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

Homological stability for quotients of mapping class groups of surfaces by the Johnson subgroups

Tomáš Zeman
Abstract

We study quotients of mapping class groups Γg,1\Gamma_{g,1} of oriented surfaces with one boundary component by terms of their Johnson filtrations, and we show that the homology of these quotients with suitable systems of twisted coefficients stabilises as the genus of the surface goes to infinity. We also compute the stable (co)homology with constant rational coefficients for one family of such quotients.

1 Introduction and statement of results

Let Sg,nS_{g,n} be an oriented surface of genus gg with nn boundary components, and let Γg,n\Gamma_{g,n} be its mapping class group, i.e. the group of isotopy classes of orientation-preserving homeomorphisms of Sg,kS_{g,k} that fix a neighbourhood of the boundary pointwise.

Gluing on a pair of pants S0,3S_{0,3} along one or two boundary components yields embeddings Sg,nSg,n+1S_{g,n}\hookrightarrow S_{g,n+1} and Sg,nSg+1,n1S_{g,n}\hookrightarrow S_{g+1,n-1} respectively, and these in turn give rise to homomorphisms Γg,nΓg,n+1\Gamma_{g,n}\to\Gamma_{g,n+1} and Γg,nΓg+1,n1\Gamma_{g,n}\to\Gamma_{g+1,n-1} which extend a given mapping class by the identity on S0,3S_{0,3}. Harer and Ivanov showed that these homomorphisms induce isomorphisms on group homology in degrees which are small compared to the genus. In particular, the sequence S0,1S1,1S_{0,1}\hookrightarrow S_{1,1}\hookrightarrow\cdots of embeddings constructed by attaching S1,2S_{1,2} to the unique boundary component of S,1S_{\ast,1} yields a sequence of homomorphisms Γ0,1Γ1,1\Gamma_{0,1}\to\Gamma_{1,1}\to\ldots which satisfies homological stability in the sense that

Hi(Γg1,1;)Hi(Γg,1;)H_{i}(\Gamma_{g-1,1};\mathbb{Z})\to H_{i}(\Gamma_{g,1};\mathbb{Z})

is an isomorphism whenever i23g43i\leqslant\frac{2}{3}g-\frac{4}{3} (see e.g. Wahl’s survey [16]).

If we fix a base-point Sg,1\ast\in\partial S_{g,1}, the fundamental group ππ1(Sg,1,)\pi\coloneqq\pi_{1}(S_{g,1},\ast) enjoys an action of Γg,1\Gamma_{g,1}. This action preserves the lower central series222We use the indexing convention γ1(π)=π\gamma_{1}(\pi)=\pi. {γk(π)}k1\{\gamma_{k}(\pi)\}_{k\geqslant 1} of π\pi, and the Johnson filtration {g,1(k)}k1\{\mathscr{I}_{g,1}(k)\}_{k\geqslant 1} of Γg,1\Gamma_{g,1} consists of kernels of the induced actions of Γg,1\Gamma_{g,1} on the quotients π/γk+1(π)\pi/\gamma_{k+1}(\pi). It is easy to show that the above homomorphisms Γg1,1Γg,1\Gamma_{g-1,1}\to\Gamma_{g,1} preserve the Johnson filtration. Hence a natural question arises whether the induced homomorphisms Γg1,1/g1,1(k)Γg,1/g,1(k)\Gamma_{g-1,1}/\mathscr{I}_{g-1,1}(k)\to\Gamma_{g,1}/\mathscr{I}_{g,1}(k) are also homology isomorphisms when gg is sufficiently large.

Theorem A.

Let k1k\geqslant 1, and let M0M1M_{0}\to M_{1}\to\cdots be a good module of the sequence of groups Γ0,1/0,1(k)Γ1,11,1(k)\Gamma_{0,1}/\mathscr{I}_{0,1}(k)\to\Gamma_{1,1}\mathscr{I}_{1,1}(k)\to\cdots. Then for every ii, the homomorphisms in the sequence

Hi(Γg1,1/g1,1(k);Mg1)Hi(Γg,1/g,1(k);Mg)\cdots\to H_{i}(\Gamma_{g-1,1}/\mathscr{I}_{g-1,1}(k);M_{g-1})\to H_{i}(\Gamma_{g,1}/\mathscr{I}_{g,1}(k);M_{g})\to\cdots

are eventually (for large gg) isomorphisms.

See Definition 5.7 for the precise meaning of a good module of a sequence of groups; in essence, these are modules that come from well-behaved sequences of representations of the symplectic groups Sp(2g,)\mathrm{Sp}(2g,\mathbb{Z}). In particular, for any abelian group AA, the constant coefficients in AA constitute a good module.

Our proof is by induction, bootstrapping known results about homological stability of the symplectic groups Sp(2g,)Γg,1/g,1(1)\mathrm{Sp}(2g,\mathbb{Z})\cong\Gamma_{g,1}/\mathscr{I}_{g,1}(1) to the higher quotients Γg,1/g,1(k)\Gamma_{g,1}/\mathscr{I}_{g,1}(k), k>1k>1, by means of a spectral sequence argument applied to the extensions

0g,1(k1)/g,1(k)Γg,1/g,1(k)Γg,1/g,1(k1)1.0\to\mathscr{I}_{g,1}(k-1)/\mathscr{I}_{g,1}(k)\to\Gamma_{g,1}/\mathscr{I}_{g,1}(k)\to\Gamma_{g,1}/\mathscr{I}_{g,1}(k-1)\to 1.

Our approach has a drawback in that we get no hold on the stable range: the heart of our argument is Theorem 4.7 which says that certain 𝖥𝖨\mathsf{FI}-modules which encapsulate the successive quotients g,1(k1)/g,1(k)\mathscr{I}_{g,1}(k-1)/\mathscr{I}_{g,1}(k) are finitely generated. Its proof relies on the noetherian property of the category of 𝖥𝖨\mathsf{FI}-modules over \mathbb{Z} which was proved by Church, Ellenberg, Farb and Nagpal in [5]. This property is non-constructive in the sense that it guarantees the existence of a finite generating set of a sub-𝖥𝖨\mathsf{FI}-module of a finitely generated 𝖥𝖨\mathsf{FI}-module, but it offers no information on the size or degrees of some finite set of generators.

The inspiration to use the above bootstrapping method came from a paper of Szymik [11] where this method is applied to the closely related automorphism groups of free nilpotent groups. Szymik considers the extensions

0Kn(k)AutNn(k)AutNn(k1)10\to K_{n}(k)\to\operatorname{Aut}N_{n}(k)\to\operatorname{Aut}N_{n}(k-1)\to 1

where Nn(k)N_{n}(k) is the free nilpotent group on nn generators, of nilpotency class kk. If we write 𝔽n\mathbb{F}_{n} for the free group on nn generators, then Nn(k)𝔽n/γk+1(𝔽n)N_{n}(k)\cong\mathbb{F}_{n}/\gamma_{k+1}(\mathbb{F}_{n}). The kernels Kn(k)K_{n}(k) in the above extensions are relatively easy to understand, allowing for extraction of linear stability bounds from Szymik’s argument. Indeed, one can show that

Kn(k)Hom(n,γk(𝔽n)/γk+1(𝔽n))K_{n}(k)\cong\operatorname{Hom}(\mathbb{Z}^{n},\gamma_{k}(\mathbb{F}_{n})/\gamma_{k+1}(\mathbb{F}_{n}))

which is a free abelian group. The quotients g,1(k1)/g,1(k)\mathscr{I}_{g,1}(k-1)/\mathscr{I}_{g,1}(k), on the other hand, seem more difficult to analyse, and while they become subgroups of K2g(k)K_{2g}(k) upon fixing an isomorphism π1(Sg,1,)𝔽2g\pi_{1}(S_{g,1},\ast)\cong\mathbb{F}_{2g}, it is not clear to us which subgroups. It is for this reason that we use the techniques of representation stability and consequently lose information about the stable range.

The pair-of-pants product endows the space g0BΓg,1\coprod_{g\geqslant 0}B\Gamma_{g,1} with a monoid structure, and the group completion of this monoid is (weakly equivalent to) ×BΓ+\mathbb{Z}\times B\Gamma_{\infty}^{+} by the group-completion theorem, where Γ=colimgΓg,1\Gamma_{\infty}=\operatorname{colim}_{g}\Gamma_{g,1} and X+X^{+} is the Quillen plus construction on XX with respect to the maximal perfect subgroup of π1(X)\pi_{1}(X). Tillmann [13, 14] showed that ×BΓ+\mathbb{Z}\times B\Gamma_{\infty}^{+} is in fact an infinite loop space, and that the obvious map ×BΓ+KSp()\mathbb{Z}\times B\Gamma_{\infty}^{+}\to K_{\mathrm{Sp}}(\mathbb{Z}) to the Hermitian KK-theory of \mathbb{Z} is a map of infinite loop spaces. We show that analogous constructions for the quotients Γg,1/g,1(k)\Gamma_{g,1}/\mathscr{I}_{g,1}(k) also yield infinite loop spaces and that we get an infinite tower of infinite loop spaces interpolating between ×BΓ+\mathbb{Z}\times B\Gamma_{\infty}^{+} and KSp()K_{\mathrm{Sp}}(\mathbb{Z}).

Theorem B.

The obvious quotient homomorphisms of groups give rise to a commutative diagram

×BΓ+{\mathbb{Z}\times B\Gamma_{\infty}^{+}}{\cdots}×BΓ(3)+{\mathbb{Z}\times B\Gamma_{\infty}(3)^{+}}×BΓ(2)+{\mathbb{Z}\times B\Gamma_{\infty}(2)^{+}}×BΓ(1)+{\mathbb{Z}\times B\Gamma_{\infty}(1)^{+}}KSp(){K_{\mathrm{Sp}}(\mathbb{Z})}{\simeq}

of maps of infinite loop spaces.

Outline of paper

In Section 2, we discuss 𝒞\mathscr{C}-modules, i.e. functors from a small category 𝒞\mathscr{C} to the category of abelian groups, and we introduce the two choices of 𝒞\mathscr{C} that will be of interest to us, namely the category 𝖥𝖨\mathsf{FI} of finite sets and injections, and the category 𝖲𝖨\mathsf{SI} of finitely generated free symplectic spaces over \mathbb{Z} and symplectic maps. In Sections 3 and 4, we recall some facts about automorphism groups of free nilpotent groups, we construct the 𝖲𝖨\mathsf{SI}-modules Q(k)Q(k) which encapsulate the successive quotients g,1(k1)/g,1(k)\mathscr{I}_{g,1}(k-1)/\mathscr{I}_{g,1}(k) as Sp(2g,)\mathrm{Sp}(2g,\mathbb{Z})-modules, and we prove the key Theorem 4.7. Sections 5 and 6 are devoted to the proofs of Theorem A and B, respectively.

1.1 Acknowledgements

I would like to thank my supervisor Ulrike Tillmann for her constant encouragement and many useful discussions, as well as for her suggestions which greatly improved the exposition of this paper.

1.2 Notation

We use the following conventions: {0,1,2,}\mathbb{N}\coloneqq\{0,1,2,\ldots\} is the set of natural numbers including zero; 𝖦𝗋𝗉\mathsf{Grp} and 𝖠𝖻\mathsf{Ab} are the categories of groups and abelian groups, respectively; Σn\Sigma_{n} is the symmetric group on nn letters; Sp(2n,R)\mathrm{Sp}(2n,R) is the group of 2n×2n2n\times 2n symplectic matrices with entries in a ring RR — for us, R=R=\mathbb{Z} most of the time. In a group GG, we write [x,y]=xyx1y1[x,y]=xyx^{-1}y^{-1} for the commutator of x,yGx,y\in G.

2 Modules of finite degree

2.1 Modules over small categories

Let 𝒞\mathscr{C} be a small category with objects (canonically identified with) the natural numbers \mathbb{N}.

Definition 2.1.

A 𝒞\mathscr{C}-object in a category 𝒜\mathscr{A} is a functor 𝒞𝒜\mathscr{C}\to\mathscr{A}. A morphism of 𝒞\mathscr{C}-objects in 𝒜\mathscr{A} is a natural transformation of functors.

Notation.

We write FnF_{n} for the value of a 𝒞\mathscr{C}-object FF at n𝒞n\in\mathscr{C}. When f:nmf:n\to m is an arrow in 𝒞\mathscr{C}, we write f:FnFmf_{\ast}:F_{n}\to F_{m} for the induced morphism in 𝒜\mathscr{A}. When ϕ:FF\phi:F\to F^{\prime} is a morphism of 𝒞\mathscr{C}-objects, we write ϕn:FnFn\phi_{n}:F_{n}\to F^{\prime}_{n} for its component at n𝒞n\in\mathscr{C}.

Definition 2.2.

A 𝒞\mathscr{C}-module is a 𝒞\mathscr{C}-object in 𝖠𝖻\mathsf{Ab}. We write 𝒞–Mod\mathscr{C}\mbox{--Mod} for the category of 𝒞\mathscr{C}-modules and natural transformations.

Remark 2.3.

𝒞–Mod\mathscr{C}\mbox{--Mod} is abelian, with kernels and cokernels computed pointwise. Hence it makes good sense to talk of sub-𝒞\mathscr{C}-modules, quotient 𝒞\mathscr{C}-modules etc.

Now suppose further that the category 𝒞\mathscr{C} is strict monoidal, where the monoidal product \oplus is given by nm=n+mn\oplus m=n+m on objects, and that the monoidal unit 0 is initial in 𝒞\mathscr{C}.

Definition 2.4.

We will call such 𝒞\mathscr{C} a category with objects the naturals, or CON for short.

Consider the functor R1():𝒞𝒞R\coloneqq 1\oplus(-):\mathscr{C}\to\mathscr{C}. Since 0 is initial, the unique arrow 010\to 1 induces a natural transformation ρ:IdR\rho:\operatorname{Id}\Rightarrow R from the identity Id=0()\operatorname{Id}=0\oplus(-) to RR.

Definition 2.5.

Let FF be a 𝒞\mathscr{C}-module. The suspension of FF is ΣFFR\Sigma F\coloneqq F\circ R. We define two new 𝒞\mathscr{C}-modules, called the kernel and cokernel of FF and denoted KerF\operatorname{Ker}F and CokerF\operatorname{Coker}F, to be the kernel and cokernel of the morphism Fρ:FΣFF\rho:F\to\Sigma F of 𝒞\mathscr{C}-modules.

Notation.

To avoid any confusion, ker\ker and coker\operatorname{coker} with small initial letters will always refer to the usual kernel and cokernel in 𝒞–Mod\mathscr{C}\mbox{--Mod} (or elsewhere) as an abelian category, whereas Ker\operatorname{Ker} and Coker\operatorname{Coker} with capitals will always denote the above constructions.

Example 2.6.

