ^n_{g,1/2};\mathbf {Z})\xrightarrow {(s_F,p)}\mathrm {H}_2(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}(\mathbf {Z});\mathbf {Z})\overset{\chi ^2}{\longrightarrow }\mathbf {Z}, \end{aligned}$$
where the first morphism is induced by acting on a stable framing F of \(W_{g,1}\) as in Sect. 2. A priori, this requires three choices: a stable framing, a generator \(\pi _n\mathrm {SO}\cong \mathbf {Z}\), and a symplectic basis \(H(g)\cong \mathbf {Z}^{2g}\). However, the composition turns out to not be affected by these choices and the following proposition shows that it is related to the invariant of \((W_{g,1},D^{2n-1})\)-bundles explained in Example 3.1.
Proposition 3.18
For \(n\equiv 3\ (\mathrm {mod}\ 4)\), the composition
sends the class of a \((W_{g,1},D^{2n-1})\)-bundle \(\pi :E\rightarrow S\) to \(\chi ^2(T_\pi E)\).
Proof
The relative Serre spectral sequence of
$$\begin{aligned} (W_{g,1},\partial W_{g,1})\longrightarrow (E,\partial E)\overset{\pi }{\longrightarrow } S \end{aligned}$$
induces canonical isomorphisms
$$\begin{aligned} \mathrm {H}^{2n+2}(E,\partial E;\mathbf {Z})\cong \mathrm {H}^2(S;\mathbf {Z})\quad \text {and}\quad \mathrm {H}^{n+1}(E,\partial E;\mathbf {Z})\cong \mathrm {H}^1(S;f^*\mathcal {H}(g)), \end{aligned}$$
where f denotes the composition
$$\begin{aligned} S\longrightarrow \mathrm {BDiff}_{\partial /2}(W_{g,1})\longrightarrow \mathrm {B}\Gamma _{g,1/2}^n\overset{p}{\longrightarrow }\mathrm {BSp}_{2g}(\mathbf {Z}). \end{aligned}$$
By the compatibility of the Serre spectral sequence with the cup-product and after identifying \(\mathrm {H}^1(S;f^*\mathcal {H}(g))\) with \(\mathrm {H}^1(\pi _1(S;*);H(g)\otimes S\pi _n\mathrm {SO}(n))\), it suffices to show that the second isomorphism sends \(\chi (T_\pi E)\in \mathrm {H}^{n+1}(E,\partial E;\mathbf {Z})\) up to signs to the class represented by the cocycle
$$\begin{aligned} \pi _1(S;*) \longrightarrow \Gamma ^n_{g,1/2}\overset{s_F}{\longrightarrow }[W_{g,1},\mathrm {SO}]_*\cong H(g)\otimes \pi _n\mathrm {SO}, \end{aligned}$$
(3.11)
involving the choice of stable framing \(F:TW_{g,1}\oplus \varepsilon ^k\cong \varepsilon ^{2n+k}\) as in Sect. 2. As a first step, we describe this isomorphism more explicitly:
Note that \(\mathrm {H}^{n+1}(E,\partial E;\pi _n\mathrm {SO})\cong \mathrm {H}^{n+1}(E;\pi _n\mathrm {SO})\) as \(\mathrm {H}^{*}(\partial E;\pi _n\mathrm {SO})\) is trivial for \(*=n,n+1\). Unwinding the construction of the Serre spectral sequence using a skeletal filtration of S, one sees that after fixing an identification \(W_{g,1}\cong \pi ^{-1}(*)\subset E\), the image of a class \(x\in \mathrm {H}^{n+1}(E,\pi _n\mathrm {SO})\) under the isomorphism in question is represented by the cocycle \(\pi _1(S;*)\rightarrow H(g)\otimes \pi _n\mathrm {SO}\) which maps a loop \(\omega :([0,1],\{0,1\})\rightarrow (S,*)\) to the class obtained from a choice of lift \(\widetilde{x}\in \mathrm {H}^{n+1}(E,\pi ^{-1}(*);\pi _n\mathrm {SO})\) by pulling it back along
$$\begin{aligned} (W_{g,1}\times [0,1],W_{g,1}\times \{0,1\})\longrightarrow (\omega ^*E,W_{g,1}\times \{0,1\})\longrightarrow (E,\pi ^{-1}(*)), \end{aligned}$$
where the second morphism is induced by pulling back the bundle \(\pi :E\rightarrow S\) along \(\omega \) and the first morphism is the unique (up to homotopy) trivialisation \(\omega ^*E\cong [0,1]\times W_{g,1}\) relative to \(W_{g,1}\times \{0\}\) of the pullback bundle over [0, 1]; here we used the canonical isomorphism \(\mathrm {H}^{n+1}(W_{g,1}\times [0,1],W_{g,1}\times \{0,1\};\pi _n\mathrm {SO})\cong H(g)\otimes \pi _n\mathrm {SO}\).
