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

Cutting and Pasting in the Torelli subgroup of Out(Fn)\operatorname{Out}(F_{n})

Jacob Landgraf111This material is based upon work supported by the National Science Foundation under Grant Number DMS-1547292
Abstract

Using ideas from 3-manifolds, Hatcher–Wahl defined a notion of automorphism groups of free groups with boundary. We study their Torelli subgroups, adapting ideas introduced by Putman for surface mapping class groups. Our main results show that these groups are finitely generated, and also that they satisfy an appropriate version of the Birman exact sequence.

1 Introduction

Let Fn=x1,,xnF_{n}=\langle x_{1},\ldots,x_{n}\rangle be the free group on nn letters, and let Out(Fn)\operatorname{Out}(F_{n}) be the group of outer automorphisms of FnF_{n}. In many ways, Out(Fn)\operatorname{Out}(F_{n}) behaves very similarly to Mod(Σg,b)\operatorname{Mod}(\Sigma_{g,b}), the mapping class group of the surface Σg,b\Sigma_{g,b} of genus gg with bb boundary components. For an overview of some of these similarities, see [7].

One such connection is that they both contain a Torelli subgroup. In the mapping class group, the Torelli subgroup (Σg,b)Mod(Σg,b)\operatorname{\mathcal{I}}(\Sigma_{g,b})\subset\operatorname{Mod}(\Sigma_{g,b}) is defined to be the kernel of the action on H1(Σn,b;)H_{1}(\Sigma_{n,b};\mathbb{Z}) for b=0,1b=0,1. In Out(Fn)\operatorname{Out}(F_{n}), we define a similar subgroup222It is also common to see this group denoted by IAnIA_{n}, but we wish to reserve this notation for the analogous subgroup of Aut(Fn)\operatorname{Aut}(F_{n}), denoted IOnIO_{n}, as the kernel of the action of Out(Fn)\operatorname{Out}(F_{n}) on H1(Fn;)=nH_{1}(F_{n};\mathbb{Z})=\mathbb{Z}^{n}.

On surfaces with multiple boundary components, there are many possible definitions one might use to define a Torelli subgroup of Mod(Σg,b)\operatorname{Mod}(\Sigma_{g,b}). In [22], Putman defines a Torelli subgroup (Σg,b,P)\operatorname{\mathcal{I}}(\Sigma_{g,b},P) for b>1b>1 requiring the additional data of a partition PP of the boundary components. The goal of the current paper is to mirror Putman’s procedure to define an “IOnIO_{n} with boundary.”

Let Mn,b=#n(S1×S2)(b open 3-disks)M_{n,b}=\#_{n}(S^{1}\times S^{2})\setminus(b\text{ open 3-disks}). For simplicity, we will write MnM_{n} if b=0b=0. A key property of Mn,bM_{n,b} is that it has fundamental group FnF_{n}. Fix such an identification. The mapping class group Mod(Mn,b)\operatorname{Mod}(M_{n,b}) is the group of orientation-preserving diffeomorphisms of Mn,bM_{n,b} fixing the boundary pointwise modulo isotopies fixing the boundary pointwise. Letting Diff+(Mn,b,Mn,b)\operatorname{Diff^{+}}(M_{n,b},\partial M_{n,b}) be the topological group of diffeomorphisms fixing the boundary pointwise, we can also write Mod(Mn,b)=π0(Diff+(Mn,b,Mn,b))\operatorname{Mod}(M_{n,b})=\pi_{0}(\operatorname{Diff^{+}}(M_{n,b},\partial M_{n,b})). By a theorem of Laudenbach [19], there is an exact sequence