When 𝒞\mathscr{C} is the poset \mathbb{N} with the evident monoidal structure, a 𝒞\mathscr{C}-module FF is just a composable sequence F0F1F_{0}\to F_{1}\to\cdots of homomorphisms of abelian groups. The morphism Fρ:FΣFF\rho:F\to\Sigma F is

F0{F_{0}}F1{F_{1}}{\cdots}F1{F_{1}}F2{F_{2}}{\cdots}

KerF\operatorname{Ker}F consists of the kernels of the homomorphisms in FF and zero maps, and similarly for the cokernel.

Observe that Ker\operatorname{Ker} and Coker\operatorname{Coker} are in fact functors 𝒞–Mod𝒞–Mod\mathscr{C}\mbox{--Mod}\to\mathscr{C}\mbox{--Mod}. We have the following easy lemma.

Lemma 2.7.

If 0F𝜄F𝜋F′′00\to F^{\prime}\xrightarrow{\iota}F\xrightarrow{\pi}F^{\prime\prime}\to 0 is a short exact sequence of 𝒞\mathscr{C}-modules, there is an exact sequence

0KerFKerιKerFKerπKerF′′CokerFCokerιCokerFCokerπCokerF′′0.0\to\operatorname{Ker}F^{\prime}\xrightarrow{\operatorname{Ker}\iota}\operatorname{Ker}F\xrightarrow{\operatorname{Ker}\pi}\operatorname{Ker}F^{\prime\prime}\to\operatorname{Coker}F^{\prime}\xrightarrow{\operatorname{Coker}\iota}\operatorname{Coker}F\xrightarrow{\operatorname{Coker}\pi}\operatorname{Coker}F^{\prime\prime}\to 0.
Proof.

Immediate from the snake lemma. ∎

Let 𝒟\mathscr{D} be another CON, and let X:𝒟𝒞X:\mathscr{D}\to\mathscr{C} be a strict monoidal functor which is the identity on objects. XX induces a pullback functor X:𝒞–Mod𝒟–ModX^{\ast}:\mathscr{C}\mbox{--Mod}\to\mathscr{D}\mbox{--Mod} taking FF to FXF\circ X. Moreover, we have the following lemma.

Lemma 2.8.

XX^{\ast} commutes with the functors Σ\Sigma, Ker\operatorname{Ker} and Coker\operatorname{Coker}. In other words, for a 𝒞\mathscr{C}-module FF, X(ΣF)=Σ(XF)X^{\ast}(\Sigma F)=\Sigma(X^{\ast}F), etc.

Proof.

This is immediate once we note that since XX is strict monoidal and the identity on objects, we have R𝒞X=XR𝒟R_{\mathscr{C}}X=XR_{\mathscr{D}} and ρ𝒞X=Xρ𝒟\rho_{\mathscr{C}}X=X\rho_{\mathscr{D}}. Here R𝒞R_{\mathscr{C}} and ρ𝒞\rho_{\mathscr{C}} denote the above functor and natural transformation in 𝒞\mathscr{C}, and similarly for 𝒟\mathscr{D}. ∎

Definition 2.9.

We will call a strict monoidal functor X:𝒟𝒞X:\mathscr{D}\to\mathscr{C} which is the identity on objects a functor of categories with objects the naturals, or a functor of CONs for short.

2.2 The degree of a 𝒞\mathscr{C}-module

Let 𝒞\mathscr{C} be a CON, and let FF be a 𝒞\mathscr{C}-module.

Definition 2.10.

FF has degree 1-1 if Fn=0F_{n}=0 for all sufficiently large nn. For d0d\geqslant 0, FF has degree dd if KerF\operatorname{Ker}F has degree 1-1 and CokerF\operatorname{Coker}F has degree d1d-1. FF has finite degree if FF has degree dd for some d1d\geqslant-1.

The key observation of this section is the following lemma.

Lemma 2.11.

Suppose X:𝒟𝒞X:\mathscr{D}\to\mathscr{C} is a functor of CONs. Then a 𝒞\mathscr{C}-module FF has degree dd if and only if the 𝒟\mathscr{D}-module XFX^{\ast}F has degree dd.

Proof.

This follows from an easy induction on dd using Lemma 2.8. Since XX is the identity on objects, (XF)n=Fn(X^{\ast}F)_{n}=F_{n}, so FF has degree 1-1 if and only if XFX^{\ast}F does. Assuming the result for d1d-1, the following are equivalent:

  • FF has degree dd;

  • KerF\operatorname{Ker}F has degree 1-1 and CokerF\operatorname{Coker}F has degree d1d-1;

  • XKerF=KerXFX^{\ast}\operatorname{Ker}F=\operatorname{Ker}X^{\ast}F has degree 1-1 and XCokerF=CokerXFX^{\ast}\operatorname{Coker}F=\operatorname{Coker}X^{\ast}F has degree d1d-1;

  • XFX^{\ast}F has degree dd.

2.3 The categories 𝖥𝖨\mathsf{FI} and 𝖲𝖨\mathsf{SI}

In this section we introduce two more examples of CONs beyond \mathbb{N}, namely the categories 𝖥𝖨\mathsf{FI} of finite sets and injections, and 𝖲𝖨\mathsf{SI} of free finitely generated abelian groups with symplectic forms and symplectic injections.

Definition 2.12.

Let 𝖥𝖨\mathsf{FI} be the category of sets of the form 𝐧{1,2,,n}\mathbf{n}\coloneqq\{1,2,\ldots,n\}, n0n\geqslant 0, and set injections.

𝖥𝖨\mathsf{FI} is strict monoidal with the monoidal product \sqcup coming from disjoint union of sets: 𝐧𝐦=𝐧+𝐦={1,2,,n+m}\mathbf{n}\sqcup\mathbf{m}=\mathbf{n+m}=\{1,2,\ldots,n+m\} on objects and

(fg)(i)={f(i)if ing(in)+mif i>n(f\sqcup g)(i)=\begin{cases}f(i)&\mbox{if }i\leqslant n\\ g(i-n)+m&\mbox{if }i>n\end{cases}

on morphisms f:𝐧𝐦f:\mathbf{n}\to\mathbf{m} and g:𝐤𝐥g:\mathbf{k}\to\mathbf{l}. It is easily seen that this is indeed a CON.

Definition 2.13.

Let VnV_{n}, n0n\geqslant 0, be the free abelian group with basis a1,b1,a2,b2,,an,bna_{1},b_{1},a_{2},b_{2},\ldots,a_{n},b_{n}, equipped with the symplectic form (,)(\cdot,\cdot) given by

(ai,aj)=0=(bi,bj)\displaystyle(a_{i},a_{j})=0=(b_{i},b_{j})
(ai,bj)=δij=(bj,ai)\displaystyle(a_{i},b_{j})=\delta_{ij}=-(b_{j},a_{i})

for all 1i,jn1\leqslant i,j\leqslant n, where δij\delta_{ij} is the Kronecker delta. Let 𝖲𝖨\mathsf{SI} be the category with objects the VnV_{n} and arrows the form-preserving homomorphisms.

The following lemma is standard linear algebra. We quote it here for later reference.

Lemma 2.14.
  1. (i)

    End𝖲𝖨(Vn)=Aut𝖲𝖨(Vn)=Sp(2n,)\operatorname{End}_{\mathsf{SI}}(V_{n})=\operatorname{Aut}_{\mathsf{SI}}(V_{n})=\mathrm{Sp}(2n,\mathbb{Z})

  2. (ii)

    For every f:VnVmf:V_{n}\to V_{m} in 𝖲𝖨\mathsf{SI}, there is a unique CVmC\leqslant V_{m} such that Vm(imf)CV_{m}\cong(\operatorname{im}f)\oplus C as abelian groups with bilinear forms (where the form on CC is the restriction of the form on VmV_{m}). Moreover, CVmnC\cong V_{m-n}.

  3. (iii)

    If ff and CC are as above and sAut𝖲𝖨(Vm)s\in\operatorname{Aut}_{\mathsf{SI}}(V_{m}) satisfies f=sff=s\circ f, then ss preserves the above direct sum decomposition and s=idss=\operatorname{id}\oplus s^{\prime} for some sAut(C)s^{\prime}\in\operatorname{Aut}(C).

𝖲𝖨\mathsf{SI} is strict monoidal with monoidal product \boxtimes coming from direct sum. Writing VnVmV_{n}\oplus V_{m} for the usual direct sum of abelian groups with bilinear forms, let αn,m:VnVmVn+m\alpha_{n,m}:V_{n}\oplus V_{m}\xrightarrow{\sim}V_{n+m} be the obvious form-preserving isomorphism which takes the bases of VnV_{n} and VmV_{m} to the first 2n2n and last 2m2m basis elements of Vn+mV_{n+m}, respectively, in an order-preserving fashion. We define VnVmVn+mV_{n}\boxtimes V_{m}\coloneqq V_{n+m} and we let fgf\boxtimes g be the composition

Vn+kαn,k1VnVkfgVmVlαm,lVm+lV_{n+k}\xrightarrow{\alpha_{n,k}^{-1}}V_{n}\oplus V_{k}\xrightarrow{f\oplus g}V_{m}\oplus V_{l}\xrightarrow{\alpha_{m,l}}V_{m+l}

for morphisms f:VnVmf:V_{n}\to V_{m}, g:VkVlg:V_{k}\to V_{l}. Again it is easily checked that this makes 𝖲𝖨\mathsf{SI} into a CON.

Definition 2.15.

Let X:𝖥𝖨𝖲𝖨X:\mathsf{FI}\to\mathsf{SI} be the functor which sends 𝐧\mathbf{n} to VnV_{n} and f:𝐧𝐦f:\mathbf{n}\to\mathbf{m} to the homomorphism Xf:VnVmXf:V_{n}\to V_{m} defined on the standard basis by Xf(ai)=af(i)Xf(a_{i})=a_{f(i)} and Xf(bi)=bf(i)Xf(b_{i})=b_{f(i)}.

It is readily seen that XX is a functor of CONs, and so Lemma 2.11 applies.

Remark.

The main reasons why the definition of VnV_{n} contains a reference to a specific basis is that it gives us for free a coherent choice of the isomorphisms αn,m\alpha_{n,m}, and the construction of the above functor XX becomes rather easy. If we defined the VnV_{n} simply as free abelian groups of rank 2n2n equipped with a symplectic form, we would have to do more work to construct \boxtimes and XX.

2.4 𝖥𝖨\mathsf{FI}-modules

In this section, we investigate 𝖥𝖨\mathsf{FI}-modules in greater detail, and give a complete characterisation of 𝖥𝖨\mathsf{FI}-modules of finite degree.

Basic notions

𝖥𝖨\mathsf{FI}-modules have been much studied in recent years in connection to stable representation theory of symmetric groups, and there is a wealth of literature on them. We define all the relevant terms and state the results we will be using later on, but see e.g. [5] for a more detailed introduction.

Definition 2.16.

For n0n\geqslant 0, the nn-th principal projective PnP^{n} is the 𝖥𝖨\mathsf{FI}-module which sends 𝐦\mathbf{m} to the free abelian group generated by the set 𝖥𝖨(𝐧,𝐦)\mathsf{FI}(\mathbf{n},\mathbf{m}), and similarly for arrows.

An 𝖥𝖨\mathsf{FI}-module is free if it is a direct sum of principal projectives.

The PnP^{n} are indeed projective objects of 𝖥𝖨–Mod\mathsf{FI}\mbox{--Mod}. By the Yoneda lemma, there is an isomorphism 𝖥𝖨–Mod(Pn,F)Fn\mathsf{FI}\mbox{--Mod}(P^{n},F)\cong F_{n}, natural in F𝖥𝖨–ModF\in\mathsf{FI}\mbox{--Mod}, which sends ϕ:PnF\phi:P^{n}\to F to the image under ϕn:PnnFn\phi_{n}:P_{n}^{n}\to F_{n} of the identity id𝐧Pnn\operatorname{id}_{\mathbf{n}}\in P_{n}^{n}. The image of ϕ\phi in FF can be seen to be the least sub-𝖥𝖨\mathsf{FI}-module of FF which contains ϕn(id𝐧)Fn\phi_{n}(\operatorname{id}_{\mathbf{n}})\in F_{n}.

Similarly, homomorphisms

iIPniF\bigoplus_{i\in I}P^{n_{i}}\to F

from a fixed free 𝖥𝖨\mathsf{FI}-module biject naturally with II-indexed sequences (xiFni)iI(x_{i}\in F_{n_{i}})_{i\in I}, and the image of such a homomorphism is the least sub-𝖥𝖨\mathsf{FI}-module of FF containing xiFnix_{i}\in F_{n_{i}} for all ii ([5]). This motivates the following definition.

Definition 2.17.

Let FF be an 𝖥𝖨\mathsf{FI}-module. We say FF is generated in degrees k\leqslant k if FF receives a surjection

iIPniF\bigoplus_{i\in I}P^{n_{i}}\twoheadrightarrow F (1)

from a free module with nikn_{i}\leqslant k for all iIi\in I. FF is generated in finite degree if it is generated in degrees k\leqslant k for some kk.

FF is finitely generated if there is a surjection (1) with II finite.

We say FF is generated in degrees k\leqslant k and related in degrees d\leqslant d if there is an exact sequence

jJPmjiIPniF0\bigoplus_{j\in J}P^{m_{j}}\to\bigoplus_{i\in I}P^{n_{i}}\to F\to 0 (2)

where each nikn_{i}\leqslant k and each mjdm_{j}\leqslant d. FF is generated and related in finite degrees if it is generated in degrees k\leqslant k and related in degrees d\leqslant d for some kk and dd.

Remark 2.18.

Note that a surjection iIPniF\bigoplus_{i\in I}P^{n_{i}}\twoheadrightarrow F extends to an exact sequence as in (2) if and only if its kernel is generated in degrees d\leqslant d. Hence an equivalent definition of FF being generated in degrees k\leqslant k and related in degrees d\leqslant d is that there is a short exact sequence

0KiIPniF00\to K\to\bigoplus_{i\in I}P^{n_{i}}\to F\to 0

where each nikn_{i}\leqslant k and where KK is generated in degrees d\leqslant d.

The main result of [5] (and the reason why 𝖥𝖨\mathsf{FI}-modules are useful to us) is the following theorem.

Theorem 2.19 ([5]).

The category 𝖥𝖨–Mod\mathsf{FI}\mbox{--Mod} is noetherian: every sub-𝖥𝖨\mathsf{FI}-module of a finitely generated 𝖥𝖨\mathsf{FI}-module is finitely generated.

𝖥𝖨\mathsf{FI}-modules of finite degree

We record the following result from [3, Section 4] which will be useful later on.

Lemma 2.20.

An 𝖥𝖨\mathsf{FI}-module FF is generated in finite degree if and only if CokernF\operatorname{Coker}^{n}F has degree 1-1 for some nn (where Coker\operatorname{Coker} is the functor from Definition 2.5).

Proof.