Recall from Example 3.1 that the class \(\chi (T_\pi E)\in \mathrm {H}^{n+1}(E;\pi _n\mathrm {SO})\) is the primary obstruction to extending the canonical \(\tau _{>n}\mathrm {BO}\)-structure on \(T_\pi E\) to a \(\tau _{>n+1}\mathrm {BO}\)-structure. The choice of framing F induces such an extension on the fibre \(W_{g,1}\cong \pi ^{-1}(*)\subset E\) and thus induces a lift of \(\chi (T_\pi E)\) to a relative class in \(\mathrm {H}^{n+1}(E,\pi ^{-1}(*);\pi _n\mathrm {SO})\). This uses that the \(\tau _{>n}\mathrm {BO}\) structure on \(W_{g,1}\) induced by the framing agrees with the restriction of the \(\tau _{>n}\mathrm {BO}\)-structure on \(T_\pi E\) to \(\pi _{-1}(*)\cong W_{g,1}\) by obstruction theory. Using the above description, we see that the image of \(\chi (T_\pi E)\) under the isomorphism in question is represented by the cocyle \(\pi _1(S;*)\rightarrow H(g)\otimes \pi _n\mathrm {SO}\) that sends a loop \(\omega \) to the primary obstruction in \(\mathrm {H}^{n}(W_{g,1};\pi _n\mathrm {SO})\cong H(g)\otimes \pi _n\mathrm {SO}\) to solving the lifting problem
which agrees with the corresponding obstruction when replacing \(\mathrm {BSO}\) by \(\tau _{>n}\mathrm {BSO}\). Here the left square in the diagram is given by the trivialisation explained earlier. There is a useful alternative description of this obstruction class: relative to the subspace \(W_{g,1}\times \{0\}\subset W_{g,1}\times \{0,1\}\) there is a unique (up to homotopy) lift in (3.12), so the obstruction to finding a lift relative to the subspace \(W_{g,1}\times \{0,1\}\) can be seen as an element in the group of path-components of the fibre of the principal fibration \(\mathrm {Map}_*(W_{g,1},\tau _{>n+1}\mathrm {BSO})\rightarrow \mathrm {Map}_*(W_{g,1},\mathrm {BSO})\), which is exactly \([W_{g,1},\tau _{\le n}\mathrm {SO}]_*\cong H(g)\otimes \pi _n\mathrm {SO}\).
To see that the cocyle we just described agrees with (3.11), note that the function \(s_F:\Gamma ^n_{g,1/2}\rightarrow H(g)\otimes \pi _n\mathrm {SO}\cong [W_{g,1},\tau _{\le n}\mathrm {SO}]_*\) induced by acting on the stable framing F arises as the connecting map \(\pi _1(\mathrm {BDiff}^{\tau _{>n+1}}_{\partial /2}(W_{g,1});[F])\rightarrow \pi _0\mathrm {Map}_*(W_{g,1},\tau _{\le n}\mathrm {SO})\) of the fibration
$$\begin{aligned} \mathrm {Map}_*(W_{g,1},\tau _{\le n}\mathrm {SO})\longrightarrow \mathrm {BDiff}^{\tau _{>n+1}}_{\partial /2}(W_{g,1})\longrightarrow \mathrm {BDiff}_{\partial /2}(W_{g,1}), \end{aligned}$$
where \(\mathrm {BDiff}^{\tau _{>n+1}}_{\partial /2}(W_{g,1})\) is the space that classifies \((W_{g,1},D^{2n-1})\)-bundles with a \(\tau _{>n+1}\mathrm {BO}\)-structure on the vertical tangent bundle extending the given \(\tau _{>n+1}\mathrm {BO}\)-structure on the restriction to the trivial \(D^{2n-1}\)-subbundle induced by the standard framing of \(D^{2n-1}\); here we identified the space of \(\tau _{>n+1}\mathrm {BO}\)-structures of \(W_{g,1}\) relative to \(D^{2n-1}\) with the mapping space \(\mathrm {Map}_*(W_{g,1},\tau _{\le n}\mathrm {SO})\) by using the choice of stable framing F, which also induces the basepoint \([F]\in \mathrm {BDiff}^{\tau _{>n+1}}_{\partial /2}(W_{g,1})\). This shows that the value of (3.11) on a loop \([\omega ]\in \pi _1(S;*)\) is given by the component in \([W_{g,1},\tau _{\le n}\mathrm {SO}]_*\cong H(g)\otimes \pi _n\mathrm {SO}\) obtained by evaluating a choice of path-lift
at the end point. Such a path-lift precisely classifies a lift as in (3.12) relative to the subspace \(W_{g,1}\times \{0\}\subset W_{g,1}\times \{0,1\}\), so the claim follows from the second description of the obstruction to solving the lifting problem (3.12) mentioned above. \(\square \)
Lemma 3.19
Let \(g\ge 1\).
