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

Cohomology of automorphism groups of free groups with twisted coefficients

Oscar Randal-Williams o.randal-williams@dpmms.cam.ac.uk Centre for Mathematical Sciences
Wilberforce Road
Cambridge CB3 0WB
UK
(Date: August 13, 2025)
Abstract.

We compute the groups H(Aut(Fn);M)H^{*}(\mathrm{Aut}(F_{n});M) and H(Out(Fn);M)H^{*}(\mathrm{Out}(F_{n});M) in a stable range, where MM is obtained by applying a Schur functor to HH_{\mathbb{Q}} or HH^{*}_{\mathbb{Q}}, respectively the first rational homology and cohomology of FnF_{n}. The answer may be described in terms of stable multiplicities of irreducibles in the plethysm SymkSyml\mathrm{Sym}^{k}\circ\mathrm{Sym}^{l} of symmetric powers.

We also compute the stable integral cohomology groups of Aut(Fn)\mathrm{Aut}(F_{n}) with coefficients in HH or HH^{*}.

Key words and phrases:
Automorphisms of free groups, homology stability
2010 Mathematics Subject Classification:
20F28, 20J06, 57R20
Supported by ERC Advanced Grant No. 228082, the Danish National Research Foundation through the Centre for Symmetry and Deformation, and EPSRC grant EP/M027783/1.

1. Statement of results

Galatius [12] has proved the remarkable theorem that the natural homomorphisms

ΣnAut(Fn)Out(Fn)\Sigma_{n}\longrightarrow\mathrm{Aut}(F_{n})\longrightarrow\mathrm{Out}(F_{n})

both induce homology isomorphisms in degrees 2n32*\leq n-3 with integral coefficients. His approach is to model BOut(Fn)B\mathrm{Out}(F_{n}) as the space 𝒢n\mathcal{G}_{n} of graphs of the homotopy type of nS1\vee^{n}S^{1}, and BAut(Fn)B\mathrm{Aut}(F_{n}) as the space 𝒢n1\mathcal{G}^{1}_{n} of pointed graphs of the same homotopy type. He then produces a natural map from such spaces of graphs to the infinite loop space Q0(S0)Q_{0}(S^{0}), which he shows has a certain homological connectivity. One consequence of this is that Hi(Aut(Fn);)=0H^{i}(\mathrm{Aut}(F_{n});\mathbb{Q})=0 for 0<in320<i\leq\tfrac{n-3}{2}.

At the same time, Satoh [26, 27] has studied the low dimensional (co)homology of Aut(Fn)\mathrm{Aut}(F_{n}) and Out(Fn)\mathrm{Out}(F_{n}) with coefficients in the module H:=H1(Fn;)H\mathrel{\mathop{\mathchar 12346\relax}}=H_{1}(F_{n};\mathbb{Z}) given by the abelianisation of FnF_{n}, and in the dual module H:=H1(Fn;)H^{*}\mathrel{\mathop{\mathchar 12346\relax}}=H^{1}(F_{n};\mathbb{Z}). His methods are those of combinatorial group theory, and proceed by calculation with a presentation of these groups.

Our goal is to show that the stable cohomology of Aut(Fn)\mathrm{Aut}(F_{n}) and Out(Fn)\mathrm{Out}(F_{n}) with twisted coefficients may also be approached with the geometric techniques used by Galatius, along with a little representation theory. The method we shall introduce is quite general and may be applied whenever a suitable Madsen–Weiss-type theorem has been proved; in Appendix B we will show how to use it to recover a theorem of Looijenga [20] on the stable cohomology of mapping class groups with twisted coefficients.

For Aut(Fn)\mathrm{Aut}(F_{n}) and Out(Fn)\mathrm{Out}(F_{n}) we will consider cohomology with coefficients in the modules H:=H1(Fn;)H^{*}_{\mathbb{Q}}\mathrel{\mathop{\mathchar 12346\relax}}=H^{1}(F_{n};\mathbb{Q}) and H:=H1(Fn;)H_{\mathbb{Q}}\mathrel{\mathop{\mathchar 12346\relax}}=H_{1}(F_{n};\mathbb{Q}), and more generally with coefficients in Sλ(H)S_{\lambda}(H^{*}_{\mathbb{Q}}) and Sλ(H)S_{\lambda}(H_{\mathbb{Q}}), where Sλ()S_{\lambda}(-) is the Schur functor associated to a partition λq\lambda\vdash q, which we think of as being given by a Young diagram. To define this, recall that to such a partition there is an associated irreducible [Σq]\mathbb{Q}[\Sigma_{q}]-module SλS^{\lambda}, the Specht module. For a \mathbb{Q}-vector space VV we may consider VqV^{\otimes q} as a [Σq]\mathbb{Q}[\Sigma_{q}]-module by permuting the factors, and we may hence form Sλ(V):=Hom[Σq](Sλ,Vq)S_{\lambda}(V)\mathrel{\mathop{\mathchar 12346\relax}}=\mathrm{Hom}_{\mathbb{Q}[\Sigma_{q}]}(S^{\lambda},V^{\otimes q}). This construction defines the Schur functor SλS_{\lambda}. It is a basic result that Sλ(V)S_{\lambda}(V) is an irreducible representation of GL(V)GL(V).

In fact, our result is best expressed as calculating H(Aut(Fn);Hq)H^{*}(\mathrm{Aut}(F_{n});H_{\mathbb{Q}}^{\otimes q}) as a [Σq]\mathbb{Q}[\Sigma_{q}]-module. The result for Sλ(H)S_{\lambda}(H_{\mathbb{Q}}) may then be extracted as

H(Aut(Fn);Sλ(H))=HomΣq(Sλ,H(Aut(Fn);Hq)).H^{*}(\mathrm{Aut}(F_{n});S_{\lambda}(H_{\mathbb{Q}}))=\mathrm{Hom}_{\Sigma_{q}}(S^{\lambda},H^{*}(\mathrm{Aut}(F_{n});H_{\mathbb{Q}}^{\otimes q})).

We write \mathbb{Q}^{-} for the sign representation of Σq\Sigma_{q}.

Theorem A.

  1. (i)

    H(Aut(Fn);(H)q)=0H^{*}(\mathrm{Aut}(F_{n});(H^{*})^{\otimes q})=0 for 2nq32*\leq n-q-3.

  2. (ii)

    H(Aut(Fn);Hq)=0H^{*}(\mathrm{Aut}(F_{n});H_{\mathbb{Q}}^{\otimes q})=0 for 2nq32*\leq n-q-3 if q*\neq q, and Hq(Aut(Fn);Hq)H^{q}(\mathrm{Aut}(F_{n});H_{\mathbb{Q}}^{\otimes q})\otimes\mathbb{Q}^{-} is the permutation module on the set of partitions of {1,2,,q}\{1,2,\ldots,q\}.

Theorem A (i) may also be deduced from work of Djament–Vespa [9]. We believe that Theorem A (ii) may also be obtained by combining work of Djament [8] and Vespa [28].

However, more important than these particular results is our technique, which is of very general applicability. For example, it can easily be modified to obtain results for Out(Fn)\mathrm{Out}(F_{n}).

Theorem B.

  1. (i)

    H(Out(Fn);(H)q)=0H^{*}(\mathrm{Out}(F_{n});(H^{*}_{\mathbb{Q}})^{\otimes q})=0 for 2nq32*\leq n-q-3.

  2. (ii)

    H(Out(Fn);Hq)=0H^{*}(\mathrm{Out}(F_{n});H_{\mathbb{Q}}^{\otimes q})=0 for 2nq32*\leq n-q-3 if q*\neq q, and as long as n4q+3n\geq 4q+3 Hq(Out(Fn);Hq)H^{q}(\mathrm{Out}(F_{n});H_{\mathbb{Q}}^{\otimes q})\otimes\mathbb{Q}^{-} is the permutation module on the set of partitions of {1,2,,q}\{1,2,\ldots,q\} having no parts of size 1.

We give tables listing the dimensions of the groups H|λ|(Aut(Fn);Sλ(H))H^{|\lambda|}(\mathrm{Aut}(F_{n});S_{\lambda}(H_{\mathbb{Q}})) and H|λ|(Out(Fn);Sλ(H))H^{|\lambda|}(\mathrm{Out}(F_{n});S_{\lambda}(H_{\mathbb{Q}})) for |λ|6|\lambda|\leq 6 in Appendix C.

Each partition of {1,2,,q}\{1,2,\ldots,q\} may be expressed as a partition of a smaller set with no parts of size 1 along with the set of parts of size 1, which translates to the expression

Hq(Aut(Fn);Hq)qIndΣ×ΣqΣq(H(Out(Fn);H))H^{q}(\mathrm{Aut}(F_{n});H_{\mathbb{Q}}^{\otimes q})\cong\bigoplus_{\ell\leq q}\mathrm{Ind}_{\Sigma_{\ell}\times\Sigma_{q-\ell}}^{\Sigma_{q}}(H^{\ell}(\mathrm{Out}(F_{n});H_{\mathbb{Q}}^{\otimes\ell})\otimes\mathbb{Q}^{-})

as long as n4q+3n\geq 4q+3. Applying HomΣq(Sλ,)\mathrm{Hom}_{\Sigma_{q}}(S^{\lambda},-) and using the Pieri rule gives the pleasant formula

(1.1) H|λ|(Aut(Fn);Sλ(H))μρ(λ)H|μ|(Out(Fn);Sμ(H))H^{|\lambda|}(\mathrm{Aut}(F_{n});S_{\lambda}(H_{\mathbb{Q}}))\cong\bigoplus_{\mu\in\rho(\lambda)}H^{|\mu|}(\mathrm{Out}(F_{n});S_{\mu}(H_{\mathbb{Q}}))

as long as n4|λ|+3n\geq 4|\lambda|+3, where ρ(λ)\rho(\lambda) denotes the set of Young diagrams which may be obtained from λ\lambda by removing at most one box from each row.

1.1. Stable plethysm of symmetric powers

For a Young diagram μ\mu and integers kk and ll such that 2|μ|kl2|\mu|\leq kl, let (kl|μ|,μ)(kl-|\mu|,\mu) be the Young diagram obtained by adding a row of length kl|μ|kl-|\mu| to the top of μ\mu, and let νk,l(μ)\nu^{k,l}(\mu) denote the multiplicity of the irreducible GL(V)GL(V)-representation S(kl|μ|,μ)(V)S_{(kl-|\mu|,\mu)}(V) in Symk(Syml(V))\mathrm{Sym}^{k}(\mathrm{Sym}^{l}(V)). Manivel has shown [21] that the numbers νk,l(μ)\nu^{k,l}(\mu) are increasing and eventually constant functions of both kk and ll, and we write ν(μ)\nu^{\infty}(\mu) for the stable value. This stable value is attained as soon as lμ1l\geq\mu_{1} and k|μ|k\geq|\mu|. Using the work of Manivel, we are able to relate these stable multiplicities directly to the cohomology of Out(Fn)\mathrm{Out}(F_{n}), as follows.

Theorem C.

As long as n4|λ|+3n\geq 4|\lambda|+3 we have

dimH|λ|(Out(Fn);Sλ(H))=ν(λ)\dim_{\mathbb{Q}}H^{|\lambda|}(\mathrm{Out}(F_{n});S_{\lambda}(H_{\mathbb{Q}}))=\nu^{\infty}(\lambda^{\prime})

and as long as n2|λ|+3n\geq 2|\lambda|+3 we have

dimH|λ|(Aut(Fn);Sλ(H))=μρ(λ)ν(μ).\dim_{\mathbb{Q}}H^{|\lambda|}(\mathrm{Aut}(F_{n});S_{\lambda}(H_{\mathbb{Q}}))=\sum_{\mu\in\rho(\lambda)}\nu^{\infty}(\mu^{\prime}).

It follows directly from [21, Proposition 4.4.1] that H|λ|(Out(F);Sλ(H))=0H^{|\lambda|}(\mathrm{Out}(F_{\infty});S_{\lambda}(H_{\mathbb{Q}}))=0 if 2λ1>|λ|2\lambda_{1}>|\lambda|, or if 2λ1=|λ|2\lambda_{1}=|\lambda| and λ\lambda does not consist of two rows of equal length.

1.2. Symmetric and exterior powers

In order to demonstrate how our theorems may be used, we compute the dimension of the associated cohomology groups for the modules q(H)\wedge^{q}(H_{\mathbb{Q}}) and Symq(H)\mathrm{Sym}^{q}(H_{\mathbb{Q}}).

Corollary D.

  1. (i)

    For n2q+3n\geq 2q+3, Hq(Aut(Fn);q(H))H^{q}(\mathrm{Aut}(F_{n});\wedge^{q}(H_{\mathbb{Q}})) has dimension given by the number of partitions of qq; for n4q+3n\geq 4q+3, Hq(Out(Fn);q(H))H^{q}(\mathrm{Out}(F_{n});\wedge^{q}(H_{\mathbb{Q}})) has dimension given by the number of partitions of qq into pieces none of which are 1.

  2. (ii)

    Let q2q\geq 2. For n2i+q+3n\geq 2i+q+3,

    Hi(Aut(Fn);Symq(H))=Hi(Out(Fn);Symq(H))=0.H^{i}(\mathrm{Aut}(F_{n});\mathrm{Sym}^{q}(H_{\mathbb{Q}}))=H^{i}(\mathrm{Out}(F_{n});\mathrm{Sym}^{q}(H_{\mathbb{Q}}))=0.
Proof.

The dimension of Hq(Aut(Fn);q(H))H^{q}(\mathrm{Aut}(F_{n});\wedge^{q}(H_{\mathbb{Q}})) is the multiplicity of the sign representation in Hq(Aut(Fn);Hq)H^{q}(\mathrm{Aut}(F_{n});H_{\mathbb{Q}}^{\otimes q}), which by Theorem A is the multiplicity of the trivial representation in the permutation module for the set of partitions of {1,2,,q}\{1,2,\ldots,q\}: this is the number of partitions of qq.

The dimension of Hq(Aut(Fn);Symq(H))H^{q}(\mathrm{Aut}(F_{n});\mathrm{Sym}^{q}(H_{\mathbb{Q}})) is the multiplicity of the trivial representation in Hq(Aut(Fn);Hq)H^{q}(\mathrm{Aut}(F_{n});H_{\mathbb{Q}}^{\otimes q}), which by Theorem A is the multiplicity of the sign representation in the permutation module for the set of partitions of {1,2,,q}\{1,2,\ldots,q\}. This may be computed via inner product of characters as 1q!\tfrac{1}{q!} times

σΣq#{partitions P such that σ(P)=P}sign(σ)=partitionsPσΣqσ(P)=Psign(σ),\sum_{\sigma\in\Sigma_{q}}\#\{\text{partitions $P$ such that $\sigma(P)=P$}\}\cdot\mathrm{sign}(\sigma)=\sum_{\begin{subarray}{c}\text{partitions}\\ P\end{subarray}}\sum_{\begin{subarray}{c}\sigma\in\Sigma_{q}\\ \sigma(P)=P\end{subarray}}\mathrm{sign}(\sigma),

but, writing ΣPΣq\Sigma_{P}\leq\Sigma_{q} for the stabiliser of a partition PP, we have that σΣPsign(σ)\sum_{\sigma\in\Sigma_{P}}\mathrm{sign}(\sigma) is |ΣP||\Sigma_{P}| times the multiplicity of the sign representation in the trivial representation of ΣP\Sigma_{P}, which is zero (as ΣP\Sigma_{P} always contains at least one transposition if q2q\geq 2).