[3] proves the following closely related statement: FF is generated in finite degree if and only if CokernF=0\operatorname{Coker}^{n}F=0 for some nn. Our statement is equivalent: the zero 𝖥𝖨\mathsf{FI}-module certainly has degree 1-1. Conversely, if CokernF\operatorname{Coker}^{n}F has degree 1-1, then there exists MM such that (CokernF)m=0(\operatorname{Coker}^{n}F)_{m}=0 for all mMm\geqslant M. But then Cokern+MF=0\operatorname{Coker}^{n+M}F=0. ∎

It can be seen that the definition of an 𝖥𝖨\mathsf{FI}-module FF having finite degree does not depend on the structure of FF in any finite set of degrees: FF has degree 1-1 if it is eventually 0, and higher degrees are defined recursively. Hence the following definition will be useful in our investigation of modules of finite degree.

Definition 2.21.

We say that a morphism ϕ:FF\phi:F\to F^{\prime} of 𝖥𝖨\mathsf{FI}-modules is eventually isomorphic if ϕn:FnFn\phi_{n}:F_{n}\to F^{\prime}_{n} is an isomorphism for all sufficiently large nn.

Suppose that 0KP𝜋F00\to K\to P\xrightarrow{\pi}F\to 0 is a short exact sequence of 𝖥𝖨\mathsf{FI}-modules. Coker\operatorname{Coker} is right-exact, so for every i0i\geqslant 0, the sequence

CokeriKCokeriPCokeriπCokeriF0\operatorname{Coker}^{i}K\to\operatorname{Coker}^{i}P\xrightarrow{\operatorname{Coker}^{i}\pi}\operatorname{Coker}^{i}F\to 0

is exact. Let K(i)K^{(i)} be the kernel of the morphism Cokeriπ\operatorname{Coker}^{i}\pi, so K(0)=KK^{(0)}=K and there is an induced morphism ui:CokeriKK(i)u_{i}:\operatorname{Coker}^{i}K\to K^{(i)}.

Lemma 2.22.

Let 0KPF00\to K\to P\to F\to 0 and K(i)K^{(i)} be as above, and suppose that FF has finite degree. Then for every ii, ui:CokeriKK(i)u_{i}:\operatorname{Coker}^{i}K\to K^{(i)} is eventually isomorphic.

Proof.

We proceed by induction on ii. When i=0i=0, there is nothing to prove.

Suppose the claim has been established for i1i-1. Applying Coker\operatorname{Coker} to the short exact sequence 0K(i1)Cokeri1PCokeri1F00\to K^{(i-1)}\to\operatorname{Coker}^{i-1}P\to\operatorname{Coker}^{i-1}F\to 0 yields the exact sequence

KerCokeri1F{\operatorname{Ker}\operatorname{Coker}^{i-1}F}CokerK(i1){\operatorname{Coker}K^{(i-1)}}CokeriP{\operatorname{Coker}^{i}P}CokeriF{\operatorname{Coker}^{i}F}0{0}K(i){K^{(i)}}q\scriptstyle{q}

by Lemma 2.7. Since FF has finite degree, KerCokeri1F\operatorname{Ker}\operatorname{Coker}^{i-1}F has degree 1-1, so the morphism q:CokerK(i1)K(i)q:\operatorname{Coker}K^{(i-1)}\to K^{(i)} is eventually isomorphic. Cokerui1\operatorname{Coker}u_{i-1} is eventually isomorphic by the inductive hypothesis. Hence uiu_{i}, which is equal to the composition

CokeriKCokerui1CokerK(i1)𝑞K(i),\operatorname{Coker}^{i}K\xrightarrow{\operatorname{Coker}u_{i-1}}\operatorname{Coker}K^{(i-1)}\xrightarrow{q}K^{(i)},

is also eventually isomorphic. ∎

Now we are ready to characterise 𝖥𝖨\mathsf{FI}-modules of finite degree.

Proposition 2.23.

Let FF be an 𝖥𝖨\mathsf{FI}-module. Then FF has finite degree if and only if FF is generated and related in finite degrees.

Proof.

The statement that 𝖥𝖨\mathsf{FI}-modules which are generated and related in finite degrees have finite degree, is precisely the content of [9, Proposition 4.18].

To prove the other direction, suppose FF has finite degree. Then FF is generated in finite degree by Lemma 2.20, so there is a short exact sequence

0KLF00\to K\to L\to F\to 0

where LL is free and generated in finite degree. If we can show that KK is generated in finite degree, we are done by Remark 2.18.

Using the notation of Lemma 2.22, there are eventual isomorphisms CokeriKK(i)\operatorname{Coker}^{i}K\to K^{(i)}. But by Lemma 2.20, CokeriL\operatorname{Coker}^{i}L has degree 1-1 for some ii. Since K(i)K^{(i)} is a sub-𝖥𝖨\mathsf{FI}-module of CokeriL\operatorname{Coker}^{i}L, it also has degree 1-1, and so CokeriK\operatorname{Coker}^{i}K has degree 1-1 by virtue of being eventually isomorphic to K(i)K^{(i)}. Hence KK is generated in finite degree. ∎

Corollary 2.24.

Finitely generated 𝖥𝖨\mathsf{FI}-modules have finite degree.

Proof.

Suppose FF is a finitely generated 𝖥𝖨\mathsf{FI}-module, i.e. there is a surjection as in (1) with II finite. The kernel of this surjection is finitely generated by Theorem 2.19, so FF is generated and related in finite degrees. ∎

To prove our second corollary, we need the following technical lemma.

Lemma 2.25.

For every n,mn,m, the 𝖥𝖨\mathsf{FI}-module PnPm:𝖥𝖨Pn×Pm𝖠𝖻×𝖠𝖻𝖠𝖻P^{n}\otimes P^{m}:\mathsf{FI}\xrightarrow{P^{n}\times P^{m}}\mathsf{Ab}\times\mathsf{Ab}\xrightarrow{\otimes}\mathsf{Ab} is free and finitely generated in degrees n+m\leqslant n+m.

Proof.

Write FPnPmF\coloneqq P^{n}\otimes P^{m}. For every kk, Fk=PknPkmF_{k}=P_{k}^{n}\otimes P_{k}^{m} is a free abelian group with basis all pairs (f:𝐧𝐤,g:𝐦𝐤)(f:\mathbf{n}\hookrightarrow\mathbf{k},g:\mathbf{m}\hookrightarrow\mathbf{k}) of injections. To each such pair, we associate the following data:

N\displaystyle N f1(imfimg)𝐧\displaystyle\coloneqq f^{-1}(\operatorname{im}f\cap\operatorname{im}g)\subset\mathbf{n}
M\displaystyle M g1(imfimg)𝐦\displaystyle\coloneqq g^{-1}(\operatorname{im}f\cap\operatorname{im}g)\subset\mathbf{m}
σ\displaystyle\sigma N𝑓imfimgg1M.\displaystyle\coloneqq N\xrightarrow{f}\operatorname{im}f\cap\operatorname{im}g\xrightarrow{g^{-1}}M.

Now observe that if h:𝐤𝐥h:\mathbf{k}\to\mathbf{l} is a morphism in 𝖥𝖨\mathsf{FI}, then h(f,g)=(hf,hg)Flh_{\ast}(f,g)=(h\circ f,h\circ g)\in F_{l} has the same associated data as (f,g)(f,g). It follows that FF splits as a direct sum of factors T(N,M,σ)T^{(N,M,\sigma)} indexed by triples (N,M,σ)(N,M,\sigma) where N𝐧N\subset\mathbf{n}, M𝐦M\subset\mathbf{m} and σ:NM\sigma:N\xrightarrow{\sim}M is an isomorphism: Tk(N,M,σ)T_{k}^{(N,M,\sigma)} is the subgroup of FkF_{k} generated by precisely those basis elements (f,g)(f,g) whose associated data are the prescribed (N,M,σ)(N,M,\sigma).

Now we claim that each T(N,M,σ)T^{(N,M,\sigma)} is in fact a principal projective. Indeed, generators (f,g)(f,g) of Tk(N,M,σ)T_{k}^{(N,M,\sigma)} biject naturally (in 𝐤𝖥𝖨\mathbf{k}\in\mathsf{FI}) with maps 𝐧σ𝐦𝐤\mathbf{n}\cup_{\sigma}\mathbf{m}\to\mathbf{k}, where 𝐧σ𝐦\mathbf{n}\cup_{\sigma}\mathbf{m} is the pushout of 𝐧N𝜎𝐦\mathbf{n}\hookleftarrow N\xrightarrow{\sigma}\mathbf{m}. Hence T(N,M,σ)T^{(N,M,\sigma)} is isomorphic to the principal projective generated in degree |𝐧σ𝐦|=n+m|N|n+m|\mathbf{n}\cup_{\sigma}\mathbf{m}|=n+m-|N|\leqslant n+m. ∎

Remark.

Note that rankPkn=n!(kn)\operatorname{rank}_{\mathbb{Z}}P_{k}^{n}=n!\binom{k}{n}. By computing the rank of PknPkmP_{k}^{n}\otimes P_{k}^{m} in two ways using the above proof, we obtain the following rather nice identity

aa!(n+ma)!(na)(ma)(kn+ma)=n!m!(kn)(km)\sum_{a}a!(n+m-a)!\binom{n}{a}\binom{m}{a}\binom{k}{n+m-a}=n!m!\binom{k}{n}\binom{k}{m}

where a!(na)(ma)a!\binom{n}{a}\binom{m}{a} is the number of different triples (N,M,σ)(N,M,\sigma) with |N|=a|N|=a.

Corollary 2.26.

If F,FF,F^{\prime} are 𝖥𝖨\mathsf{FI}-modules of finite degree, then so is the degree-wise tensor product FF:𝖥𝖨F×F𝖠𝖻×𝖠𝖻𝖠𝖻F\otimes F^{\prime}:\mathsf{FI}\xrightarrow{F\times F^{\prime}}\mathsf{Ab}\times\mathsf{Ab}\xrightarrow{\otimes}\mathsf{Ab}.

Proof.

By Proposition 2.23, there are exact sequences

L1L0F0\displaystyle L_{1}\to L_{0}\to F\to 0
L1L0F0\displaystyle L^{\prime}_{1}\to L^{\prime}_{0}\to F^{\prime}\to 0

where the Li,LiL_{i},L^{\prime}_{i} are free and generated in finite degree. Considering the total complex of the bicomplex LLL_{\ast}\otimes L^{\prime}_{\ast} yields the exact sequence

(L0L1)(L1L0)L0L0FF0.(L_{0}\otimes L^{\prime}_{1})\oplus(L_{1}\otimes L^{\prime}_{0})\to L_{0}\otimes L^{\prime}_{0}\to F\otimes F^{\prime}\to 0.

But the first two terms in the sequence are free and generated in finite degree by the previous lemma, and so FFF\otimes F^{\prime} has finite degree. ∎

Corollary 2.27 (of Lemma 2.25; see also [5]).

If F,FF,F^{\prime} are finitely generated 𝖥𝖨\mathsf{FI}-modules, then so is FFF\otimes F^{\prime}.

We end this section with the following lemma. Using the monoidal structure of 𝖥𝖨\mathsf{FI}, we can define the ‘duplication functor’

D:𝖥𝖨Δ𝖥𝖨×𝖥𝖨𝖥𝖨D:\mathsf{FI}\xrightarrow{\Delta}\mathsf{FI}\times\mathsf{FI}\xrightarrow{\sqcup}\mathsf{FI}

where Δ\Delta is the usual diagonal functor. DD is clearly not a functor of CONs, so Lemma 2.11 does not apply, but we still have the following result.

Lemma 2.28.

D:𝖥𝖨–Mod𝖥𝖨–ModD^{\ast}:\mathsf{FI}\mbox{--Mod}\to\mathsf{FI}\mbox{--Mod} takes finitely generated 𝖥𝖨\mathsf{FI}-modules to finitely generated 𝖥𝖨\mathsf{FI}-modules.

Proof.

DD^{\ast} clearly preserves epimorphisms (in fact it is exact). Hence it suffices to prove that DPnD^{\ast}P^{n} is finitely generated for every nn. We show DPnD^{\ast}P^{n} is generated in degrees n\leqslant n. Since for every mm, (DPn)m=P2mn(D^{\ast}P^{n})_{m}=P_{2m}^{n} is a finitely generated abelian group, this will complete the proof.

Let f:𝐧𝟐𝐦f:\mathbf{n}\to\mathbf{2m} be a generator of (DPn)m(D^{\ast}P^{n})_{m} where m>nm>n. Then there is some i𝐦i\in\mathbf{m} such that both i,i+m𝟐𝐦i,i+m\in\mathbf{2m} are not in the image of ff. Let g:𝐦𝟏𝐦g:\mathbf{m-1}\to\mathbf{m} be the unique order-preserving injection which misses ii. Define f:𝐧𝟐𝐦𝟐f^{\prime}:\mathbf{n}\to\mathbf{2m-2} by

f(j)={f(j)if f(j)<i,f(j)1if i<f(j)<i+m,f(j)2if i+m<f(j).f^{\prime}(j)=\begin{cases}f(j)&\mbox{if }f(j)<i,\\ f(j)-1&\mbox{if }i<f(j)<i+m,\\ f(j)-2&\mbox{if }i+m<f(j).\end{cases}

Then f=(gg)ff=(g\sqcup g)\circ f^{\prime}, so g:(DPn)m1(DPn)mg_{\ast}:(D^{\ast}P^{n})_{m-1}\to(D^{\ast}P^{n})_{m} takes ff^{\prime} to ff. ∎

3 Automorphisms of free nilpotent groups

Lower central series

Recall that the lower central series {γk(G)}k1\{\gamma_{k}(G)\}_{k\geqslant 1} of a group GG is defined inductively by setting γ1(G)=G\gamma_{1}(G)=G and γ(G)=[G,γk1(G)]\gamma_{(}G)=[G,\gamma_{k-1}(G)] for k>1k>1. The successive quotients grkGγk(G)/γk+1(G)\operatorname{gr}_{k}G\coloneqq\gamma_{k}(G)/\gamma_{k+1}(G) are abelian. In fact, [γk(G),γl(G)]γk+l(G)[\gamma_{k}(G),\gamma_{l}(G)]\leqslant\gamma_{k+l}(G), and so the commutator bracket descends to a Lie bracket on the associated graded group grG=k1grkG\operatorname{gr}G=\bigoplus_{k\geqslant 1}\operatorname{gr}_{k}G. These constructions give rise to functors γk:𝖦𝗋𝗉𝖦𝗋𝗉\gamma_{k}:\mathsf{Grp}\to\mathsf{Grp} and gr:𝖦𝗋𝗉𝗀𝗋𝖫𝗂𝖾\operatorname{gr}:\mathsf{Grp}\to\mathsf{grLie}_{\mathbb{Z}}, where 𝗀𝗋𝖫𝗂𝖾\mathsf{grLie}_{\mathbb{Z}} is the category of graded Lie algebras over \mathbb{Z}.