-
(i)
The image of the composition induced by inclusion and \(\chi ^2\)
$$\begin{aligned} \mathrm {H}_2(\mathbf {Z}^{2g};\mathbf {Z})\longrightarrow \mathrm {H}_2(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}^q_{2g}(\mathbf {Z});\mathbf {Z})\longrightarrow \mathrm {H}_2(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}(\mathbf {Z});\mathbf {Z})\overset{\chi ^2}{\longrightarrow }\mathbf {Z}\end{aligned}$$
contains \(2\cdot \mathbf {Z}\) and agrees with \(2\cdot \mathbf {Z}\) for \(g=1\).
-
(ii)
We have
$$\begin{aligned} \mathrm {im}\left( \mathrm {H}_2(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}^{q}_{2g}(\mathbf {Z}) ;\mathbf {Z})\xrightarrow {(\mathrm {sgn},\chi ^2)}\mathbf {Z}\oplus \mathbf {Z}\right)= & {} {\left\{ \begin{array}{ll}\langle (0,2)\rangle &{}\quad \text{ if } g=1\\ \langle (8,0),(0,2)\rangle &{}\quad \text{ if } g\ge 2. \end{array}\right. }\\ \mathrm {im}\left( \mathrm {H}_2(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}(\mathbf {Z}) ;\mathbf {Z})\xrightarrow {(\mathrm {sgn},\chi ^2)}\mathbf {Z}\oplus \mathbf {Z}\right)= & {} {\left\{ \begin{array}{ll}\langle (0,2)\rangle &{}\quad \text{ if } g=1\\ \langle (4,0),(0,1)\rangle &{}\quad \text{ if } g\ge 2. \end{array}\right. } \end{aligned}$$
-
(iii)
For \(n=3,7\), we have
$$\begin{aligned} \mathrm {im}\left( \mathrm {H}_2(\Gamma _{g,1/2}^n ;\mathbf {Z})\overset{(s_F,p)_*}{\longrightarrow }\mathrm {H}_2(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}(\mathbf {Z}) ;\mathbf {Z})\xrightarrow {\chi ^2-\mathrm {sgn}}\mathbf {Z}\right) =8\cdot \mathbf {Z}. \end{aligned}$$
Proof
By the compatibility of \(\chi ^2\) with the stabilisation maps, it suffices to show the first part for \(g=1\), which follows from checking that the image of a generator in \(\mathrm {H}_2(\mathbf {Z}^{2} ;\mathbf {Z})\cong \mathbf {Z}\) is mapped to \(\pm 2\) under \(\chi ^2\) by chasing through the definition. As the signature morphism pulls back from \(\mathrm {Sp}_{2g}(\mathbf {Z})\), the second part follows from the first part and Lemma 3.15 by showing that the image of \(\chi ^2\) is for \(\mathrm {Sp}_{2g}^q(\mathbf {Z})\) always divisible by \(2\cdot \mathbf {Z}\) and for \(\mathrm {Sp}_{2g}(\mathbf {Z})\) divisible by 2 if and only if \(g=1\). In the case of \(\mathrm {Sp}_{2g}^q(\mathbf {Z})\), this can be shown “geometrically” as in the proof of Lemma 3.15: choose \(n\equiv 3\ (\mathrm {mod}\ 4), n\ne 3,7\) and consider the composition