The arguments for Out(Fn)\mathrm{Out}(F_{n}) are identical. ∎

1.3. Integral and torsion results

Finally, our technique can be made to give integral and local information as well. It is not hard to show that Hi(Aut(Fn);H)=0H^{i}(\mathrm{Aut}(F_{n});H^{*})=0 for 2in42i\leq n-4 (see Proposition 2.3), but we also have the following.

Theorem E.

  1. (i)

    H(Aut(F);H)H^{*}(\mathrm{Aut}(F_{\infty});H) is a free H(Aut(F);)H^{*}(\mathrm{Aut}(F_{\infty});\mathbb{Z})-module (on a single generator in degree 1).

  2. (ii)

    For a partition λq\lambda\vdash q and a prime number p>qp>q, H(Aut(F);Sλ(H(p)))H^{*}(\mathrm{Aut}(F_{\infty});S_{\lambda}(H\otimes\mathbb{Z}_{(p)})) is a free H(Aut(F);(p))H^{*}(\mathrm{Aut}(F_{\infty});\mathbb{Z}_{(p)})-module (on generators in degree qq, the number of which may be deduced from Theorem A).

That Theorem E (ii) might hold was suggested to the author by Aurélien Djament upon hearing of Theorem E (i). We refer to Section 6 for a description of what we mean by the Schur functor SλS_{\lambda} in the pp-local setting.

One may deduce from Theorem E (i) that Hi(Out(Fn);H[1n1])=0H^{i}(\mathrm{Out}(F_{n});H\otimes\mathbb{Z}[\tfrac{1}{n-1}])=0 for 2in42i\leq n-4. In Appendix A we describe how the calculation H1(Out(Fn);H)/(n1)H^{1}(\mathrm{Out}(F_{n});H)\cong\mathbb{Z}/(n-1) for n9n\geq 9 follows from a reasonable-seeming conjecture about spaces of graphs.

Acknowledgements

This paper is an update to a 2010 preprint [23] in which I proved, inter alia, Corollary D subject to a sequence of conjectures. Since that time, in joint work with Nathalie Wahl [25] we proved Conjecture A (and I explain in this paper how a version of Conjecture B follows from it), and in 2014 Søren Galatius explained to me a proof of Conjecture C. Thus these conjectural calculations from my 2010 preprint hold.

Recent work of Aurélien Djament [8] and Christine Vespa [28] obtains these calculations by very different means. I was motivated by their results to revisit these techniques and to clarify the status of the conjectures from my 2010 preprint. I would like to thank all of the above named for their interest in, and useful comments on, the content of this note.

2. An observation regarding homology stability

The observations of this section are no doubt known to some experts. The groups Aut(Fn)\mathrm{Aut}(F_{n}) and Out(Fn)\mathrm{Out}(F_{n}) fit into a more general family of groups denoted Γn,s\Gamma_{n,s} by Hatcher–Vogtmann [16], where Aut(Fn)=Γn,1\mathrm{Aut}(F_{n})=\Gamma_{n,1} and Out(Fn)=Γn,0\mathrm{Out}(F_{n})=\Gamma_{n,0}. Classifying spaces for these may be taken to be the spaces 𝒢ns\mathcal{G}_{n}^{s} of graphs of the homotopy type of nS1\vee^{n}S^{1} equipped with ss distinct ordered marked points.

Hatcher and Vogtmann [16] prove that the map 𝒢ns𝒢n+1s\mathcal{G}_{n}^{s}\to\mathcal{G}_{n+1}^{s} (defined for s>0s>0) that adds a loop at a particular marked point induces an integral homology isomorphism in degrees 2n22*\leq n-2 (and induces a rational homology isomorphism in degrees 54n105*\leq 4n-10). Furthermore, the map 𝒢ns𝒢ns1\mathcal{G}_{n}^{s}\to\mathcal{G}_{n}^{s-1} that forgets a marked point is an integral homology isomorphism in degrees 2n32*\leq n-3 (or 2n42*\leq n-4 if it is the last marked point).

We can make an immediate observation regarding cohomology with coefficients in HH^{*} from this homology stability result. There is an extension FnAut(Fn)Out(Fn)F_{n}\to\mathrm{Aut}(F_{n})\to\mathrm{Out}(F_{n}), and the corresponding Leray–Hochschild–Serre spectral sequence has two rows. However, as the projection is a homology equivalence in a range of degrees we deduce

Proposition 2.1.

The groups H(Out(Fn);H)H^{*}(\mathrm{Out}(F_{n});H^{*}) are zero for 2n62*\leq n-6.

Similarly, there is a fibration with section nS1𝒢n2𝒢n1\vee^{n}S^{1}\to\mathcal{G}_{n}^{2}\to\mathcal{G}_{n}^{1} and the projection map is a homology equivalence in a range of degrees, so

Proposition 2.2.

The groups H(Aut(Fn);H)H^{*}(\mathrm{Aut}(F_{n});H^{*}) are zero for 2n42*\leq n-4.

More generally, the map 𝒢nk+1𝒢n1\mathcal{G}_{n}^{k+1}\to\mathcal{G}_{n}^{1} has fibre (nS1)k(\vee^{n}S^{1})^{k} and Aut(Fn)\mathrm{Aut}(F_{n}) acts on its homology diagonally. Thus for k=2k=2 the Serre spectral sequence has three rows, with

E2,1=H(Aut(Fn);HH)=0for 2n4 by Proposition 2.2E_{2}^{*,1}=H^{*}(\mathrm{Aut}(F_{n});H^{*}\oplus H^{*})=0\quad\text{for $2*\leq n-4$ by Proposition \ref{prop:AutFnHomologyRep}}

and

E2,2=H(Aut(Fn);HH).E_{2}^{*,2}=H^{*}(\mathrm{Aut}(F_{n});H^{*}\otimes H^{*}).

Using the fact that the projection map is a homology equivalence in degrees 2n22*\leq n-2, we deduce that H(Aut(Fn);HH)=0H^{*}(\mathrm{Aut}(F_{n});H^{*}\otimes H^{*})=0 for 2n82*\leq n-8. Continuing in this way for higher kk, we establish the following proposition.

Proposition 2.3.

For all q1q\geq 1, H(Aut(Fn);(H)q)H^{*}(\mathrm{Aut}(F_{n});(H^{*})^{\otimes q}) is zero for 2n4q2*\leq n-4q.

3. Homology stability with coefficient systems

By Galatius’ theorem, the groups Aut(Fn)\mathrm{Aut}(F_{n}) are closely related to the symmetric groups, but also share many properties with mapping class groups of surfaces. These three families of groups are known to exhibit homological stability for integral homology, but symmetric groups and mapping class groups also exhibit homological stability for certain systems of coefficients, those of “finite degree”, a notion that is originally due to Dwyer [10] in his study of homological stability for general linear groups with coefficient systems.

This notion of degree may be formalised, in the context of free groups, as follows.

Definition 3.1.

Let 𝔊𝔯\mathfrak{Gr} denote the category whose objects are the finitely generated free groups, and where a morphism from GG to HH is given by a pair

(f:GH,XH)(f\mathrel{\mathop{\mathchar 12346\relax}}G\to H,X\leq H)

consisting of an injective group homomorphism ff and a finitely-generated (free) subgroup XHX\leq H such that H=f(G)XH=f(G)*X.

A coefficient system is a covariant functor V:𝔊𝔯𝐀𝐛V\mathrel{\mathop{\mathchar 12346\relax}}\mathfrak{Gr}\to\mathbf{Ab} to the category of abelian groups. We declare the constant functors to be polynomial of degree 0, and more generally we declare a functor VV to be polynomial of degree k\leq k if

  1. (i)

    VV sends the the canonical morphism sG:=(inc:GG,)s_{G}\mathrel{\mathop{\mathchar 12346\relax}}=(\mathrm{inc}\mathrel{\mathop{\mathchar 12346\relax}}G\to G*\mathbb{Z},\mathbb{Z}) to an injection, and

  2. (ii)

    the new coefficient system GCoker(sG:V(G)V(G))G\mapsto\mathrm{Coker}(s_{G}\mathrel{\mathop{\mathchar 12346\relax}}V(G)\to V(G*\mathbb{Z})) is polynomial of degree k1\leq k-1.

In a 2010 preprint [23], we conjectured (based on the analogy with general linear groups [10] and mapping class groups [17]) that the groups H(Aut(Fn);V(Fn))H_{*}(\mathrm{Aut}(F_{n});V(F_{n})) should exhibit homological stability in degrees 2nk22*\leq n-k-2 when VV is a polynomial coefficient system of degree k\leq k. Since then, in joint work with Wahl [25] we have established a quite general homological stability theorem with polynomial coefficients, and using the highly-connected simplicial complexes of [15] it applies in this case. The result obtained is as follows.

Theorem 3.2 (Randal-Williams–Wahl [25]).

If VV is a polynomial coefficient system of degree k\leq k, then the natural map

H(Aut(Fn);V(Fn))H(Aut(Fn+1);V(Fn+1))H_{*}(\mathrm{Aut}(F_{n});V(F_{n}))\longrightarrow H_{*}(\mathrm{Aut}(F_{n+1});V(F_{n+1}))

induces an epimorphism for 2nk12*\leq n-k-1 and an isomorphism for 2nk32*\leq n-k-3.

One advantage of the category 𝔊𝔯\mathfrak{Gr} over the more naïve category 𝔤𝔯\mathfrak{gr} of finitely generated free groups and injective homomorphisms is that there are more functors out of it.

Definition 3.3.

  1. (i)

    Let H:𝔊𝔯𝐀𝐛H\mathrel{\mathop{\mathchar 12346\relax}}\mathfrak{Gr}\to\mathbf{Ab} be the coefficient system sending GG to H1(G;)=GabH_{1}(G;\mathbb{Z})=G^{ab}, and sending a morphism (f:GH,X)(f\mathrel{\mathop{\mathchar 12346\relax}}G\to H,X) to ff_{*}. There is an exact sequence

    0H(G)H(sG)H(G)00\longrightarrow H(G)\overset{H(s_{G})}{\longrightarrow}H(G*\mathbb{Z})\longrightarrow\mathbb{Z}\longrightarrow 0

    so this coefficient system is polynomial of degree 1.

  2. (ii)

    Let H:𝔊𝔯𝐀𝐛H^{*}\mathrel{\mathop{\mathchar 12346\relax}}\mathfrak{Gr}\to\mathbf{Ab} be the coefficient system sending GG to H1(G;)=Hom(Gab,)H^{1}(G;\mathbb{Z})=\mathrm{Hom}(G^{ab},\mathbb{Z}) , and sending a morphism (f:GH,X)(f\mathrel{\mathop{\mathchar 12346\relax}}G\to H,X) to the linear dual of

    H1(H;)=H1(f(G)X;)H1(f(G);)H1(G;).H_{1}(H;\mathbb{Z})=H_{1}(f(G)*X;\mathbb{Z})\longrightarrow H_{1}(f(G);\mathbb{Z})\cong H_{1}(G;\mathbb{Z}).

    (Note that this does not define a functor on 𝔤𝔯\mathfrak{gr}!) There is an exact sequence

    0H(G)H(sG)H(G)00\longrightarrow H^{*}(G)\overset{H^{*}(s_{G})}{\longrightarrow}H^{*}(G*\mathbb{Z})\longrightarrow\mathbb{Z}\longrightarrow 0

    so this coefficient system is polynomial of degree 1.

Remark 3.4.

Not only can more coefficient systems be defined on 𝔊𝔯\mathfrak{Gr} than on 𝔤𝔯\mathfrak{gr}, but it follows from recent work of Djament–Vespa [9] that the most homologically interesting functors must be defined here: they show that if VV is a polynomial coefficient system which is reduced (i.e. V({e})=0V(\{e\})=0) and factors through 𝔤𝔯\mathfrak{gr}, then

colimnH(Aut(Fn);V(Fn))=0.\underset{n\to\infty}{\mathrm{colim\,}}H_{*}(\mathrm{Aut}(F_{n});V(F_{n}))=0.

The coefficient system HH^{*} is reduced, but does not factor through 𝔤𝔯\mathfrak{gr}: indeed, Satoh [26] has shown that colimnH1(Aut(Fn);H)=\underset{n\to\infty}{\mathrm{colim\,}}H_{1}(\mathrm{Aut}(F_{n});H^{*})=\mathbb{Z}, and we will show how to recover (the dual version of) this in Section 6.

There are some easy consequences of the definition of polynomiality which allow us to compute or estimate degrees of coefficient systems. Firstly, a summand of a polynomial functor of degree k\leq k is again polynomial of degree k\leq k, and an extension of two polynomial functors of degree k\leq k is again polynomial of degree k\leq k. Secondly, if VV and WW are coefficient systems which are polynomial of degree kk and \ell respectively, and if they both take values in flat \mathbb{Z}-modules, then VWV\otimes_{\mathbb{Z}}W is polynomial of degree (k+)\leq(k+\ell). More generally, if 𝔽\mathbb{F} is a field and V,W:𝔊𝔯𝔽-mod𝐀𝐛V,W\mathrel{\mathop{\mathchar 12346\relax}}\mathfrak{Gr}\to\mathbb{F}\text{-mod}\to\mathbf{Ab} are coefficient systems which are polynomial of degree kk and \ell respectively and factor through the category of 𝔽\mathbb{F}-modules, then V𝔽WV\otimes_{\mathbb{F}}W is polynomial of degree (k+)\leq(k+\ell).

We will often be interested in cohomology rather than homology, for which we will use the following result of Universal Coefficient-type, which is surely standard but for which we could not find a reference.

Lemma 3.5.

Let GG be a group, RR be a PID, and MM a left R[G]R[G]-module, and write M=HomR(M,R)M^{*}=\mathrm{Hom}_{R}(M,R), a right R[G]R[G]-module. There is a natural short exact sequence

0ExtR1(Hi1(G;M),R)Hi(G;M)HomR(Hi(G;M),R)0.0\longrightarrow\mathrm{Ext}_{R}^{1}(H_{i-1}(G;M),R)\longrightarrow H^{i}(G;M^{*})\longrightarrow\mathrm{Hom}_{R}(H_{i}(G;M),R)\longrightarrow 0.
Proof.

Analogous to [3, Proposition 7.1]. Let PRP_{\bullet}\to R be a projective right R[G]R[G]-module resolution, so C:=HomR[G](P,M)C^{\bullet}\mathrel{\mathop{\mathchar 12346\relax}}=\mathrm{Hom}_{R[G]}(P_{\bullet},M^{*}) is a cochain complex of RR-modules which computes H(G;M)H^{*}(G;M^{*}). Writing

C=HomR(P,M)GHomR(PRM,R)GHomR(PR[G]M,R)C^{\bullet}=\mathrm{Hom}_{R}(P_{\bullet},M^{*})^{G}\cong\mathrm{Hom}_{R}(P_{\bullet}\otimes_{R}M,R)^{G}\cong\mathrm{Hom}_{R}(P_{\bullet}\otimes_{R[G]}M,R)

and applying the Universal Coefficient Theorem (e.g. [4, p. 114]), using the fact that a submodule of a free module over a PID is free, gives the desired sequence. ∎

Using the stability theorem and Lemma 3.5, we may improve Proposition 2.3 to the following.