Hence for an 𝖥𝖨\mathsf{FI}-group222I.e. an 𝖥𝖨\mathsf{FI}-object in the category 𝖦𝗋𝗉\mathsf{Grp}. G:𝖥𝖨𝖦𝗋𝗉G:\mathsf{FI}\to\mathsf{Grp}, we obtain a filtration γk(G)\gamma_{k}(G) of GG by sub-𝖥𝖨\mathsf{FI}-groups, and a graded 𝖥𝖨\mathsf{FI}-Lie algebra grG\operatorname{gr}G consisting of 𝖥𝖨\mathsf{FI}-modules grkG\operatorname{gr}_{k}G.

Definition 3.1.

Let 𝔽:𝖥𝖨𝖦𝗋𝗉\mathbb{F}:\mathsf{FI}\to\mathsf{Grp} be the restriction to 𝖥𝖨\mathsf{FI} of the free group functor 𝖲𝖾𝗍𝖦𝗋𝗉\mathsf{Set}\to\mathsf{Grp}.

Thus 𝔽n\mathbb{F}_{n} is the free group generated by the set 𝐧\mathbf{n}, although we will denote its generators by x1,,xnx_{1},\ldots,x_{n} to avoid confusion.

Lemma 3.2.

Each grk𝔽\operatorname{gr}_{k}\mathbb{F} is a finitely generated 𝖥𝖨\mathsf{FI}-module taking values in free abelian groups.

Proof.

The fact that (grk𝔽)n=γk(𝔽n)/γk+1(𝔽n)(\operatorname{gr}_{k}\mathbb{F})_{n}=\gamma_{k}(\mathbb{F}_{n})/\gamma_{k+1}(\mathbb{F}_{n}) is a finitely generated free abelian group is standard — see e.g. [10].

We prove by induction on kk that grk𝔽\operatorname{gr}_{k}\mathbb{F} is generated in degrees k\leqslant k. The case k=1k=1 is clear: gr1𝔽P1\operatorname{gr}_{1}\mathbb{F}\cong P^{1} is the abelianisation of 𝔽\mathbb{F}, and therefore generated by the image of the generator x1𝔽1x_{1}\in\mathbb{F}_{1}\cong\mathbb{Z}.

Assume the claim has been established for all k<k0k<k_{0}. The graded Lie algebra gr𝔽n\operatorname{gr}\mathbb{F}_{n} is generated by gr1𝔽n\operatorname{gr}_{1}\mathbb{F}_{n}, so grk0𝔽n\operatorname{gr}_{k_{0}}\mathbb{F}_{n} is spanned over \mathbb{Z} by elements of the form [x,y][x,y] where xgrk𝔽nx\in\operatorname{gr}_{k}\mathbb{F}_{n} and ygrk0k𝔽ny\in\operatorname{gr}_{k_{0}-k}\mathbb{F}_{n} for some 0<k<k00<k<k_{0}. If nk0n\geqslant k_{0}, then by the induction hypothesis,

x\displaystyle x =f:𝐤𝐧f(xf)\displaystyle=\sum_{f:\mathbf{k}\to\mathbf{n}}f_{\ast}(x_{f})
y\displaystyle y =g:𝐤𝟎𝐤𝐧g(yg)\displaystyle=\sum_{g:\mathbf{k_{0}-k}\to\mathbf{n}}g_{\ast}(y_{g})

for some xfgrk𝔽kx_{f}\in\operatorname{gr}_{k}\mathbb{F}_{k} and yggrk0k𝔽k0ky_{g}\in\operatorname{gr}_{k_{0}-k}\mathbb{F}_{k_{0}-k}. For every pair (f:𝐤𝐧,g:𝐤𝟎𝐤𝐧)(f:\mathbf{k}\to\mathbf{n},g:\mathbf{k_{0}-k}\to\mathbf{n}), form the pushout Pf,g𝖥𝖨P_{f,g}\in\mathsf{FI} of the pullback of ff and gg to obtain the commutative diagram

𝐤{\mathbf{k}}𝐤𝟎𝐤{\mathbf{k_{0}-k}}Pf,g{P_{f,g}}𝐧{\mathbf{n}}if,g\scriptstyle{i_{f,g}}f\scriptstyle{f}jf,g\scriptstyle{j_{f,g}}g\scriptstyle{g}f+g\scriptstyle{f+g}

Clearly n(f,g)|Pf,g|k0n(f,g)\coloneqq|P_{f,g}|\leqslant k_{0}. We also have

[x,y]\displaystyle[x,y] =f:𝐤𝐧g:𝐤𝟎𝐤𝐧[f(xf),g(yg)]\displaystyle=\sum_{\begin{subarray}{c}f:\mathbf{k}\to\mathbf{n}\\ g:\mathbf{k_{0}-k}\to\mathbf{n}\end{subarray}}[f_{\ast}(x_{f}),g_{\ast}(y_{g})]
=f:𝐤𝐧g:𝐤𝟎𝐤𝐧(f+g)[(if,g)(xf),(jf,g)(yg)],\displaystyle=\sum_{\begin{subarray}{c}f:\mathbf{k}\to\mathbf{n}\\ g:\mathbf{k_{0}-k}\to\mathbf{n}\end{subarray}}(f+g)_{\ast}[(i_{f,g})_{\ast}(x_{f}),(j_{f,g})_{\ast}(y_{g})],

where each bracket in the latter sum is an element of grk0𝔽n(f,g)\operatorname{gr}_{k_{0}}\mathbb{F}_{n(f,g)}. Hence grk0𝔽\operatorname{gr}_{k_{0}}\mathbb{F} is generated in degrees k0\leqslant k_{0}. ∎

Free nilpotent groups

Write Nn(k)𝔽n/γk+1(𝔽n)N_{n}(k)\coloneqq\mathbb{F}_{n}/\gamma_{k+1}(\mathbb{F}_{n}) for k1k\geqslant 1. This is the free nilpotent group on nn generators of nilpotency class kk. In particular, Nn(1)nN_{n}(1)\cong\mathbb{Z}^{n}; N2(2)N_{2}(2) is the so-called Heisenberg group. We will denote the generators of Nn(k)N_{n}(k) coming from the generators x1,,xn𝔽nx_{1},\ldots,x_{n}\in\mathbb{F}_{n} by the same letters.

Nn(k)N_{n}(k) for a fixed kk and varying nn fit into the 𝖥𝖨\mathsf{FI}-group N(k)𝔽/γk+1(𝔽)N(k)\coloneqq\mathbb{F}/\gamma_{k+1}(\mathbb{F}). There are extensions of 𝖥𝖨\mathsf{FI}-groups

0grk𝔽N(k)N(k1)10\to\operatorname{gr}_{k}\mathbb{F}\to N(k)\to N(k-1)\to 1

for all k>1k>1. For every i<ki<k, there is a canonical isomorphism γi(𝔽)/γk+1(𝔽)γi(N(k))\gamma_{i}(\mathbb{F})/\gamma_{k+1}(\mathbb{F})\cong\gamma_{i}(N(k)) induced by the quotient morphism 𝔽N(k)\mathbb{F}\to N(k).

Automorphism groups

Consider now the groups AutNn(k)\operatorname{Aut}N_{n}(k) of automorphisms of Nn(k)N_{n}(k). For a fixed kk, these also form an 𝖥𝖨\mathsf{FI}-group as follows: given f:𝐧𝐦f:\mathbf{n}\hookrightarrow\mathbf{m} and ϕAutNn(k)\phi\in\operatorname{Aut}N_{n}(k), f(ϕ):Nm(k)Nm(k)f_{\ast}(\phi):N_{m}(k)\to N_{m}(k) is defined by

f(ϕ)(xi)={xiif iimf,N(k)(f)ϕ(xj)if i=f(j),j𝐧.f_{\ast}(\phi)(x_{i})=\begin{cases}x_{i}&\mbox{if }i\not\in\operatorname{im}f,\\ N(k)(f)\circ\phi(x_{j})&\mbox{if }i=f(j),j\in\mathbf{n}.\end{cases}

It is an easy check that f(ϕ)f_{\ast}(\phi) is an automorphism (indeed, f(ϕ1)f_{\ast}(\phi^{-1}) is its inverse) and that the assignment ϕf(ϕ):AutNn(k)AutNm(k)\phi\mapsto f_{\ast}(\phi):\operatorname{Aut}N_{n}(k)\to\operatorname{Aut}N_{m}(k) is a homomorphism.

There are morphisms of 𝖥𝖨\mathsf{FI}-groups ρ(k):AutN(k)AutN(k1)\rho(k):\operatorname{Aut}N(k)\to\operatorname{Aut}N(k-1) defined as follows: an automorphism ϕAutNn(k)\phi\in\operatorname{Aut}N_{n}(k) preserves γk(Nn(k))=γk(𝔽n)/γk+1(𝔽n)\gamma_{k}(N_{n}(k))=\gamma_{k}(\mathbb{F}_{n})/\gamma_{k+1}(\mathbb{F}_{n}) and therefore descends to an endomorphism ϕ\phi^{\prime} of Nn(k1)N_{n}(k-1). But ϕ1\phi^{-1} descends to the inverse of ϕ\phi^{\prime}, so ϕAutNn(k1)\phi^{\prime}\in\operatorname{Aut}N_{n}(k-1). The assignment ϕϕ\phi\mapsto\phi^{\prime} then defines the map ρ(k)\rho(k) of 𝖥𝖨\mathsf{FI}-groups.

The kernel of ρ(k)\rho(k)

Suppose that ϕAutNn(k)\phi\in\operatorname{Aut}N_{n}(k) lies in the kernel of ρn(k)\rho_{n}(k). Then the set map Nn(k)γk(Nn(k))grk𝔽nN_{n}(k)\to\gamma_{k}(N_{n}(k))\cong\operatorname{gr}_{k}\mathbb{F}_{n} sending xx to ϕ(x)x1\phi(x)x^{-1} is in fact a homomorphism since γk(Nn(k))\gamma_{k}(N_{n}(k)) is central in Nn(k)N_{n}(k). This map descends to a homomorphism ϕ:ngrk𝔽n\phi^{\prime}:\mathbb{Z}^{n}\to\operatorname{gr}_{k}\mathbb{F}_{n}, where n\mathbb{Z}^{n} is the abelianisation of Nn(k)N_{n}(k). We obtain a homomorphism

ψn(k):kerρn(k)\displaystyle\psi_{n}(k):\ker\rho_{n}(k) Hom(n,grk𝔽n)\displaystyle\to\operatorname{Hom}(\mathbb{Z}^{n},\operatorname{gr}_{k}\mathbb{F}_{n})
ϕ\displaystyle\phi ϕ\displaystyle\mapsto\phi^{\prime}

It is well known that ψn(k)\psi_{n}(k) is an isomorphism of abelian groups.

Let :𝖥𝖨op𝖠𝖻\mathbb{Z}^{\bullet}:\mathsf{FI}^{op}\to\mathsf{Ab} be the 𝖥𝖨op\mathsf{FI}^{op}-module taking 𝐧\mathbf{n} to n\mathbb{Z}^{n} and f:𝐧𝐦f:\mathbf{n}\hookrightarrow\mathbf{m} to the projection

f:m\displaystyle f^{\ast}:\mathbb{Z}^{m} n\displaystyle\to\mathbb{Z}^{n}
xi\displaystyle x_{i} {0if iimfxjif i=f(j)\displaystyle\mapsto\begin{cases}0&\mbox{if }i\notin\operatorname{im}f\\ x_{j}&\mbox{if }i=f(j)\end{cases}

Observe that the dual 𝖥𝖨\mathsf{FI}-module ()\left(\mathbb{Z}^{\bullet}\right)^{\ast} is canonically isomorphic to gr1𝔽\operatorname{gr}_{1}\mathbb{F}. Let K(k)K(k) be the 𝖥𝖨\mathsf{FI}-module given by

K(k):𝖥𝖨()op×grk𝔽𝖠𝖻op×𝖠𝖻Hom(,)𝖠𝖻.K(k):\mathsf{FI}\xrightarrow{\left(\mathbb{Z}^{\bullet}\right)^{op}\times\operatorname{gr}_{k}\mathbb{F}}\mathsf{Ab}^{op}\times\mathsf{Ab}\xrightarrow{\operatorname{Hom}(-,-)}\mathsf{Ab}.

We leave it to the reader to verify that the ψn(k)\psi_{n}(k) are components of a natural isomorphism ψ(k):kerρ(k)K(k)\psi(k):\ker\rho(k)\to K(k).

Lemma 3.3.

For every k1k\geqslant 1, kerρ(k)\ker\rho(k) is a finitely generated 𝖥𝖨\mathsf{FI}-module taking values in free abelian groups.

Proof.

Since kerρ(k)K(k)\ker\rho(k)\cong K(k), it is enough to prove the assertion for K(k)K(k).

Recall that there is a natural isomorphism Hom(A,B)AB\operatorname{Hom}(A,B)\cong A^{\ast}\otimes B for AA a finitely generated free abelian group and B𝖠𝖻B\in\mathsf{Ab}, where AA^{\ast} denotes the dual of AA. Hence there are isomorphisms K(k)()grk𝔽gr1𝔽grk𝔽K(k)\cong\left(\mathbb{Z}^{\bullet}\right)^{\ast}\otimes\operatorname{gr}_{k}\mathbb{F}\cong\operatorname{gr}_{1}\mathbb{F}\otimes\operatorname{gr}_{k}\mathbb{F}. But gr1𝔽\operatorname{gr}_{1}\mathbb{F} and grk𝔽\operatorname{gr}_{k}\mathbb{F} are finitely generated and take values in free abelian groups by Lemma 3.2. The result follows from Corollary 2.27. ∎

Johnson filtration

There is also a homomorphism Aut𝔽nAutNn(k)\operatorname{Aut}\mathbb{F}_{n}\to\operatorname{Aut}N_{n}(k) for every kk and nn. Fix nn; then the kernels

IAn(k){ϕAut𝔽nϕ(x)x1γk+1(𝔽n) for all x𝔽n}I\!A_{n}(k)\coloneqq\left\{\phi\in\operatorname{Aut}\mathbb{F}_{n}\mid\phi(x)x^{-1}\in\gamma_{k+1}(\mathbb{F}_{n})\mbox{ for all }x\in\mathbb{F}_{n}\right\}

of these homomorphisms form a filtration of Aut𝔽n\operatorname{Aut}\mathbb{F}_{n}. We record the following theorem, originally due to Andreadakis [1], for later use.

Theorem 3.4 ([1, Theorem 1.1]).

The filtration {IAn(k)}k1\{I\!A_{n}(k)\}_{k\geqslant 1} of IAn(1)I\!A_{n}(1) is central.

4 Quotients of mapping class groups

Mapping class groups

Let Sg,nS_{g,n} be an oriented surface of genus gg with nn boundary components. Fix collars [0,ϵ)×S1Sg,n\left[0,\epsilon\right)\times S^{1}\hookrightarrow S_{g,n} of the boundary components.