1(/2)nMod(Mn)Out(Fn)1,1\to(\mathbb{Z}/2)^{n}\to\operatorname{Mod}(M_{n})\to\operatorname{Out}(F_{n})\to 1, (1)

where the map Mod(Mn)Out(Fn)\operatorname{Mod}(M_{n})\to\operatorname{Out}(F_{n}) is given by the action (up to conjugation) on π1(Mn)\pi_{1}(M_{n}), and the (/2)n(\mathbb{Z}/2)^{n} is generated by sphere twists about nn disjointly embedded 22-spheres (see Section 2 for the definition and relevant properties of sphere twists). Recent work of Brendle, Broaddus, and Putman [6] shows that this sequence actually splits as a semidirect product. This exact sequence implies that, modulo a finite group, Out(Fn)\operatorname{Out}(F_{n}) acts on MnM_{n} up to isotopy. Therefore, MnM_{n} plays almost the same role for Out(Fn)\operatorname{Out}(F_{n}) that Σg,b\Sigma_{g,b} plays for Mod(Σg,b)\operatorname{Mod}(\Sigma_{g,b}).

Adding boundary components.

From Laudenbach’s sequence (1), we see that Out(Fn)Mod(Mn)/STwist(Mn)\operatorname{Out}(F_{n})\cong\operatorname{Mod}(M_{n})/\text{STwist}(M_{n}), where STwist(Mn)(/2)n\text{STwist}(M_{n})\cong(\mathbb{Z}/2)^{n} is the subgroup of Mod(Mn)\operatorname{Mod}(M_{n}) generated by sphere twists. Now that we have related Out(Fn)\operatorname{Out}(F_{n}) to a geometrically defined group, we can start introducing boundary components. Extending the relationship given by Laudenbach’s sequence, we define “Out(Fn)\operatorname{Out}(F_{n}) with boundary” as

Out(Fn,b):=Mod(Mn,b)/STwist(Mn,b).\mathrm{Out}(F_{n,b}):=\operatorname{Mod}(M_{n,b})/\text{STwist}(M_{n,b}).

When b=1b=1, Laudenbach [19] also shows that Out(Fn,1)Aut(Fn)\mathrm{Out}(F_{n,1})\cong\operatorname{Aut}(F_{n}). Hatcher-Wahl [14] introduced a more general version of Out(Fn,b)\mathrm{Out}(F_{n,b}), which they denoted by An,ksA_{n,k}^{s}. The original definition of An,ksA_{n,k}^{s} has to do with classes of self-homotopy equivalences of a certain graph. However, in [14] the authors give an equivalent definition, which says that An,ksA_{n,k}^{s} is the mapping class group of MnM_{n} with ss spherical and kk toroidal boundary components, modulo sphere twists. With this definition, we see that Out(Fn,b)=An,0b\mathrm{Out}(F_{n,b})=A_{n,0}^{b}. Similar groups have been examined in the work of Jensen-Wahl [16] and Wahl [26]. Their versions, however, involve only toroidal boundary components, and thus are distinct from Out(Fn,b)\mathrm{Out}(F_{n,b}).

Torelli subgroups.

An important feature of sphere twists (discussed in Section 2) is that they act trivially on homotopy classes of embedded loops, and thus act trivially on H1(Mn)H_{1}(M_{n}). Therefore, the action of Mod(Mn,b)\operatorname{Mod}(M_{n,b}) on H1(Mn,b)H_{1}(M_{n,b}) induces an action of Out(Fn,b)\mathrm{Out}(F_{n,b}) on H1(Mn,b)H_{1}(M_{n,b}). We can then define the Torelli subgroup IOn,bOut(Fn,b)IO_{n,b}\subset\mathrm{Out}(F_{n,b}) to be the kernel of this action. However, this definition does not capture all homological information when b>1b>1, especially when Mn,bM_{n,b} is being embedded in Mm,cM_{m,c}. To see why, consider the scenario depicted in Figure 1, in which M2,2M_{2,2} has been embedded into M4M_{4}. This embedding induces a homomorphism ιM:Mod(M2,2)Mod(M4)\iota_{M}:\operatorname{Mod}(M_{2,2})\to\operatorname{Mod}(M_{4}) obtained by extending by the identity. This map sends sphere twists to sphere twists, and so we get an induced map ι:Out(F2,2)Out(F4)\iota_{*}:\mathrm{Out}(F_{2,2})\to\operatorname{Out}(F_{4}). However, this does not restrict to a map IO2,2IO4IO_{2,2}\to IO_{4} under this definition of IOn,bIO_{n,b} since elements of IO2,2IO_{2,2} are not required to fix the homology class of the subarc of α\alpha lying inside M2,2M_{2,2}. To address this issue, we will use a slightly modified homology group.