The first morphism is surjective by the second reminder at the beginning of Sect. 3.3 and the second morphism is an isomorphism as a result of Sect. 2, so it suffices to show that the composition is divisible by 2, which in turn follows from a combination of Lemma 3.18, Theorem 3.12 and the fact that in these dimensions, an n-connected almost closed \((2n+2)\)-manifold \(E'\) satisfies \(\chi ^2(E')\in 2\cdot \mathbf {Z}\) by a combination of Theorem 3.2 and (3.3). For \(\mathrm {Sp}_{2g}(\mathbf {Z})\), we argue as follows: in the case \(g=1\), we show that the morphism \(\mathrm {H}_2(\mathbf {Z}^2;\mathbf {Z})\rightarrow \mathrm {H}_2(\mathbf {Z}^{2}\rtimes \mathrm {Sp}_{2}(\mathbf {Z});\mathbf {Z})\) is surjective, which will exhibit the claimed divisibility as a consequence of (i). It follows from Lemma A.3 that the group \(\mathrm {H}_1(\mathrm {Sp}_{2}(\mathbf {Z});\mathbf {Z}^2)\) vanishes and as \(\mathrm {Sp}_2(\mathbf {Z})=\mathrm {SL}_2(\mathbf {Z})\), we also have \(\mathrm {H}_2(\mathrm {Sp}_2(\mathbf {Z});\mathbf {Z})=0\),Footnote 3 so an application of the Serre spectral sequence to \(\mathbf {Z}^{2}\rtimes \mathrm {Sp}_{2}(\mathbf {Z})\) shows the claimed surjectivity. This leaves us with proving that \(\chi ^2\) is not divisible by 2 for \(g\ge 2\) for which we use that there is class \([f:\pi _1S\rightarrow \mathrm {Sp}_{2g}(\mathbf {Z})]\in \mathrm {H}_2(\mathrm {Sp}_{2g}(\mathbf {Z});\mathbf {Z})\) of signature 4 by Lemma 3.15, so the form \(\langle -,-\rangle _f\) cannot be even and hence there is a 1-cocycle \(g:\pi _1S\rightarrow \mathbf {Z}^{2g}\) for which \(\langle [g],[g]\rangle _f\) is odd, which means that the image \(\chi ^2([g,f])\) of the class \([(g,f)]\in \mathrm {H}_2(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}(\mathbf {Z});\mathbf {Z})\) induced by the morphism \((g,f):\pi _1S\rightarrow \mathbf {Z}^2\rtimes \mathrm {Sp}_{2g}(\mathbf {Z})\) is odd. For the last part, note that the argument we gave for the divisibility in the second part for \(\mathrm {Sp}_{2g}^q(\mathbf {Z})\) shows for \(n=3,7\) that the image of the composition in (iii) is contained in \(8\cdot \mathbf {Z}\), since \(\chi ^2(E')-\mathrm {sgn}(E')\) is divisible by 8 if \(n=3,7\). Hence, to finish the proof, it suffices to establish the existence of a class in \(\mathrm {H}_2(\Gamma ^n_{g,1/2};\mathbf {Z})\) for \(n=3,7\) on which the composition evaluates to 8. To this end, we consider the square
induced by the embedding \((s_F,p)\) of the extension describing \(\Gamma _{g,1/2}^n\) into the trivial extension of \(\mathrm {Sp}_{2g}(\mathbf {Z})\) by \(\mathbf {Z}^{2g}\otimes \pi _n\mathrm {SO}\) (see Sect. 2). By the first part, there is a class \([f]\in \mathrm {H}_2(\mathbf {Z}^{2g}\otimes \pi _n\mathrm {SO};\mathbf {Z})\) with \(\chi ^2([f])=2\) and trivial signature, since the signature morphism pulls back from \(\mathrm {Sp}_{2g}(\mathbf {Z})\). As a result of Lemma 2.1, the cokernel of the left vertical map in the square is 4-torsion if \(n=3,7\), so \(4\cdot [f]\) lifts to \(\mathrm {H}_2(\mathbf {Z}^{2g}\otimes S\pi _n\mathrm {SO}(n);\mathbf {Z})\) and provides a class as desired. \(\square \)
Similar to the construction of \(\frac{\mathrm {sgn}}{8}\in \mathrm {H}^2(\mathrm {Sp}_{2g}^q(\mathbf {Z}) ;\mathbf {Z})\), we would like to lift the morphism resulting from the second part of Lemma 3.19