Let Diff+(Sg,1,)\operatorname{Diff}^{+}(S_{g,1},\partial) be the group of orientation-preserving diffeomorphisms of the surface which fix the collar pointwise. The mapping class group Γg,1\Gamma_{g,1} of Sg,1S_{g,1} is the group of isotopy classes of such diffeomorphisms:

Γg,1Diff+(Sg,1,)/Diff0+(Sg,1,)\Gamma_{g,1}\coloneqq\operatorname{Diff}^{+}(S_{g,1},\partial)/\operatorname{Diff}_{0}^{+}(S_{g,1},\partial)

where Diff0+(Sg,1,)\operatorname{Diff}_{0}^{+}(S_{g,1},\partial) is the path-component of the identity in Diff+(Sg,1,)\operatorname{Diff}^{+}(S_{g,1},\partial).

Refer to captionx1x_{1}y1y_{1}xg1x_{g-1}yg1y_{g-1}xgx_{g}ygy_{g}
Figure 1: Sg,1S_{g,1}

We fix a point Sg,1\ast\in\partial S_{g,1}, and we choose generators x1,y1,,xg,ygx_{1},y_{1},\ldots,x_{g},y_{g} of ππ1(Sg,1,)𝔽2g\pi\coloneqq\pi_{1}(S_{g,1},\ast)\cong\mathbb{F}_{2g} as in Figure 1. Then the boundary word ζπ\zeta\in\pi is equal to

i[xi,yi].\prod_{i}[x_{i},y_{i}].

π\pi admits an action of Γg,1\Gamma_{g,1}, yielding a homomorphism Γg,1Autπ\Gamma_{g,1}\to\operatorname{Aut}\pi. By the Dehn-Nielsen-Baer theorem, this homomorphism is injective with image precisely the stabiliser of ζπ\zeta\in\pi:

Γg,1{ϕAut𝔽2gϕ(ζ)=ζ}.\Gamma_{g,1}\cong\left\{\phi\in\operatorname{Aut}\mathbb{F}_{2g}\mid\phi(\zeta)=\zeta\right\}.

Johnson filtration

Just like in the case of Aut𝔽n\operatorname{Aut}\mathbb{F}_{n}, we can filter Γg,1\Gamma_{g,1} by the kernels g,1(k)\mathscr{I}_{g,1}(k) of its action on the quotients 𝔽2g/γk+1(𝔽2g)\mathbb{F}_{2g}/\gamma_{k+1}(\mathbb{F}_{2g}):

g,1(k)={ϕΓg,1ϕ(x)x1γk+1(𝔽2g) for all x𝔽2g}=Γg,1IA2g(k).\mathscr{I}_{g,1}(k)=\left\{\phi\in\Gamma_{g,1}\mid\phi(x)x^{-1}\in\gamma_{k+1}(\mathbb{F}_{2g})\mbox{ for all }x\in\mathbb{F}_{2g}\right\}=\Gamma_{g,1}\cap I\!A_{2g}(k).

This is the Johnson filtration Γg,1=g,1(0)g,1(1)\Gamma_{g,1}=\mathscr{I}_{g,1}(0)\rhd\mathscr{I}_{g,1}(1)\rhd\cdots, defined by D. Johnson in the 80s to study the Torelli group g,1=g,1(1)\mathscr{I}_{g,1}=\mathscr{I}_{g,1}(1). We write

Γg,1(k)Γg,1/g,1(k)\Gamma_{g,1}(k)\coloneqq\Gamma_{g,1}/\mathscr{I}_{g,1}(k)

for the quotient. Γg,1(k)\Gamma_{g,1}(k) is canonically isomorphic to the image of Γg,1\Gamma_{g,1} in AutN2g(k)\operatorname{Aut}N_{2g}(k).

In the case k=1k=1, π/γ2(π)\pi/\gamma_{2}(\pi) is naturally isomorphic to HgH1(Sg,1;)2gH_{g}\coloneqq H_{1}(S_{g,1};\mathbb{Z})\cong\mathbb{Z}^{2g}. The action of Γg,1\Gamma_{g,1} on HgH_{g} preserves the algebraic intersection pairing, so the image of Γg,1\Gamma_{g,1} in AutN2g(1)GL(Hg)\operatorname{Aut}N_{2g}(1)\cong\mathrm{GL}(H_{g}) lies inside Sp(Hg)Sp(2g,)\mathrm{Sp}(H_{g})\cong\mathrm{Sp}(2g,\mathbb{Z}). It is well known that this image is in fact the whole of Sp(Hg)\mathrm{Sp}(H_{g}).

Lemma 4.1.

The filtration {g,1(k)}k1\{\mathscr{I}_{g,1}(k)\}_{k\geqslant 1} of g,1\mathscr{I}_{g,1} is central.

Proof.

This is a classical result (it also follows immediately from Theorem 3.4). ∎

𝖲𝖨\mathsf{SI}-modules from the Johnson filtration

Suppose ϕ:Sg,1Sg,1\phi:S_{g,1}\hookrightarrow S_{g^{\prime},1} is an orientation-preserving embedding of surfaces. We can extend a given mapping class on Sg,1imϕS_{g,1}\cong\operatorname{im}\phi by the identity on the complement Sg,1imϕS_{g^{\prime},1}\setminus\operatorname{im}\phi to obtain a mapping class on Sg,1S_{g^{\prime},1}. This construction clearly produces a well-defined homomorphism Γg,1Γg,1\Gamma_{g,1}\to\Gamma_{g^{\prime},1}.

Definition 4.2.

We write ϕ:Γg,1Γg,1\phi_{\flat}:\Gamma_{g,1}\to\Gamma_{g^{\prime},1} for the above homomorphism.

When we picture surfaces Sg,1S_{g,1} and Sg,1S_{g^{\prime},1} with ggg\leqslant g^{\prime} as in Figure 1, there is an obvious embedding T:Sg,1Sg,1T:S_{g,1}\hookrightarrow S_{g^{\prime},1} as the subsurface consisting of the gg rightmost tori. Algebraically, T:Γg,1Γg,1T_{\flat}:\Gamma_{g,1}\to\Gamma_{g^{\prime},1} extends a given mapping class on Sg,1S_{g,1} (thought of as an automorphism of 𝔽2g\mathbb{F}_{2g}) to an automorphism of 𝔽2g\mathbb{F}_{2g^{\prime}} by the identity on the remaining generators xi,yix_{i},y_{i}, i>gi>g. It is easy to see that TT_{\flat} preserves the Johnson filtration. In fact, this is true for any embedding:

Lemma 4.3.

Let ϕ:Sg,1Sg,1\phi:S_{g,1}\hookrightarrow S_{g^{\prime},1} be an embedding. Then ϕ\phi_{\flat} preserves the Johnson filtration.

Proof.

First of all observe that ϕ\phi_{\flat} only depends on the isotopy class of ϕ\phi, whence we may assume that imϕ\operatorname{im}\phi is disjoint from some collar of Sg,1\partial S_{g^{\prime},1} and therefore that Sg,1imϕS_{g^{\prime},1}\setminus\operatorname{im}\phi is connected. Let SS be the closure of Sg,1imϕS_{g^{\prime},1}\setminus\operatorname{im}\phi; by the classification of oriented surfaces, SSgg,2S\cong S_{g^{\prime}-g,2}. Hence we can extend ϕ\phi to a diffeomorphism ϕ~\widetilde{\phi} of Sg,1S_{g^{\prime},1} such that ϕ=ϕ~T\phi=\widetilde{\phi}\circ T. But then ϕ\phi_{\flat} and TT_{\flat} are conjugate by [ϕ~][\widetilde{\phi}] in Γg,1\Gamma_{g^{\prime},1}. ∎

Hence ϕ\phi_{\flat} descends to well-defined homomorphisms

ϕ:g,1(k1)/g,1(k)g,1(k1)/g,1(k)\phi_{\flat}:\mathscr{I}_{g,1}(k-1)/\mathscr{I}_{g,1}(k)\to\mathscr{I}_{g^{\prime},1}(k-1)/\mathscr{I}_{g^{\prime},1}(k)

for each k>1k>1. We call these homomorphisms ϕ\phi_{\flat} as well, since there is no danger of confusion.

Lemma 4.4.
  1. (i)

    Every symplectic map f:HgHgf:H_{g}\to H_{g^{\prime}} is induced by an orientation-preserving embedding Sg,1Sg,1S_{g,1}\hookrightarrow S_{g^{\prime},1} of surfaces.

  2. (ii)

    If two orientation-preserving embeddings ϕ,ϕ:Sg,1Sg,1\phi,\phi^{\prime}:S_{g,1}\hookrightarrow S_{g^{\prime},1} induce the same map on the first homology group, then there is some ψDiff+(Sg,1,)\psi\in\operatorname{Diff}^{+}(S_{g^{\prime},1},\partial) which represents a class in g,1\mathscr{I}_{g^{\prime},1} and such that ϕ\phi^{\prime} and ψϕ\psi\circ\phi are isotopic.

Proof.
  1. (i)

    There is certainly an embedding T:Sg,1Sg,1T:S_{g,1}\hookrightarrow S_{g^{\prime},1} which induces the standard inclusion T:HgHgT_{\ast}:H_{g}\to H_{g^{\prime}} that sends [xi][x_{i}] to [xi][x_{i}] and [yi][y_{i}] to [yi][y_{i}]. But by Lemma 2.14, Sp(Hg)\mathrm{Sp}(H_{g^{\prime}}) acts transitively on the space of symplectic maps HgHgH_{g}\to H_{g^{\prime}}, so there is some sSp(Hg)s\in\mathrm{Sp}(H_{g^{\prime}}) such that sT=fs\circ T_{\ast}=f. If ϕDiff+(Sg,1,)\phi\in\operatorname{Diff}^{+}(S_{g^{\prime},1},\partial) represents a lift of ss to Γg,1\Gamma_{g^{\prime},1}, then clearly ϕT:Sg,1Sg,1\phi\circ T:S_{g,1}\hookrightarrow S_{g^{\prime},1} is an embedding and (ϕT)=f\left(\phi\circ T\right)_{\ast}=f.

  2. (ii)

    This is the content of [4, Lemma 4.1(ii)].

Write Qg(k)g,1(k1)/g,1(k)Q_{g}(k)\coloneqq\mathscr{I}_{g,1}(k-1)/\mathscr{I}_{g,1}(k). We show that these groups fit into an 𝖲𝖨\mathsf{SI}-module Q(k)Q(k).

Remark 4.5.

Observe that the objects VgV_{g} of 𝖲𝖨\mathsf{SI} are isomorphic (as symplectic spaces) to the HgH_{g}. In the rest of this section, we assume the HgH_{g} are the objects of 𝖲𝖨\mathsf{SI}, since it makes the construction of Q(k)Q(k) more natural.

Proposition 4.6.

For every k>1k>1, there is an 𝖲𝖨\mathsf{SI}-module Q(k)Q(k) which takes HgH_{g} to Qg(k)Q_{g}(k) and f:HgHgf:H_{g}\to H_{g^{\prime}} to ϕ:Qg(k)Qg(k)\phi_{\flat}:Q_{g}(k)\to Q_{g^{\prime}}(k), where ϕ\phi is as in Lemma 4.4(i).

Proof.

We define Q(k)Q(k) as follows: Q(k)(Hg)=Qg(k)Q(k)(H_{g})=Q_{g}(k). Given a symplectic map f:HgHgf:H_{g}\to H_{g^{\prime}}, let ϕ:Sg,1Sg,1\phi:S_{g,1}\hookrightarrow S_{g^{\prime},1} be an embedding with ϕ=f\phi_{\ast}=f. By Lemma 4.3, the induced homomorphism ϕ\phi_{\flat} gives rise to a homomorphism Qg(k)Qg(k)Q_{g}(k)\to Q_{g^{\prime}}(k). We claim that this induced map is independent of the choice of ϕ\phi. Indeed, suppose that ϕ:Sg,1Sg,1\phi^{\prime}:S_{g,1}\hookrightarrow S_{g^{\prime},1} is another embedding with ϕ=f\phi^{\prime}_{\ast}=f. Let ψ\psi be as in Lemma 4.4. Then ϕ=(ψϕ)\phi^{\prime}_{\flat}=\left(\psi\circ\phi\right)_{\flat}, i.e. ϕ\phi_{\flat} and ϕ\phi^{\prime}_{\flat} are conjugate by [ψ][\psi] in Γg,1\Gamma_{g^{\prime},1}. But [ψ]g,1[\psi]\in\mathscr{I}_{g^{\prime},1} and the Johnson filtration is central in g,1\mathscr{I}_{g^{\prime},1}, so ϕ\phi_{\flat} and ϕ\phi^{\prime}_{\flat} induce the same homomorphism Qg(k)Qg(k)Q_{g}(k)\to Q_{g^{\prime}}(k). Hence Q(k)Q(k) is well defined. It is easy to check that Q(k)Q(k) is a functor. ∎

Recall the functor X:𝖥𝖨𝖲𝖨X:\mathsf{FI}\to\mathsf{SI} from Definition 2.15: it sends 𝐧\mathbf{n} to HnH_{n} and f:𝐧𝐦f:\mathbf{n}\to\mathbf{m} to the symplectic map

Xf:Hn\displaystyle Xf:H_{n} Hm\displaystyle\to H_{m}
[xi]\displaystyle[x_{i}] [xf(i)]\displaystyle\mapsto[x_{f(i)}]
[yi]\displaystyle[y_{i}] [yf(i)]\displaystyle\mapsto[y_{f(i)}]
Theorem 4.7.

For every k>1k>1, XQ(k)X^{\ast}Q(k) is a finitely generated 𝖥𝖨\mathsf{FI}-module. It takes values in free abelian groups.

Proof.

Write FXQ(k)F\coloneqq X^{\ast}Q(k). For every nn, n,1(k)\mathscr{I}_{n,1}(k) is the kernel of the action of Γn,1\Gamma_{n,1} on N2n(k)N_{2n}(k), so there are injections jn:Γn,1(k)AutN2n(k)j_{n}:\Gamma_{n,1}(k)\hookrightarrow\operatorname{Aut}N_{2n}(k). Fn=n,1(k1)/n,1(k)F_{n}=\mathscr{I}_{n,1}(k-1)/\mathscr{I}_{n,1}(k) is mapped into the kernel kerρ2n(k)\ker\rho_{2n}(k) which is free abelian by Lemma 3.3. Hence FnF_{n} is also free abelian.

Moreover, we claim that the injections jn:Fnkerρ2n(k)j_{n}:F_{n}\hookrightarrow\ker\rho_{2n}(k) are the components of a morphism j:FDkerρ(k)j:F\hookrightarrow D^{\ast}\ker\rho(k) of 𝖥𝖨\mathsf{FI}-modules, where D:𝖥𝖨𝖥𝖨D:\mathsf{FI}\to\mathsf{FI} is the duplication functor defined at the end of Section 2.4. Observe that we only need to check naturality with respect to the inclusions 𝐧𝟏𝐧\mathbf{n-1}\hookrightarrow\mathbf{n} and automorphisms of 𝐧\mathbf{n}, as these arrows generate all the arrows in 𝖥𝖨\mathsf{FI}.

