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

Operations on stable moduli spaces

Søren Galatius galatius@math.ku.dk Department of Mathematics
University of Copenhagen
Denmark
 and  Oscar Randal-Williams o.randal-williams@dpmms.cam.ac.uk Centre for Mathematical Sciences
Wilberforce Road
Cambridge CB3 0WB
UK
Abstract.

We construct certain operations on stable moduli spaces and use them to compare cohomology of moduli spaces of closed manifolds with tangential structure. We obtain isomorphisms in a stable range provided the pp-adic valuation of the Euler characteristics agree, for all primes pp not invertible in the coefficients for cohomology.

2010 Mathematics Subject Classification:
55P47, 55R40, 57S05, 57R15, 57R90

1. Introduction

An influential theorem of Harer [Har85] shows that the cohomology of the moduli stack g\mathcal{M}_{g} of genus gg Riemann surfaces is independent of gg in a range of degrees called the stable range, even though there is no direct map between the moduli spaces for different genera. With rational coefficients the cohomology in the stable range is a polynomial ring, but with more general coefficients it is best described via infinite loop spaces, as shown by [Til97, MT01, MW07]. In earlier papers ([GRW14, GRW18, GRW17], see also [GRW19] for a survey) we have studied moduli spaces of higher dimensional manifolds, and in some cases have again shown that different moduli spaces have isomorphic cohomology in a range of degrees. For n>1n>1 one can in most cases not make an integral comparison of moduli spaces of manifolds related by connected sum with copies of Sn×SnS^{n}\times S^{n}, at least not by an obvious generalization of the n=1n=1 case, where a zig-zag of integral homology equivalences can be defined using manifolds with boundary. In this paper we show that a comparison is possible after all, although not with all coefficient modules. We also give examples showing that assumptions on the coefficients are necessary.

1.1. Comparing moduli spaces of closed manifolds

All manifolds in this paper will be smooth, compact, connected, and without boundary. If WW denotes such a manifold then there is a moduli space (W)\mathcal{M}(W) classifying smooth fibre bundles whose fibres are diffeomorphic to WW. As a model we may take (W)=BDiff(W)\mathcal{M}(W)=B\mathrm{Diff}(W), the classifying space of the diffeomorphism group Diff(W)\mathrm{Diff}(W) of WW, equipped with the CC^{\infty} topology. Then for AA an abelian group Hi((W);A)H^{i}(\mathcal{M}(W);A) is the group of Hi(;A)H^{i}(-;A)-valued characteristic classes of such fibre bundles.

Now let d=2nd=2n and WW be a dd-manifold. The connected sum W#(Sn×Sn)W\#(S^{n}\times S^{n}) is then well defined up to (non-canonical) diffeomorphism, as Sn×SnS^{n}\times S^{n} admits an orientation-reversing diffeomorphism, and we write W#g(Sn×Sn)W\#g(S^{n}\times S^{n}) for the gg-fold iteration of this operation. Two manifolds WW and WW^{\prime} are called stably diffeomorphic if W#g(Sn×Sn)W\#g(S^{n}\times S^{n}) is diffeomorphic to W#g(Sn×Sn)W^{\prime}\#g^{\prime}(S^{n}\times S^{n}) for some g,gg,g^{\prime}\in\mathbb{N}. For example, any two orientable connected surfaces are stably diffeomorphic, while two non-orientable connected surfaces are stably diffeomorphic if and only if their Euler characteristic have the same parity.

In this paper we shall ask about the relationship between H((W);A)H^{*}(\mathcal{M}(W);A) and H((W);A)H^{*}(\mathcal{M}(W^{\prime});A) when WW and WW^{\prime} are stably diffeomorphic. As a special case our main result will provide a canonical isomorphism

Hi((W);(p))Hi((W);(p))H^{i}(\mathcal{M}(W);\mathbb{Z}_{(p)})\cong H^{i}(\mathcal{M}(W^{\prime});\mathbb{Z}_{(p)})

as long as these manifolds are simply-connected and of dimension 2n>42n>4, and both (1)nχ(W)(-1)^{n}\chi(W) and (1)nχ(W)(-1)^{n}\chi(W^{\prime}) are large compared with ii and have the same pp-adic valuation.

The precise statement of our main result applies more generally, and before giving it we first explain its natural setting. If WW is given an orientation λ\lambda then there is a corresponding moduli space or(W,λ)\mathcal{M}^{\mathrm{or}}(W,\lambda) classifying smooth fibre bundles with oriented fibres which are oriented diffeomorphic to (W,λ)(W,\lambda), and a forgetful map or(W,λ)(W)\mathcal{M}^{\mathrm{or}}(W,\lambda)\to\mathcal{M}(W). Then the connected sum W#g(Sn×Sn)W\#g(S^{n}\times S^{n}) inherits an orientation, well defined up to oriented diffeomorphism, and we say that (W,λ)(W,\lambda) is oriented stably diffeomorphic to (W,λ)(W^{\prime},\lambda^{\prime}) provided W#g(Sn×Sn)W\#g(S^{n}\times S^{n}) is oriented diffeomorphic to W#g(Sn×Sn)W^{\prime}\#g^{\prime}(S^{n}\times S^{n}) for some g,gg,g^{\prime}\in\mathbb{N}. In this situation our result will also imply a canonical isomorphism Hi(or(W,λ);(p))Hi(or(W,λ);(p))H^{i}(\mathcal{M}^{\mathrm{or}}(W,\lambda);\mathbb{Z}_{(p)})\cong H^{i}(\mathcal{M}^{\mathrm{or}}(W^{\prime},\lambda^{\prime});\mathbb{Z}_{(p)}), under the same hypotheses.

More generally, for a space Λ\Lambda equipped with a continuous action of GLd+1()\mathrm{GL}_{d+1}(\mathbb{R}) a Λ\Lambda-structure on a dd-manifold WW is a GLd()\mathrm{GL}_{d}(\mathbb{R})-equivariant map λ:Fr(TW)Λ\lambda\mathrel{\mathop{\mathchar 58\relax}}\mathrm{Fr}(TW)\to\Lambda, or, equivalently, a GLd+1()\mathrm{GL}_{d+1}(\mathbb{R})-equivariant map Fr(ε1TW)Λ\mathrm{Fr}(\varepsilon^{1}\oplus TW)\to\Lambda. For example, if Λ={±1}\Lambda=\{\pm 1\} on which GLd+1()\mathrm{GL}_{d+1}(\mathbb{R}) acts by multiplication by the sign of the determinant, then a Λ\Lambda-structure λ:Fr(TW){±1}\lambda\mathrel{\mathop{\mathchar 58\relax}}\mathrm{Fr}(TW)\to\{\pm 1\} is the same thing as an orientation: it distinguishes oriented frames from non-oriented ones. Two Λ\Lambda-structures on the same manifold are homotopic if they are homotopic through equivariant maps, and (W,λ)(W,\lambda) is Λ\Lambda-diffeomorphic to (W,λ)(W^{\prime},\lambda^{\prime}) if there exists a diffeomorphism ϕ:WW\phi\mathrel{\mathop{\mathchar 58\relax}}W\to W^{\prime} such that λDϕ\lambda\circ D\phi is homotopic to λ\lambda^{\prime}. The usual embedding of Sn×Sn2n+1S^{n}\times S^{n}\subset\mathbb{R}^{2n+1} as the boundary of a thickened Sn×{0}n+1×nS^{n}\times\{0\}\subset\mathbb{R}^{n+1}\times\mathbb{R}^{n} gives a trivialisation of ε1T(Sn×Sn)\varepsilon^{1}\oplus T(S^{n}\times S^{n}) and a Λ\Lambda-structure on WW extends to one on W#(Sn×Sn)W\#(S^{n}\times S^{n}), canonically up to Λ\Lambda-diffeomorphism. For two pairs (W,λ)(W,\lambda) and (W,λ)(W^{\prime},\lambda^{\prime}) consisting of a manifold and a Λ\Lambda-structure, we say that they are stably Λ\Lambda-diffeomorphic if W#g(Sn×Sn)W\#g(S^{n}\times S^{n}) is Λ\Lambda-diffeomorphic to W#g(Sn×Sn)W^{\prime}\#g^{\prime}(S^{n}\times S^{n}) for some g,gg,g^{\prime}\in\mathbb{N}.

There is a moduli space Λ(W,λ)\mathcal{M}^{\Lambda}(W,\lambda) parametrising smooth fibre bundles π:EX\pi\mathrel{\mathop{\mathchar 58\relax}}E\to X with dd-dimensional fibres, and where the fibrewise tangent bundle TπET_{\pi}E is equipped with an equivariant map Fr(ε1TπE)Λ\mathrm{Fr}(\varepsilon^{1}\oplus T_{\pi}E)\to\Lambda, such that all fibres of π\pi are Λ\Lambda-diffeomorphic to (W,λ)(W,\lambda). Our main result is then as follows.

Theorem 1.1.