$$\begin{aligned} \chi ^2/2:\mathrm {H}_2(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}^q(\mathbf {Z}) ;\mathbf {Z})\rightarrow \mathbf {Z}\end{aligned}$$
to a class \(\frac{\chi ^2}{2}\in \mathrm {H}^2(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}^q(\mathbf {Z}) ;\mathbf {Z})\). To this end, observe that \(\mathrm {Sp}_{2g}^q(\mathbf {Z})\subset \mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}^q(\mathbf {Z})\) induces an isomorphism on abelianisations for \(g\ge 2\) since the coinvariants \((\mathbf {Z}^{2g})_{\mathrm {Sp}_{2g}^q(\mathbf {Z})}\) vanish in this range by Lemma A.2. The morphism
$$\begin{aligned} a:\mathbf {Z}/4\longrightarrow \mathrm {Sp}_2^q(\mathbf {Z})\subset \mathbf {Z}^{2g}\rtimes \mathrm {Sp}^q_{2g}(\mathbf {Z}) \end{aligned}$$
considered in Sect. 3.4 thus induces a splitting
$$\begin{aligned} a^*\oplus h:\mathrm {H}^2(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}^q(\mathbf {Z}) )\xrightarrow {\cong } \mathrm {H}^2(\mathbf {Z}/4 )\oplus \mathrm {Hom}(\mathrm {H}_2(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}^q(\mathbf {Z}) ),\mathbf {Z}), \end{aligned}$$
(3.13)
for \(g\ge 3\), analogous to the splitting (3.10). As before, this splitting is compatible with the inclusions \(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}^q(\mathbf {Z})\subset \mathbf {Z}^{2g+2}\rtimes \mathrm {Sp}_{2g+2}^q(\mathbf {Z})\), so the following is valid.
Definition 3.20
-
(i)
Define the class
$$\begin{aligned} \textstyle {\frac{\chi ^2}{2}\in \mathrm {H}^2(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}^q(\mathbf {Z});\mathbf {Z})} \end{aligned}$$
for \(g\gg 0\) via the splitting (3.13) by declaring its image in the first summand to be trivial and to be \(\chi ^2/2\) in the second. For small g, the class \(\frac{\chi ^2}{2}\) is defined as the pullback of the class for \(g\gg 0\).
-
(ii)
Define the class
$$\begin{aligned} \textstyle {\frac{\chi ^2}{2}\in \mathrm {H}^2(\Gamma ^n_{g,1};\mathbf {Z})\quad \text {for }n\ne 1,3,7\text { and }n\equiv 3\ (\mathrm {mod}\ 4)}\end{aligned}$$
as the pullback of the same-named class along the map
$$\begin{aligned} \Gamma ^n_{g,1/2}\overset{(s_F,p)}{\longrightarrow } \big (H(g)\otimes S\pi _n\mathrm {SO}(n)\big )\rtimes G_g\cong \mathbf {Z}^{2g}\rtimes \mathrm {Sp}_{2g}^q(\mathbf {Z}). \end{aligned}$$
-
(iii)
Define the class
$$\begin{aligned} \textstyle {\frac{\chi ^2-\mathrm {sgn}}{8}\in \mathrm {H}^2(\Gamma ^n_{g,1/2};\mathbf {Z})\cong \mathrm {Hom}(\mathrm {H}_2(\Gamma ^n_{g,1/2} ;\mathbf {Z}),\mathbf {Z})\quad \text {for }n=3,7} \end{aligned}$$
for \(g\gg 0\) as image of \((\chi ^2-\mathrm {sgn})/8\) ensured by Lemma 3.19, and for small g as the pullback of \(\frac{\chi ^2-\mathrm {sgn}}{8}\in \mathrm {H}^2(\Gamma _{g+h,1/2};\mathbf {Z})\) for \(h\gg 0\).