Naturality with respect to the inclusions 𝐧𝟏𝐧\mathbf{n-1}\hookrightarrow\mathbf{n} follows easily from the definitions. Given a permutation σAut𝖥𝖨(𝐧)\sigma\in\operatorname{Aut}_{\mathsf{FI}}(\mathbf{n}), let σ~Γn,1\widetilde{\sigma}\in\Gamma_{n,1} be a lift of XσAut𝖲𝖨(Hg)=Sp(Hg)X\sigma\in\operatorname{Aut}_{\mathsf{SI}}(H_{g})=\mathrm{Sp}(H_{g}). Let also sAut𝔽2ns\in\operatorname{Aut}\mathbb{F}_{2n} be the obvious lift of XσX\sigma, namely the automorphism sending xixσ(i)x_{i}\mapsto x_{\sigma(i)} and yiyσ(i)y_{i}\mapsto y_{\sigma(i)}. In general, ss will not be in Γn,1\Gamma_{n,1} because it does not preserve the boundary word i[xi,yi]\prod_{i}[x_{i},y_{i}]. We want to show the following diagram commutes

n,1(k1)/n,1(k){\mathscr{I}_{n,1}(k-1)/\mathscr{I}_{n,1}(k)}kerρ2n(k){\ker\rho_{2n}(k)}n,1(k1)/n,1(k){\mathscr{I}_{n,1}(k-1)/\mathscr{I}_{n,1}(k)}kerρ2n(k){\ker\rho_{2n}(k)}jn\scriptstyle{j_{n}}σ\scriptstyle{\sigma_{\ast}}σ\scriptstyle{\sigma_{\ast}}jn\scriptstyle{j_{n}}

where the left vertical arrow is conjugation by σ~\widetilde{\sigma} while the right arrow is conjugation by ss.

Observe that s1σ~IA2n(1)s^{-1}\widetilde{\sigma}\in I\!A_{2n}(1). Indeed, σ~(zi)zσ(i)1γ1(𝔽2n)\widetilde{\sigma}(z_{i})z_{\sigma(i)}^{-1}\in\gamma_{1}(\mathbb{F}_{2n}), where zz stands for either xx or yy. Hence

s1(σ~(zi)zσ(i)1)=(s1σ~)(zi)zi1γ1(𝔽2n),s^{-1}\left(\widetilde{\sigma}(z_{i})z_{\sigma(i)}^{-1}\right)=(s^{-1}\widetilde{\sigma})(z_{i})z_{i}^{-1}\in\gamma_{1}(\mathbb{F}_{2n}),

so by Theorem 3.4, [s1σ~,ϕ]IA2n(k)[s^{-1}\widetilde{\sigma},\phi]\in I\!A_{2n}(k) for every ϕn,1(k1)\phi\in\mathscr{I}_{n,1}(k-1). Therefore sϕs1s\phi s^{-1} and σ~ϕσ~1\widetilde{\sigma}\phi\widetilde{\sigma}^{-1} induce the same automorphism of N2n(k)N_{2n}(k), which proves the claim.

But Dkerρ(k)D^{\ast}\ker\rho(k) is finitely generated by Lemma 2.28 since kerρ(k)\ker\rho(k) is finitely generated. Hence FF is finitely generated by Theorem 2.19. ∎

Remark 4.8.

In particular, Q(k)Q(k) has finite degree as an 𝖲𝖨\mathsf{SI}-module by Lemma 2.11. We need, however, the stronger statement we made in the theorem, since it will allow us to argue that the exterior powers ΛnQ(k)\Lambda^{n}Q(k) have finite degree as well.

5 Homological stability for the groups Γg,1(k)\Gamma_{g,1}(k)

5.1 A general framework

We will deduce Theorem A from Theorem 4.7 and homological stability for the groups Sp(2n,)\mathrm{Sp}(2n,\mathbb{Z}) using a bootstrapping argument.

Definition 5.1.
  1. (i)

    A sequence of groups is an \mathbb{N}-object in the category 𝖦𝗋𝗉\mathsf{Grp}, i.e. a composable sequence G0G1G_{0}\to G_{1}\to\cdots of group homomorphisms. A morphism of a sequence of groups is a natural transformation.

  2. (ii)

    A GG-module MM for a sequence of groups GG is a composable sequence M0M1M_{0}\to M_{1}\to\cdots of abelian group homomorphisms, together with a GnG_{n}-action on MnM_{n} for each nn such that the connecting homomorphisms Mn1MnM_{n-1}\to M_{n} are equivariant with respect to Gn1GnG_{n-1}\to G_{n}.

  3. (iii)

    A sequence of groups GG has homological stability with coefficients in a GG-module MM if, for every ii, the induced homomorphism

    Hi(Gn1;Mn1)Hi(Gn;Mn)H_{i}(G_{n-1};M_{n-1})\to H_{i}(G_{n};M_{n})

    is an isomorphism for all nn(i)n\geqslant n(i) for some n(i)1n(i)\geqslant 1.

Remark 5.2.

The category 𝖦𝗋𝗉\mathsf{Grp}^{\mathbb{N}} of sequences of groups is complete, with limits computed pointwise, and the sequence 111\to 1\to\cdots is a null object. In particular, the category has all kernels. Also, a morphism f:GHf:G\to H is monic (resp. epic) if and only if every component fn:GnHnf_{n}:G_{n}\to H_{n} is monic (resp. epic).

Example 5.3 (Sequences of groups).

The homomorphisms T:Γg,1Γg,1T_{\flat}:\Gamma_{g,1}\to\Gamma_{g^{\prime},1} for ggg\leqslant g^{\prime} defined in the previous section give rise to a sequence of groups

Γ(Γ0,1Γ1,1).\Gamma\coloneqq\left(\Gamma_{0,1}\to\Gamma_{1,1}\to\cdots\right).

By Lemma 4.3, Γ\Gamma is filtered by sequences of groups arising from the Johnson filtration

(k)(0,1(k)1,1(k))\mathscr{I}(k)\coloneqq\left(\mathscr{I}_{0,1}(k)\to\mathscr{I}_{1,1}(k)\to\cdots\right)

and the quotients also fit into sequences of groups

Γ(k)(Γ0,1(k)Γ1,1(k))\Gamma(k)\coloneqq\left(\Gamma_{0,1}(k)\to\Gamma_{1,1}(k)\to\cdots\right)

for each k1k\geqslant 1.

Example 5.4 (Modules of sequences of groups).

Consider the sequence Γ(1)\Gamma(1) which consists of the groups Γn,1(1)Sp(2n,)\Gamma_{n,1}(1)\cong\mathrm{Sp}(2n,\mathbb{Z}). The connecting homomorphisms Sp(2(n1),)Sp(2n,)\mathrm{Sp}(2(n-1),\mathbb{Z})\to\mathrm{Sp}(2n,\mathbb{Z}) take a symplectic (2n2)×(2n2)\left(2n-2\right)\times\left(2n-2\right) matrix AA to the symplectic 2n×2n2n\times 2n matrix

(A0[1pt/2pt]0I_2)\left(\begin{array}[]{c;{1pt/2pt}c}A&0\\ \hline\cr[1pt/2pt]0&I_2\end{array}\right)

where I2I_{2} is the 2×22\times 2 identity matrix. Another way to think about the connecting homomorphisms is in the context of the category 𝖲𝖨\mathsf{SI}: we have Sp(2n,)=Aut𝖲𝖨(Vn)\mathrm{Sp}(2n,\mathbb{Z})=\operatorname{Aut}_{\mathsf{SI}}(V_{n}), and the connecting homomorphism is induced by the monoidal product: Aut𝖲𝖨(Vn1)()V1Aut𝖲𝖨(Vn)\operatorname{Aut}_{\mathsf{SI}}(V_{n-1})\xrightarrow{(-)\boxtimes V_{1}}\operatorname{Aut}_{\mathsf{SI}}(V_{n}).

Hence an 𝖲𝖨\mathsf{SI}-module FF gives rise to a Γ(1)\Gamma(1)-module

F0(i0)F1(i1)F_{0}\xrightarrow{\left(i_{0}\right)_{\ast}}F_{1}\xrightarrow{\left(i_{1}\right)_{\ast}}\cdots

where in:VnVn+1i_{n}:V_{n}\hookrightarrow V_{n+1} is the inclusion ‘on the left’, i.e. the arrow Vn(V0V1)V_{n}\boxtimes\left(V_{0}\to V_{1}\right). We will abuse notation and write FF for this Γ(1)\Gamma(1)-module as well.

Bootstrapping stability

If f:GHf:G\to H is a morphism of sequences of groups and MM is an HH-module, we obtain a GG-module fMf^{\ast}M by restricting the action of HnH_{n} on MnM_{n} along fn:GnHnf_{n}:G_{n}\to H_{n}. Suppose we have an extension

1G𝜄GG′′11\to G^{\prime}\xrightarrow{\iota}G\to G^{\prime\prime}\to 1

of sequences of groups and a GG-module MM. Then for every ii, the sequence

Hi(Gn1;ιMn1)Hi(Gn;ιMn)\cdots\to H_{i}(G^{\prime}_{n-1};\iota^{\ast}M_{n-1})\to H_{i}(G^{\prime}_{n};\iota^{\ast}M_{n})\to\cdots

forms a G′′G^{\prime\prime}-module Hi(G;ιM)H_{i}(G^{\prime};\iota^{\ast}M), where the module structure comes from the conjugation action of Gn′′G^{\prime\prime}_{n} on the homology of GnG^{\prime}_{n}. We can deduce from the Lyndon-Hochschild-Serre spectral sequence the following theorem.

Theorem 5.5.

Let 1G𝜄GG′′11\to G^{\prime}\xrightarrow{\iota}G\to G^{\prime\prime}\to 1 and MM be as above. If for every ii, G′′G^{\prime\prime} has homological stability with coefficients in Hi(G;ιM)H_{i}(G^{\prime};\iota^{\ast}M), then GG has homological stability with coefficients in MM.

See e.g. [11] for a proof.

5.2 Proof of the main result

First of all recall the following well-known theorem which says that the groups Sp(2n,)\mathrm{Sp}(2n,\mathbb{Z}) enjoy homological stability with suitable twisted coefficients.

Theorem 5.6.

If a module MM of the sequence Γ(1)=(1Sp(2,)Sp(4,))\Gamma(1)=\left(1\to\mathrm{Sp}(2,\mathbb{Z})\to\mathrm{Sp}(4,\mathbb{Z})\to\cdots\right) comes from an 𝖲𝖨\mathsf{SI}-module of finite degree as in Example 5.4, then Γ(1)\Gamma(1) has homological stability with coefficients in MM.

See e.g. [9] for a proof.

Definition 5.7.

We say a Γ(1)\Gamma(1)-module is good if it satisfies the hypothesis of Theorem 5.6.

We say a Γ(k)\Gamma(k)-module, k>1k>1, is good if it is the restriction of a good Γ(1)\Gamma(1)-module along the quotient morphism Γ(k)Γ(1)\Gamma(k)\to\Gamma(1).

Now we are ready to prove Theorem A.

Theorem A.

For all k1k\geqslant 1, Γ(k)\Gamma(k) enjoys homological stability with coefficients in every good Γ(k)\Gamma(k)-module.

Proof.

We proceed by induction on kk. The case k=1k=1 is precisely the content of Theorem 5.6.

Suppose the claim has been established for Γ(k1)\Gamma(k-1) for some k>1k>1, and let MM be a good Γ(k)\Gamma(k)-module. There is an extension of sequences of groups

0Q(k)Γ(k)Γ(k1)10\to Q(k)\to\Gamma(k)\to\Gamma(k-1)\to 1

where Q(k)=(Q0(k)Q1(k))Q(k)=\left(Q_{0}(k)\to Q_{1}(k)\to\cdots\right) is the obvious subobject of the sequence of groups Γ(k)\Gamma(k). Since Q(k)Q(k) lies in the kernel of Γ(k)Γ(1)\Gamma(k)\to\Gamma(1) and MM is pulled back along this map, the action of Qn(k)Q_{n}(k) on MnM_{n} is trivial for each nn. Each Qn(k)Q_{n}(k) is a finitely generated free abelian group by Theorem 4.7, and so by the Universal Coefficient Theorem,

Hi(Qn(k);Mn)MnΛiQn(k)H_{i}(Q_{n}(k);M_{n})\cong M_{n}\otimes_{\mathbb{Z}}\Lambda^{i}Q_{n}(k)

where Γn,1(k1)\Gamma_{n,1}(k-1) acts diagonally.

Since MM is good, it is induced from some 𝖲𝖨\mathsf{SI}-module of finite degree which we also call MM. Then the Γ(k1)\Gamma(k-1)-module Hi(Q(k);M)H_{i}(Q(k);M) is induced from the 𝖲𝖨\mathsf{SI}-module MΛiQ(k)M\otimes\Lambda^{i}Q(k). Once we show that this 𝖲𝖨\mathsf{SI}-module is good, we are done: by the induction hypothesis, Γ(k1)\Gamma(k-1) then enjoys homological stability with coefficients in Hi(Q(k);M)H_{i}(Q(k);M) and by Theorem 5.5, Γ(k)\Gamma(k) has homological stability with coefficients in MM.

By Lemma 2.11, it suffices to show that the pulled-back 𝖥𝖨\mathsf{FI}-module X(MΛiQ(k))(XM)Λi(XQ(k))X^{\ast}\left(M\otimes\Lambda^{i}Q(k)\right)\cong\left(X^{\ast}M\right)\otimes\Lambda^{i}\left(X^{\ast}Q(k)\right) has finite degree. XQ(k)X^{\ast}Q(k) is finitely generated by Theorem 4.7, so (XQ(k))i\left(X^{\ast}Q(k)\right)^{\otimes i} is also finitely generated by Corollary 2.27. Quotients of finitely generated 𝖥𝖨\mathsf{FI}-modules are clearly finitely generated, so Λi(XQ(k))\Lambda^{i}\left(X^{\ast}Q(k)\right) is finitely generated. Hence by Corollary 2.26, (XM)Λi(XQ(k))\left(X^{\ast}M\right)\otimes\Lambda^{i}\left(X^{\ast}Q(k)\right) has finite degree. ∎

5.3 Towards a computation of the stable homology

An obvious question to ask now is: what is the stable homology of the Γg,1(k)=Γg,1/g,1(k)\Gamma_{g,1}(k)=\Gamma_{g,1}/\mathcal{I}_{g,1}(k) for k>1k>1, at least rationally? In a recent preprint [12], Szymik shows that in the case of automorphism groups of free nilpotent groups (see Section 3), the canonical homomorphisms AutNnkGL(n,)\operatorname{Aut}N_{n}^{k}\to\mathrm{GL}(n,\mathbb{Z}) are rational homology isomorphisms in the stable range. He does so by showing that for every kk, the homology Lyndon-Hochschild-Serre spectral sequence for the extension

1KnkAutNnkGL(n,)11\to K_{n}^{k}\to\operatorname{Aut}N_{n}^{k}\to\mathrm{GL}(n,\mathbb{Z})\to 1

where KnkAutNnkK_{n}^{k}\lhd\operatorname{Aut}N_{n}^{k} is the kernel, stably collapses on the second page, with all rows except the 0th stably vanishing. Hence the edge homomorphism H(AutNnk;)H(GL(n,);)H_{\ast}(\operatorname{Aut}N_{n}^{k};\mathbb{Q})\to H_{\ast}(\mathrm{GL}(n,\mathbb{Z});\mathbb{Q}) is stably an isomorphism.