Definition.

Fix a partition PP of the boundary components of Mn,bM_{n,b}.

  1. (a)

    Two boundary components 1,2\partial_{1},\partial_{2} of Mn,bM_{n,b} are PP-adjacent if there is some pPp\in P such that {1,2}p\{\partial_{1},\partial_{2}\}\subset p.

  2. (b)

    Let H1P(Mn,b)H_{1}^{P}(M_{n,b}) be the subgroup of H1(Mn,b,Mn,b)H_{1}(M_{n,b},\partial M_{n,b}) spanned by

    {[h]H1(Mn,b,Mn,b)\displaystyle\{[h]\in H_{1}(M_{n,b},\partial M_{n,b})\mid either hh is a simple closed curve or
    hh is a properly embedded arc with endpoints
    in distinct P-adjacent boundary components}.\displaystyle\text{ in distinct $P$-adjacent boundary components}\}.
  3. (c)

    There is a natural action of Out(Fn,b)\mathrm{Out}(F_{n,b}) on H1P(Mn,b)H_{1}^{P}(M_{n,b}), and we define the Torelli subgroup IOn,bPOut(Fn,b)IO_{n,b}^{P}\subset\mathrm{Out}(F_{n,b}) to be the kernel of this action.

Returning to Figure 1, let PP be the trivial partition of the boundary components of M2,2M_{2,2} with a single PP-adjacency class. With this choice of partition, we see that [αM2,2]H1P(M2,2)[\alpha\cap M_{2,2}]\in H_{1}^{P}(M_{2,2}). If fIO2,2Pf\in IO_{2,2}^{P}, then it follows that ι(f)Out(F4)\iota_{*}(f)\in\operatorname{Out}(F_{4}) preserves the homology class of α\alpha. Therefore, ι(f)IO4\iota_{*}(f)\in IO_{4}, and so ι\iota_{*} restricts to a map IO2,2PIO4IO_{2,2}^{P}\to IO_{4}.

\labellist
\hair

2pt \pinlabelα\alpha at 210 125 \pinlabelα\alpha at 360 125 \pinlabelM2,2M_{2,2} at 130 125 \pinlabelM1,2M_{1,2} at 445 125 \endlabellistRefer to caption

Figure 1: A copy of M2,2M_{2,2} and M1,2M_{1,2} glued together to obtain M4M_{4}. We realize M2,2M_{2,2} as a 3-sphere with the six indicated open balls removed, then the boundaries of these removed balls are identified according to the arrows (and similarly for M1,2M_{1,2}). The class [α][\alpha] need not be fixed by elements of IO2,2IO_{2,2} with the naïve definition.

Restriction.

As we discussed in the last paragraph, given an embedding ι:Mn,bMm\iota:M_{n,b}\hookrightarrow M_{m}, we can extend by the identity to get a map ι:Out(Fn,b)Out(Fm)\iota_{*}:\mathrm{Out}(F_{n,b})\to\operatorname{Out}(F_{m}). In general, ι\iota_{*} may not be injective. However, it is injective if no connected component of MmMn,bM_{m}\setminus M_{n,b} is diffeomorphic to D3D^{3} (see Appendix A). Moreover, such an embedding induces a natural partition of the boundary components of Mn,bM_{n,b} as follows.

Definition.

Fix an embedding ι:Mn,bMm\iota:M_{n,b}\hookrightarrow M_{m}. Let NN be a connected component of Mmint(Mn,b)M_{m}\setminus\operatorname{int}(M_{n,b}), and let pNp_{N} be the set of boundary components of Mn,bM_{n,b} shared with NN. Then the partition PP of the boundary components of Mn,bM_{n,b} induced by ι\iota is defined to be

P={pNN a connected component of MmMn,b}.P=\{p_{N}\mid N\text{ a connected component of }M_{m}\setminus M_{n,b}\}.