The isomorphism \(\mathrm {H}^2(\Gamma ^n_{g,1/2};\mathbf {Z})\cong \mathrm {Hom}(\mathrm {H}_2(\Gamma ^n_{g,1/2};\mathbf {Z}),\mathbf {Z})\) for \(n=3,7\) in the previous definition is assured by Corollary 2.4 and Lemma A.1, as the abelianisation \(\mathrm {H}_1(\Gamma ^n_{g,1/2};\mathbf {Z})\) vanishes for \(g\gg 0\).
We finish this subsection with an auxiliary lemma convenient for later purposes.
Lemma 3.21
For \(g=1\), the class \(\frac{\mathrm {sgn}}{8}\in \mathrm {H}^2(\mathrm {Sp}^q_{2g}(\mathbf {Z});\mathbf {Z})\) and the pullback of the class \(\frac{\chi ^2}{2}\in \mathrm {H}^2(\mathbf {Z}^{2g}\rtimes \mathrm {Sp}^q_{2g}(\mathbf {Z});\mathbf {Z})\) to \(\mathrm {H}^2(\mathrm {Sp}^q_{2g}(\mathbf {Z});\mathbf {Z})\) are trivial.
Proof
Both classes evaluate trivially on \(\mathrm {H}_2(\mathrm {Sp}_{2g}^{q}(\mathbf {Z});\mathbf {Z})\). For the first class, this is a consequence of Lemma 3.15, for the second this holds by construction. Moreover, both classes pull back trivially to \(\mathrm {H}^2(\mathbf {Z}/4;\mathbf {Z})\) along the morphism \(a:\mathbf {Z}/4\rightarrow \mathrm {Sp}^q_{2g}(\mathbf {Z})\) by definition. Although this morphism does not induce an isomorphism on abelianisations for \(g=1\), it still induces one on their torsion subgroups by Lemma A.1 and this is sufficient to deduce the assertion from the universal coefficient theorem. \(\square \)
The proof of Theorem B
We are ready to prove our main result Theorem B, which we state equivalently in terms of the central extension
$$\begin{aligned} 0\longrightarrow \Theta _{2n+1}\longrightarrow \Gamma ^n_{g}\longrightarrow \Gamma ^n_{g,1/2}\longrightarrow 0, \end{aligned}$$
explained in Sect. 1.3. Our description of its extension class in \(\mathrm {H}^2(\Gamma _{g,1/2}^n;\Theta _{2n+1})\) involves the cohomology classes in \(\mathrm {H}^2(\Gamma _{g,1/2}^n;\mathbf {Z})\) of Definitions 3.17 and 3.20, and the two homotopy spheres \(\Sigma _P\) and \(\Sigma _Q\) in the subgroup \(\mathrm {bA}_{2n+2}\subset \Theta _{2n+1}\) examined in Sect. 3.2.1. We write \((-)\cdot \Sigma \in \mathrm {H}^2(\Gamma _{g,1/2}^n;\mathbf {Z})\rightarrow \mathrm {H}^2(\Gamma _{g,1/2}^n;\Theta _{2n+1})\) for the change of coefficients induced by \(\Sigma \in \Theta _{2n+1}\).
Theorem 3.22
For \(n\ge 3\) odd, the extension
$$\begin{aligned} 0\longrightarrow \Theta _{2n+1}\longrightarrow \Gamma _{g,1}^n\longrightarrow \Gamma _{g,1/2}^n\longrightarrow 0 \end{aligned}$$
is classified by the class
$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{\mathrm {sgn}}{8}\cdot \Sigma _P&{}\quad \text{ if } n\equiv 1\ (\mathrm {mod}\ 4)\\ \frac{\mathrm {sgn}}{8}\cdot \Sigma _P+\frac{\chi ^2}{2}\cdot \Sigma _Q&{}\quad \text{ if } n\equiv 3\ (\mathrm {mod}\ 4),n\ne 3,7\\ \frac{\chi ^2-\mathrm {sgn}}{8}\cdot \Sigma _Q&{}\quad \text{ if } n=3,7. \end{array}\right. } \end{aligned}$$
and its induced differential \(d_2:\mathrm {H}_2(\Gamma _{g,1/2}^n)\rightarrow \Theta _{2n+1}\) has image
$$\begin{aligned} \mathrm {im}\left( \mathrm {H}_2(\Gamma ^n_{g,1/2})\overset{d_2}{\longrightarrow }\Theta _{2n+1}\right) = {\left\{ \begin{array}{ll} \mathrm {bA}_{2n+2}&{}\quad \text{ if } g\ge 2\\ \langle \Sigma _Q \rangle &{}\quad \text{ if } g=1. \end{array}\right. }. \end{aligned}$$
Moreover, the extension splits if and only if \(g=1\) and \(n\equiv 1\ (\mathrm {mod}\ 4)\).