It would be tempting to think that an analogous statement might hold in the case of the quotients Γg,1(k)\Gamma_{g,1}(k), but this is in fact not true. Indeed, consider the extensions

1Qg(2)Γg,1(2)Sp(2g,)1.1\to Q_{g}(2)\to\Gamma_{g,1}(2)\to\mathrm{Sp}(2g,\mathbb{Z})\to 1. (3)

By a classical result of Johnson, Qg(2)=g,1/g,1(2)Λ3HgQ_{g}(2)=\mathscr{I}_{g,1}/\mathscr{I}_{g,1}(2)\cong\Lambda^{3}H_{g}. Hence

H(Qg(2);)Λ(Λ3H^g)H^{\ast}(Q_{g}(2);\mathbb{Q})\cong\Lambda^{\ast}\left(\Lambda^{3}\widehat{H}_{g}^{\ast}\right)

where H^gHg\widehat{H}_{g}\coloneqq H_{g}\otimes\mathbb{Q} is the standard rational representation of Sp(2g,)\mathrm{Sp}(2g,\mathbb{Q}) and H^g\widehat{H}_{g}^{\ast} denotes its \mathbb{Q}-linear dual. Hence the Sp(2g,)\mathrm{Sp}(2g,\mathbb{Z})-action on H(Qg(2);)H^{\ast}(Q_{g}(2);\mathbb{Q}) comes from a natural Sp(2g,)\mathrm{Sp}(2g,\mathbb{Q})-action. Hence by the complete reducibility of finite-dimensional Sp(2g,)\mathrm{Sp}(2g,\mathbb{Q})-representations and by Borel’s theorem [2] on the vanishing of the cohomology of Sp(2g,)\mathrm{Sp}(2g,\mathbb{Z}) with coefficients in non-trivial irreducible Sp(2g,)\mathrm{Sp}(2g,\mathbb{Q})-representations,

Hp(Sp(2g,);Hq(Qg(2);))Hp(Sp(2g,);)Hq(Qg(2);)Sp(2g,)H^{p}(\mathrm{Sp}(2g,\mathbb{Z});H^{q}(Q_{g}(2);\mathbb{Q}))\cong H^{p}(\mathrm{Sp}(2g,\mathbb{Z});\mathbb{Q})\otimes H^{q}(Q_{g}(2);\mathbb{Q})^{\mathrm{Sp}(2g,\mathbb{Z})}

when gg is large compared to pp.

The stable rational cohomology of Sp(2g,)\mathrm{Sp}(2g,\mathbb{Z}) is the polynomial algebra

[c1,c3,c5,deg(ci)=2i].\mathbb{Q}\left[c_{1},c_{3},c_{5},\ldots\mid\deg(c_{i})=2i\right].

Kawazumi and Morita [8] show that H(Qg(2);)Sp(2g,)H^{\ast}(Q_{g}(2);\mathbb{Q})^{\mathrm{Sp}(2g,\mathbb{Z})} is stably isomorphic to the polynomial algebra over \mathbb{Q} generated by the set 𝒢\mathcal{G} of finite non-empty connected trivalent graphs (with loops and multiple edges); the degree of a generator Γ𝒢\Gamma\in\mathcal{G} is the order of Γ\Gamma.

Let EE_{\bullet}^{\ast\ast} be the Lyndon-Hochschild-Serre spectral sequence in rational cohomology associated to the extension (3). By the above, the second page is stably given by

E2pq[c1,c3,c5,][𝒢].E_{2}^{pq}\cong\mathbb{Q}[c_{1},c_{3},c_{5},\ldots]\otimes\mathbb{Q}[\mathcal{G}].

But this is concentrated in even bidegrees, so the spectral sequence stably collapses on the second page and we obtain the following result.

Proposition 5.8.

The stable rational cohomology ring of Γg,1(2)\Gamma_{g,1}(2) is isomorphic to

[𝒢{c1,c3,c5,}].\mathbb{Q}[\mathcal{G}\cup\{c_{1},c_{3},c_{5},\ldots\}].

Recall that the stable cohomology ring H(Γ;)H^{\ast}(\Gamma_{\infty};\mathbb{Q}) of the Γg,1\Gamma_{g,1} is a polynomial algebra on the Miller-Morita-Mumford classes eiH2i(Γ;)e_{i}\in H^{2i}(\Gamma_{\infty};\mathbb{Q}). The classes cic_{i} can be chosen so that they pull back to eie_{i} along the projection Γg,1Sp(2g,)\Gamma_{g,1}\to\mathrm{Sp}(2g,\mathbb{Z}). Also, Kawazumi and Morita [8] show that each Γ𝒢\Gamma\in\mathcal{G} (viewed as a stable cohomology class on Γg,1(2)\Gamma_{g,1}(2)) pulls back to ±e|Γ|\pm e_{|\Gamma|}. Hence we see that the projections Γg,1Γg,1(2)Sp(2g,)\Gamma_{g,1}\twoheadrightarrow\Gamma_{g,1}(2)\twoheadrightarrow\mathrm{Sp}(2g,\mathbb{Z}) induce these maps on stable cohomology:

[c1,c3,c5,][𝒢{c1,c3,c5,}]\displaystyle\mathbb{Q}[c_{1},c_{3},c_{5},\ldots]\hookrightarrow\mathbb{Q}[\mathcal{G}\cup\{c_{1},c_{3},c_{5},\ldots\}] [e1,e2,e3,]\displaystyle\twoheadrightarrow\mathbb{Q}[e_{1},e_{2},e_{3},\ldots]
ci\displaystyle c_{i} ei\displaystyle\mapsto e_{i}
Γ\displaystyle\Gamma ±e|Γ|\displaystyle\mapsto\pm e_{|\Gamma|}

6 Infinite loop-space structure

There are ‘pair-of-pants products’ Γg,1×Γh,1Γg+h,1\Gamma_{g,1}\times\Gamma_{h,1}\to\Gamma_{g+h,1} which in a sense generalise the connecting homomorphisms of Definition 4.2. They are constructed as follows: gluing Sg,1S_{g,1} and Sh,1S_{h,1} to two connected components of the pair of pants S0,3S_{0,3} yields a copy of Sg+h,1S_{g+h,1}. The product takes a mapping class on Sg,1S_{g,1} and one on Sh,1S_{h,1} and extends them by the identity on S0,3S_{0,3}. These products make g0BΓg,1\coprod_{g\geqslant 0}B\Gamma_{g,1} into a monoid, and we have

ΩB(g0BΓg,1)×BΓ+\Omega B\left(\coprod_{g\geqslant 0}B\Gamma_{g,1}\right)\simeq\mathbb{Z}\times B\Gamma_{\infty}^{+}

by the group-completion theorem. Here ++ is the Quillen plus-construction and ΓcolimgΓg,1\Gamma_{\infty}\coloneqq\operatorname{colim}_{g}\Gamma_{g,1} is the stable mapping class group. Tillmann [14] proved that this is in fact an infinite loop-space.

We can likewise put a monoid structure on the space Xk:g0BΓg,1(k)X_{k}\coloneqq:\coprod_{g\geqslant 0}B\Gamma_{g,1}(k), and the obvious maps XkXkX_{k}\to X_{k^{\prime}} for 1kk1\leqslant k^{\prime}\leqslant k\leqslant\infty are monoid homomorphisms. In this section we prove the following result (where Γg,1()Γg,1\Gamma_{g,1}(\infty)\coloneqq\Gamma_{g,1}).

Theorem B.

For every k1k\geqslant 1, ΩBXk×BΓ(k)+\Omega BX_{k}\simeq\mathbb{Z}\times B\Gamma_{\infty}(k)^{+} is an infinite loop-space. For every 1kk1\leqslant k^{\prime}\leqslant k\leqslant\infty, the quotient map XkXkX_{k}\to X_{k^{\prime}} group-completes to a map of infinite loop-spaces.

6.1 The surface operad

There is a coherence issue which we swept under the rug in the above definition of monoid structure on XkX_{k}. If we pick the identifications Sg,1S0,3Sh,1Sg+h,1S_{g,1}\cup_{\partial}S_{0,3}\cup_{\partial}S_{h,1}\cong S_{g+h,1} independently of one another, there is no reason why the resulting product structure on XkX_{k} should be associative. It is, however, possible to replace XkX_{k} with an equivalent space which is a strict monoid, and moreover an algebra over the surface operad \mathscr{M}. By a theorem of Tillmann [14], this monoid then group-completes to an infinite loop-space.

We briefly recall the construction of the operad \mathscr{M} from [14] as it will be useful in the proof of Theorem B. From now on, every surface Sg,n+1S_{g,n+1} comes with a marked boundary component. We fix three ‘atomic’ surfaces: a disc DS0,1D\coloneqq S_{0,1}, a pair of pants PS0,3P\coloneqq S_{0,3} and a torus with two boundary components TS1,2T\coloneqq S_{1,2}.

Let g,n\mathscr{E}_{g,n} be the groupoid with objects (S,σ)(S,\sigma) where SSg,n+1S\cong S_{g,n+1} is a surface which is built from the atomic surfaces by identifying the marked boundary on one surface with an unmarked boundary on another surface using the fixed parametrisation, and σ\sigma is an ordering of the free boundary components of SS. The set of morphisms (S,σ)(S,σ)(S,\sigma)\to(S^{\prime},\sigma^{\prime}) in g,n\mathscr{E}_{g,n} is π0Diff(S,S;)\pi_{0}\operatorname{Diff}(S,S^{\prime};\partial), where Diff(S,S;)\operatorname{Diff}(S,S^{\prime};\partial) is the group of diffeomorphisms SSS\to S^{\prime} which are compatible with the fixed collars, which take the marked boundary to the marked boundary and which preserve the ordering of the free boundary components. (Note that such diffeomorphisms must necessarily preserve orientation.) We will suppress σ\sigma from the notation.

Gluing of surfaces yields functors

γ:g,n×h1,i1××hn,ing+h,i\gamma:\mathscr{E}_{g,n}\times\mathscr{E}_{h_{1},i_{1}}\times\cdots\times\mathscr{E}_{h_{n},i_{n}}\to\mathscr{E}_{g+h,i}

where h=jhjh=\sum_{j}h_{j} and i=jiji=\sum_{j}i_{j}. These functors almost define the structure of an operad on the sequence of spaces {gBg,n}n0\{\coprod_{g}B\mathscr{E}_{g,n}\}_{n\geqslant 0} — they are associative and equivariant with respect to the appropriate actions of symmetric groups, but there is no unit in gBg,1\coprod_{g}B\mathscr{E}_{g,1}. This was remedied by Tillmann [14] (with a minor mistake corrected by Wahl [15]):

Theorem 6.1.

There are full subgroupoids 𝒮g,ng,n\mathscr{S}_{g,n}\hookrightarrow\mathscr{E}_{g,n}, retractions R:g,n𝒮g,nR:\mathscr{E}_{g,n}\to\mathscr{S}_{g,n} and functors γ¯\overline{\gamma} which make the diagrams

g,n×h1,i1××hn,in{\mathscr{E}_{g,n}\times\mathscr{E}_{h_{1},i_{1}}\times\cdots\times\mathscr{E}_{h_{n},i_{n}}}g+h,i{\mathscr{E}_{g+h,i}}𝒮g,n×𝒮h1,i1××𝒮hn,in{\mathscr{S}_{g,n}\times\mathscr{S}_{h_{1},i_{1}}\times\cdots\times\mathscr{S}_{h_{n},i_{n}}}𝒮g+h,i{\mathscr{S}_{g+h,i}}γ\scriptstyle{\gamma}R\scriptstyle{R}R\scriptstyle{R}γ¯\scriptstyle{\overline{\gamma}}

commute and which make the sequence {gB𝒮g,n}n0\{\coprod_{g}B\mathscr{S}_{g,n}\}_{n\geqslant 0} of spaces into an operad. Moreover, this data can be chosen in such a way that the pair-of-pants product γ¯(P;,)\overline{\gamma}(P;-,-) is associative in this operad.

Note that provided the functors γ¯\overline{\gamma} exist, they are uniquely specified by the commutativity condition alone.

Definition 6.2.

The operad {gB𝒮g,n}n0\{\coprod_{g}B\mathscr{S}_{g,n}\}_{n\geqslant 0} of the above theorem is the surface operad \mathscr{M}.

The pair-of-pants product makes \mathscr{M}-algebras into strict monoids. The key result of [14] can now be stated as follows.

Theorem 6.3.

Group-completion is a functor from \mathscr{M}-algebras to infinite loop-spaces.

6.2 The \mathscr{M}-algebras B𝒮0(k)B\mathscr{S}_{0}(k)

We construct a sequence of \mathscr{M}-algebras B𝒮0(k)B\mathscr{S}_{0}(k) such that each B𝒮0(k)B\mathscr{S}_{0}(k) is homotopy-equivalent to gBΓg,1(k)\coprod_{g}B\Gamma_{g,1}(k).

Fixing a base-point on S1S^{1} yields, via the fixed parametrisations, a choice of base-point on every boundary component of every object of n=gg,n\mathscr{E}_{n}=\coprod_{g}\mathscr{E}_{g,n}. We define the base-point of SnS\in\mathscr{E}_{n} to be the base-point on the marked boundary, and π1S\pi_{1}S will always be understood as the fundamental group at this base-point.

We now restrict our attention to the groupoid 0\mathscr{E}_{0} of surfaces with a unique boundary component. The elements of Diff(S,S;)\operatorname{Diff}(S,S^{\prime};\partial) preserve base-points for all S,S0S,S^{\prime}\in\mathscr{E}_{0}, so we get a well-defined set map

0(S,S)Iso(π1(S),π1(S)).\mathscr{E}_{0}(S,S^{\prime})\to\operatorname{Iso}(\pi_{1}(S),\pi_{1}(S^{\prime})).

This map is in fact injective by the Dehn-Nielsen-Baer theorem. This motivates the following definition.

Definition 6.4.