With this definition, one might guess that ι1(IOn)=IOn,bP\iota_{*}^{-1}(IO_{n})=IO_{n,b}^{P}. This turns out to be the case, and this is our first main theorem, which we prove in Section 3.

Theorem A (Restriction Theorem).

Let ι:Mn,bMm\iota:M_{n,b}\hookrightarrow M_{m} be an embedding, ι:Out(Fn,b)Out(Fm)\iota_{*}:\mathrm{Out}(F_{n,b})\to\operatorname{Out}(F_{m}) the induced map, and PP the induced partition of the boundary components of Mn,bM_{n,b}. Then IOn,bP=ι1(IOm)IO_{n,b}^{P}=\iota_{*}^{-1}(IO_{m}).

Birman exact sequence.

From here, we move on to exploring the parallels between these Torelli subgroups and those of mapping class groups. There is a well-known relationship between the mapping class groups of surfaces with a different number of boundary components called the Birman exact sequence (see [12]):

1π1(UT(Σn,b1))Mod(Σg,b)Mod(Σg,b1)1.1\to\pi_{1}(UT(\Sigma_{n,b-1}))\to\operatorname{Mod}(\Sigma_{g,b})\to\operatorname{Mod}(\Sigma_{g,b-1})\to 1.

Here, UT(Σn,b1)UT(\Sigma_{n,b-1}) is the unit tangent bundle of Σn,b1\Sigma_{n,b-1}, the map π1(UT(Σn,b1))Mod(Σg,b)\pi_{1}(UT(\Sigma_{n,b-1}))\to\operatorname{Mod}(\Sigma_{g,b}) is given by pushing a boundary component around a loop, and the map Mod(Σg,b)Mod(Σg,b1)\operatorname{Mod}(\Sigma_{g,b})\to\operatorname{Mod}(\Sigma_{g,b-1}) is given by attaching a disk onto this boundary component. In Section 4, we will prove versions of the Birman exact sequence for Mod(Mn,b)\operatorname{Mod}(M_{n,b}) and Out(Fn,b)\mathrm{Out}(F_{n,b}), all culminating in the following sequence for IOn,bPIO_{n,b}^{P}.

Theorem B (Birman exact sequence).

Fix n,b>0n,b>0 such that (n,b)(1,1)(n,b)\neq(1,1), and let Mn,bMn,b1M_{n,b}\hookrightarrow M_{n,b-1} be an embedding obtained by gluing a ball to the boundary component \partial. Fix xMn,b1Mn,bx\in M_{n,b-1}\setminus M_{n,b}. Let PP be a partition of the boundary components of Mn,bM_{n,b}, let PP^{\prime} be the induced partition of the boundary components of Mn,b1M_{n,b-1}, and let pPp\in P be the set containing \partial. We then have an exact sequence

1LIOn,bPιIOn,b1P1,1\to L\to IO_{n,b}^{P}\overset{\iota_{*}}{\to}IO_{n,b-1}^{P^{\prime}}\to 1,

where LL is equal to:

  1. (a)

    π1(Mn,b1,x)Fn\pi_{1}(M_{n,b-1},x)\cong F_{n} if p={}p=\{\partial\}.

  2. (b)

    [π1(Mn,b1,x),π1(Mn,b1,x)][Fn,Fn][\pi_{1}(M_{n,b-1},x),\pi_{1}(M_{n,b-1},x)]\cong[F_{n},F_{n}] if p{}p\neq\{\partial\}.

Moreover, this sequence splits if b2b\geq 2.

Remark.

This theorem may seem superficially similar to results proven by Day-Putman in [9] and [11]. However, we consider a very different notion of “automorphisms with boundary,” and so these results are unrelated.

Finite generation.