Corollary 3.6.

For all q1q\geq 1, H(Aut(Fn);(H)q)=0H^{*}(\mathrm{Aut}(F_{n});(H^{*})^{\otimes q})=0 for 2nq32*\leq n-q-3.

This proves Theorem A (i). We remark that Theorem A (i) also follows from [9, Théorème 1] and Lemma 3.5, because the functor Hq:𝔊𝔯𝔤𝔯𝐀𝐛H^{\otimes q}\mathrel{\mathop{\mathchar 12346\relax}}\mathfrak{Gr}\to\mathfrak{gr}\to\mathbf{Ab} is polynomial and vanishes on the trivial group (for q1q\geq 1), so by that theorem H(Aut(Fn);Hq)H_{*}(\mathrm{Aut}(F_{n});H^{\otimes q}) vanishes in the stable range.

Finally, we record the stability range for cohomology with coefficients in Sλ(H)S_{\lambda}(H_{\mathbb{Q}}) with λq\lambda\vdash q. This follows from Sλ(H)Sλ(H)S_{\lambda}(H_{\mathbb{Q}})\cong S_{\lambda}(H_{\mathbb{Q}}^{*})^{*}, that Sλ(H)S_{\lambda}(H_{\mathbb{Q}}^{*}) is polynomial of degree q\leq q as it is a summand of (H)q(H_{\mathbb{Q}}^{*})^{\otimes q}, the stability theorem, and Lemma 3.5.

Corollary 3.7.

The groups Hi(Aut(Fn);Sλ(H))H^{i}(\mathrm{Aut}(F_{n});S_{\lambda}(H_{\mathbb{Q}})) are independent of nn for 2inq32i\leq n-q-3.

4. Graphs labeled by a space XX

Let (X,x0)(X,x_{0}) be a based space. Let us write 𝒢X\mathcal{G}^{X} for the topological category whose objects are finite sets inside \mathbb{R}^{\infty}, and whose morphisms from SS\subset\mathbb{R}^{\infty} to TT\subset\mathbb{R}^{\infty} consist of a real number t>0t>0 and a graph Γ[0,t]×\Gamma\subset[0,t]\times\mathbb{R}^{\infty} with leaves {0}×S\{0\}\times S and {t}×T\{t\}\times T, equipped with a continuous map f:(Γ,ST)(X,x0)f\mathrel{\mathop{\mathchar 12346\relax}}(\Gamma,S\cup T)\to(X,x_{0}). This may be given a topology following [12]. Choosing once and for all an embedding {1,2,,s}\{1,2,\ldots,s\}\subset\mathbb{R}^{\infty} the space 𝒢ns(X)\mathcal{G}_{n}^{s}(X) may be defined as the subspace of 𝒢X(,{1,2,,s})\mathcal{G}^{X}(\emptyset,\{1,2,\ldots,s\}) consisting of those graphs which are connected and homotopy equivalent to nS1\vee^{n}S^{1}.

Let 𝒢ns(X)𝒢ns(X)×+1\mathcal{G}_{n}^{s}(X)_{*}\subset\mathcal{G}_{n}^{s}(X)\times\mathbb{R}^{\infty+1} be the subspace given by tuples (t,Γ,f;p)(t,\Gamma,f;p) where pΓp\in\Gamma. Forgetting the point pp gives a map

π:𝒢ns(X)𝒢ns(X)\pi\mathrel{\mathop{\mathchar 12346\relax}}\mathcal{G}_{n}^{s}(X)_{*}\longrightarrow\mathcal{G}_{n}^{s}(X)

which is a fibration with fibre over the point (t,Γ,f)𝒢ns(X)(t,\Gamma,f)\in\mathcal{G}_{n}^{s}(X) given by the graph Γ\Gamma. Furthermore, sending (t,Γ,f,p)(t,\Gamma,f,p) to f(p)Xf(p)\in X defines a map

e:𝒢ns(X)X.e\mathrel{\mathop{\mathchar 12346\relax}}\mathcal{G}_{n}^{s}(X)_{*}\longrightarrow X.

As the map π\pi is a fibration with homotopy-finite fibres, it admits a Becker–Gottlieb transfer map trfπ:Σ+𝒢ns(X)Σ+𝒢ns(X)\mathrm{trf}_{\pi}\mathrel{\mathop{\mathchar 12346\relax}}\Sigma^{\infty}_{+}\mathcal{G}_{n}^{s}(X)\to\Sigma^{\infty}_{+}\mathcal{G}_{n}^{s}(X)_{*}. Composing this with Σ+e\Sigma^{\infty}_{+}e and taking the adjoint gives a map

τns:𝒢ns(X)Q1n(X+)\tau_{n}^{s}\mathrel{\mathop{\mathchar 12346\relax}}\mathcal{G}_{n}^{s}(X)\longrightarrow Q_{1-n}(X_{+})

to the free infinite loop space on XX, landing in the path component indexed by 1n1-n, the Euler characteristic of nS1\vee^{n}S^{1}. This map is natural in XX, and factors through the map τn0:𝒢n(X)Q1n(X+)\tau^{0}_{n}\mathrel{\mathop{\mathchar 12346\relax}}\mathcal{G}_{n}(X)\to Q_{1-n}(X_{+}) up to homotopy.

Theorem 4.1.

The induced map

τ1:hocolimn𝒢n1(X)Q0(X+)\tau_{\infty}^{1}\mathrel{\mathop{\mathchar 12346\relax}}\underset{n\to\infty}{\mathrm{hocolim\,}}\mathcal{G}^{1}_{n}(X)\longrightarrow Q_{0}(X_{+})

is an integral homology equivalence as long as XX is path-connected.

The point of this theorem is that the right-hand side may be computed: the rational cohomology of Q1n(X+)Q_{1-n}(X_{+}) may be described compactly as S(H~(X;))S^{*}(\widetilde{H}^{*}(X;\mathbb{Q})), the free graded-commutative algebra on the reduced cohomology of XX.

Proof sketch.

Following [12], and keeping track of the role that the maps to XX play, one may show that B𝒢XQ(S1X+)B\mathcal{G}^{X}\simeq Q(S^{1}\wedge X_{+}). Furthermore, under this equivalence the composition

𝒢n(X)𝒢X(,)ΩB𝒢XQ(X+)\mathcal{G}_{n}(X)\subset\mathcal{G}^{X}(\emptyset,\emptyset)\longrightarrow\Omega B\mathcal{G}^{X}\simeq Q(X_{+})

is weakly homotopic to τn0\tau^{0}_{n} by the analogue (with maps to XX) of the discussion in Section 5.3 of [12].

Now we claim that the natural map

hocolimn𝒢n1(X)Ω[,{0}]B𝒢X\underset{n\to\infty}{\mathrm{hocolim\,}}\mathcal{G}^{1}_{n}(X)\longrightarrow\Omega_{[\emptyset,\{0\}]}B\mathcal{G}^{X}

is an integral homology equivalence as long as XX is path-connected. This may be proved following [12] and also using ideas from [13]. The crucial point is to consider the category 𝒢X\mathcal{G}^{X}_{\bullet} where objects SS contain the origin 00\in\mathbb{R}^{\infty} and morphisms Γ\Gamma contain the interval [0,t]×{0}[0,t]\times\{0\} on which the function ff is constant. The inclusion 𝒢X𝒢X\mathcal{G}^{X}_{\bullet}\to\mathcal{G}^{X} may be seen to induce an equivalence on classifying spaces as in [13, Lemma 4.6]. One then considers the subcategory 𝒢,cX𝒢X\mathcal{G}^{X}_{\bullet,c}\subset\mathcal{G}^{X}_{\bullet} having only object {0}\{0\} and in which the morphisms are required to be connected. We claim that this inclusion induces an equivalence on classifying spaces. To make morphisms connected use a move similar to that of [12, Lemma 4.24] which connects an arbitrary path component to the standard stick ×{0}×\mathbb{R}\times\{0\}\subset\mathbb{R}\times\mathbb{R}^{\infty} (this requires XX to be path-connected). To reduce to a single object use a move similar to that of [13, Section 4] to make objects consist of a single point, then isotope them into standard position as in [13, Proposition 4.26].

The category 𝒢,cX\mathcal{G}^{X}_{\bullet,c} is therefore a monoid, and, as we can slide edges along the standard interval [0,t]×{0}[0,t]\times\{0\}, it is a homotopy commutative monoid. We may thus apply the group-completion theorem to it, showing that

hocolim𝒢,cX({0},{0})ΩB𝒢,cX,\mathrm{hocolim\,}\mathcal{G}^{X}_{\bullet,c}(\{0\},\{0\})\longrightarrow\Omega B\mathcal{G}^{X}_{\bullet,c},

is a homology isomorphism, where the homotopy colimit is formed by left multiplication with a connected graph of genus 1. By precomposing with a morphism {0}𝒢X\emptyset\leadsto\{0\}\in\mathcal{G}^{X} given by an interval, this map is easily compared with that in the statement of the theorem. ∎

The main technical result we will require is the following homological stability theorem for the spaces 𝒢ns(X)\mathcal{G}_{n}^{s}(X). We will deduce it from (two) arguments of Cohen–Madsen in the analogous situation of surfaces with maps to a background space.

Theorem 4.2.

Suppose that XX is simply-connected.

  1. (i)

    The map 𝒢n1(X)𝒢n+11(X)\mathcal{G}^{1}_{n}(X)\to\mathcal{G}^{1}_{n+1}(X) induces a homology isomorphism in degrees 2n32*\leq n-3.

  2. (ii)

    The map 𝒢n1(X)𝒢n(X)\mathcal{G}^{1}_{n}(X)\to\mathcal{G}_{n}(X) induces an isomorphism on homology with [1n1]\mathbb{Z}[\frac{1}{n-1}]-module coefficients in degrees 2n32*\leq n-3.

Proof.

For part (i) we follow the argument of Cohen and Madsen [6]. There is a map of homotopy fibre sequences

map(nS1,X)\textstyle{\mathrm{map}_{*}(\vee^{n}S^{1},X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒢n1(X)\textstyle{\mathcal{G}^{1}_{n}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒢n1()\textstyle{\mathcal{G}^{1}_{n}(*)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}map(n+1S1,X)\textstyle{\mathrm{map}_{*}(\vee^{n+1}S^{1},X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒢n+11(X)\textstyle{\mathcal{G}^{1}_{n+1}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒢n+11()\textstyle{\mathcal{G}^{1}_{n+1}(*)}

so it is enough to show that Aut(Fn)Hi(map(nS1,X);)\mathrm{Aut}(F_{n})\circlearrowleft H_{i}(\mathrm{map}_{*}(\vee^{n}S^{1},X);\mathbb{Z}) is part of a polynomial coefficient system of degree i\leq i. In this case the map of Serre spectral sequences will induce an isomorphism on Es,t2E^{2}_{s,t} for 2snt32s\leq n-t-3 and an epimorphism for 2snt12s\leq n-t-1; in particular it induces an isomorphism for 2(s+t)n32(s+t)\leq n-3 and an epimorphism for 2(s+t)n12(s+t)\leq n-1. It follows from the spectral sequence comparison theorem that the map H(𝒢n1(X))H(𝒢n+11(X))H_{*}(\mathcal{G}^{1}_{n}(X))\to H_{*}(\mathcal{G}^{1}_{n+1}(X)) in an isomorphism in degrees 2n32*\leq n-3 and an epimorphism in degrees 2n12*\leq n-1.

We define a coefficient system ViX:𝔊𝔯𝐀𝐛V_{i}^{X}\mathrel{\mathop{\mathchar 12346\relax}}\mathfrak{Gr}\to\mathbf{Ab} on objects by

ViX(G):=Hi(map(BG,X);)V_{i}^{X}(G)\mathrel{\mathop{\mathchar 12346\relax}}=H_{i}(\mathrm{map}_{*}(BG,X);\mathbb{Z})

and on a morphism (f:GH,X)(f\mathrel{\mathop{\mathchar 12346\relax}}G\to H,X) we use precomposition by

BH=B(f(G)X)B(f(G))BG.BH=B(f(G)*X)\longrightarrow B(f(G))\cong BG.

This defines a coefficient system, and since B(G)BGB(G*\mathbb{Z})\to BG is split surjective the stabilisation maps ViX(sG)V_{i}^{X}(s_{G}) are all split injective. The coefficient system V0XV_{0}^{X} agrees with the constant coefficient system \mathbb{Z}, as XX has been assumed to be simply-connected: in particular it is polynomial of degree 0.

Let us suppose for an induction that VjXV_{j}^{X} has degree j\leq j for all j<ij<i. Consider the homotopy fibre sequence

map(BG,X)𝜄map(B(G),X)map(B,X),\mathrm{map}_{*}(BG,X)\overset{\iota}{\longrightarrow}\mathrm{map}_{*}(B(G*\mathbb{Z}),X)\longrightarrow\mathrm{map}_{*}(B\mathbb{Z},X),

where the map ι\iota induces ViX(sG)V_{i}^{X}(s_{G}) on iith homology. As ι\iota is split injective, the Serre spectral sequence for this fibration (which is over a path-connected base) collapses, and we find that Coker(ViX(G)ViX(G))\mathrm{Coker}(V_{i}^{X}(G)\to V_{i}^{X}(G*\mathbb{Z})) has a filtration with associated graded

{Hij(ΩX;VjX(G))}j=0i1.\{H_{i-j}(\Omega X;V_{j}^{X}(G))\}_{j=0}^{i-1}.

Each VjX()V_{j}^{X}(-) is polynomial of degree i1\leq i-1 so Hij(ΩX;VjX())H_{i-j}(\Omega X;V_{j}^{X}(-)) is too; as degree is preserved under extensions it follows that Coker(ViX(G)ViX(G))\mathrm{Coker}(V_{i}^{X}(G)\to V_{i}^{X}(G*\mathbb{Z})) has degree i1\leq i-1, hence ViXV_{i}^{X} has degree i\leq i.

For part (ii), we follow a different argument of Cohen and Madsen [7]. We have a diagram

nS1\textstyle{\vee^{n}S^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒢n1(X)\textstyle{\mathcal{G}_{n}^{1}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}𝒢n(X)\textstyle{\mathcal{G}_{n}(X)_{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}e\scriptstyle{e}X\textstyle{X}𝒢n(X)\textstyle{\mathcal{G}_{n}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τn0\scriptstyle{\tau_{n}^{0}}Q1n(X+)\textstyle{Q_{1-n}(X_{+})}

in which the row and column are fibrations, and the map τn0\tau_{n}^{0} is the adjoint to the map of spectra Σ+𝒢n(X)trfπΣ+𝒢n(X)𝑒Σ+X\Sigma^{\infty}_{+}\mathcal{G}_{n}(X)\overset{\mathrm{trf}_{\pi}}{\to}\Sigma^{\infty}_{+}\mathcal{G}_{n}(X)_{*}\overset{e}{\to}\Sigma^{\infty}_{+}X given by the composition of the Becker–Gottlieb transfer for the map π\pi followed by the map ee given by evaluating the map to XX at the marked point.