For each k1k\geqslant 1, we define the equivalence relation k\sim_{k} on the hom-sets of 0\mathscr{E}_{0} as follows: ϕ,ϕ:SS\phi,\phi^{\prime}:S\to S^{\prime} have ϕkϕ\phi\sim_{k}\phi^{\prime} if and only if ϕ\phi and ϕ\phi^{\prime} induce the same isomorphism

π1S/γk(π1S)π1S/γk(π1S).\pi_{1}S/\gamma_{k}(\pi_{1}S)\to\pi_{1}S^{\prime}/\gamma_{k}(\pi_{1}S^{\prime}).

It can be seen that these relations respect composition in 0\mathscr{E}_{0}. Hence they give rise to quotient groupoids 0(k)0/k\mathscr{E}_{0}(k)\coloneqq\mathscr{E}_{0}/\sim_{k} and 𝒮0(k)𝒮0/k\mathscr{S}_{0}(k)\coloneqq\mathscr{S}_{0}/\sim_{k}.

Observe that for every kk and gg, 𝒮g,0/k\mathscr{S}_{g,0}/\sim_{k} (and g,0/k\mathscr{E}_{g,0}/\sim_{k}) is equivalent to Γg,1(k)\Gamma_{g,1}(k), so we have

B𝒮0(k)gBΓg,1(k).B\mathscr{S}_{0}(k)\simeq\coprod_{g}B\Gamma_{g,1}(k).

We claim that B𝒮0(k)B\mathscr{S}_{0}(k) is an \mathscr{M}-algebra. First we need the following proposition which relates the functors γ\gamma and k\sim_{k}.

Proposition 6.5.

Suppose that ϕ:AA\phi:A\to A^{\prime} is a morphism in n\mathscr{E}_{n} and ψikψi:BiBi\psi_{i}\sim_{k}\psi^{\prime}_{i}:B_{i}\to B^{\prime}_{i} are pairs of k\sim_{k}-equivalent morphisms in 0\mathscr{E}_{0} for 1in1\leqslant i\leqslant n. Then

γ(ϕ;ψ1,,ψn)kγ(ϕ;ψ1,,ψn).\gamma(\phi;\psi_{1},\ldots,\psi_{n})\sim_{k}\gamma(\phi;\psi^{\prime}_{1},\ldots,\psi^{\prime}_{n}).
Proof.

Let us first assume that n=1n=1, ϕ=idA\phi=\operatorname{id}_{A} and B1=B1B_{1}=B^{\prime}_{1}. We drop the subscripts from BB and ψ,ψ\psi,\psi^{\prime}.

By the van Kampen theorem, π1(γ(A;B))π1(A)π\pi_{1}(\gamma(A;B))\cong\pi_{1}(A)\ast_{\partial}\pi, where π\pi is a subgroup of π1(γ(A;B))\pi_{1}(\gamma(A;B)) isomorphic to π1(B)\pi_{1}(B) and \ast_{\partial} denotes amalgamation over the identified boundaries of AA and BB, and similarly π1(γ(A;B))π1(A)π\pi_{1}(\gamma(A;B^{\prime}))\cong\pi_{1}(A)\ast_{\partial}\pi^{\prime}. The isomorphisms ππ1B\pi\cong\pi_{1}B and ππ1B\pi^{\prime}\cong\pi_{1}B^{\prime} are not canonical, but instead depend on a choice of paths in AA from the base-point of AA to the base-point on its free boundary. If we let them both be induced by the same path, we obtain the diagram

π1(A)π1(B){\pi_{1}(A)\ast_{\partial}\pi_{1}(B)}π1(γ(A;B)){\pi_{1}(\gamma(A;B))}π1(A)π1(B){\pi_{1}(A)\ast_{\partial}\pi_{1}(B^{\prime})}π1(γ(A;B)){\pi_{1}(\gamma(A;B^{\prime}))}\scriptstyle{\sim}idψ\scriptstyle{\operatorname{id}\ast_{\partial}\psi_{\ast}}γ(A;ψ)\scriptstyle{\gamma(A;\psi)_{\ast}}\scriptstyle{\sim}

which commutes since the class γ(A;ψ)\gamma(A;\psi) of diffeomorphisms γ(A;B)γ(A;B)\gamma(A;B)\to\gamma(A;B^{\prime}) can be represented by a diffeomorphism which restricts to the identity on AA.

We have an analogous diagram where ψ\psi is replaced with ψ\psi^{\prime}. Upon applying the functor GG/γk(G):𝖦𝗋𝗉𝖦𝗋𝗉G\mapsto G/\gamma_{k}(G):\mathsf{Grp}\to\mathsf{Grp} to both diagrams, the left vertical arrows become equal since ψkψ\psi\sim_{k}\psi^{\prime}. The horizontal arrows are also equal (they are equal even before we quotient out by γk\gamma_{k}), and so γ(A;ψ)kγ(A;ψ)\gamma(A;\psi)\sim_{k}\gamma(A;\psi^{\prime}).

Now suppose BBB\neq B^{\prime}. If ψ,ψ:BB\psi,\psi^{\prime}:B\to B^{\prime} are k\sim_{k}-equivalent, then so are id,ψ1ψ:BB\operatorname{id},\psi^{-1}\psi^{\prime}:B\to B. Hence by the above, γ(A;id)kγ(A;ψ1ψ)\gamma(A;\operatorname{id})\sim_{k}\gamma(A;\psi^{-1}\psi^{\prime}). But we clearly have γ(A;id)=idγ(A;B)\gamma(A;\operatorname{id})=\operatorname{id}_{\gamma(A;B)} and γ(A;ψ1ψ)=γ(A;ψ)1γ(A;ψ)\gamma(A;\psi^{-1}\psi^{\prime})=\gamma(A;\psi)^{-1}\gamma(A;\psi^{\prime}), so γ(A;ψ)kγ(A;ψ)\gamma(A;\psi)\sim_{k}\gamma(A;\psi^{\prime}).

For the most general case, note that

γ(ϕ;ψ1,,ψn)=γ(ϕ;B1,,Bn)i=1nγ(A;B1,,Bi1,ψi,Bi+1,,Bn)\gamma(\phi;\psi_{1},\ldots,\psi_{n})=\gamma(\phi;B^{\prime}_{1},\ldots,B^{\prime}_{n})\prod_{i=1}^{n}\gamma(A;B_{1},\ldots,B_{i-1},\psi_{i},B^{\prime}_{i+1},\ldots,B^{\prime}_{n})

and

γ(A;B1,,Bi1,ψi,Bi+1,,Bn)=γ(Ai;ψi)\gamma(A;B_{1},\ldots,B_{i-1},\psi_{i},B^{\prime}_{i+1},\ldots,B^{\prime}_{n})=\gamma(A_{i};\psi_{i})

where Aiγ(A;B1,,Bi1,,Bi+1,,Bn)A_{i}\coloneqq\gamma(A;B_{1},\ldots,B_{i-1},-,B^{\prime}_{i+1},\ldots,B^{\prime}_{n}). By the above, γ(Ai,ψi)kγ(Ai;ψi)\gamma(A_{i},\psi_{i})\sim_{k}\gamma(A_{i};\psi^{\prime}_{i}), finishing the proof. ∎

Now we are ready to prove Theorem B; it follows easily using Theorem 6.3 from the following result.

Proposition 6.6.

For every kk, the structure functors γ¯:𝒮n×𝒮0n𝒮0\overline{\gamma}:\mathscr{S}_{n}\times\mathscr{S}_{0}^{n}\to\mathscr{S}_{0} descend to functors

θ:𝒮n×𝒮0(k)n𝒮0(k)\theta:\mathscr{S}_{n}\times\mathscr{S}_{0}(k)^{n}\to\mathscr{S}_{0}(k)

along the projection 𝒮0𝒮0(k)\mathscr{S}_{0}\to\mathscr{S}_{0}(k). This induces an \mathscr{M}-algebra structure on each B𝒮0(k)B\mathscr{S}_{0}(k). For every 1kk1\leqslant k^{\prime}\leqslant k\leqslant\infty, the quotient functor 𝒮0(k)𝒮0(k)\mathscr{S}_{0}(k)\to\mathscr{S}_{0}(k^{\prime}) gives rise to a map of \mathscr{M}-algebras upon realisation.

Proof.

Since the projection functor 𝒮0𝒮0(k)\mathscr{S}_{0}\to\mathscr{S}_{0}(k) is full (and the identity on objects), for every nn there is at most one θ\theta making the right square in the diagram

n×0n{\mathscr{E}_{n}\times\mathscr{E}_{0}^{n}}𝒮n×𝒮0n{\mathscr{S}_{n}\times\mathscr{S}_{0}^{n}}𝒮n×𝒮0(k)n{\mathscr{S}_{n}\times\mathscr{S}_{0}(k)^{n}}0{\mathscr{E}_{0}}𝒮0{\mathscr{S}_{0}}𝒮0(k){\mathscr{S}_{0}(k)}R\scriptstyle{R}γ\scriptstyle{\gamma}γ¯\scriptstyle{\overline{\gamma}}θ\scriptstyle{\theta}R\scriptstyle{R} (4)

commute. It is easily seen that if θ\theta exist for all nn, they satisfy the axioms for an algebra over an operad, and any quotient map 𝒮0(k)𝒮0(k)\mathscr{S}_{0}(k)\to\mathscr{S}_{0}(k^{\prime}) commutes with the θ\theta, giving rise to an \mathscr{M}-algebra map B𝒮0(k)B𝒮0(k)B\mathscr{S}_{0}(k)\to B\mathscr{S}_{0}(k^{\prime}). So we only need to check existence.

RR, being a retraction of the inclusion of a full subgroupoid, is fully faithful, so for every Sg,0S\in\mathscr{E}_{g,0}, there is a unique ΦS:SRS\Phi_{S}:S\to RS which is mapped to the identity on RSRS by RR. But then every morphism ϕ:SS\phi:S\to S^{\prime} in g,0\mathscr{E}_{g,0} has

Rϕ=ΦSϕΦS1.R\phi=\Phi_{S^{\prime}}\circ\phi\circ\Phi_{S}^{-1}.

Hence for a parallel pair of morphisms ϕ,ϕ:SS\phi,\phi^{\prime}:S\to S^{\prime}, ϕkϕ\phi\sim_{k}\phi^{\prime} if and only if RϕkRϕR\phi\sim_{k}R\phi^{\prime}.

To show that θ\theta is well defined, let ψ:AA\psi:A\to A^{\prime} be an arrow in 𝒮n\mathscr{S}_{n} and let ϕikϕi:BiBi\phi_{i}\sim_{k}\phi^{\prime}_{i}:B_{i}\to B^{\prime}_{i} be nn pairs of k\sim_{k}-equivalent morphisms in 𝒮0\mathscr{S}_{0}. Then γ(ψ;ϕ1,,ϕn)kγ(ψ;ϕ1,,ϕn)\gamma(\psi;\phi_{1},\ldots,\phi_{n})\sim_{k}\gamma(\psi;\phi^{\prime}_{1},\ldots,\phi^{\prime}_{n}) in 0\mathscr{E}_{0} by the preceding proposition. Hence by the above observation about RR and using the fact that on 𝒮n×𝒮0n\mathscr{S}_{n}\times\mathscr{S}_{0}^{n}, γ¯=Rγ\overline{\gamma}=R\circ\gamma, we see that

γ¯(ψ;ϕ1,,ϕn)kγ¯(ψ;ϕ1,,ϕn),\overline{\gamma}(\psi;\phi_{1},\ldots,\phi_{n})\sim_{k}\overline{\gamma}(\psi;\phi^{\prime}_{1},\ldots,\phi^{\prime}_{n}),

and so γ¯\overline{\gamma} descends to a well-defined functor θ\theta. ∎

References

  • Andreadakis [1965] S. Andreadakis. On the automorphisms of free groups and free nilpotent groups. Proc. London Math. Soc., 3(15):239–268, 1965. doi:10.1112/plms/s3-15.1.239.
  • Borel [1981] A. Borel. Stable real cohomology of arithmetic groups II. In Manifolds and Lie groups: papers in honor of Yozo Matsushima, volume 14 of Progress in Mathematics, pages 21–55. Birkhäuser, 1981.
  • Church and Ellenberg [2016] T. Church and J. S. Ellenberg. Homology of 𝖥𝖨\mathsf{FI}-modules. arXiv: 1506.01022v2, 2016.
  • Church and Putman [2015] T. Church and A. Putman. Generating the Johnson filtration. Geom. Topol., 19(4):2217–2255, 2015. doi:10.2140/gt.2015.19.2217.
  • Church et al. [2014] T. Church, J. S. Ellenberg, B. Farb, and R. Nagpal. 𝖥𝖨\mathsf{FI}-modules over Noetherian rings. Geom. Topol., 18(5):2951–2984, 2014. doi:10.2140/gt.2014.18.2951.
  • Galatius [2011] S. Galatius. Stable homology of automorphism groups of free groups. Ann. of Math., 173(2):705–768, 2011. doi:10.4007/annals.2011.173.2.3.
  • Galatius et al. [2009] S. Galatius, I. Madsen, U. Tillmann, and M. Weiss. The homotopy type of the cobordism category. Acta Math., 202(2):195–239, 2009. doi:10.1007/s11511-009-0036-9.
  • Kawazumi and Morita [1996] N. Kawazumi and S. Morita. The primary approximation to the cohomology of the moduli space of curves and cocycles for the stable characteristic classes. Math. Res. Lett., 3(5):629–641, 1996. doi:10.4310/MRL.1996.v3.n5.a6.
  • Randal-Williams and Wahl [2015] O. Randal-Williams and N. Wahl. Homological stability for automorphism groups. arXiv: 1409.3541v3, 2015.
  • Serre [2006] J.-P. Serre. Lie algebras and Lie groups. Number 1500 in Lecture Notes in Mathematics. Springer, 2006.
  • Szymik [2014] M. Szymik. Twisted homological stability for extensions and automorphism groups of free nilpotent groups. Journal of K-theory: K-theory and its Applications to Algebra, Geometry, and Topology, 14(1):185–201, 2014. doi:10.1017/is014005031jkt267.
  • Szymik [2016] M. Szymik. The rational stable homology of mapping class groups of universal nil-manifolds. arXiv: 1605.06508, 2016.
  • Tillmann [1997] U. Tillmann. On the homotopy of the stable mapping class group. Invent. Math., 130(2):257–275, 1997. doi:10.1007/s002220050184.
  • Tillmann [2000] U. Tillmann. Higher genus surface operad detects infinite loop spaces. Math. Ann., 317(3):613–628, 2000. doi:10.1007/pl00004416.
  • Wahl [2004] N. Wahl. Infinite loop space structure(s) on the stable mapping class group. Topology, 43(2):343–368, 2004. doi:10.1016/S0040-9383(03)00045-4.
  • Wahl [2013] N. Wahl. Homological stability for mapping class groups of surfaces. In Handbook of moduli, Vol. III, number 26 in Adv. Lect. Math., pages 547–583. 2013. URL http://www.math.ku.dk/~wahl/MCGsurvey.pdf.