Once we have established this version of the Birman exact sequence, in Section 5, we will define a generating set for IOn,bPIO_{n,b}^{P}. This generating set will be inspired by the generating set for IOnIO_{n} found by Magnus [21] in 1935.

Theorem 1.1 (Magnus).

Let Fn=x1,,xnF_{n}=\langle x_{1},\ldots,x_{n}\rangle. The group IOnIO_{n} is generated by the Out(Fn)\operatorname{Out}(F_{n})-classes of the automorphisms

Mij:xixjxixj1,Mijk:xixi[xj,xk],M_{ij}:x_{i}\mapsto x_{j}x_{i}x_{j}^{-1},\qquad M_{ijk}:x_{i}\mapsto x_{i}[x_{j},x_{k}],

for all distinct i,j,k{1,,n}i,j,k\in\{1,\ldots,n\} with j<kj<k. Here, the automorphisms are understood to fix xx_{\ell} for i\ell\neq i.

Throughout this paper, we will use the convention [a,b]=aba1b1[a,b]=aba^{-1}b^{-1}. Since we defined IOn,bPIO_{n,b}^{P} to be a subgroup of Mod(Mn,b)/STwist(Mn,b)\operatorname{Mod}(M_{n,b})/\text{STwist}(M_{n,b}), our generators will be defined geometrically rather than algebraically. However, in the case of b=0b=0, they will reduce directly to Magnus’s generators. In Section 6, we will show that these elements do indeed generate IOn,bPIO_{n,b}^{P}.

Theorem C.

The group IOn,bPIO_{n,b}^{P} is finitely generated for n1n\geq 1, b0b\geq 0.

This is rather striking because the analogous result for Torelli subgroups of mapping class groups with multiple boundary components is still open. We will prove this theorem by using the Birman exact sequence to reduce to b=0b=0 and applying Magnus’s theorem. Unfortunately, the tools we have constructed do not seem strong enough to give a novel proof of Magnus’s theorem. We will, however, prove a weaker version in Section 7. The original proof of Magnus’s Theorem 1.1 comes in two steps: showing that the given automorphisms Out(Fn)\operatorname{Out}(F_{n})-normally generate IOnIO_{n}, and then showing that the subgroup they generate is normal in Out(Fn)\operatorname{Out}(F_{n}). We will give a proof of the first step in our setting. For alternative proofs of the first step, as well as more information on the second step in this context, see [5] and [10].

Theorem D.

The group IOnIO_{n} is Out(Fn)\operatorname{Out}(F_{n})-normally generated by the automorphisms MijM_{ij} and MijkM_{ijk}, where i,j,k{1,,n}i,j,k\in\{1,\ldots,n\} and j<kj<k.

Abelianization.

Once we have a finite generating set for IOn,bPIO_{n,b}^{P}, a natural question arises: how does the cardinality of this generating set compare to the rank of H1(IOn,bP)H_{1}(IO_{n,b}^{P})? For b1b\leq 1, this question is answered by a result of Andreadakis [1] and Bachmuth [3].

Theorem 1.2 (Andreadakis, Bachmuth).

The abelianization of IOn,bIO_{n,b} is torsion-free of rank n(n2)nn\cdot\binom{n}{2}-n if n=0n=0, and rank n(n2)n\cdot\binom{n}{2} if n=1n=1.

This theorem was proved using a version of the Johnson homomorphism

τ:IAnHom(H,2H),\tau:IA_{n}\to\operatorname{Hom}(H,\wedge^{2}H),

where H=H1(Fn)=nH=H_{1}(F_{n})=\mathbb{Z}^{n}. We will recall the definition of this homomorphism in Section 8, along with the proof of Theorem 1.2. We then move on to computing the rank of H1(IOn,bP)H_{1}(IO_{n,b}^{P}) for b>1b>1. To do this, we choose an embedding Mn,bMm,1M_{n,b}\hookrightarrow M_{m,1}, which induces an injection IOn,bPIOm,1=IAmIO_{n,b}^{P}\to IO_{m,1}=IA_{m}. Composing this map with τ\tau gives a map τ:IOn,bPHom(H,2H)\tau_{*}:IO_{n,b}^{P}\to\operatorname{Hom}(H,\wedge^{2}H). We then compute the image of our generators under τ\tau_{*}, and use this to count the rank of τ(IOn,bP)\tau_{*}(IO_{n,b}^{P}).