By part (i), Theorem 4.1, and the analogue with maps to XX of the discussion in Section 5.3 of [12], the composition τn0fτn1:𝒢n1(X)Q1n(X+)\tau_{n}^{0}\circ f\simeq\tau^{1}_{n}\mathrel{\mathop{\mathchar 12346\relax}}\mathcal{G}_{n}^{1}(X)\to Q_{1-n}(X_{+}) is an isomorphism in degrees 2n32*\leq n-3, and in particular τn0\tau^{0}_{n} is surjective on homology in this range. It follows from Leray–Hirsch that the map

e×(τn0π):𝒢n(X)X×Q1n(X+)e\times(\tau_{n}^{0}\circ\pi)\mathrel{\mathop{\mathchar 12346\relax}}\mathcal{G}_{n}(X)_{*}\longrightarrow X\times Q_{1-n}(X_{+})

is also an isomorphism in this range.

Now we have the commutative diagram

H(𝒢n(X))\textstyle{H_{*}(\mathcal{G}_{n}(X))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}trfπ\scriptstyle{\mathrm{trf}_{\pi}}H(𝒢n(X))\textstyle{H_{*}(\mathcal{G}_{n}(X)_{*})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}e×τn0π\scriptstyle{e\times\tau^{0}_{n}\circ\pi}H(X×Q1n(X+))\textstyle{H_{*}(X\times Q_{1-n}(X_{+}))}H(𝒢n(X))\textstyle{H_{*}(\mathcal{G}_{n}(X))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τn0\scriptstyle{\tau^{0}_{n}}H(Q1n(X+))\textstyle{H_{*}(Q_{1-n}(X_{+}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Δ\scriptstyle{\Delta}H(Q1n(X+)×Q1n(X+)),\textstyle{H_{*}(Q_{1-n}(X_{+})\times Q_{1-n}(X_{+}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces,}σId\scriptstyle{\sigma\otimes\mathrm{Id}}

where σ\sigma denotes the homology suspension. The transfer map trfπ\mathrm{trf}_{\pi} is (split) injective with [1n1]\mathbb{Z}[\frac{1}{n-1}]-module coefficients, so in degrees 2n32*\leq n-3 the map τn0\tau_{n}^{0} is injective too. Hence τn0\tau^{0}_{n} is an isomorphism in this range, so ff is too. ∎

5. Proof of the main theorems

5.1. Recollections on Schur–Weyl duality

We shall need a small amount of representation theory, but nothing beyond e.g. the first few parts of Chapter 9 of [22]. Recall that for a vector space VV (over \mathbb{Q}) and a partition λq\lambda\vdash q we have defined Sλ(V)S_{\lambda}(V) as HomΣq(Sλ,Vq)\mathrm{Hom}_{\Sigma_{q}}(S^{\lambda},V^{\otimes q}). If VV is finite-dimensional then as the Specht modules SλS^{\lambda} give a complete set of irreducible representations of Σq\Sigma_{q} we may write

VqλqSλ(V)SλV^{\otimes q}\cong\bigoplus_{\lambda\vdash q}S_{\lambda}(V)\otimes S^{\lambda}

as GL(V)×ΣqGL(V)\times\Sigma_{q}-modules. The Sλ(V)S_{\lambda}(V) are non-zero as long as dimVq\dim_{\mathbb{Q}}V\geq q, in which case they are irreducible GL(V)GL(V)-modules.

5.2. Labeled partitions

Let us fix a sequence of \mathbb{Q}-vector spaces W1,W2,W_{1},W_{2},\ldots, and consider

F(V):=Sym(W1Sym1(V)W2Sym2(V)W3Sym3(V)).F(V)\mathrel{\mathop{\mathchar 12346\relax}}=\mathrm{Sym}^{*}(W_{1}\otimes\mathrm{Sym}^{1}(V)\oplus W_{2}\otimes\mathrm{Sym}^{2}(V)\oplus W_{3}\otimes\mathrm{Sym}^{3}(V)\oplus\cdots).

If we consider this as a functor landing in graded vector spaces, by letting VV have degree 1, then the homogenous pieces of FF are each polynomial functors of VV (see [22, §9.7] for this notion). Thus the Σq\Sigma_{q}-module HomGL(V)(Vq,F(V))\mathrm{Hom}_{GL(V)}(V^{\otimes q},F(V)) is independent of VV as long as dimVq\dim_{\mathbb{Q}}V\geq q. We wish to identify this Σq\Sigma_{q}-module.

Choose bases Ωi\Omega_{i} for the vector spaces WiW_{i}, and let 𝒫q(Ω1,Ω2,)\mathcal{P}_{q}(\Omega_{1},\Omega_{2},\ldots) be the set of partitions PP of {1,2,,q}\{1,2,\ldots,q\} equipped with a labeling of each part XX of size ii with an element XΩi\ell_{X}\in\Omega_{i}. For each such datum (P,)(P,\ell) there is a map of GL(V)GL(V)-modules

ϕP,:Vq\displaystyle\phi_{P,\ell}\mathrel{\mathop{\mathchar 12346\relax}}V^{\otimes q} Sym(W1Sym1(V)W2Sym2(V))\displaystyle\longrightarrow\mathrm{Sym}^{*}(W_{1}\otimes\mathrm{Sym}^{1}(V)\oplus W_{2}\otimes\mathrm{Sym}^{2}(V)\oplus\cdots)
v1vq\displaystyle v_{1}\otimes\cdots\otimes v_{q} XP[XiXvi].\displaystyle\longmapsto\prod_{X\subset P}\left[\ell_{X}\cdot\prod_{i\in X}v_{i}\right].

If σΣq\sigma\in\Sigma_{q} then ϕP,σ=ϕσ(P,)\phi_{P,\ell}\circ\sigma=\phi_{\sigma(P,\ell)}, giving a Σq\Sigma_{q}-equivariant map

ϕ:{𝒫q(Ω1,Ω2,)}HomGL(V)(Vq,F(V)).\phi\mathrel{\mathop{\mathchar 12346\relax}}\mathbb{Q}\{\mathcal{P}_{q}(\Omega_{1},\Omega_{2},\ldots)\}\longrightarrow\mathrm{Hom}_{GL(V)}(V^{\otimes q},F(V)).
Proposition 5.1.

The map ϕ\phi is an isomorphism as long as dimVq\dim_{\mathbb{Q}}V\geq q.

Proof.

Let us call an orbit of Σq\Sigma_{q} acting on 𝒫q(Ω1,Ω2,)\mathcal{P}_{q}(\Omega_{1},\Omega_{2},\ldots) a type of partition. It consists of a partition λq\lambda\vdash q along with an unordered list of labels LiL_{i} in Ωi\Omega_{i} for the parts of size ii, and we write 𝒫(λ;L1,L2,)\mathcal{P}(\lambda;L_{1},L_{2},\ldots) for this orbit.

The source of ϕ\phi splits as a direct sum of (cyclic) modules {𝒫(λ;L1,L2,)}\mathbb{Q}\{\mathcal{P}(\lambda;L_{1},L_{2},\ldots)\} one for each type of partition. For a type of partition 𝒫(λ;L1,L2,)\mathcal{P}(\lambda;L_{1},L_{2},\ldots) and a ωΩi\omega\in\Omega_{i}, let us write |ω|=i|\omega|=i and aωa_{\omega} for the number of parts labeled by ω\omega. The target of ϕ\phi splits as a direct sum of terms

M(λ;L1,L2,):=HomGL(V)(Vq,ωiΩiSymaω({ω}Sym|ω|(V)))M(\lambda;L_{1},L_{2},\ldots)\mathrel{\mathop{\mathchar 12346\relax}}=\mathrm{Hom}_{GL(V)}\left(V^{\otimes q},\bigotimes_{\omega\in\cup_{i}\Omega_{i}}\mathrm{Sym}^{a_{\omega}}(\{\omega\}\otimes\mathrm{Sym}^{|\omega|}(V))\right)

one for each type of partition. The map ϕ\phi restricts to a [Σq]\mathbb{Q}[\Sigma_{q}]-module map

ϕ𝒫(λ;L1,L2,):{𝒫(λ;L1,L2,)}M(λ;L1,L2,),\phi_{\mathcal{P}(\lambda;L_{1},L_{2},\ldots)}\mathrel{\mathop{\mathchar 12346\relax}}\mathbb{Q}\{\mathcal{P}(\lambda;L_{1},L_{2},\ldots)\}\longrightarrow M(\lambda;L_{1},L_{2},\ldots),

and it is enough to show that each of these is an isomorphism.

As GL(V)GL(V)-representations we have a surjection

ωiΩi({ω}V|ω|)aωωiΩiSymaω({ω}Sym|ω|(V))\bigotimes_{\omega\in\cup_{i}\Omega_{i}}(\{\omega\}\otimes V^{\otimes|\omega|})^{\otimes a_{\omega}}\longrightarrow\bigotimes_{\omega\in\cup_{i}\Omega_{i}}\mathrm{Sym}^{a_{\omega}}(\{\omega\}\otimes\mathrm{Sym}^{|\omega|}(V))

which is split by the standard symmetrisers. Furthermore if we choose a (P,)𝒫(λ;L1,L2,)(P,\ell)\in\mathcal{P}(\lambda;L_{1},L_{2},\ldots) then we obtain an isomorphism

Vq\displaystyle V^{\otimes q} ωiΩi({ω}V|ω|)aω\displaystyle\overset{\sim}{\longrightarrow}\bigotimes_{\omega\in\cup_{i}\Omega_{i}}(\{\omega\}\otimes V^{\otimes|\omega|})^{\otimes a_{\omega}}
v1vq\displaystyle v_{1}\otimes\cdots\otimes v_{q} XP(XiXvi)\displaystyle\longmapsto\bigotimes_{X\subset P}\left(\ell_{X}\otimes\bigotimes_{i\in X}v_{i}\right)

(where these terms must be suitably permuted to be put in the right form).

In total this identifies M(λ;L1,L2,)M(\lambda;L_{1},L_{2},\ldots) with a summand of HomGL(V)(Vq,Vq)\mathrm{Hom}_{GL(V)}(V^{\otimes q},V^{\otimes q}), which by the first fundamental theorem of invariant theory for GL(V)GL(V) is a free left [Σq]\mathbb{Q}[\Sigma_{q}]-module generated by the identity map of VqV^{\otimes q} (as long as dimVq\dim_{\mathbb{Q}}V\geq q). Thus the [Σq]\mathbb{Q}[\Sigma_{q}]-module M(λ;L1,L2,)M(\lambda;L_{1},L_{2},\ldots) may be identified with the cyclic left submodule of [Σq]\mathbb{Q}[\Sigma_{q}] generated by the idempotent γ:=1|S|σSσ\gamma\mathrel{\mathop{\mathchar 12346\relax}}=\tfrac{1}{|S|}\sum_{\sigma\in S}\sigma, where SΣqS\leq\Sigma_{q} is the stabiliser of (P,)𝒫q(Ω1,Ω2,)(P,\ell)\in\mathcal{P}_{q}(\Omega_{1},\Omega_{2},\ldots).

Under this identification we have ϕ𝒫(λ;L1,L2,)(P,)=γ\phi_{\mathcal{P}(\lambda;L_{1},L_{2},\ldots)}(P,\ell)=\gamma, so ϕ𝒫(λ;L1,L2,)\phi_{\mathcal{P}(\lambda;L_{1},L_{2},\ldots)} is surjective. On the other hand, we claim that the submodule [Σq]γ[Σq]\mathbb{Q}[\Sigma_{q}]\cdot\gamma\leq\mathbb{Q}[\Sigma_{q}] is isomorphic to [Σq/S]\mathbb{Q}[\Sigma_{q}/S], and hence has dimension the size of the orbit 𝒫(λ;L1,L2,)\mathcal{P}(\lambda;L_{1},L_{2},\ldots). This implies that ϕ𝒫(λ;L1,L2,)\phi_{\mathcal{P}(\lambda;L_{1},L_{2},\ldots)} is an isomorphism. To prove the claim, note that the surjective module map [Σq]γ[Σq](1γ)=[Σq][Σq/S]\mathbb{Q}[\Sigma_{q}]\cdot\gamma\oplus\mathbb{Q}[\Sigma_{q}]\cdot(1-\gamma)=\mathbb{Q}[\Sigma_{q}]\to\mathbb{Q}[\Sigma_{q}/S] sends 1γ1-\gamma to 0, so gives a surjection [Σq]γ[Σq/S]\mathbb{Q}[\Sigma_{q}]\cdot\gamma\to\mathbb{Q}[\Sigma_{q}/S]. On the other hand the module map

aggSaggγ:[Σq/S][Σq]γ\sum a_{g}gS\mapsto\sum a_{g}g\cdot\gamma\mathrel{\mathop{\mathchar 12346\relax}}\mathbb{Q}[\Sigma_{q}/S]\longrightarrow\mathbb{Q}[\Sigma_{q}]\cdot\gamma

is well-defined and surjective. ∎

5.3. Proof of Theorem A

We have already proved Theorem A (i) in Corollary 3.6. For Theorem A (ii), first choose a functorial model for Eilenberg–MacLane spaces K(,n)K(-,n), then fix a finite-dimensional \mathbb{Q}-vector space VV and consider the fibration

K(HV,1)map(nS1,K(V,2))𝒢n1(K(V,2))𝒢n1()BAut(Fn)K(H^{*}\otimes_{\mathbb{Z}}V^{*},1)\simeq\mathrm{map}_{*}(\vee^{n}S^{1},K(V^{*},2))\longrightarrow\mathcal{G}_{n}^{1}(K(V^{*},2))\longrightarrow\mathcal{G}_{n}^{1}(*)\simeq B\mathrm{Aut}(F_{n})

and its associated Serre spectral sequence with \mathbb{Q}-coefficients

(5.1) E2p,q:=Hp(Aut(Fn);q(HV))Hp+q(𝒢n1(K(V,2));).E_{2}^{p,q}\mathrel{\mathop{\mathchar 12346\relax}}=H^{p}(\mathrm{Aut}(F_{n});\wedge^{q}(H_{\mathbb{Q}}\otimes V))\Longrightarrow H^{p+q}(\mathcal{G}_{n}^{1}(K(V^{*},2));\mathbb{Q}).

The action of GL(V)GL(V) on K(V,2)K(V^{*},2), and hence on the fibration above, make this into a spectral sequence of [GL(V)]\mathbb{Q}[GL(V)]-modules.

The proof of the following key lemma is close to an argument communicated to the author by Søren Galatius to prove [23, Conjecture C]. It was his argument that led us to think along the lines necessary to prove Theorem A.

Lemma 5.2.

The spectral sequence (5.1) collapses.

Proof.