Let Λ\Lambda be as above, and let λ\lambda and λ\lambda^{\prime} be Λ\Lambda-structures on WW and WW^{\prime} such that (W,λ)(W,\lambda) is stably Λ\Lambda-diffeomorphic to (W,λ)(W^{\prime},\lambda^{\prime}). For an abelian group AA there is a canonical isomorphism

Hi(Λ(W,λ);A)Hi(Λ(W,λ);A),H^{i}(\mathcal{M}^{\Lambda}(W,\lambda);A)\cong H^{i}(\mathcal{M}^{\Lambda}(W^{\prime},\lambda^{\prime});A),

induced by a zig-zag of maps of spaces, provided

  1. (i)

    d=2n>4d=2n>4 and WW and WW^{\prime} are simply connected,

  2. (ii)

    the integers (1)nχ(W)(-1)^{n}\chi(W) and (1)nχ(W)(-1)^{n}\chi(W^{\prime}) are both 4i+C\geq 4i+C, where

    C=6+min{(1)nχ(W0)(W0,λ0) stably Λ-diffeomorphic to (W,λ) and (W,λ)}.C=6+\min\{(-1)^{n}\chi(W_{0})\mid\text{$(W_{0},\lambda_{0})$ stably $\Lambda$-diffeomorphic to $(W,\lambda)$ and $(W^{\prime},\lambda^{\prime})$}\}.
  3. (iii)

    χ(W)\chi(W) and χ(W)\chi(W^{\prime}) are both non-zero, and vp(χ(W))=vp(χ(W))v_{p}(\chi(W))=v_{p}(\chi(W^{\prime})) for all primes pp which are not invertible in End(A)\mathrm{End}_{\mathbb{Z}}(A).

In Section 4 we give an example showing the third condition cannot be relaxed.

The main results of [GRW14, GRW18, GRW17], summarised in [GRW19], provide a map

(1.1) Λ(W,λ)(ΩMTΘ)//hAut(u)\mathcal{M}^{\Lambda}(W,\lambda)\longrightarrow(\Omega^{\infty}MT\Theta)/\!\!/\mathrm{hAut}(u)

which is an isomorphism on homology in a range of degrees, when regarded as a map to the path component which it hits. Similarly there is a map

(1.2) Λ(W,λ)(ΩMTΘ)//hAut(u)\mathcal{M}^{\Lambda}(W^{\prime},\lambda^{\prime})\longrightarrow(\Omega^{\infty}MT\Theta)/\!\!/\mathrm{hAut}(u)

which is an isomorphism on homology in a range of degrees, when regarded as a map to the path component which it hits. The definition of the codomains is recalled below. However, if χ(W)χ(W)\chi(W)\neq\chi(W^{\prime}) then these two maps land in different path components, and the problem becomes to compare the homology of these two path components.

Remark 1.2.

Using the results of Friedrich [Fri17], Theorem 1.1 can be extended to manifolds with virtually polycyclic fundamental groups. In this case the constant CC should be replaced by C+4+2hC+4+2h where hh denotes the Hirsch length of the common fundamental group of WW and WW^{\prime}.

1.2. Operations on infinite loop spaces

The data involved in defining the common target of the maps (1.1) and (1.2) is a GL2n()\mathrm{GL}_{2n}(\mathbb{R})-equivariant fibration u:ΘΛu\mathrel{\mathop{\mathchar 58\relax}}\Theta\to\Lambda with domain which is cofibrant as a GL2n()\mathrm{GL}_{2n}(\mathbb{R})-space. Letting BB denote the Borel construction Θ//GL2n()\Theta/\!\!/\mathrm{GL}_{2n}(\mathbb{R}), MTΘMT\Theta is then the Thom spectrum of the inverse of the canonical 2n2n-dimensional vector bundle over BB, and ΩMTΘ\Omega^{\infty}MT\Theta is its associated infinite loop space. By functoriality the group-like topological monoid hAut(Θ)\mathrm{hAut}(\Theta) of GL2n()\mathrm{GL}_{2n}(\mathbb{R})-equivariant homotopy equivalences f:ΘΘf\mathrel{\mathop{\mathchar 58\relax}}\Theta\to\Theta acts on the infinite loop space ΩMTΘ\Omega^{\infty}MT\Theta, so the group-like submonoid hAut(u)={fhAut(Θ)|uf=u}\mathrm{hAut}(u)=\{f\in\mathrm{hAut}(\Theta)\,|\,u\circ f=u\} does too. The target

(ΩMTΘ)//hAut(u)(\Omega^{\infty}MT\Theta)/\!\!/\mathrm{hAut}(u)

of the maps (1.1) and (1.2) is the Borel construction for this action.

In order to prove Theorem 1.1 we shall construct certain operations on the space ΩMTΘ\Omega^{\infty}MT\Theta, in the case where the GL2n()\mathrm{GL}_{2n}(\mathbb{R})-space Θ\Theta is obtained by restriction from a cofibrant GL2n+1()\mathrm{GL}_{2n+1}(\mathbb{R})-space Θ¯\overline{\Theta}. The space B¯=Θ¯//GL2n+1()\overline{B}=\overline{\Theta}/\!\!/\mathrm{GL}_{2n+1}(\mathbb{R}) carries a canonical (2n+1)(2n+1)-dimensional vector bundle, and MTΘ¯MT\overline{\Theta} denotes its associated Thom spectrum; as above, by functoriality it carries an action of the monoid hAut(Θ¯)\mathrm{hAut}(\overline{\Theta}) of GL2n+1()\mathrm{GL}_{2n+1}(\mathbb{R})-equivariant homotopy equivalences f:Θ¯Θ¯f\mathrel{\mathop{\mathchar 58\relax}}\overline{\Theta}\to\overline{\Theta}.

A key construction in this paper is a homotopy pullback diagram of infinite loop spaces, equivariant for hAut(Θ¯)\mathrm{hAut}(\overline{\Theta}), of the form

(1.3) ΩMTΘ\textstyle{\Omega^{\infty}MT\Theta\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ω1MTΘ¯\textstyle{\Omega^{\infty-1}MT\overline{\Theta}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Q(B¯+)\textstyle{Q(\overline{B}_{+})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ΩCst,\textstyle{\Omega^{\infty}C_{st},}

whose bottom right corner has π0/2\pi_{0}\cong\mathbb{Z}/2 and all higher homotopy groups are 22-power torsion, and the bottom horizontal map induces a surjection on π1\pi_{1}. It induces an isomorphism

(1.4) π0MTΘ{(χ,x)×π1MTΘ¯χmod2=w2n(x)},\pi_{0}MT\Theta\xrightarrow{\cong}\{(\chi,x)\in\mathbb{Z}\times\pi_{-1}MT\overline{\Theta}\mid\chi\bmod 2=w_{2n}(x)\},

whose first coordinate is given by the Euler class and whose second coordinate is given by the stabilisation map. To explain this claim and its notation, first note that the 2n2n-dimensional vector bundle over BB has an Euler class eH2n(B;w1)e\in H^{2n}(B;\mathbb{Z}^{w_{1}}), where the coefficients are twisted by the determinant of this vector bundle, and under the Thom isomorphism this gives a class eu2nH0(MTΘ;)e\smile u_{-2n}\in H^{0}(MT\Theta;\mathbb{Z}). Then χ\chi is the value of this spectrum cohomology class on the Hurewicz image of an element of π0MTΘ\pi_{0}MT\Theta; geometrically, it assigns to such an element the Euler characteristic of a manifold representing it. Similarly, the (2n+1)(2n+1)-dimensional vector bundle over B¯\overline{B} has a 2n2nth Stiefel–Whitney class w2nH2n(B¯;/2)w_{2n}\in H^{2n}(\overline{B};\mathbb{Z}/2), and under the Thom isomorphism this gives a class w2nu2n1H1(MTΘ¯;/2)w_{2n}\smile u_{-2n-1}\in H^{-1}(MT\overline{\Theta};\mathbb{Z}/2). Then w2n(x)w_{2n}(x) denotes the value of this spectrum cohomology class on the Hurewicz image of xx.

Theorem 1.3.

For χ\chi\in\mathbb{Z}, write ΩχMTΘ\Omega^{\infty}_{\chi}MT\Theta for the inverse image of χ\chi under the map ΩMTΘ\Omega^{\infty}MT\Theta\to\mathbb{Z} induced by the class eu2nH0(MTΘ;)e\smile u_{-2n}\in H^{0}(MT\Theta;\mathbb{Z}), i.e. the union of the path components of the form (χ,?)(\chi,?) under the bijection (1.4).