Proof
As all cohomology classes involved are compatible with the stabilisation map \(\Gamma ^n_{g,1/2}\rightarrow \Gamma ^n_{g+1,1/2}\), it is sufficient to show the first part of the claim for \(g\gg 0\) (see Sect. 1.4). We assume \(n\ne 3,7\) first. Identifying \(\Gamma ^n_{g,1/2}\) with \((H(g)\otimes \pi _n\mathrm {SO})\rtimes G_g\) via the isomorphism \((s_F,p)\) of Sect. 2, the morphism \(a:\mathbf {Z}/4\rightarrow (H(g)\otimes \pi _n\mathrm {SO})\rtimes G_g\) of the previous section induces a morphism between the sequences of the universal coefficient theorem
By the exactness of the rows and the vanishing of \(\mathrm {H}_2(\mathbf {Z}/4;\mathbf {Z})\), it is sufficient to show that the extension class in consideration agrees with the classes in the statement when mapped to \(\mathrm {Hom}(\mathrm {H}_2(\Gamma ^n_{g,1/2};\mathbf {Z}),\Theta _{2n+1})\) and \(\mathrm {H}^2(\mathbf {Z}/4,\Theta _{2n+1})\). Regarding the images in the Hom-term, it is enough to identify them after precomposition with the epimorphism
$$\begin{aligned} \mathrm {H}_2(\mathrm {BDiff}_{\partial /2}(W_{g,1});\mathbf {Z})\longrightarrow \mathrm {H}_2(\Gamma ^n_{g,1/2};\mathbf {Z}), \end{aligned}$$
so from the construction of \(\frac{\mathrm {sgn}}{8}\) and \(\frac{\chi ^2}{2}\) together with Lemmas 3.15 and 3.19, we see that it suffices to show that \(\mathrm {H}_2(\mathrm {BDiff}_{\partial /2}(W_{g,1});\mathbf {Z})\rightarrow \Theta _{2n+1}\) induced by the extension class maps the class of a bundle \(\pi :E\rightarrow S\) to \(\mathrm {sgn}(E)/8\cdot \Sigma _P\) if \(n\equiv 1\ (\mathrm {mod}\ 4)\) and to \(\mathrm {sgn}(E)/8\cdot \Sigma _P+\chi ^2(E)/2\cdot \Sigma _Q\) otherwise, which is a consequence of Theorem 3.12 combined with Proposition 3.9. By construction, the classes \(\frac{\mathrm {sgn}}{8}\cdot \Sigma _P\) and \(\frac{\chi ^2}{2}\cdot \Sigma _Q\) vanish in \(\mathrm {H}^2(\mathbf {Z}/4;\Theta _{2n+1})\), so the claim for \(n\ne 3,7\) follows from showing that the extension class is trivial in \(\mathrm {H}^2(\mathbf {Z}/4;\Theta _{2n+1})\), i.e. that the pullback of the extension to \(\mathbf {Z}/4\) splits, which is in turn equivalent to the existence of a lift
Using the standard embedding \(W_1=S^n\times S^n\subset \mathbf {R}^{n+1}\times \mathbf {R}^{n+1}\), we consider the diffeomorphism
$$\begin{aligned} \begin{array}{rcl} S^n\times S^n &{} \longrightarrow &{} S^n\times S^n \\ (x_1,\ldots ,x_{n+1},y_1,\ldots ,y_{n+1}) &{} \longmapsto &{} (-y_1,\ldots ,y_{n+1},x_1,\ldots ,x_{n+1}), \end{array} \end{aligned}$$
which is of order 4, maps to \(\left( {\begin{matrix} 0 &{} -1 \\ 1 &{} 0 \end{matrix}}\right) \in \mathrm {Sp}_2^q(\mathbf {Z})\), and has constant differential, so it vanishes in \(H(g)\otimes \pi _n\mathrm {SO}\). As the natural map \(\Gamma _{1,1}^n\rightarrow \Gamma ^n_1\) is an isomorphism by Lemma 1.1, this diffeomorphism induces a lift as required for \(g=1\), which in turn provides a lift for all \(g\ge 1\) via the stabilisation map \(\Gamma ^n_{1,1}\rightarrow \Gamma ^n_{g,1}\). For \(n=3,7\), the abelianisation of \(\Gamma ^n_{g,1/2}\) vanishes for \(g\gg 0\) due to Corollary 2.4, so it suffices to identify the extension class with the classes in the statement in \(\mathrm {Hom}(\mathrm {H}_2(\Gamma ^n_{g,1/2};\mathbf {Z}),\Theta _{2n+1})\) which follows as in the case \(n\ne 3,7\).