The action of the scalars ×GL(V)\mathbb{Q}^{\times}\leq GL(V) on VV makes it a spectral sequence of [×]\mathbb{Q}[\mathbb{Q}^{\times}]-modules, and the action of ×\mathbb{Q}^{\times} on the \mathbb{Q}-vector space q(HV)\wedge^{q}(H_{\mathbb{Q}}\otimes V) is with weight qq, i.e. u×u\in\mathbb{Q}^{\times} acts by scalar multiplication by uqu^{q}. As \mathbb{Q} has characteristic zero, distinct weights make \mathbb{Q} into distinct irreducible [×]\mathbb{Q}[\mathbb{Q}^{\times}]-modules. Thus there can be no [×]\mathbb{Q}[\mathbb{Q}^{\times}]-module maps between different rows of this spectral sequence, so it collapses. ∎

Furthermore, this argument identifies Hp(Aut(Fn);q(HV))H^{p}(\mathrm{Aut}(F_{n});\wedge^{q}_{\mathbb{Q}}(H_{\mathbb{Q}}\otimes V)) with the subspace Hp+q(𝒢n1(K(V,2));)(q)H^{p+q}(\mathcal{G}_{n}^{1}(K(V^{*},2));\mathbb{Q})^{(q)} of Hp+q(𝒢n1(K(V,2));)H^{p+q}(\mathcal{G}_{n}^{1}(K(V^{*},2));\mathbb{Q}) on which ×\mathbb{Q}^{\times} acts with weight qq.

Lemma 5.3.

If 2(p+q)n32(p+q)\leq n-3 then Hp+q(𝒢n1(K(V,2));)(q)H^{p+q}(\mathcal{G}_{n}^{1}(K(V^{*},2));\mathbb{Q})^{(q)} is zero unless p=qp=q, in which case it is isomorphic to the degree 2q2q part of Sym(Sym>0(V[2]))\mathrm{Sym}^{*}(\mathrm{Sym}^{*>0}(V[2])).

Proof.

By Theorem 4.1 and Theorem 4.2, the map

τn1:𝒢n1(K(V,2))Q1n(K(V,2)+)\tau^{1}_{n}\mathrel{\mathop{\mathchar 12346\relax}}\mathcal{G}_{n}^{1}(K(V^{*},2))\longrightarrow Q_{1-n}(K(V^{*},2)_{+})

is an isomorphism on cohomology in degrees 2n32*\leq n-3. With coefficients of characteristic zero, H~(K(V,2);)Sym>0(V[2])\widetilde{H}^{*}(K(V^{*},2);\mathbb{Q})\cong\mathrm{Sym}^{*>0}(V[2]), and taking the free infinite loop space has the effect of forming the free graded-commutative algebra, so in this case the symmetric algebra. The action of ×\mathbb{Q}^{\times} on Sym(Sym>0(V[2]))\mathrm{Sym}^{*}(\mathrm{Sym}^{*>0}(V[2])) is with weight qq precisely in degree 2q2q. ∎

Corollary 5.4.

If 2(p+q)n32(p+q)\leq n-3 then Hp(Aut(Fn);q(HV))H^{p}(\mathrm{Aut}(F_{n});\wedge^{q}_{\mathbb{Q}}(H_{\mathbb{Q}}\otimes_{\mathbb{Q}}V)) is zero unless p=qp=q, in which case it is isomorphic to the degree 2q2q part of Sym(Sym>0(V[2]))\mathrm{Sym}^{*}(\mathrm{Sym}^{*>0}(V[2])).

This proves the vanishing part of Theorem A (ii). We shall now use the fact that the identifications made so far are functorial in VV, and so are in particular GL(V)GL(V)-equivariant and can be decomposed into irreducible GL(V)GL(V)-modules. The following is a standard consequence of Schur–Weyl duality, but we explain its proof anyway.

Lemma 5.5.

As a GL(H)×GL(V)GL(H_{\mathbb{Q}})\times GL(V)-representation,

q(HV)|λ|=qSλ(H)Sλ(V),\wedge^{q}_{\mathbb{Q}}(H_{\mathbb{Q}}\otimes V)\cong\bigoplus_{|\lambda|=q}S_{\lambda}(H_{\mathbb{Q}})\otimes S_{\lambda^{\prime}}(V),

where λ\lambda^{\prime} denotes the conjugate (i.e. transpose) Young diagram to λ\lambda.

Proof.

The left-hand side is the Σq\Sigma_{q}-invariants in the GL(H)×GL(V)×ΣqGL(H_{\mathbb{Q}})\times GL(V)\times\Sigma_{q}-module (HV)qHqVq(H_{\mathbb{Q}}\otimes_{\mathbb{Q}}V)^{\otimes q}\otimes\mathbb{Q}^{-}\cong H_{\mathbb{Q}}^{\otimes q}\otimes V^{\otimes q}\otimes\mathbb{Q}^{-}. By our definition of Schur functors we have an isomorphism of GL(V)×ΣqGL(V)\times\Sigma_{q}-modules

VqλqSλ(V)SλV^{\otimes q}\cong\bigoplus_{\lambda\vdash q}S_{\lambda}(V)\otimes S^{\lambda}

where SλS^{\lambda} is the Specht module. Similarly HqμqSμ(H)SμH_{\mathbb{Q}}^{\otimes q}\cong\bigoplus_{\mu\vdash q}S_{\mu}(H_{\mathbb{Q}})\otimes S^{\mu}, so tensoring them together and taking Σq\Sigma_{q}-invariants, using (SμSλ)Σq=δμλ(S^{\mu}\otimes S^{\lambda}\otimes\mathbb{Q}^{-})^{\Sigma_{q}}=\mathbb{Q}^{\delta_{\mu\lambda^{\prime}}}, the result follows. ∎

Putting the above together, we have an isomorphism of GL(V)GL(V)-modules

|λ|=qHq(Aut(Fn);Sλ(H))Sλ(V)[Sym(Sym>0(V[2]))]2q.\bigoplus_{|\lambda|=q}H^{q}(\mathrm{Aut}(F_{n});S_{\lambda}(H_{\mathbb{Q}}))\otimes S_{\lambda^{\prime}}(V)\cong[\mathrm{Sym}^{*}(\mathrm{Sym}^{*>0}(V[2]))]_{2q}.

Choosing VV to be at least qq-dimensional, the Sμ(V)S_{\mu}(V) are then distinct non-zero irreducible GL(V)GL(V)-modules, so by Schur’s lemma applying HomGL(V)(Sλ(V),)\mathrm{Hom}_{GL(V)}(S_{\lambda^{\prime}}(V),-) gives

Hq(Aut(Fn);Sλ(H))HomGL(V)(Sλ(V),Sym(Sym>0(V)))H^{q}(\mathrm{Aut}(F_{n});S_{\lambda}(H_{\mathbb{Q}}))\cong\mathrm{Hom}_{GL(V)}(S_{\lambda^{\prime}}(V),\mathrm{Sym}^{*}(\mathrm{Sym}^{*>0}(V)))

as long as n2q+3n\geq 2q+3. Using Hq=|λ|=qSλ(H)SλH_{\mathbb{Q}}^{\otimes q}=\bigoplus_{|\lambda|=q}S_{\lambda}(H_{\mathbb{Q}})\otimes S^{\lambda}, we obtain

Hq(Aut(Fn);Hq)HomGL(V)(|λ|=qSλ(V)Sλ,Sym(Sym>0(V))),H^{q}(\mathrm{Aut}(F_{n});H_{\mathbb{Q}}^{\otimes q})\cong\mathrm{Hom}_{GL(V)}\left(\bigoplus_{|\lambda|=q}S_{\lambda^{\prime}}(V)\otimes S^{\lambda},\mathrm{Sym}^{*}(\mathrm{Sym}^{*>0}(V))\right),

and using that SλSλS^{\lambda^{\prime}}\cong S^{\lambda}\otimes\mathbb{Q}^{-} we can write the right-hand side as

HomGL(V)(Vq,Sym(Sym>0(V))).\mathrm{Hom}_{GL(V)}\left(V^{\otimes q},\mathrm{Sym}^{*}(\mathrm{Sym}^{*>0}(V))\right)\otimes\mathbb{Q}^{-}.

Along with Proposition 5.1 this finishes the proof of Theorem A.

Remark 5.6.

We could have used K(V,3)K(V^{*},3) instead of K(V,2)K(V^{*},2). In this case the argument goes through, the analogue of Lemma 5.5 is Symq(HV)|λ|=qSλ(H)Sλ(V)\mathrm{Sym}^{q}(H_{\mathbb{Q}}\otimes V)\cong\bigoplus_{|\lambda|=q}S_{\lambda}(H_{\mathbb{Q}})\otimes S_{\lambda}(V), and the result obtained is

Hq(Aut(Fn);Sλ(H))HomGL(V)(Sλ(V),S(>0(V)))H^{q}(\mathrm{Aut}(F_{n});S_{\lambda}(H_{\mathbb{Q}}))\cong\mathrm{Hom}_{GL(V)}(S_{\lambda}(V),S^{*}(\wedge^{*>0}(V)))

whenever dimVq\dim_{\mathbb{Q}}V\geq q and n2q+3n\geq 2q+3, where S()S^{*}(-) denotes the free graded-commutative algebra.

A consequence of this is that the multiplicity of Sλ(V)S_{\lambda}(V) in S(>0(V))S^{*}(\wedge^{*>0}(V)) is the same as the multiplicity of Sλ(V)S_{\lambda^{\prime}}(V) in Sym(Sym>0(V))\mathrm{Sym}^{*}(\mathrm{Sym}^{*>0}(V)), which does not seem obvious to the author.

5.4. Proof of Theorem B

We will make use of the following lemma, which follows from Kawazumi [18, Theorem 7.1]. It also follows from general principles: the Becker–Gottlieb transfer with local coefficients.

Lemma 5.7.

For any [1n1][Out(Fn)]\mathbb{Z}[\tfrac{1}{n-1}][\mathrm{Out}(F_{n})]-module MM, the map

H(Out(Fn);M)H(Aut(Fn);M)H^{*}(\mathrm{Out}(F_{n});M)\longrightarrow H^{*}(\mathrm{Aut}(F_{n});M)

is split injective, and

H(Aut(Fn);M)H(Out(Fn);M)H1(Out(Fn);HM).H^{*}(\mathrm{Aut}(F_{n});M)\cong H^{*}(\mathrm{Out}(F_{n});M)\oplus H^{*-1}(\mathrm{Out}(F_{n});H^{*}\otimes M).

Theorem B (i) and the first part of Theorem B (ii) follows immediately from this lemma, as it implies that Hi(Out(Fn);Sλ(H))H^{i}(\mathrm{Out}(F_{n});S_{\lambda}(H^{*}_{\mathbb{Q}})) and Hi(Out(Fn);Sλ(H))H^{i}(\mathrm{Out}(F_{n});S_{\lambda}(H_{\mathbb{Q}})) are summands of Hi(Aut(Fn);Sλ(H))H^{i}(\mathrm{Aut}(F_{n});S_{\lambda}(H^{*}_{\mathbb{Q}})) and Hi(Aut(Fn);Sλ(H))H^{i}(\mathrm{Aut}(F_{n});S_{\lambda}(H_{\mathbb{Q}})) respectively, so vanish under the stated assumptions.

It remains to prove the second part of Theorem B (ii). To do this, we again consider a finite-dimensional \mathbb{Q}-vector space VV and consider the diagram

K(HV,1)\textstyle{K(H^{*}\otimes_{\mathbb{Z}}V^{*},1)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}map(nS1,K(V,2))\textstyle{\mathrm{map}(\vee^{n}S^{1},K(V^{*},2))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}𝒢n(K(V,2))\textstyle{\mathcal{G}_{n}(K(V^{*},2))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}𝒢n()BOut(Fn)\textstyle{\mathcal{G}_{n}(*)\simeq B\mathrm{Out}(F_{n})}K(V,2)\textstyle{K(V^{*},2)}

where the row is a (split) fibration and the column is a (trivial) fibration. This gives a spectral sequence

(5.2) H(Out(Fn);(HV[1]))Sym(V[2])H(𝒢n(K(V,2));).H^{*}(\mathrm{Out}(F_{n});\wedge^{*}(H_{\mathbb{Q}}\otimes V[1]))\otimes\mathrm{Sym}^{*}(V[2])\Longrightarrow H^{*}(\mathcal{G}_{n}(K(V^{*},2));\mathbb{Q}).

Rather than a weight argument, which is no longer conclusive, we will deduce the degeneration of this spectral sequence from the vanishing results already established, namely that Hi(Out(Fn);Sλ(H))=0H^{i}(\mathrm{Out}(F_{n});S_{\lambda}(H_{\mathbb{Q}}))=0 in degrees 2in|λ|32i\leq n-|\lambda|-3 if i|λ|i\neq|\lambda|. By Lemma 5.5 this implies that Hp(Out(Fn);q(HV[1]))=0H^{p}(\mathrm{Out}(F_{n});\wedge^{q}(H_{\mathbb{Q}}\otimes V[1]))=0 as long as 2pnq32p\leq n-q-3 and pqp\neq q, so in bidegrees (p,q)(p,q) such that 2pnq32p\leq n-q-3 the E2E^{2}-page of the spectral sequence (5.2) is

a0b0Ha(Out(Fn);a(HV[1]))Symb(V[2]),\bigoplus_{\begin{subarray}{c}a\geq 0\\ b\geq 0\end{subarray}}H^{a}(\mathrm{Out}(F_{n});\wedge^{a}(H_{\mathbb{Q}}\otimes V[1]))\otimes\mathrm{Sym}^{b}(V[2]),

where the (a,b)(a,b)th summand lies in Ea,a+2b2E^{2}_{a,a+2b}. In particular, all summands in this range lie in even total degree, and so there can be no differentials.

Lemma 5.8.

There is an isomorphism of graded GL(V)GL(V)-representations

λH(Out(F);Sλ(H))Sλ(V[1])Sym(Sym>1(V[2])),\bigoplus_{\lambda}H^{*}(\mathrm{Out}(F_{\infty});S_{\lambda}(H_{\mathbb{Q}}))\otimes S_{\lambda^{\prime}}(V[1])\cong\mathrm{Sym}^{*}(\mathrm{Sym}^{*>1}(V[2])),

and dimH|λ|(Out(Fn);Sλ(H))\dim_{\mathbb{Q}}H^{|\lambda|}(\mathrm{Out}(F_{n});S_{\lambda}(H_{\mathbb{Q}})) is independent of nn for n4|λ|+3n\geq 4|\lambda|+3.

Proof.

Writing a(HV[1])|μ|=aSμ(H)Sμ(V[1])\wedge^{a}(H_{\mathbb{Q}}\otimes_{\mathbb{Q}}V[1])\cong\bigoplus_{|\mu|=a}S_{\mu}(H_{\mathbb{Q}})\otimes S_{\mu^{\prime}}(V[1]) using Lemma 5.5, and then identifying the abutment of the spectral sequence (5.2) as in Lemma 5.3, in total degree 2r2r we obtain

a+b=r|μ|=aHa(Out(Fn);Sμ(H))Sμ(V)Symb(V)[Sym(Sym>0(V[2]))]2r\bigoplus_{a+b=r}\bigoplus_{|\mu|=a}H^{a}(\mathrm{Out}(F_{n});S_{\mu}(H_{\mathbb{Q}}))\otimes S_{\mu^{\prime}}(V)\otimes\mathrm{Sym}^{b}(V)\cong[\mathrm{Sym}^{*}(\mathrm{Sym}^{*>0}(V[2]))]_{2r}

as long as 4rn34r\leq n-3.