For any odd number qq there exists a self-map MTΘMTΘMT\Theta\to MT\Theta inducing a map

ψq:ΩχMTΘΩqχMTΘ\psi^{q}\mathrel{\mathop{\mathchar 58\relax}}\Omega^{\infty}_{\chi}MT\Theta\longrightarrow\Omega^{\infty}_{q\chi}MT\Theta

such that

  1. (i)

    ψq\psi^{q} commutes (strictly) with the action of hAut(Θ¯)\mathrm{hAut}(\overline{\Theta}),

  2. (ii)

    ψq\psi^{q} is over the identity map of Ω1MTΘ¯\Omega^{\infty-1}MT\overline{\Theta},

  3. (iii)

    ψq\psi^{q} induces an isomorphism in homology with coefficients in any [q1]\mathbb{Z}[q^{-1}]-module.

We shall also prove a version of Theorem 1.3 for q=2q=2, although it will be marginally weaker in that rather than the map ψq\psi^{q} being defined integrally and inducing an isomorphism with coefficients in any [q1]\mathbb{Z}[q^{-1}]-module, the map ψ2\psi^{2} will only be defined after localising the spaces involved away from 22.

Theorem 1.4.

In the setup of Theorem 1.3, if χ\chi is even then there is a hAut(Θ¯)\mathrm{hAut}(\overline{\Theta})-equivariant weak equivalence of localised spaces

ψ2:(ΩχMTΘ)[12](Ω2χMTΘ)[12]\psi^{2}\mathrel{\mathop{\mathchar 58\relax}}(\Omega^{\infty}_{\chi}MT\Theta)[\tfrac{1}{2}]\longrightarrow(\Omega^{\infty}_{2\chi}MT\Theta)[\tfrac{1}{2}]

over the identity map of (Ω1MTΘ¯)[12](\Omega^{\infty-1}MT\overline{\Theta})[\frac{1}{2}].

The operations in Theorems 1.3 and 1.4 will arise from self-maps of the lower left corner in (1.3).

The proof of Theorem 1.1 will use these operations to give endomorphisms of the space (ΩMTΘ)//hAut(u)(\Omega^{\infty}MT\Theta)/\!\!/\mathrm{hAut}(u) which mix path-components, allowing us to compare the path components hit by the maps (1.1) and (1.2). This strategy is analogous to arguments of Bendersky–Miller [BM14] and Cantero–Palmer [CP15] for cohomology of configuration spaces. This strategy has also been used by Krannich [Kra19] to show that Hi(or(W,λ);A)Hi(or(W#Σ,λ);A)H^{i}(\mathcal{M}^{\mathrm{or}}(W,\lambda);A)\cong H^{i}(\mathcal{M}^{\mathrm{or}}(W\#\Sigma,\lambda);A) for (W,λ)(W,\lambda) an oriented manifold of dimension 2n>42n>4 and Σ\Sigma an exotic sphere, in a stable range of degrees when the order of [Σ]Θ2n[\Sigma]\in\Theta_{2n} is invertible in End(A)\mathrm{End}_{\mathbb{Z}}(A).

Acknowledgements

SG was supported by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement No. 682922), the Danish National Research Foundation through the Centre for Symmetry and Deformation (DNRF92), and the EliteForsk Prize. ORW was supported by EPSRC grant EP/M027783/1, the ERC under the European Union’s Horizon 2020 research and innovation programme (grant agreement No. 756444), and a Philip Leverhulme Prize from the Leverhulme Trust.

2. Proof of Theorem 1.1

We first explain how to deduce Theorem 1.1 from Theorems 1.3 and 1.4.

Let λ:Fr(ε1TW)𝜌Θ¯u¯Λ\lambda\mathrel{\mathop{\mathchar 58\relax}}\mathrm{Fr}(\varepsilon^{1}\oplus TW)\overset{\rho}{\to}\overline{\Theta}\overset{\overline{u}}{\to}\Lambda be a factorisation into an nn-connected GL2n+1()\mathrm{GL}_{2n+1}(\mathbb{R})-equivariant cofibration ρ\rho and a nn-co-connected GL2n+1()\mathrm{GL}_{2n+1}(\mathbb{R})-equivariant fibration u¯\overline{u}, and as above we write Θ\Theta for the underlying GL2n()\mathrm{GL}_{2n}(\mathbb{R})-space of Θ¯\overline{\Theta} and uu for the underlying GL2n()\mathrm{GL}_{2n}(\mathbb{R})-equivariant map of u¯\overline{u}. There is then a map

(2.1) Λ(W,λ)(ΩMTΘ)//hAut(u)\mathcal{M}^{\Lambda}(W,\lambda)\longrightarrow(\Omega^{\infty}MT\Theta)/\!\!/\mathrm{hAut}(u)

which by [GRW17, Corollary 1.9] is an isomorphism on iith (co)homology onto the path-component which it hits, as long as ig¯(W,λ)32i\leq\tfrac{\overline{g}(W,\lambda)-3}{2}. (Note that by considering a GL2n+1()\mathrm{GL}_{2n+1}(\mathbb{R})-space Λ\Lambda rather than a GL2n()\mathrm{GL}_{2n}(\mathbb{R})-space, the tangential structure Θ\Theta is “spherical” by the discussion after [GRW19, Definition 3.2], and so the stability range is as claimed.) Here g¯(W,λ)\bar{g}(W,\lambda) is the stable Λ\Lambda-genus of (W,λ)(W,\lambda), the largest gg\in\mathbb{N} for which there exists hh\in\mathbb{N} such that W#h(Sn×Sn)W\#h(S^{n}\times S^{n}) is Λ\Lambda-diffeomorphic to W0#(g+h)(Sn×Sn)W_{0}\#(g+h)(S^{n}\times S^{n}) for some (W0,λ0)(W_{0},\lambda_{0}).

Let (W0,λ0)(W_{0},\lambda_{0}) be a manifold stably Λ\Lambda-diffeomorphic to (W,λ)(W,\lambda) and minimising the quantity (1)nχ(W0)(-1)^{n}\chi(W_{0}). Such a manifold has stable Λ\Lambda-genus zero and hence for large enough hh we must have that W#h(Sn×Sn)W\#h(S^{n}\times S^{n}) is Λ\Lambda-diffeomorphic to W0#(h+g¯(W,λ))(Sn×Sn)W_{0}\#(h+\overline{g}(W,\lambda))(S^{n}\times S^{n}), so

g¯(W,λ)=(1)n(χ(W)χ(W0))/2.\overline{g}(W,\lambda)=(-1)^{n}(\chi(W)-\chi(W_{0}))/2.

It follows that (2.1) is an isomorphism on iith (co)homology as long as

(1)nχ(W)4i+(6+(1)nχ(W0)).(-1)^{n}\chi(W)\geq 4i+\left(6+(-1)^{n}\chi(W_{0})\right).

If (W,λ)(W^{\prime},\lambda^{\prime}) is stably Λ\Lambda-diffeomorphic to (W,λ)(W,\lambda) then the same analysis applies, and there is a map

(2.2) Λ(W,λ)(ΩMTΘ)//hAut(u)\mathcal{M}^{\Lambda}(W^{\prime},\lambda^{\prime})\longrightarrow(\Omega^{\infty}MT\Theta)/\!\!/\mathrm{hAut}(u)

which is an isomorphism on iith (co)homology onto the path-component which it hits, as long as

(1)nχ(W)4i+(6+(1)nχ(W0)).(-1)^{n}\chi(W^{\prime})\geq 4i+\left(6+(-1)^{n}\chi(W_{0})\right).

By assumption we may write

aχ(W)=bχ(W)a\cdot\chi(W)=b\cdot\chi(W^{\prime})

for integers aa and bb all of whose prime factors are invertible in End(A)\mathrm{End}_{\mathbb{Z}}(A). Furthermore the two Euler characteristics have the same parity, as (de)stabilisation changes the Euler characteristic by ±2\pm 2, so if either aa or bb is even then both χ(W)\chi(W) and χ(W)\chi(W^{\prime}) are even too.

By Theorems 1.3 and 1.4, writing ψx=ψx/2v2(x)(ψ2)v2(x)\psi^{x}=\psi^{x/2^{v_{2}(x)}}\circ(\psi^{2})^{v_{2}(x)}, (after perhaps implicitly localising away from 2) there are maps

Ωχ(W)MTΘ\textstyle{\Omega^{\infty}_{\chi(W)}MT\Theta\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψa\scriptstyle{\psi^{a}}Ωaχ(W)MTΘ\textstyle{\Omega^{\infty}_{a\chi(W)}MT\Theta\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωχ(W)MTΘ\textstyle{\Omega^{\infty}_{\chi(W^{\prime})}MT\Theta\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψb\scriptstyle{\psi^{b}}Ωbχ(W)MTΘ\textstyle{\Omega^{\infty}_{b\chi(W^{\prime})}MT\Theta}

which are hAut(Θ¯)\mathrm{hAut}(\overline{\Theta})-equivariant and induce isomorphisms on AA-homology, as AA is a [a1,b1]\mathbb{Z}[a^{-1},b^{-1}]-module. By construction these maps do not change the π1MTΘ¯\pi_{-1}MT\overline{\Theta}-component: we now analyse the components corresponding to WW and WW^{\prime}.