Theorem E.

The abelianization of IOn,bPIO_{n,b}^{P} is torsion-free of rank

n(n2)+(b(n2)|P|(n2))+(|P|nn).n\cdot\binom{n}{2}+\left(b\cdot\binom{n}{2}-|P|\cdot\binom{n}{2}\right)+(|P|\cdot n-n).

Acknowledgements.

I would like to thank my advisor Andy Putman for directing me to Out(Fn)\operatorname{Out}(F_{n}) and its Torelli subgroup, and for his input during the revision process. I would also like to thank Patrick Heslin and Aaron Tyrrell for helpful conversations regarding diffeomorphism groups, as well as Dan Margalit for an enlightening question which resulted in the addition of Section 8.

Outline.

In section 2, we will give a short overview of sphere twists. We then move on to proving Theorem A in Section 3. We will establish all of our versions of the Birman exact sequence (including Theorem B) in Section 4. In Section 5, we will define our candidate generators for IOn,bPIO_{n,b}^{P}, and we will prove that they generate (Theorem C) in Section 6 using the Birman exact sequence and Magnus’s Theorem 1.1. In Section 7, we will prove Theorem D. We then move on to Section 8, in which we recall the definition of the Johnson homomorphism for IAnIA_{n}, and use it to compute the rank of the abelianization of IOn,bPIO_{n,b}^{P}, proving Theorem E. Finally, we conclude with two appendices. In Appendix A, we provide conditions for a map Out(Fn,b)Out(Fm)\mathrm{Out}(F_{n,b})\to\operatorname{Out}(F_{m}) induced by an inclusion to be injective, and in Appendix B we prove a lemma which allows us to realize bases of H2(Mm)H_{2}(M_{m}) as collections of disjoint oriented spheres.

Figure conventions.

We will frequently direct the reader to figures which are intended to give some geometric intuition for the manifold Mn,bM_{n,b}. In order to assemble Mn,bM_{n,b}, we begin with one or more copies of S3S^{3}, remove a collection of open balls, and then glue the resulting boundary components together in pairs. These gluings will be indicated by double-sided arrows connecting the boundary spheres being glued. As an example, see Figure 2.

\labellist
\hair

2pt \pinlabelM1,2M_{1,2} at 88 88 \pinlabelM1,3M_{1,3} at 290 88 \pinlabelM2,1M_{2,1} at 500 88 \endlabellistRefer to caption

Figure 2: M5M_{5} realized by gluing M1,2M_{1,2}, M1,3M_{1,3}, and M2,1M_{2,1} together along their boundaries as indicated by the arrows.

2 Preliminaries

Since sphere twists play a fundamental role throughout the remainder of the paper, we will give a brief overview of them here.

Sphere twists.

Fix a smoothly embedded 2-sphere SMn,bS\subset M_{n,b}, and let US×[0,1]U\cong S\times[0,1] be a tubular neighborhood of SS. Recall that π1(SO(3),id)/2\pi_{1}(\operatorname{SO}(3),\operatorname{id})\cong\mathbb{Z}/2\mathbb{Z}, and the nontrivial element γ:[0,1]SO(3)\gamma:[0,1]\to\operatorname{SO}(3) is given by rotating 3\mathbb{R}^{3} one full revolution about any fixed axis through the origin. Fix an identification S=S23S=S^{2}\subset\mathbb{R}^{3}. Then, we define the sphere twist about SS, denoted TSMod(Mn,b)T_{S}\in\operatorname{Mod}(M_{n,b}), to be the class of the diffeomorphism which is the identity on Mn,bUM_{n,b}\setminus U and is given by (x,t)(γ(t)x,t)(x,t)\mapsto(\gamma(t)\cdot x,t) on US×[0,1]U\cong S\times[0,1]. The isotopy class of TST_{S} depends only on the isotopy class of SS. In fact, more is true: Laudenbach [19] showed that the class of TST_{S} depends only on the homotopy class of SS.