The right-hand side is independent of nn. For a partition λr\lambda\vdash r we find that

a+b=r|μ|=adimHa(Out(Fn);Sμ(H))dimHomGL(V)(Sλ(V),Sμ(V)Symb(V))\sum_{\begin{subarray}{c}a+b=r\\ |\mu|=a\end{subarray}}\dim_{\mathbb{Q}}H^{a}(\mathrm{Out}(F_{n});S_{\mu}(H_{\mathbb{Q}}))\cdot\dim_{\mathbb{Q}}\mathrm{Hom}_{GL(V)}(S_{\lambda^{\prime}}(V),S_{\mu^{\prime}}(V)\otimes\mathrm{Sym}^{b}(V))

is independent of nn as long as n4|μ|+3n\geq 4|\mu|+3. By induction we may assume that all terms with |μ|<r|\mu|<r are also independent of nn, leaving just the terms with b=0b=0, which gives dimH|λ|(Out(Fn);Sλ(H))\dim_{\mathbb{Q}}H^{|\lambda|}(\mathrm{Out}(F_{n});S_{\lambda}(H_{\mathbb{Q}})) by Schur’s lemma. This proves the second part.

Passing to the limit nn\to\infty and summing over all terms above gives an isomorphism of graded GL(V)GL(V)-representations

λH(Out(F);Sλ(H))Sλ(V[1])Sym(V[2])Sym(Sym>0(V[2])).\bigoplus_{\lambda}H^{*}(\mathrm{Out}(F_{\infty});S_{\lambda}(H_{\mathbb{Q}}))\otimes S_{\lambda^{\prime}}(V[1])\otimes\mathrm{Sym}^{*}(V[2])\cong\mathrm{Sym}^{*}(\mathrm{Sym}^{*>0}(V[2])).

Because the graded GL(V)GL(V)-representation

Sym(V[2])=Sym1(V[2])Sym2(V[2])\mathrm{Sym}^{*}(V[2])=\mathbb{Q}\oplus\mathrm{Sym}^{1}(V[2])\oplus\mathrm{Sym}^{2}(V[2])\oplus\cdots

is the trivial representation in grading zero, the isomorphism in the lemma can be established by induction on degree similarly to the argument above. More conceptually, the graded virtual representation Sym(V)\mathrm{Sym}^{*}(V) is invertible under \otimes, as it is the trivial representation in grading zero, and cancelling it from both sides gives the required isomorphism. ∎

Proceeding as in the proof of Theorem A, we obtain

Hq(Out(Fn);Hq)HomGL(V)(Vq,Sym(Sym>1(V)))H^{q}(\mathrm{Out}(F_{n});H_{\mathbb{Q}}^{\otimes q})\cong\mathrm{Hom}_{GL(V)}(V^{\otimes q},\mathrm{Sym}^{*}(\mathrm{Sym}^{*>1}(V)))\otimes\mathbb{Q}^{-}

from which Proposition 5.1 implies Theorem B.

5.5. Proof of Theorem C

By Remark 2 after Theorem 4.2.2 of [21], there is an identity

λν(λ)Sλ(V)Sym(Sym>1(V))\bigoplus_{\lambda}\nu^{\infty}(\lambda)S_{\lambda}(V)\cong\mathrm{Sym}^{*}(\mathrm{Sym}^{*>1}(V))

of GL(V)GL(V)-modules. Compare with Lemma 5.8 for the first statement; the second statement follows from this and (1.1).

6. Integral and torsion calculations

The general technique we have been using is not confined to rational coefficients. To give an example of how it may be used more generally, we shall now develop a “tame” strengthening of Theorem A, namely Theorem E (ii). Let us first explain how Theorem E (ii) implies Theorem E (i).

Proof of Theorem E (i).

By Theorem A we have that H1(Aut(F);H)=H^{1}(\mathrm{Aut}(F_{\infty});H_{\mathbb{Q}})=\mathbb{Q} and all other rational cohomology groups vanish. By Theorem E (ii), for each prime number pp, H(Aut(F);H(p))H^{*}(\mathrm{Aut}(F_{\infty});H\otimes\mathbb{Z}_{(p)}) is a free H(Aut(F);(p))H^{*}(\mathrm{Aut}(F_{\infty});\mathbb{Z}_{(p)})-module, so by the calculation above is a free module on a single generator in degree 1. Hence H1(Aut(Fn);H)H^{1}(\mathrm{Aut}(F_{n});H) is \mathbb{Z}, as it becomes a free (p)\mathbb{Z}_{(p)}-module of rank 1 when localised at any prime pp. Choosing a generator for this group gives a map

H(Aut(Fn);)H+1(Aut(Fn);H)H^{*}(\mathrm{Aut}(F_{n});\mathbb{Z})\longrightarrow H^{*+1}(\mathrm{Aut}(F_{n});H)

which must be an isomorphism, as it is so when localised at every prime. ∎

In order to prove Theorem E (ii) we first revisit some of the techniques we have used earlier and develop them in the pp-local rather than rational setting.

Let qq be fixed and pp be a prime number such that p>qp>q. Recall that to a partition λq\lambda\vdash q and a tableau TT of shape λ\lambda there is an associated Young symmetriser cT[Σq]c_{T}\in\mathbb{Z}[\Sigma_{q}], which satisfies cT2=q!dimSλcTc_{T}^{2}=\frac{q!}{\dim_{\mathbb{Q}}S^{\lambda}}c_{T}. As q!q! is a pp-local unit we may form the element eT:=dimSλq!cT(p)[Σq]e_{T}\mathrel{\mathop{\mathchar 12346\relax}}=\frac{\dim_{\mathbb{Q}}S^{\lambda}}{q!}c_{T}\in\mathbb{Z}_{(p)}[\Sigma_{q}] and this is an idempotent. If gΣqg\in\Sigma_{q} then egT=geTg1e_{gT}=ge_{T}g^{-1}, so there is a conjugacy class of idempotent associated to each λ\lambda. We write eλe_{\lambda} for the idempotent associated to the canonical tableau for λ\lambda. The Specht module is defined by Sλ:=[Σq]eλS^{\lambda}\mathrel{\mathop{\mathchar 12346\relax}}=\mathbb{Q}[\Sigma_{q}]\cdot e_{\lambda}, and by analogy we define the pp-local Specht module by S(p)λ:=(p)[Σq]eλS^{\lambda}_{(p)}\mathrel{\mathop{\mathchar 12346\relax}}=\mathbb{Z}_{(p)}[\Sigma_{q}]\cdot e_{\lambda}. It is a free (p)\mathbb{Z}_{(p)}-module of rank dimSλ\dim_{\mathbb{Q}}S^{\lambda}, and is indecomposable as a (p)[Σq]\mathbb{Z}_{(p)}[\Sigma_{q}]-module (as S(p)λ=SλS^{\lambda}_{(p)}\otimes\mathbb{Q}=S^{\lambda}). These idempotents satisfy eTeS=0e_{T}\cdot e_{S}=0 if TT and SS are tableaux of different shapes. Furthermore we have a decomposition into primitive idempotents

1=λqstandard tableauxT of shape λeT(p)[Σq].1=\sum_{\lambda\vdash q}\sum_{\begin{subarray}{c}\text{standard tableaux}\\ \text{$T$ of shape $\lambda$}\end{subarray}}e_{T}\in\mathbb{Z}_{(p)}[\Sigma_{q}].

If MM is a (p)\mathbb{Z}_{(p)}-module then MqM^{\otimes q} is a (p)[Σq]\mathbb{Z}_{(p)}[\Sigma_{q}]-module, and we define the pp-local Schur functor by Sλ(M):=eλ(Mq)S_{\lambda}(M)\mathrel{\mathop{\mathchar 12346\relax}}=e_{\lambda}(M^{\otimes q}). There is then a natural map

ϕ:λqS(p)λ(p)Sλ(M)Mq\phi\mathrel{\mathop{\mathchar 12346\relax}}\bigoplus_{\lambda\vdash q}S^{\lambda}_{(p)}\otimes_{\mathbb{Z}_{(p)}}S_{\lambda}(M)\longrightarrow M^{\otimes q}

of (p)[Σq]\mathbb{Z}_{(p)}[\Sigma_{q}]-modules given by (xeλ)eλ(m1mq)xeλ(m1mq)(x\cdot e_{\lambda})\otimes e_{\lambda}(m_{1}\otimes\cdots\otimes m_{q})\mapsto x\cdot e_{\lambda}(m_{1}\otimes\cdots\otimes m_{q}). For an inverse, define

ψ(m1mq)=λqstandard tableauxT of shape λ(gTeλ)eλ(gT1(m1mq))\psi(m_{1}\otimes\cdots\otimes m_{q})=\sum_{\lambda\vdash q}\sum_{\begin{subarray}{c}\text{standard tableaux}\\ \text{$T$ of shape $\lambda$}\end{subarray}}(g_{T}\cdot e_{\lambda})\otimes e_{\lambda}(g_{T}^{-1}(m_{1}\otimes\cdots\otimes m_{q}))

where for a tableau TT of shape λ\lambda we write eT=gTeλgT1e_{T}=g_{T}e_{\lambda}g_{T}^{-1}. This establishes the Schur–Weyl decomposition in this setting.

We now require a partial analogue of the notion of weights which we used in the proof of Lemma 5.2. For a (p)\mathbb{Z}_{(p)}-module MM and an integer tt, write M(t)M(t) for the (p)[(p)×]\mathbb{Z}_{(p)}[\mathbb{Z}_{(p)}^{\times}]-module which is the same as a (p)\mathbb{Z}_{(p)}-module and on which u(p)×u\in\mathbb{Z}_{(p)}^{\times} acts as scalar multiplication by utu^{t}; say that it is a module which is pure of weight tt.

Lemma 6.1.

If MM and NN are (p)\mathbb{Z}_{(p)}-modules then

Ext(p)[(p)×](M(t),N(t))=0\mathrm{Ext}_{\mathbb{Z}_{(p)}[\mathbb{Z}_{(p)}^{\times}]}^{*}(M(t),N(t^{\prime}))=0

if 0<|tt|<p10<|t-t^{\prime}|<p-1.

Proof.

By the natural isomorphisms X(t)(p)[(p)×](p)(s)X(t+s)X(t)\otimes_{\mathbb{Z}_{(p)}[\mathbb{Z}_{(p)}^{\times}]}\mathbb{Z}_{(p)}(s)\cong X(t+s) it is enough to establish the lemma for t=0t^{\prime}=0. In this case consider the functor

FM,t(N):=Hom(p)[(p)×](M(t),N(0)):(p)-Mod𝐀𝐛.F_{M,t}(N)\mathrel{\mathop{\mathchar 12346\relax}}=\mathrm{Hom}_{\mathbb{Z}_{(p)}[\mathbb{Z}_{(p)}^{\times}]}(M(t),N(0))\mathrel{\mathop{\mathchar 12346\relax}}\mathbb{Z}_{(p)}\text{-Mod}\longrightarrow\mathbf{Ab}.

When M=(p)M=\mathbb{Z}_{(p)} this satisfies

F(p),t(N)={xN|(ut1)x=0 for all u(p)×}.F_{\mathbb{Z}_{(p)},t}(N)=\{x\in N\,|\,(u^{t}-1)x=0\text{ for all }u\in\mathbb{Z}_{(p)}^{\times}\}.

By [1, Lemma 2.12] the gcd\gcd of the numbers kt1k^{t}-1 over all integers kk coprime to pp is itself coprime to pp as long as 0<|t|<p10<|t|<p-1, and hence F(p),t(N)=0F_{\mathbb{Z}_{(p)},t}(N)=0 under this assumption on tt. The functor NN(0)N\mapsto N(0) is exact, so taking derived functors of F(p),tF_{\mathbb{Z}_{(p)},t} shows that the claim in the lemma holds for M=(p)M=\mathbb{Z}_{(p)}, and hence for MM any free (p)\mathbb{Z}_{(p)}-module. The claim in general follows by resolving MM by free modules. ∎

We now give the proof of Theorem E (ii), which states that under our assumptions on pp and qq, H(Aut(F);Sλ(H(p)))H^{*}(\mathrm{Aut}(F_{\infty});S_{\lambda}(H_{(p)})) is a free H(Aut(F);(p))H^{*}(\mathrm{Aut}(F_{\infty});\mathbb{Z}_{(p)})-module.

Proof of Theorem E (ii).

Let kk be an odd integer, and let S(p)kS^{k}_{(p)} be a choice of model for the pp-local kk-sphere. For each unit u(p)×u\in\mathbb{Z}_{(p)}^{\times} we may find a map fu:S(p)kS(p)kf_{u}\mathrel{\mathop{\mathchar 12346\relax}}S^{k}_{(p)}\to S^{k}_{(p)} inducing multiplication by uu on Hk(S(p)k;)H_{k}(S^{k}_{(p)};\mathbb{Z}), and these satisfy fufvfuvf_{u}\circ f_{v}\simeq f_{u\cdot v}. Set Y:=(S(p)k)qY\mathrel{\mathop{\mathchar 12346\relax}}=(S^{k}_{(p)})^{q}, and consider the space 𝒢n1(Y)\mathcal{G}_{n}^{1}(Y), the Serre fibration

(6.1) map(nS1,Y)𝒢n1(Y)𝒢n1()BAut(Fn),\mathrm{map}_{*}(\vee^{n}S^{1},Y)\longrightarrow\mathcal{G}_{n}^{1}(Y)\longrightarrow\mathcal{G}_{n}^{1}(*)\simeq B\mathrm{Aut}(F_{n}),

and its associated Serre spectral sequence.

Recall that the (p)\mathbb{Z}_{(p)}-cohomology ring of ΩS(p)k\Omega S^{k}_{(p)} is the divided power algebra Γ(p)(xk1)\Gamma_{\mathbb{Z}_{(p)}}^{*}(x_{k-1}) on the class obtained by looping a generator of Hk(Sk;(p))H^{k}(S^{k};\mathbb{Z}_{(p)}). This identifies the (p)\mathbb{Z}_{(p)}-cohomology ring of map(nS1,Y)\mathrm{map}_{*}(\vee^{n}S^{1},Y) with Γ(p)(H(p)[k1])q\Gamma_{\mathbb{Z}_{(p)}}^{*}(H_{(p)}[k-1])^{\otimes q}, so, taking the limit nn\to\infty, there is a spectral sequence

E2,=H(Aut(F);Γ(p)(H(p)[k1])q)H(Q0(S0)×QY;(p)).E_{2}^{*,*}=H^{*}(\mathrm{Aut}(F_{\infty});\Gamma_{\mathbb{Z}_{(p)}}^{*}(H_{(p)}[k-1])^{\otimes q})\Longrightarrow H^{*}(Q_{0}(S^{0})\times QY;\mathbb{Z}_{(p)}).