We now claim that ψa([W,ρ])=ψb([W,ρ])π0(ΩMTΘ)\psi^{a}([W,\rho])=\psi^{b}([W^{\prime},\rho^{\prime}])\in\pi_{0}(\Omega^{\infty}MT\Theta) for a suitable choice of ρ:Fr(ε1TW)Θ¯\rho^{\prime}\mathrel{\mathop{\mathchar 58\relax}}\mathrm{Fr}(\varepsilon^{1}\oplus TW^{\prime})\to\overline{\Theta} lifting λ\lambda^{\prime}. Since these two elements of π0(MTΘ)\pi_{0}(MT\Theta) have the same Euler characteristic, it suffices to arrange that they also have the same π1MTΘ¯\pi_{-1}MT\overline{\Theta}-component. The stable Λ\Lambda-diffeomorphism from (W,λ)(W,\lambda) to (W,λ)(W^{\prime},\lambda^{\prime}) gives a Λ\Lambda-cobordism

X:W#g(Sn×Sn)W#g(Sn×Sn)X\mathrel{\mathop{\mathchar 58\relax}}W\#g(S^{n}\times S^{n})\leadsto W^{\prime}\#g^{\prime}(S^{n}\times S^{n})

which is furthermore an hh-cobordism. We can therefore extend the Θ¯\overline{\Theta}-structure given by (W,ρ)(W,\rho), stabilised, to a Θ¯\overline{\Theta}-structure on XX lifting the given Λ\Lambda-structure, and hence obtain a Θ¯\overline{\Theta}-manifold (W#g(Sn×Sn),ρ′′)(W^{\prime}\#g^{\prime}(S^{n}\times S^{n}),\rho^{\prime\prime}) whose underlying Λ\Lambda-manifold (W#g(Sn×Sn),uρ′′)(W^{\prime}\#g^{\prime}(S^{n}\times S^{n}),u\circ\rho^{\prime\prime}) is the stabilisation of (W,λ)(W^{\prime},\lambda^{\prime}). Now the Θ¯\overline{\Theta}-manifolds

(2.3) (W#g(Sn×Sn),ρ′′) and (W,ρ)#g(Sn×Sn)(W^{\prime}\#g^{\prime}(S^{n}\times S^{n}),\rho^{\prime\prime})\text{ and }(W^{\prime},\rho^{\prime})\#g^{\prime}(S^{n}\times S^{n})

need not be Θ¯\overline{\Theta}-diffeomorphic, but must differ by an equivalence f:Θ¯Θ¯f\mathrel{\mathop{\mathchar 58\relax}}\overline{\Theta}\to\overline{\Theta} over Λ\Lambda (see [GRW17, Lemma 9.2]). However the Θ¯\overline{\Theta}-structure ρ\rho^{\prime} on WW^{\prime} is merely a choice of lift of λ\lambda^{\prime} along u¯\overline{u}, and by re-choosing it to be fρf\circ\rho^{\prime} we may then suppose that the manifolds (2.3) are indeed Θ¯\overline{\Theta}-diffeomorphic. With this choice we therefore have the desired

[W,ρ]=[W,ρ]π1MTΘ¯,[W,\rho]=[W^{\prime},\rho^{\prime}]\in\pi_{-1}MT\overline{\Theta},

using the Θ¯\overline{\Theta}-cobordism XX and the fact that this cobordism theory is insensitive to stabilisation by standard Sn×SnS^{n}\times S^{n}’s.

Denoting by [[W,λ]]π0MTΘ[[W,\lambda]]\subset\pi_{0}MT\Theta the π0hAut(u¯)\pi_{0}\mathrm{hAut}(\overline{u})-orbit of [W,ρ][W,\rho], and similarly [[W,λ]][[W^{\prime},\lambda^{\prime}]], and using the forgetful homomorphism hAut(u¯)hAut(Θ¯)\mathrm{hAut}(\overline{u})\to\mathrm{hAut}(\overline{\Theta}) to let the monoid hAut(u¯)\mathrm{hAut}(\overline{u}) act on ΩMTΘ\Omega^{\infty}MT\Theta, we therefore have a zig-zag of maps

(2.4) (Ω[[W,λ]]MTΘ)//hAut(u¯)(Ω[[W,λ]]MTΘ)//hAut(u¯)\left(\Omega^{\infty}_{[[W,\lambda]]}MT\Theta\right)/\!\!/\mathrm{hAut}(\overline{u})\longrightarrow\cdot\longleftarrow\left(\Omega^{\infty}_{[[W^{\prime},\lambda^{\prime}]]}MT\Theta\right)/\!\!/\mathrm{hAut}(\overline{u})

which induce isomorphisms on homology with AA-coefficients. The argument is completed by the following lemma.

Lemma 2.1.

The natural map hAut(u¯)hAut(u)\mathrm{hAut}(\overline{u})\to\mathrm{hAut}(u) is a weak equivalence.

Proof.

Working in the categories of GL2n()\mathrm{GL}_{2n}(\mathbb{R})-spaces over Λ\Lambda, or GL2n+1()\mathrm{GL}_{2n+1}(\mathbb{R})-spaces over Λ\Lambda, we have

mapGL2n()/Λ(Θ,Θ)=mapGL2n+1()/Λ(GL2n+1()×GL2n()Θ,Θ¯)\mathrm{map}^{/\Lambda}_{\mathrm{GL}_{2n}(\mathbb{R})}(\Theta,\Theta)=\mathrm{map}^{/\Lambda}_{\mathrm{GL}_{2n+1}(\mathbb{R})}(\mathrm{GL}_{2n+1}(\mathbb{R})\times_{\mathrm{GL}_{2n}(\mathbb{R})}\Theta,\overline{\Theta})

but the natural GL2n+1()\mathrm{GL}_{2n+1}(\mathbb{R})-equivariant map GL2n+1()×GL2n()ΘΘ¯\mathrm{GL}_{2n+1}(\mathbb{R})\times_{\mathrm{GL}_{2n}(\mathbb{R})}\Theta\to\overline{\Theta} has homotopy fibre GL2n+1()/GL2n()S2n\mathrm{GL}_{2n+1}(\mathbb{R})/\mathrm{GL}_{2n}(\mathbb{R})\simeq S^{2n} so is 2n2n-connected, whereas u¯:Θ¯Λ\overline{u}\mathrel{\mathop{\mathchar 58\relax}}\overline{\Theta}\to\Lambda is nn-co-connected, so the restriction map

mapGL2n+1()/Λ(Θ¯,Θ¯)mapGL2n+1()/Λ(GL2n+1()×GL2n()Θ,Θ¯)\mathrm{map}^{/\Lambda}_{\mathrm{GL}_{2n+1}(\mathbb{R})}(\overline{\Theta},\overline{\Theta})\longrightarrow\mathrm{map}^{/\Lambda}_{\mathrm{GL}_{2n+1}(\mathbb{R})}(\mathrm{GL}_{2n+1}(\mathbb{R})\times_{\mathrm{GL}_{2n}(\mathbb{R})}\Theta,\overline{\Theta})

is an equivalence. The claim now follows by restricting to the path-components of homotopy equivalences. ∎

Remark 2.2.

This argument also gives a conclusion about homology with certain local coefficients. The maps (2.1) and (2.2) are in fact acyclic in a range of degrees [GRW17, Corollary 1.9], and the maps ψq\psi^{q} are acyclic with [q1]\mathbb{Z}[q^{-1}]-module coefficients (as they are infinite loop maps which are isomorphisms on homology with these coefficients) so remain so after taking homotopy orbits by hAut(u¯)\mathrm{hAut}(\overline{u}).

So if 𝒜\mathcal{A} is a system of local coefficients on the middle space of the zig-zag (2.4), with typical fibre AA and having vp(χ(W))=vp(χ(W))v_{p}(\chi(W))=v_{p}(\chi(W^{\prime})) for all primes pp which are not invertible in End(A)\mathrm{End}_{\mathbb{Z}}(A), then there is also an isomorphism Hi(Λ(W,λ);𝒜)Hi(Λ(W,λ);𝒜)H^{i}(\mathcal{M}^{\Lambda}(W,\lambda);\mathcal{A})\cong H^{i}(\mathcal{M}^{\Lambda}(W^{\prime},\lambda^{\prime});\mathcal{A}) in a range of degrees.