Lemma 3.19 (iii) implies that the image of \(d_2\) for \(n=3,7\) is generated by \(\Sigma _Q\) for all \(g\ge 1\). For \(n\ne 3,7\), the map \(\Gamma _{g,1/2}^n\rightarrow (H(g)\otimes \pi _n\mathrm {SO})\rtimes G_g\) is an isomorphism (see Sect. 2), so Lemma 3.19 tells us that the image of \(d_2\) for \(n\equiv 3\ (\mathrm {mod}\ 4)\) is generated by \(\Sigma _P\) and \(\Sigma _Q\) if \(g\ge 2\) and by \(\Sigma _Q\) if \(g=1\). For \(n\equiv 1\ (\mathrm {mod}\ 4)\), it follows from Lemma 3.15 that the image of \(d_2\) is generated by \(\Sigma _P\) if \(g\ge 2\) and that it is trivial for \(g=1\). In sum, this implies the second part of the claim by Corollary 3.5, and also that the differential \(d_2\) does not vanish for \(g\ge 2\), so the extension is nontrivial in these cases. For \(n\equiv 3\ (\mathrm {mod}\ 4)\), the homotopy sphere \(\Sigma _Q\) is nontrivial by Theorem 3.3, so \(d_2\) does not vanish for \(g=1\) either. Finally, in the case \(n\equiv 1\ (\mathrm {mod}\ 4)\), the extension is classified by \(\frac{\mathrm {sgn}}{8}\cdot \Sigma _P\), which is trivial for \(g=1\) by Lemma 3.21, so the extension splits and the proof is finished. \(\square \)
It is time to make good for the missing part of the proof of Theorem 2.2.
Proof of Theorem 2.2for \(n=3,7\). We have \(G_g=\mathrm {Sp}_{2g}(\mathbf {Z})\), so the case \(g=1\) follows from the fact that \(\mathrm {H}^2(\mathrm {Sp}_{2g}(\mathbf {Z});\mathbf {Z}^{2g}\otimes S\pi _n\mathrm {SO}(n))\) vanishes by Lemma A.3. To prove the case \(g\ge 2\), note that a hypothetical splitting \(s:\mathrm {Sp}_{2g}(\mathbf {Z})\rightarrow \Gamma ^n_{g,1/2}\) of the upper row of the commutative diagram
induces a splitting \((s_F,p)\circ s\) of the lower row, which agrees with the canonical splitting of the lower row up to conjugation with \(\mathbf {Z}^{2g}\otimes \pi _n\mathrm {SO}\), because such splittings up to conjugation are a torsor for \(\mathrm {H}^1(\mathrm {Sp}_{2g}(\mathbf {Z});\mathbf {Z}^{2g}\otimes \pi _n\mathrm {SO})\) which vanishes by Lemma A.3. Lemma 3.15 on the other hand ensures that there is a class \([f]\in \mathrm {H}_2(\mathrm {Sp}_{2g}(\mathbf {Z});\mathbf {Z})\) with signature 4, so \(s_*[f]\in \mathrm {H}_2(\Gamma ^n_{g,1/2};\mathbf {Z})\) satisfies \(\mathrm {sgn}(s_*[f])=4\) and \(\chi ^2(s_*[f])=0\), which contradicts Lemma 3.19 (iii).\(\square \)