It follows from the results of [5, p. 40] that the map

Sym(p)(H~(Y;(p)))H(QY;(p))\mathrm{Sym}^{*}_{\mathbb{Z}_{(p)}}(\widetilde{H}_{*}(Y;\mathbb{Z}_{(p)}))\longrightarrow H_{*}(QY;\mathbb{Z}_{(p)})

from the free commutative algebra, induced by the map YQ(Y)Y\to Q(Y), is an isomorphism on homology in degrees pk*\leq pk. By the Künneth theorem we have

H(Y;(p))=((p)[0](p)[k])q,H_{*}(Y;\mathbb{Z}_{(p)})=(\mathbb{Z}_{(p)}[0]\oplus\mathbb{Z}_{(p)}[k])^{\otimes q},

which is free as a (p)\mathbb{Z}_{(p)}-module. This shows that H(QY;(p))H^{*}(QY;\mathbb{Z}_{(p)}) is a free (p)\mathbb{Z}_{(p)}-module in degrees pk*\leq pk, and shows that it is supported in degrees divisible by kk. Using Galatius’ theorem and the Künneth theorem again we may therefore identify the target of the spectral sequence with H(Aut(F);(p))(p)H(QY;(p))H^{*}(\mathrm{Aut}(F_{\infty});\mathbb{Z}_{(p)})\otimes_{\mathbb{Z}_{(p)}}H^{*}(QY;\mathbb{Z}_{(p)}) in a range of degrees. In total degrees pk*\leq pk the spectral sequence takes the form

E2,=H(Aut(F);Γ(p)(H(p)[k1])q)H(Aut(F);(p))H(QY;(p))E_{2}^{*,*}=H^{*}(\mathrm{Aut}(F_{\infty});\Gamma_{\mathbb{Z}_{(p)}}^{*}(H_{(p)}[k-1])^{\otimes q})\Rightarrow H^{*}(\mathrm{Aut}(F_{\infty});\mathbb{Z}_{(p)})\otimes H^{*}(QY;\mathbb{Z}_{(p)})

and is a spectral sequence of H(Aut(F);(p))H^{*}(\mathrm{Aut}(F_{\infty});\mathbb{Z}_{(p)})-modules.

For u(p)×u\in\mathbb{Z}_{(p)}^{\times} the homotopy equivalence fuq:YYf_{u}^{q}\mathrel{\mathop{\mathchar 12346\relax}}Y\to Y induces a map of spectral sequences by functoriality, making it into a spectral sequence of (p)[(p)×]\mathbb{Z}_{(p)}[\mathbb{Z}_{(p)}^{\times}]-modules. The induced map on Γ(p)i(H(p)[k1])q\Gamma_{\mathbb{Z}_{(p)}}^{i}(H_{(p)}[k-1])^{\otimes q} is given by scalar multiplication by uiqu^{iq}, so the j(k1)j(k-1)st row of the spectral sequence is pure of weight jj. Furthermore, the fibration (6.1) has a section so there are no differentials entering the bottom row, and this row splits off the filtration. The rows (k1),2(k1),,(p1)(k1)(k-1),2(k-1),\ldots,(p-1)(k-1) have different weights which differ by at most (p2)(p-2), so by Lemma 6.1 there are no differentials in total degree <p(k1)*<p(k-1) and the associated filtration of H(Aut(F);(p))H(QY;(p))H^{*}(\mathrm{Aut}(F_{\infty});\mathbb{Z}_{(p)})\otimes H^{*}(QY;\mathbb{Z}_{(p)}) splits as (p)[(p)×]\mathbb{Z}_{(p)}[\mathbb{Z}_{(p)}^{\times}]-modules.

The map induced by fuqf_{u}^{q} on Hkj(QY;(p))H^{kj}(QY;\mathbb{Z}_{(p)}) is multiplication by uju^{j}, so this is pure of weight jj. It follows that the q(k1)q(k-1)st row of the spectral sequence may be identified in a range of degrees with H(Aut(F);(p))Hkq(QY;(p))H^{*}(\mathrm{Aut}(F_{\infty});\mathbb{Z}_{(p)})\otimes H^{kq}(QY;\mathbb{Z}_{(p)}) and so is a free H(Aut(F);(p))H^{*}(\mathrm{Aut}(F_{\infty});\mathbb{Z}_{(p)})-module. The Aut(Fn)\mathrm{Aut}(F_{n})-module Γ(p)(H(p)[k1])q\Gamma_{\mathbb{Z}_{(p)}}^{*}(H_{(p)}[k-1])^{\otimes q} contains (H(p)[k1])q(H_{(p)}[k-1])^{\otimes q} as a summand in degree q(k1)q(k-1), so contains Sλ(H(p))S_{\lambda}(H_{(p)}) as a summand in this degree too. Thus the q(k1)q(k-1)st row of the spectral sequence contains H(Aut(F);Sλ(H(p)))H^{*}(\mathrm{Aut}(F_{\infty});S_{\lambda}(H_{(p)})) as a summand, so this is a projective H(Aut(F);(p))H^{*}(\mathrm{Aut}(F_{\infty});\mathbb{Z}_{(p)})-module in degrees <p(k1)q(k1)=(pq)(k1)*<p(k-1)-q(k-1)=(p-q)(k-1), so in all degrees as p>qp>q and kk was arbitrary. Finally, as H(Aut(F);(p))H^{*}(\mathrm{Aut}(F_{\infty});\mathbb{Z}_{(p)}) is a connected graded algebra over the local ring (p)\mathbb{Z}_{(p)}, projective graded modules which are finitely-generated in each degree are free. ∎

Appendix A An integral calculation for Out(Fn)\mathrm{Out}(F_{n})

It seems reasonable to suppose that Theorem 4.2 (ii) holds with integral and not just [1n1]\mathbb{Z}[\tfrac{1}{n-1}]-module coefficients, that is, that

Conjecture A.

The map 𝒢n1(X)𝒢n(X)\mathcal{G}^{1}_{n}(X)\to\mathcal{G}_{n}(X) induces an isomorphism on homology in degrees 2n32*\leq n-3.

Putting this together with Theorem 4.1, it follows that τn0:𝒢n(X)Q1n(X+)\tau^{0}_{n}\mathrel{\mathop{\mathchar 12346\relax}}\mathcal{G}_{n}(X)\to Q_{1-n}(X_{+}) is an isomorphism on integral homology in degrees 2n32*\leq n-3. Assuming this conjecture, we may make the following calculation.

Proposition A.1.

We have

H1(Out(Fn);H)\displaystyle H^{1}(\mathrm{Out}(F_{n});H) =0 for n7\displaystyle=0\quad\quad\quad\quad\text{\,\,\,\,for $n\geq 7$}
H2(Out(Fn);H)\displaystyle H^{2}(\mathrm{Out}(F_{n});H) =/(n1)for n9.\displaystyle=\mathbb{Z}/(n-1)\quad\text{for $n\geq 9$.}

We emphasise that this proposition holds without assuming Conjecture A: it follows from Theorem B that these groups are torsion, Satoh [26] has computed that H1(Out(Fn);H)/(n1)H_{1}(\mathrm{Out}(F_{n});H^{*})\cong\mathbb{Z}/(n-1) for n4n\geq 4, and one easily computes that H0(Out(Fn);H)=0H_{0}(\mathrm{Out}(F_{n});H^{*})=0. Our purpose here is to give another proof of this proposition using Conjecture A.

This proposition should be contrasted with with a theorem of Bridson and Vogtmann [2, Theorem B], who show that the extension

H=Fn/FnAut(Fn)/FnOut(Fn)H=F_{n}/F_{n}^{\prime}\longrightarrow\mathrm{Aut}(F_{n})/F_{n}^{\prime}\longrightarrow\mathrm{Out}(F_{n})

is non-trivial for all n2n\geq 2, and hence gives a non-trivial class ζH2(Out(Fn);H)\zeta\in H^{2}(\mathrm{Out}(F_{n});H). Our calculation H2(Out(Fn);H)=/(n1)H^{2}(\mathrm{Out}(F_{n});H)=\mathbb{Z}/(n-1) along with the result of Bridson and Vogtmann [2] that their class ζ\zeta remains non-trivial in the group H2(Out(Fn);H/rH)H^{2}(\mathrm{Out}(F_{n});H/rH) for any rr not coprime to (n1)(n-1) implies that the class ζ\zeta generates H2(Out(Fn);H)H^{2}(\mathrm{Out}(F_{n});H) as long as n9n\geq 9.

Proof of Proposition A.1 assuming Conjecture A.

Consider the diagram

BH\textstyle{BH^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}map(nS1,)\textstyle{\mathrm{map}(\vee^{n}S^{1},\mathbb{C}\mathbb{P}^{\infty})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}𝒢n()\textstyle{\mathcal{G}_{n}(\mathbb{C}\mathbb{P}^{\infty})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}𝒢n()BOut(Fn)\textstyle{\mathcal{G}_{n}(*)\simeq B\mathrm{Out}(F_{n})}\textstyle{\mathbb{C}\mathbb{P}^{\infty}}

where the row and column are fibrations. Note that pp is a trivial fibration and is split via the inclusion s:map(nS1,)s\mathrel{\mathop{\mathchar 12346\relax}}\mathbb{C}\mathbb{P}^{\infty}\to\mathrm{map}(\vee^{n}S^{1},\mathbb{C}\mathbb{P}^{\infty}) of the constant maps, and there is an inclusion ι:𝒢n()×𝒢n()\iota\mathrel{\mathop{\mathchar 12346\relax}}\mathcal{G}_{n}(*)\times\mathbb{C}\mathbb{P}^{\infty}\to\mathcal{G}_{n}(\mathbb{C}\mathbb{P}^{\infty}) of the graphs with constant maps to \mathbb{C}\mathbb{P}^{\infty}. The Leray–Serre spectral sequence for the horizontal fibration is

(A.1) E¯2p,:=Hp(Out(Fn);H)[a]H(𝒢n();)\bar{E}_{2}^{p,*}\mathrel{\mathop{\mathchar 12346\relax}}=H^{p}(\mathrm{Out}(F_{n});\wedge^{*}H)\otimes\mathbb{Z}[a]\Longrightarrow H^{*}(\mathcal{G}_{n}(\mathbb{C}\mathbb{P}^{\infty});\mathbb{Z})

where aa is the canonical class in H2(;)H^{2}(\mathbb{C}\mathbb{P}^{\infty};\mathbb{Z}), so has bidegree (p,q)=(0,2)(p,q)=(0,2).

We first claim that the map

𝑠map(nS1,)𝒢n()\mathbb{C}\mathbb{P}^{\infty}\overset{s}{\longrightarrow}\mathrm{map}(\vee^{n}S^{1},\mathbb{C}\mathbb{P}^{\infty})\longrightarrow\mathcal{G}_{n}(\mathbb{C}\mathbb{P}^{\infty})

has image (n1)=H2(;)(n-1)\mathbb{Z}\subset\mathbb{Z}=H^{2}(\mathbb{C}\mathbb{P}^{\infty};\mathbb{Z}) on second cohomology. By our conjecture, it is enough to prove this after composing with the map τn0:𝒢n()Q0(+)\tau_{n}^{0}\mathrel{\mathop{\mathchar 12346\relax}}\mathcal{G}_{n}(\mathbb{C}\mathbb{P}^{\infty})\to Q_{0}(\mathbb{C}\mathbb{P}^{\infty}_{+}) as long as n7n\geq 7. Up to translation of components, this map is given by the Becker–Gottlieb transfer Q1n(×(nS1)+)\mathbb{C}\mathbb{P}^{\infty}\to Q_{1-n}(\mathbb{C}\mathbb{P}^{\infty}\times(\vee^{n}S^{1})_{+}) for the trivial graph bundle over \mathbb{C}\mathbb{P}^{\infty} composed with projection to Q1n(+)Q_{1-n}(\mathbb{C}\mathbb{P}^{\infty}_{+}). By standard properties of the transfer, this is (1n)(1-n) times the standard inclusion, which on second cohomology induces multiplication by (1n)(1-n), as required.

This describes the edge homomorphism of the spectral sequence (A.1). We now use the homotopy equivalence

Q0(+)Q0(S0)×Q()Q_{0}(\mathbb{C}\mathbb{P}^{\infty}_{+})\simeq Q_{0}(S^{0})\times Q(\mathbb{C}\mathbb{P}^{\infty})

and that S2Q()S^{2}\to Q(\mathbb{C}\mathbb{P}^{\infty}) is 3-connected (by the Freudenthal suspension theorem) to describe the cohomology of Q0(+)Q_{0}(\mathbb{C}\mathbb{P}^{\infty}_{+}) in low degrees. Using that Hi(Q0(S0);)H^{i}(Q_{0}(S^{0});\mathbb{Z}) is ,0,/2,/2\mathbb{Z},0,\mathbb{Z}/2,\mathbb{Z}/2 for i=0,1,2,3i=0,1,2,3, it follows that Hi(Q0(+);)H^{i}(Q_{0}(\mathbb{C}\mathbb{P}^{\infty}_{+});\mathbb{Z}) is ,0,/2,/2\mathbb{Z},0,\mathbb{Z}\oplus\mathbb{Z}/2,\mathbb{Z}/2 for i=0,1,2,3i=0,1,2,3. As the spectral sequence (A.1) converges to zero for positive Leray filtration in total degree 33, the differential d2:=E¯20,2E¯22,1d_{2}\mathrel{\mathop{\mathchar 12346\relax}}\mathbb{Z}=\bar{E}_{2}^{0,2}\to\bar{E}_{2}^{2,1} must be onto (so E¯22,1\bar{E}_{2}^{2,1} is cyclic) and the kernel is (n1)(n-1)\mathbb{Z}, so E¯22,1=H2(Out(Fn);H)/(n1)\bar{E}_{2}^{2,1}=H^{2}(\mathrm{Out}(F_{n});H)\cong\mathbb{Z}/(n-1). On the other hand, in total degree 22 we see (n1)=E¯0,2(n-1)\mathbb{Z}=\bar{E}_{\infty}^{0,2} and /2=E¯2,0\mathbb{Z}/2=\bar{E}_{\infty}^{2,0}, and it converges to /2\mathbb{Z}/2\oplus\mathbb{Z}, so observing the direction of the Leray filtration we see that H1(Out(Fn);H)=E¯21,1=0H^{1}(\mathrm{Out}(F_{n});H)=\bar{E}_{2}^{1,1}=0. ∎

Appendix B Mapping class groups of surfaces

Let Γg\Gamma_{g} denote the mapping class group of a surface Σg\Sigma_{g} of genus gg, and consider H:=H1(Σg;)H_{\mathbb{Q}}\mathrel{\mathop{\mathchar 12346\relax}}=H_{1}(\Sigma_{g};\mathbb{Q}) as a Γg\Gamma_{g}-module. In this case Poincaré duality gives HHH^{*}_{\mathbb{Q}}\cong H_{\mathbb{Q}}. Looijenga has already computed H(Γg;Sλ(H))H^{*}(\Gamma_{g};S_{\lambda}(H_{\mathbb{Q}})) in the stable range [20], but we wish to explain here how the computation may also be performed using the techniques of this paper.

Cohen and Madsen [6] have introduced spaces 𝒮g(X)\mathcal{S}_{g}(X) of surfaces diffeomorphic to Σg\Sigma_{g} equipped with a map to XX, and have identified the stable (co)homology of these spaces with that of the infinite loop space Ω(MTSO(2)X+)\Omega^{\infty}(\mathrm{MTSO}(2)\wedge X_{+}) as long as XX is simply-connected. Furthermore, there is a fibration sequence

map(Σg,X)𝒮g(X)BΓg\mathrm{map}(\Sigma_{g},X)\longrightarrow\mathcal{S}_{g}(X)\longrightarrow B\Gamma_{g}

so an associated Serre spectral sequence.