3. Proof of Theorems 1.3 and 1.4

The proof of Theorem 1.3 is by an explicit construction of ψq\psi^{q} as a map of spectra. The main ingredient is a certain commutative diagram of spectra, which we first describe informally. It is

ΣB¯+\textstyle{\Sigma^{\infty}\overline{B}_{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}MTΘ\textstyle{{MT\Theta}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}sz\scriptstyle{sz}S1MTΘ¯\textstyle{S^{1}\wedge MT\overline{\Theta}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ΣB¯+\textstyle{\Sigma^{\infty}\overline{B}_{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}st\scriptstyle{st}ΣB¯+\textstyle{\Sigma^{\infty}\overline{B}_{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Cst\textstyle{C_{st}}

where s:BB¯s\mathrel{\mathop{\mathchar 58\relax}}B\to\overline{B} is the natural map of Borel constructions. The map ss is homotopy equivalent to a smooth fibre bundle with fibres S2nS^{2n} so we have a Becker–Gottlieb transfer t:ΣB¯+ΣB+t\mathrel{\mathop{\mathchar 58\relax}}\Sigma^{\infty}\overline{B}_{+}\to\Sigma^{\infty}B_{+}, factoring as a pre-transfer p:ΣB¯+MTΘp\mathrel{\mathop{\mathchar 58\relax}}\Sigma^{\infty}\overline{B}_{+}\to MT\Theta composed with a map z:MTΘΣB+z\mathrel{\mathop{\mathchar 58\relax}}MT\Theta\to\Sigma^{\infty}B_{+} induced by the zero section of θ\theta. The spectrum CstC_{st} is defined to be the homotopy cofibre of stst, and both rows are cofibre sequences. It follows that the right square in the diagram is a homotopy pullback, and hence we get the homotopy pullback diagram of infinite loop spaces (1.3) mentioned in the introduction. On spectrum homology the map stst induces multiplication by χ(S2n)=2\chi(S^{2n})=2, from which it follows that the homology and hence homotopy groups of CstC_{st} are 2-power torsion. The space B¯\overline{B} is path connected, because WW is, so π0(ΣB¯+)=H0(ΣB¯+;)=\pi_{0}(\Sigma^{\infty}\overline{B}_{+})=H_{0}(\Sigma^{\infty}\overline{B}_{+};\mathbb{Z})=\mathbb{Z}. Thus π0(Cst)=/2\pi_{0}(C_{st})=\mathbb{Z}/2, and the map ΣB¯+Cst\Sigma^{\infty}\overline{B}_{+}\to C_{st} is surjective on π1\pi_{1} because stst is injective on π0\pi_{0}.

To produce an endomorphism of ΩMTΘ\Omega^{\infty}MT\Theta satisfying part (ii) of the theorem, it therefore suffices to produce an endomorphism of ΣB¯+\Sigma^{\infty}\overline{B}_{+} over CstC_{st}. For q=1+2kq=1+2k, we may use the map id+kst:ΣB¯+ΣB¯+\mathrm{id}+kst\mathrel{\mathop{\mathchar 58\relax}}\Sigma^{\infty}\overline{B}_{+}\to\Sigma^{\infty}\overline{B}_{+} which is obviously over CstC_{st}, at least in the homotopy category, since CstC_{st} is the cofibre of the map stst. In spectrum homology, stst multiplies by χ(S2n)=2\chi(S^{2n})=2 and hence id+kst\mathrm{id}+kst induces multiplication by 1+2k=q1+2k=q on H(ΣB¯+;)H_{*}(\Sigma^{\infty}\overline{B}_{+};\mathbb{Z}) ensuring part (iii) of the theorem. Furthermore it acts by multiplication by qq on π0ΣB¯+=π0Q(B¯+)=\pi_{0}\Sigma^{\infty}\overline{B}_{+}=\pi_{0}Q(\overline{B}_{+})=\mathbb{Z}, so indeed sends ΩχMTΘ\Omega^{\infty}_{\chi}MT\Theta to ΩqχMTΘ\Omega^{\infty}_{q\chi}MT\Theta.

It remains to explain how to achieve part (i) of the theorem, that the continuous action of the topological monoid hAut(Θ¯)\mathrm{hAut}(\overline{\Theta}) on the space ΩMTΘ\Omega^{\infty}MT\Theta commutes with ψq\psi^{q}. It is not sufficient that ψq\psi^{q} commutes up to homotopy with the action of individual elements of hAut(Θ¯)\mathrm{hAut}(\overline{\Theta}), since we want to descend ψq\psi^{q} to the homotopy orbit space. To give a convincing proof, it seems best to spell out a point-set model for the square (1.3).

Proof of Theorem 1.3.

As explained above, it remains to give a point-set model for the diagram (1.3) and the self-map id+kst\mathrm{id}+kst of Q(B¯+)Q(\overline{B}_{+}) over ΩCst\Omega^{\infty}C_{st}, all of which commutes strictly with the action of hAut(Θ¯)\mathrm{hAut}(\overline{\Theta}).

We must adopt some conventions. Let us consider GL2n()\mathrm{GL}_{2n}(\mathbb{R}) as lying inside GL2n+1()\mathrm{GL}_{2n+1}(\mathbb{R}) using the last 2n2n coordinates. Let us consider N1\mathbb{R}^{N-1} as lying inside N\mathbb{R}^{N} as the subspace of vectors whose last coordinate is 0, and take \mathbb{R}^{\infty} to be the direct limit. To form the Borel constructions we shall take EGL2n():=Fr2n()E\mathrm{GL}_{2n}(\mathbb{R})\mathrel{\mathop{\mathchar 58\relax}}=\mathrm{Fr}_{2n}(\mathbb{R}^{\infty}), and similarly take EGL2n+1():=Fr2n+1()E\mathrm{GL}_{2n+1}(\mathbb{R})\mathrel{\mathop{\mathchar 58\relax}}=\mathrm{Fr}_{2n+1}(\mathbb{R}\oplus\mathbb{R}^{\infty}). The map Fr2n()Fr2n+1()\mathrm{Fr}_{2n}(\mathbb{R}^{\infty})\to\mathrm{Fr}_{2n+1}(\mathbb{R}\oplus\mathbb{R}^{\infty}) which adds the basis vector of the first \mathbb{R}-summand as the first element of the (2n+1)(2n+1)-frame is then equivariant for the inclusion GL2n()GL2n+1()\mathrm{GL}_{2n}(\mathbb{R})\subset\mathrm{GL}_{2n+1}(\mathbb{R}).

Then we have BGL2n+1()=Gr2n+1()B\mathrm{GL}_{2n+1}(\mathbb{R})=\mathrm{Gr}_{2n+1}(\mathbb{R}\oplus\mathbb{R}^{\infty}), which we may filter in the usual way by Gr2n+1(N1)\mathrm{Gr}_{2n+1}(\mathbb{R}\oplus\mathbb{R}^{N-1}). Pulling back this filtration along the map θ¯:B¯Gr2n+1()\overline{\theta}\mathrel{\mathop{\mathchar 58\relax}}\overline{B}\to\mathrm{Gr}_{2n+1}(\mathbb{R}^{\infty}), we set B¯N:=(θ¯)1(Gr2n+1(N1))\overline{B}_{N}\mathrel{\mathop{\mathchar 58\relax}}=(\overline{\theta})^{-1}(\mathrm{Gr}_{2n+1}(\mathbb{R}\oplus\mathbb{R}^{N-1})). There is an induced map θ¯N:B¯NGr2n+1(N1)\overline{\theta}_{N}\mathrel{\mathop{\mathchar 58\relax}}\overline{B}_{N}\to\mathrm{Gr}_{2n+1}(\mathbb{R}\oplus\mathbb{R}^{N-1}) and we shall write θ¯Nγ=θ¯Nγ2n+1,N\overline{\theta}_{N}^{*}\gamma^{\perp}=\overline{\theta}_{N}^{*}\gamma_{2n+1,N}^{\perp} for the pullback of the (N2n1)(N-2n-1)-dimensional bundle of orthogonal complements. Then MTΘ¯MT\overline{\Theta} is the spectrum with NNth space given by the Thom space (B¯N)θ¯Nγ(\overline{B}_{N})^{\overline{\theta}_{N}^{*}\gamma^{\perp}}, so that

Ω1MTΘ¯=colimNΩN1(B¯N)θ¯Nγ.\Omega^{\infty-1}MT\overline{\Theta}=\operatorname*{colim}_{N\to\infty}\Omega^{N-1}(\overline{B}_{N})^{\overline{\theta}_{N}^{*}\gamma^{\perp}}.

We similarly define θN:BNGr2n(N)\theta_{N}\mathrel{\mathop{\mathchar 58\relax}}B_{N}\to\mathrm{Gr}_{2n}(\mathbb{R}^{N}), and hence the spectrum MTΘMT\Theta. There is a map

(3.1) Gr2n(N1)Gr2n+1(N1),\mathrm{Gr}_{2n}(\mathbb{R}^{N-1})\hookrightarrow\mathrm{Gr}_{2n+1}(\mathbb{R}\oplus\mathbb{R}^{N-1}),

given by direct sum with the 1-dimensional vector space given by the first \mathbb{R}-summand, which induces a map BN1B¯NB_{N-1}\to\overline{B}_{N}. The map (3.1) is 2n2n-connected, but is covered by an (N2)(N-2)-connected map Gr2n(N1)S(γ2n+1,N)\mathrm{Gr}_{2n}(\mathbb{R}^{N-1})\to S(\gamma_{2n+1,N}) and hence gives a (N2)(N-2)-connected map BN1S(θ¯Nγ2n+1,N)B_{N-1}\to S(\overline{\theta}_{N}^{*}\gamma_{2n+1,N}). Passing to Thom spaces this gives a (2N2n2)(2N-2n-2)-connected map

S1(BN1)(θ¯N|BN1)γ2n,N1S(θ¯Nγ2n+1,N)θ¯Nγ2n+1,N.S^{1}\wedge(B_{N-1})^{(\overline{\theta}_{N}|_{B_{N-1}})^{*}\gamma_{2n,N-1}^{\perp}}\longrightarrow S(\overline{\theta}_{N}^{*}\gamma_{2n+1,N})^{\overline{\theta}_{N}^{*}\gamma_{2n+1,N}^{\perp}}.

These combine to define a map from MTΘMT\Theta to the spectrum whose (N1)(N-1)st space is S(θ¯Nγ2n+1,N)θ¯Nγ2n+1,NS(\overline{\theta}_{N}^{*}\gamma_{2n+1,N})^{\overline{\theta}_{N}^{*}\gamma_{2n+1,N}^{\perp}}, and this map is a weak equivalence. This map is also hAut(Θ¯)\mathrm{hAut}(\overline{\Theta})-equivariant. (This weak equivalence does not come with a spectrum map in the other direction, let alone an equivariant one.)

The square (1.3) will be assembled from a square of spaces fibred over B¯N\overline{B}_{N}, and we first explain the constructions on fibres. Let VGr2n+1(N)V\in\mathrm{Gr}_{2n+1}(\mathbb{R}^{N}) and write S(V)S(V) for the unit sphere of VV and SVS^{V} for the one-point compactification. If xNx\in\mathbb{R}^{N} we shall write πV(x)V\pi_{V}(x)\in V for the orthogonal projection. If xV0x\in V\setminus 0 we shall write πS(x)=x/|x|S(V)\pi_{S}(x)=x/|x|\in S(V) for the nearest point in the sphere. We will describe certain explicit maps p(V):SVS(V)ε1p(V)\mathrel{\mathop{\mathchar 58\relax}}S^{V}\to S(V)^{\varepsilon^{1}} and z(V):S(V)ε1S(V)+SVz(V)\mathrel{\mathop{\mathchar 58\relax}}S(V)^{\varepsilon^{1}}\to S(V)_{+}\wedge S^{V}, and explain how the composition z(V)p(V)z(V)\circ p(V) gives rise to a model for the Becker–Gottlieb transfer for a linear sphere bundle (indeed, we will just unwrap the definition of [BG75, Section 3] in this case).

The map

p(V):SVS(V)ε1,p(V)\mathrel{\mathop{\mathchar 58\relax}}S^{V}\longrightarrow S(V)^{\varepsilon^{1}},

is induced by the Pontryagin–Thom construction applied to the embedding S(V)VS(V)\subset V. In formulas, we can take e.g.

p(V)(x)=(πS(x),log|x|)S(V)+S1=S(V)ε1p(V)(x)=(\pi_{S}(x),\log|x|)\in S(V)_{+}\wedge S^{1}=S(V)^{\varepsilon^{1}}

when x0,SVx\neq 0,\infty\in S^{V}. The Thom space S(V)ε1S(V)^{\varepsilon^{1}} is homeomorphic to the quotient SV/S0S^{V}/S^{0}, and under this identification the map p(V)p(V) is the quotient map.

The map

z(V):S(V)ε1S(V)TS(V)ε1=S(V)+SVz(V)\mathrel{\mathop{\mathchar 58\relax}}S(V)^{\varepsilon^{1}}\longrightarrow S(V)^{TS(V)\oplus\varepsilon^{1}}=S(V)_{+}\wedge S^{V}

is given by the zero section of the tangent bundle of S(V)S(V). In formulas, it sends (x,t)S(V)×S(V)ε1(x,t)\in S(V)\times\mathbb{R}\subset S(V)^{\varepsilon^{1}} to (x,tx)S(V)×VS(V)+SV(x,tx)\in S(V)\times V\subset S(V)_{+}\wedge S^{V}.

If we compose these two maps and smash with SVS^{V^{\perp}}, we get

SN=SVSVp(V)idS(V)ε1SVz(V)idS(V)+SVSV=S(V)+SN.S^{N}=S^{V}\wedge S^{V^{\perp}}\xrightarrow{p(V)\wedge\mathrm{id}}S(V)^{\varepsilon^{1}}\wedge S^{V^{\perp}}\xrightarrow{z(V)\wedge\mathrm{id}}S(V)_{+}\wedge S^{V}\wedge S^{V^{\perp}}=S(V)_{+}\wedge S^{N}.

Finally, we write s(V):S(V)+SNSNs(V)\mathrel{\mathop{\mathchar 58\relax}}S(V)_{+}\wedge S^{N}\to S^{N} for the map induced by collapsing S(V)S(V) to a point. Then the composition

b(V)=s(V)(z(V)id)(p(V)id):SNSNb(V)=s(V)\circ(z(V)\wedge\mathrm{id})\circ(p(V)\wedge\mathrm{id})\mathrel{\mathop{\mathchar 58\relax}}S^{N}\longrightarrow S^{N}

is a continuous map of degree χ(S2n)=2\chi(S^{2n})=2 (by the Poincaré–Hopf theorem, see [BG75, Theorem 2.4]), depending continuously on the point VGr2n+1(N)V\in\mathrm{Gr}_{2n+1}(\mathbb{R}^{N}). The resulting continuous map b:Gr2n+1(N)ΩNSNb\mathrel{\mathop{\mathchar 58\relax}}\mathrm{Gr}_{2n+1}(\mathbb{R}^{N})\to\Omega^{N}S^{N} in the limit gives a map BGL2n+1()QS0B\mathrm{GL}_{2n+1}(\mathbb{R})\to QS^{0} which is a model for the Becker–Gottlieb transfer of the sphere bundle over BGL2n+1()BO(2n+1)B\mathrm{GL}_{2n+1}(\mathbb{R})\simeq B\mathrm{O}(2n+1).

Now consider the diagram

SN\textstyle{S^{N}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p(V)\scriptstyle{p(V)}S(V)ε1SV\textstyle{{S(V)^{\varepsilon^{1}}\wedge S^{V^{\perp}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}sz\scriptstyle{sz}Cp(V)\textstyle{C_{p(V)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}SN\textstyle{S^{N}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}st(V)\scriptstyle{st(V)}SN\textstyle{S^{N}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Cst(V),\textstyle{C_{st(V)},}

where the entries in the right column are the mapping cylinders. Since p(V)p(V) induces a homeomorphism SN/SVS(V)εSVS^{N}/S^{V^{\perp}}\to S(V)^{\varepsilon}\wedge S^{V^{\perp}}, it follows from the Puppe sequence that there is a canonical induced homeomorphism Cp(V)S1SVC_{p(V)}\cong S^{1}\wedge S^{V^{\perp}}. Since st(V)st(V) has degree 2, there is a homotopy equivalence from Cst(V)C_{st(V)} to a mod 2 Moore space, but this is not quite sufficiently canonical for our purposes (since we get a different mod 2 Moore space for each VV). We have proved that for each VGr2n+1(N)V\in\mathrm{Gr}_{2n+1}(\mathbb{R}^{N}) there is a canonical commutative diagram

(3.2) S(V)ε1SV\textstyle{{S(V)^{\varepsilon^{1}}\wedge S^{V^{\perp}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}sz\scriptstyle{sz}S1SV\textstyle{S^{1}\wedge S^{V^{\perp}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}SN\textstyle{S^{N}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Cst(V),\textstyle{C_{st(V)},}

which is a pushout and homotopy pushout.

There is a canonical homotopy from the composition of st(V):SNSNst(V)\mathrel{\mathop{\mathchar 58\relax}}S^{N}\to S^{N} and SNCst(V)S^{N}\to C_{st(V)} to the constant map. Suspending once, S1SNS1SNS1Cst(V)S^{1}\wedge S^{N}\to S^{1}\wedge S^{N}\to S^{1}\wedge C_{st(V)} is canonically null homotopic. If k0k\geq 0 is an integer, we may use the S1S^{1} coordinate to form the sum of the identity map 1:S1SNS1SN1\mathrel{\mathop{\mathchar 58\relax}}S^{1}\wedge S^{N}\to S^{1}\wedge S^{N} and kk copies of the map st(V):S1SNS1SNst(V)\mathrel{\mathop{\mathchar 58\relax}}S^{1}\wedge S^{N}\to S^{1}\wedge S^{N}. We obtain a diagram

(3.3) S1SN\textstyle{{S^{1}\wedge S^{N}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1+kst(V)\scriptstyle{1+kst(V)}S1Cst(V)\textstyle{{S^{1}\wedge C_{st(V)}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}S1SN\textstyle{{S^{1}\wedge S^{N}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}S1Cst(V),\textstyle{{S^{1}\wedge C_{st(V)},}}

which commutes up to a canonical homotopy. (The canonical nullhomotopy of each stst gives a homotopy from 1+kst1+kst to the sum of the identity map and kk copies of the constant map; this is in turn canonically homotopic to the identity map.) The homotopy class of the map 1+kst(V):SNSN1+kst(V)\mathrel{\mathop{\mathchar 58\relax}}S^{N}\to S^{N} is determined by its degree which is 2k+12k+1, but the actual map depends in a non-trivial way on VGr2n+1(N)V\in\mathrm{Gr}_{2n+1}(\mathbb{R}^{N}).