Considering X=K(V,4)X=K(V^{*},4), the argument of Section 5.4 goes through without change and identifies

λH(Γg;Sλ(H))Sλ(V)[3|λ|]Sym(V[2])Sym(V[4])\bigoplus_{\lambda}H^{*}(\Gamma_{g};S_{\lambda}(H_{\mathbb{Q}}))\otimes S_{\lambda^{\prime}}(V)[3|\lambda|]\otimes\mathrm{Sym}^{*}(V[2])\otimes\mathrm{Sym}^{*}(V[4])

with

H(Ω0(MTSO(2)K(V,4)+);)=Sym([WSym(V[4])]>0)H^{*}(\Omega^{\infty}_{0}(\mathrm{MTSO}(2)\wedge K(V^{*},4)_{+});\mathbb{Q})=\mathrm{Sym}^{*}([W_{*}\otimes\mathrm{Sym}^{*}(V[4])]_{>0})

in a range of degrees increasing with gg, where W=H(MTSO(2);)W_{*}=H^{*}(\mathrm{MTSO}(2);\mathbb{Q}) is the graded vector space which is \mathbb{Q} in degrees 2,0,2,4,6,-2,0,2,4,6,\ldots and zero otherwise. This may be written as

Sym(W>0)Sym(V[2])Sym(V[4])Sym(W>0V[4]WSym>1(V[4]))\mathrm{Sym}^{*}(W_{*>0})\otimes\mathrm{Sym}^{*}(V[2])\otimes\mathrm{Sym}^{*}(V[4])\otimes\mathrm{Sym}^{*}(W_{*>0}\otimes V[4]\oplus W_{*}\otimes\mathrm{Sym}^{*>1}(V[4]))

and the first term Sym(W>0)\mathrm{Sym}^{*}(W_{*>0}) is H(Γg;)H^{*}(\Gamma_{g};\mathbb{Q}) in the stable range, so we obtain

λH(Γg;Sλ(H))Sλ(V)[3|λ|]H(Γg;)Sym(W>0V[4]WSym>1(V[4]))\bigoplus_{\lambda}H^{*}(\Gamma_{g};S_{\lambda}(H_{\mathbb{Q}}))\otimes S_{\lambda^{\prime}}(V)[3|\lambda|]\cong H^{*}(\Gamma_{g};\mathbb{Q})\otimes\mathrm{Sym}^{*}(W_{*>0}\otimes V[4]\oplus W_{*}\otimes\mathrm{Sym}^{*>1}(V[4]))

in a range of degrees increasing with gg.

It follows that each H(Γg;Sλ(H))H^{*}(\Gamma_{g};S_{\lambda}(H_{\mathbb{Q}})) is a free H(Γg;)H^{*}(\Gamma_{g};\mathbb{Q})-module in the stable range. To describe the space of module generators we can proceed as in the proof of Theorem A, and hence identify H(Γg;Hq)H^{*}(\Gamma_{g};H_{\mathbb{Q}}^{\otimes q}) as a graded Σq\Sigma_{q}-module with

H(Γg;)HomGL(V)(Vq,Sym(W>0V[4]WSym>1(V[4])))[3q].H^{*}(\Gamma_{g};\mathbb{Q})\otimes\mathrm{Hom}_{GL(V)}(V^{\otimes q},\mathrm{Sym}^{*}(W_{*>0}\otimes V[4]\oplus W_{*}\otimes\mathrm{Sym}^{*>1}(V[4])))[-3q]\otimes\mathbb{Q}^{-}.

Proposition 5.1 identifies the space of GL(V)GL(V)-module homomorphisms with the permutation module on the set of the following data: a partition PP of {1,2,,q}\{1,2,\ldots,q\}, a labeling of each part of size 1 in the set {x2,x4,x6,}\{x_{2},x_{4},x_{6},\ldots\}, and a labeling of each part of size >1>1 in the set {x2,x0,x2,x4,x6,}\{x_{-2},x_{0},x_{2},x_{4},x_{6},\ldots\}. Such a datum is given grading qq plus the sum of the degrees of the labels (which are given by their subscripts).

Example B.1.

When q=1q=1 we find that H(Γg;H)H^{*}(\Gamma_{g};H_{\mathbb{Q}}) is a free H(Γg;)H^{*}(\Gamma_{g};\mathbb{Q})-module with generators in degrees {3,5,7,}\{3,5,7,\ldots\}.

When q=2q=2 the partition {1,2}\{1,2\} may be labeled by {x2,x0,x2,x4,x6,}\{x_{-2},x_{0},x_{2},x_{4},x_{6},\ldots\}, and the partition {{1},{2}}\{\{1\},\{2\}\} may have each part labeled by {x2,x4,x6,}\{x_{2},x_{4},x_{6},\ldots\}. As a Σ2\Sigma_{2}-set the action is trivial on the first type of elements, and the second type form trivial orbits

{{1;x2i},{2;x2i}}\{\{1;x_{2i}\},\{2;x_{2i}\}\}

and free orbits

{{1;x2i},{2;x2j}},{{1;x2j},{2;x2i}}\{\{1;x_{2i}\},\{2;x_{2j}\}\},\{\{1;x_{2j}\},\{2;x_{2i}\}\}

with iji\neq j. Thus H(Γg;2(H))H^{*}(\Gamma_{g};\wedge^{2}(H_{\mathbb{Q}})) has H(Γg;)H^{*}(\Gamma_{g};\mathbb{Q})-module generators given by the multiplicities of the trivial representation in the indicated permutation module, so in degrees {0,2,4,6,;6,8,10,;10,12,14,;14,16,18,;}\{0,2,4,6,\ldots;6,8,10,\ldots;10,12,14,\ldots;14,16,18,\ldots;\ldots\}. Similarly H(Γg;Sym2(H))H^{*}(\Gamma_{g};\mathrm{Sym}^{2}(H_{\mathbb{Q}})) has H(Γg;)H^{*}(\Gamma_{g};\mathbb{Q})-module generators given by the multiplicities of the sign representation in the indicated permutation module, so in degrees {8,10,12,;12,14,16,;16,18,20,;}\{8,10,12,\ldots;12,14,16,\ldots;16,18,20,\ldots;\ldots\}.

See [11, Theorem G] for a complete calculation in the case of exterior powers, and [24, Proposition 5.2] for a complete calculation in the case of symmetric powers.

By considering the analogous space 𝒮g,1(X)\mathcal{S}_{g,1}(X) of surfaces of genus gg with one boundary equipped with maps to X=K(V,4)X=K(V^{*},4) which map the boundary to the basepoint, one identifies λH(Γg,1;Sλ(H))Sλ(V)[3|λ|]\bigoplus_{\lambda}H^{*}(\Gamma_{g,1};S_{\lambda}(H_{\mathbb{Q}}))\otimes S_{\lambda^{\prime}}(V)[3|\lambda|] with

H(Γg,1;)Sym(W>2V[4]WSym>0(V[4]))H^{*}(\Gamma_{g,1};\mathbb{Q})\otimes\mathrm{Sym}^{*}(W_{*>-2}\otimes V[4]\oplus W_{*}\otimes\mathrm{Sym}^{*>0}(V[4]))

in a range of degrees increasing with gg, from which H(Γg,1;Hq)H^{*}(\Gamma_{g,1};H_{\mathbb{Q}}^{\otimes q}) may be computed as above. There is a similar description in terms of partitions, the only difference being that now pieces of a partition of size 1 may also be labeled by x0x_{0}. For example, one finds that H(Γg,1;H)H^{*}(\Gamma_{g,1};H_{\mathbb{Q}}) is a free H(Γg,1;)H^{*}(\Gamma_{g,1};\mathbb{Q})-module with generators in degrees {1,3,5,7,}\{1,3,5,7,\ldots\}. This description of H(Γg,1;Hq)H^{*}(\Gamma_{g,1};H_{\mathbb{Q}}^{\otimes q}) (in terms of labelled partitions) has previously been obtained by Kawazumi [19], cf. Theorem 1.B of that paper, even with integral as opposed to just rational coefficients. However Kawazumi does not establish this description as Σq\Sigma_{q}-modules, because his method of proof breaks the Σq\Sigma_{q}-symmetry.

Appendix C Tables

The result of Theorems A and B for |λ|6|\lambda|\leq 6 is compiled in the tables below, which was computed with the SchurRings package for Macaulay2 [14], and checked via Theorem C and the table at the end of Section 4.3 of [21].

Dimensions of H|λ|(Aut(Fn);Sλ(H))H^{|\lambda|}(\mathrm{Aut}(F_{n});S_{\lambda}(H_{\mathbb{Q}})) for |λ|6|\lambda|\leq 6 and n2|λ|+3n\geq 2|\lambda|+3.
()() (1)(1) (12)(1^{2}) (2)(2) (13)(1^{3}) (21)(21) (3)(3) (14)(1^{4}) (212)(21^{2}) (22)(2^{2}) (31)(31) (4)(4)
1 1 2 0 3 1 0 5 2 2 0 0
(15)(1^{5}) (213)(21^{3}) (221)(2^{2}1) (312)(31^{2}) (32)(32) (41)(41) (5)(5)
7 5 4 0 1 0 0
(16)(1^{6}) (214)(21^{4}) (2212)(2^{2}1^{2}) (23)(2^{3}) (313)(31^{3}) (321)(321) (32)(3^{2}) (412)(41^{2}) (42)(42) (51)(51) (6)(6)
11 8 10 2 1 2 2 0 0 0 0
Dimensions of H|λ|(Out(Fn);Sλ(H))H^{|\lambda|}(\mathrm{Out}(F_{n});S_{\lambda}(H_{\mathbb{Q}})) for |λ|6|\lambda|\leq 6 and n4|λ|+3n\geq 4|\lambda|+3.
()() (1)(1) (11)(11) (2)(2) (111)(111) (21)(21) (3)(3) (1111)(1111) (211)(211) (22)(22) (31)(31) (4)(4)
1 0 1 0 1 0 0 2 0 1 0 0
(11111)(11111) (2111)(2111) (221)(221) (311)(311) (32)(32) (41)(41) (5)(5)
2 1 1 0 0 0 0
(16)(1^{6}) (214)(21^{4}) (2212)(2^{2}1^{2}) (23)(2^{3}) (313)(31^{3}) (321)(321) (32)(3^{2}) (412)(41^{2}) (42)(42) (51)(51) (6)(6)
4 1 3 0 0 0 1 0 0 0 0

References

  • [1] J. F. Adams. On the groups J(X)J(X). II. Topology, 3:137–171, 1965.
  • [2] Martin R. Bridson and Karen Vogtmann. Abelian covers of graphs and maps between outer automorphism groups of free groups. Math. Ann., 353(4):1069–1102, 2012.
  • [3] Kenneth S. Brown. Cohomology of groups, volume 87 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1994. Corrected reprint of the 1982 original.
  • [4] Henri Cartan and Samuel Eilenberg. Homological algebra. Princeton University Press, Princeton, N. J., 1956.
  • [5] Frederick R. Cohen, Thomas J. Lada, and J. Peter May. The homology of iterated loop spaces. Springer-Verlag, Berlin, 1976. Lecture Notes in Mathematics, Vol. 533.
  • [6] Ralph Cohen and Ib Madsen. Surfaces in a background space and the homology of mapping class group. Proc. Symp. Pure Math., 80(1):43–76, 2009.
  • [7] Ralph Cohen and Ib Madsen. Stability for closed surfaces in a background space. Homology Homotopy Appl., 13(2):301–313, 2011.
  • [8] Aurélien Djament. Décomposition de Hodge pour l’homologie stable des groupes d’automorphismes des groupes libres. arXiv:1510.03546, 2015.
  • [9] Aurélien Djament and Christine Vespa. Sur l’homologie des groupes d’automorphismes des groupes libres à coefficients polynomiaux. Comment. Math. Helv., 90(1):33–58, 2015.
  • [10] W. G. Dwyer. Twisted homological stability for general linear groups. Ann. of Math. (2), 111(2):239–251, 1980.
  • [11] Johannes Ebert and Oscar Randal-Williams. Stable cohomology of the universal Picard varieties and the extended mapping class group. Doc. Math., 17:417–450, 2012.
  • [12] Søren Galatius. Stable homology of automorphism groups of free groups. Ann. of Math. (2), 173(2):705–768, 2011.
  • [13] Søren Galatius and Oscar Randal-Williams. Monoids of moduli spaces of manifolds. Geom. Topol., 14(3):1243–1302, 2010.
  • [14] Daniel R. Grayson and Michael E. Stillman. Macaulay2, a software system for research in algebraic geometry. Available at http://www.math.uiuc.edu/Macaulay2.
  • [15] Allen Hatcher and Karen Vogtmann. Cerf theory for graphs. J. London Math. Soc. (2), 58(3):633–655, 1998.
  • [16] Allen Hatcher and Karen Vogtmann. Homology stability for outer automorphism groups of free groups. Algebr. Geom. Topol., 4:1253–1272 (electronic), 2004.
  • [17] Nikolai V. Ivanov. On the homology stability for Teichmüller modular groups: closed surfaces and twisted coefficients. In Mapping class groups and moduli spaces of Riemann surfaces (Göttingen, 1991/Seattle, WA, 1991), volume 150 of Contemp. Math., pages 149–194. Amer. Math. Soc., Providence, RI, 1993.
  • [18] Nariya Kawazumi. Cohomological aspects of Magnus expansions. arXiv:math/0505497, 2005.
  • [19] Nariya Kawazumi. On the stable cohomology algebra of extended mapping class groups for surfaces. In Groups of diffeomorphisms, volume 52 of Adv. Stud. Pure Math., pages 383–400. Math. Soc. Japan, Tokyo, 2008.
  • [20] Eduard Looijenga. Stable cohomology of the mapping class group with symplectic coefficients and of the universal Abel-Jacobi map. J. Algebraic Geom., 5(1):135–150, 1996.
  • [21] L. Manivel. Gaussian maps and plethysm. In Algebraic geometry (Catania, 1993/Barcelona, 1994), volume 200 of Lecture Notes in Pure and Appl. Math., pages 91–117. Dekker, New York, 1998.
  • [22] Claudio Procesi. Lie groups. Universitext. Springer, New York, 2007.
  • [23] Oscar Randal-Williams. The stable cohomology of automorphisms of free groups with coefficients in the homology representation. arXiv:1012.1433, 2010.
  • [24] Oscar Randal-Williams. The space of immersed surfaces in a manifold. Math. Proc. Cambridge Philos. Soc., 154(3):419–438, 2013.
  • [25] Oscar Randal-Williams and Nathalie Wahl. Homological stability for automorphism groups. arXiv:1409.3541, 2014.
  • [26] Takao Satoh. Twisted first homology groups of the automorphism group of a free group. J. Pure Appl. Algebra, 204(2):334–348, 2006.
  • [27] Takao Satoh. Twisted second homology groups of the automorphism group of a free group. J. Pure Appl. Algebra, 211(2):547–565, 2007.
  • [28] Christine Vespa. Extensions between functors from groups. arXiv:1511.03098, 2015.