All spaces in the diagram “vary continuously in VV”, in the sense that they are fibres over VV of fibre bundles over Gr2n+1(N)\mathrm{Gr}_{2n+1}(\mathbb{R}^{N}). The commutative diagram (3.2) in the category of spaces over Gr2n+1(N)\mathrm{Gr}_{2n+1}(\mathbb{R}^{N}) may be pulled back along θ¯N:B¯NGr2n+1(N)\overline{\theta}_{N}\mathrel{\mathop{\mathchar 58\relax}}\overline{B}_{N}\to\mathrm{Gr}_{2n+1}(\mathbb{R}^{N}) to form a diagram

(3.4) S(θ¯Nγ)ε1θ¯Nγ\textstyle{{S(\overline{\theta}_{N}^{*}\gamma)^{\varepsilon^{1}\oplus\overline{\theta}_{N}^{*}\gamma^{\perp}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}sz\scriptstyle{sz}S1B¯Nθ¯Nγ\textstyle{S^{1}\wedge\overline{B}_{N}^{\overline{\theta}_{N}^{*}\gamma^{\perp}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}SN(B¯N)+\textstyle{S^{N}\wedge(\overline{B}_{N})_{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}CstB¯N,\textstyle{C_{st}^{\overline{B}_{N}},}

which is again a pushout and homotopy pushout, where CstB¯NC_{st}^{\overline{B}_{N}} is the mapping cylinder of the map SN(B¯N)+SN(B¯N)+S^{N}\wedge(\overline{B}_{N})_{+}\to S^{N}\wedge(\overline{B}_{N})_{+} given on (v,x)SN×B¯N(v,x)\in S^{N}\times\overline{B}_{N} by st(v,x)=(st(f(x))v,x)st(v,x)=(st(f(x))v,x).

Similarly, the diagrams (3.3) assemble over VV to a diagram

(3.5) S1SN(B¯N)+\textstyle{{S^{1}\wedge S^{N}\wedge(\overline{B}_{N})_{+}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1+kst\scriptstyle{1+kst}S1CstB¯N\textstyle{{S^{1}\wedge C_{st}^{\overline{B}_{N}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}S1SN(B¯N)+\textstyle{{S^{1}\wedge S^{N}\wedge(\overline{B}_{N})_{+}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}S1CstB¯N,\textstyle{{S^{1}\wedge C_{st}^{\overline{B}_{N}},}}

which commutes up to a canonical homotopy.

Applying ΩN+1S1()\Omega^{N+1}S^{1}\wedge(-) to the diagram (3.4) and letting NN\to\infty we get a model for (1.3). The monoid hAut(Θ¯)\mathrm{hAut}(\overline{\Theta}) acts on the whole diagram (3.4), since it acts on B¯N\overline{B}_{N} over Gr2n+1(N)\mathrm{Gr}_{2n+1}(\mathbb{R}^{N}). This gives a weak equivalence from ΩMTΘ\Omega^{\infty}MT\Theta to the homotopy pullback in (1.3), which is also an hAut(Θ¯)\mathrm{hAut}(\overline{\Theta}) equivariant map. The monoid hAut(Θ¯)\mathrm{hAut}(\overline{\Theta}) also acts on the diagram (3.5), including the homotopy, and after applying ΩN+1\Omega^{N+1} and taking NN\to\infty we obtain a self-map of Q(B¯+)Q(\overline{B}_{+}) which is over ΩCst\Omega^{\infty}C_{st} up to a specified homotopy. Again this self-map and the specified homotopy commutes strictly with the action of hAut(Θ¯)\mathrm{hAut}(\overline{\Theta}) since both the map and the homotopy arose from fibrewise constructions over Gr2n+1(N)\mathrm{Gr}_{2n+1}(\mathbb{R}^{N}).

Finally, the self-map of Q(B¯+)Q(\overline{B}_{+}) induces an hAut(Θ¯)\mathrm{hAut}(\overline{\Theta})-equivariant self-map of the homotopy pullback of Q(B¯+)ΩCstΩ1MTΘ¯Q(\overline{B}_{+})\rightarrow\Omega^{\infty}C_{st}\leftarrow\Omega^{\infty-1}MT\overline{\Theta}, and we have seen that this pullback is weakly equivalent to ΩMTΘ\Omega^{\infty}MT\Theta by an hAut(Θ¯)\mathrm{hAut}(\overline{\Theta})-equivariant map. ∎

Proof of Theorem 1.4.

We continue with the notation developed above. The spectrum homology of CstC_{st} is all 22-torsion, so the localisation Cst[12]C_{st}[\frac{1}{2}] as a spectrum is contractible. However, the localised space (ΩCst)[12](\Omega^{\infty}C_{st})[\frac{1}{2}] is not contractible since it has two components. Instead, there is a spectrum map w2n:CstH𝔽2w_{2n}\mathrel{\mathop{\mathchar 58\relax}}C_{st}\to H\mathbb{F}_{2} which becomes an isomorphism in homology of infinite loop spaces with coefficients in any [12]\mathbb{Z}[\frac{1}{2}]-module. Similarly, the map

ΩMTΘQ(B¯+)×ΩH𝔽2Ω1MTΘ¯\Omega^{\infty}MT\Theta\longrightarrow Q(\overline{B}_{+})\times_{\Omega^{\infty}H\mathbb{F}_{2}}\Omega^{\infty-1}MT\overline{\Theta}

induces an isomorphism in homology with coefficients in any [12]\mathbb{Z}[\frac{1}{2}]-module, and hence a weak equivalence of localized spaces. The spectrum map 2:S0S02\mathrel{\mathop{\mathchar 58\relax}}S^{0}\to S^{0} induces a self-map of Q(B¯+)Q(\overline{B}_{+}) commuting with the action of hAut(Θ¯)\mathrm{hAut}(\overline{\Theta}) and whose restriction to the even-degree path components commutes with the map to ΩH𝔽2\Omega^{\infty}H\mathbb{F}_{2}. This self-map can be used in place of 1+kst1+kst to produce ψ2\psi^{2}. ∎

4. An example

In this section we will give an example to show that in Theorem 1.1 it is indeed necessary to take homology with certain primes inverted. We will take as an example the 6-manifolds VdV_{d} given by a smooth degree dd hypersurface in 4\mathbb{CP}^{4}, which we have studied in detail in [GRW19, Section 5.3]. Any unattributed claims about these manifolds may be found there. We will also consider their stabilisations

Vd,g:=Vd#g(S3×S3)V_{d,g}\mathrel{\mathop{\mathchar 58\relax}}=V_{d}\#g(S^{3}\times S^{3})

obtained by connect-sum of VdV_{d} with gg copies of S3×S3S^{3}\times S^{3}, which contain

g(Vd,g)=g+12(d45d3+10d210d+4)g(V_{d,g})=g+\tfrac{1}{2}(d^{4}-5d^{3}+10d^{2}-10d+4)

copies of S3×S3S^{3}\times S^{3}.

Theorem 4.1.

Let p7p\geq 7 be a prime number, and suppose that g(Vd,g)9g(V_{d,g})\geq 9. Then

H3(or(Vd,g);(p))(p)/gcd(d,g).H^{3}(\mathcal{M}^{\mathrm{or}}(V_{d,g});\mathbb{Z}_{(p)})\cong\mathbb{Z}_{(p)}/\gcd(d,g).

The formula χ=χ(Vd,g)=d(1010d+5d2d3)2g\chi=\chi(V_{d,g})=d(10-10d+5d^{2}-d^{3})-2g implies that gcd(d,g)=gcd(d,χ)\gcd(d,g)=\gcd(d,\chi), so the theorem may also be written

H3(or(Vd,g);(p))/pmin(vp(d),vp(χ)).H^{3}(\mathcal{M}^{\mathrm{or}}(V_{d,g});\mathbb{Z}_{(p)})\cong\mathbb{Z}/p^{\min(v_{p}(d),v_{p}(\chi))}\mathbb{Z}.

Hence the moduli spaces for the oriented stably diffeomorphic manifolds Vd,gV_{d,g} and Vd,gV_{d,g^{\prime}} have isomorphic H3(;(p))H^{3}(-;\mathbb{Z}_{(p)}) if and only if vp(χ(Vd,g))=vp(χ(Vd,g))v_{p}(\chi(V_{d,g}))=v_{p}(\chi(V_{d,g^{\prime}})), provided those pp-adic valuations are at most vp(d)v_{p}(d).

Proof of Theorem 4.1.

In [GRW19, Section 5.3] we computed the \mathbb{Q}-cohomology of or(Vd,g)\mathcal{M}^{\mathrm{or}}(V_{d,g}) in a stable range. We will refer to details of the notation from that discussion, which differs slightly from the notation used earlier in this note.

Firstly, the \mathbb{Q}-cohomology calculation goes through without significant changes for or(Vd,g)\mathcal{M}^{\mathrm{or}}(V_{d,g}), because Vd,gV_{d,g} and VdV_{d} have the same Moore–Postnikov 33-stage, and because any orientation preserving diffeomorphism of Vd,gV_{d,g} must also act trivially on H2(Vd,g;)H^{2}(V_{d,g};\mathbb{Z}). The only difference is that the formula for the d3d_{3}-differential now involves characteristic numbers of Vd,gV_{d,g}, which can be calculated to give

d3(κp2)\displaystyle d_{3}(\kappa_{p_{2}}) =0\displaystyle=0
d3(κp12)\displaystyle d_{3}(\kappa_{p_{1}^{2}}) =0\displaystyle=0
d3(κte)\displaystyle d_{3}(\kappa_{te}) =κe=χ(Vd,g)=d(1010d+5d2d3)2g\displaystyle=\kappa_{e}=\chi(V_{d,g})=d(10-10d+5d^{2}-d^{3})-2g
d3(κt2p1)\displaystyle d_{3}(\kappa_{t^{2}p_{1}}) =2κtp1=2d(5d2)\displaystyle=2\kappa_{tp_{1}}=2d(5-d^{2})
d3(κt4)\displaystyle d_{3}(\kappa_{t^{4}}) =4κt3=4d.\displaystyle=4\kappa_{t_{3}}=4d.

Secondly, the \mathbb{Q}-cohomology calculation yields an analogous (p)\mathbb{Z}_{(p)}-cohomology calculation for large enough primes pp. Specifically the spectrum MTθdMT\theta_{d} is (6)(-6)-connected, so by the Atiyah–Hirzebruch spectral sequence the Hurewicz map

πi(MTθd)(p)Hi(MTθd;(p))Hi+6(Bd;(p))\pi_{i}(MT\theta_{d})_{(p)}\longrightarrow H_{i}(MT\theta_{d};\mathbb{Z}_{(p)})\cong H_{i+6}(B_{d};\mathbb{Z}_{(p)})

is an isomorphism as long as i<2p36i<2p-3-6, so as long as i5i\leq 5 since we have assumed that p7p\geq 7. As pp is odd we have

H(Bd;(p))=H(BSO(6)×K(,2);(p))=(p)[p1,p2,e,t].H^{*}(B_{d};\mathbb{Z}_{(p)})=H^{*}(B\mathrm{SO}(6)\times K(\mathbb{Z},2);\mathbb{Z}_{(p)})=\mathbb{Z}_{(p)}[p_{1},p_{2},e,t].

Thus we have π1(Ω0MTθd)(p)=0\pi_{1}(\Omega^{\infty}_{0}MT\theta_{d})_{(p)}=0, π2(Ω0MTθd)(p)(p)5\pi_{2}(\Omega^{\infty}_{0}MT\theta_{d})_{(p)}\cong\mathbb{Z}_{(p)}^{5} with the isomorphism given by the tautological classes κp2,κp12,κte,κt2p1,κt4\kappa_{p_{2}},\kappa_{p_{1}^{2}},\kappa_{te},\kappa_{t^{2}p_{1}},\kappa_{t^{4}}, and π3(Ω0MTθd)(p)=0\pi_{3}(\Omega^{\infty}_{0}MT\theta_{d})_{(p)}=0. Therefore

Hi(θd(Vd,g,Vd,g);(p))={(p)i=00i=1(p){κp2,κp12,κte,κt2p1,κt4}i=20i=3.H^{i}(\mathcal{M}^{\theta_{d}}(V_{d,g},\ell_{V_{d,g}});\mathbb{Z}_{(p)})=\begin{cases}\mathbb{Z}_{(p)}&i=0\\ 0&i=1\\ \mathbb{Z}_{(p)}\{\kappa_{p_{2}},\kappa_{p_{1}^{2}},\kappa_{te},\kappa_{t^{2}p_{1}},\kappa_{t^{4}}\}&i=2\\ 0&i=3.\end{cases}

The submonoid GhAut(u)G\leq\mathrm{hAut}(u) of those path components which stabilise [Vd,g,Vd,g][V_{d,g},\ell_{V_{d,g}}] is path connected, and as the map u:BdBSO(6)×K(,2)u\mathrel{\mathop{\mathchar 58\relax}}B_{d}\to B\mathrm{SO}(6)\times K(\mathbb{Z},2) is a (p)\mathbb{Z}_{(p)}-homology equivalence, since pp is odd, we also have that πi(G)(p)=0\pi_{i}(G)\otimes\mathbb{Z}_{(p)}=0 for i>0i>0. Thus the map θd(Vd,g,Vd,g)μ(Vd,g,uVd,g)\mathcal{M}^{\theta_{d}}(V_{d,g},\ell_{V_{d,g}})\to\mathcal{M}^{\mu}(V_{d,g},u\circ\ell_{V_{d,g}}) is a (p)\mathbb{Z}_{(p)}-homology equivalence.

It remains to study the Serre spectral sequence for the fibration sequence

μ(Vd,g,uVd,g)or(Vd,g)K(,3),\mathcal{M}^{\mu}(V_{d,g},u\circ\ell_{V_{d,g}})\longrightarrow\mathcal{M}^{\mathrm{or}}(V_{d,g})\longrightarrow K(\mathbb{Z},3),

which in low degrees has a single differential

d3:E30,2=(p){κp2,κp12,κte,κt2p1,κt4}E33,0=H3(K(,3);(p))=(p)d_{3}\mathrel{\mathop{\mathchar 58\relax}}E_{3}^{0,2}=\mathbb{Z}_{(p)}\{\kappa_{p_{2}},\kappa_{p_{1}^{2}},\kappa_{te},\kappa_{t^{2}p_{1}},\kappa_{t^{4}}\}\longrightarrow E_{3}^{3,0}=H^{3}(K(\mathbb{Z},3);\mathbb{Z}_{(p)})=\mathbb{Z}_{(p)}

given by the formula above, so H3(or(Vd,g);(p))H^{3}(\mathcal{M}^{\mathrm{or}}(V_{d,g});\mathbb{Z}_{(p)}) is given by the cokernel of this differential. The claim now follows by the identity of ideals

(4d,2d(5d2),d(1010d+5d2d3)2g)=(d,g)(4d,2d(5-d^{2}),d(10-10d+5d^{2}-d^{3})-2g)=(d,g)

of (p)\mathbb{Z}_{(p)}, using again that pp is odd. ∎

References

  • [BG75] J. C. Becker and D. H. Gottlieb, The transfer map and fiber bundles, Topology 14 (1975), 1–12.
  • [BM14] M. Bendersky and J. Miller, Localization and homological stability of configuration spaces, Q. J. Math. 65 (2014), no. 3, 807–815.
  • [CP15] F. Cantero and M. Palmer, On homological stability for configuration spaces on closed background manifolds, Doc. Math. 20 (2015), 753–805.
  • [Fri17] N. Friedrich, Homological stability of automorphism groups of quadratic modules and manifolds, Doc. Math. 22 (2017), 1729–1774.
  • [GRW14] S. Galatius and O. Randal-Williams, Stable moduli spaces of high-dimensional manifolds, Acta Math. 212 (2014), no. 2, 257–377.
  • [GRW17] by same author, Homological stability for moduli spaces of high dimensional manifolds. II, Ann. of Math. (2) 186 (2017), no. 1, 127–204.
  • [GRW18] by same author, Homological stability for moduli spaces of high dimensional manifolds. I, J. Amer. Math. Soc. 31 (2018), no. 1, 215–264.
  • [GRW19] by same author, Moduli spaces of manifolds: a user’s guide, Handbook of homotopy theory, Chapman & Hall/CRC, CRC Press, Boca Raton, FL, 2019, pp. 445–487.
  • [Har85] J. L. Harer, Stability of the homology of the mapping class groups of orientable surfaces, Ann. of Math. (2) 121 (1985), no. 2, 215–249.
  • [Kra19] M. Krannich, On characteristic classes of exotic manifold bundles, Mathematische Annalen (2019), arXiv:1802.02609.
  • [MT01] I. Madsen and U. Tillmann, The stable mapping class group and Q(+)Q(\mathbb{C}\,\mathbb{P}^{\infty}_{+}), Invent. Math. 145 (2001), no. 3, 509–544.
  • [MW07] I. Madsen and M. Weiss, The stable moduli space of Riemann surfaces: Mumford’s conjecture, Ann. of Math. (2) 165 (2007), no. 3, 843–941.
  • [Til97] U. Tillmann, On the homotopy of the stable mapping class group, Invent. Math. 130 (1997), no. 2, 257–275.