Action on curves and surfaces.

Since π1(SO(3),id)/2\pi_{1}(\operatorname{SO}(3),\operatorname{id})\cong\mathbb{Z}/2\mathbb{Z}, we see that sphere twists have order at most two. However, it is tricky to show that sphere twists are actually nontrivial because they act trivially on homotopy classes of embedded arcs and surfaces. To see why this is true, let SMn,bS\subset M_{n,b} be an embedded 2-sphere, and let U=S×[0,1]U=S\times[0,1] be a tubular neighborhood of SS. Suppose that α\alpha is an arc or surface embedded in Mn,bM_{n,b}. We can homotope α\alpha such that it is either disjoint from UU or intersects UU transversely. Let pSp\in S be one of points in SS which lies on the axis of rotation used to construct TST_{S}. We can homotope α\alpha such that αU\alpha\cap U collapses into p[0,1]p\in[0,1]. Note that this process is not an isotopy, and α\alpha is no longer embedded in Mn,bM_{n,b}. This is not an issue because a result of Laudenbach [19] shows that if α\alpha is fixed up to homotopy, then it is fixed up to isotopy. Since TST_{S} fixes p×[0,1]p\times[0,1] pointwise, it follows that TST_{S} fixes α\alpha up to homotopy. The upshot of this is that a more sophisticated invariant must be constructed to detect the nontriviality of TST_{S}. In [18, 19], Laudenbach uses framed cobordisms to show that for b=0,1b=0,1, the sphere twist TST_{S} is trivial if and only if SS is separating. In the case of no boundary components, Brendle, Broaddus, and Putman [6] give another proof of this fact by showing that sphere twists act nontrivially on a trivialization of the tangent bundle of MnM_{n} up to isotopy.

Sphere twist subgroup.

Let STwist(Mn,b)Mod(Mn,b)\text{STwist}(M_{n,b})\subset\operatorname{Mod}(M_{n,b}) be the subgroup generated by sphere twists. Given 𝔣Mod(Mn,b)\mathfrak{f}\in\operatorname{Mod}(M_{n,b}) and a sphere twists TST_{S}, we have the “change of coordinates” formula

𝔣TS𝔣1=T𝔣(S).\mathfrak{f}T_{S}\mathfrak{f}^{-1}=T_{\mathfrak{f}(S)}.

This shows that STwist(Mn,b)\text{STwist}(M_{n,b}) is a normal subgroup of Mod(Mn,b)\operatorname{Mod}(M_{n,b}). In fact, even more is true. Letting 𝔣=TS\mathfrak{f}=T_{S^{\prime}} in the above formula and using the fact that sphere twists act trivially on embedded surfaces up to isotopy, we find that

TSTSTS1=TTS(S)=TS,T_{S^{\prime}}T_{S}T_{S^{\prime}}^{-1}=T_{T_{S^{\prime}}(S)}=T_{S},

which implies STwist(Mn,b)\text{STwist}(M_{n,b}) is actually abelian. Since nontrivial sphere twists have order two, it follows that STwist(Mn,b)\text{STwist}(M_{n,b}) is isomorphic to a product of copies of /2\mathbb{Z}/2\mathbb{Z}. For b=0,1b=0,1, another result of Laudenbach shows that STwist(Mn,b)(/2)n\text{STwist}(M_{n,b})\cong(\mathbb{Z}/2\mathbb{Z})^{n} and is generated by the sphere twists about the nn core spheres ×S2*\times S^{2} in each S1×S2S^{1}\times S^{2} summand. For b>1b>1, one can show that STwist(Mn,b)(/2)n+b1\text{STwist}(M_{n,b})\cong(\mathbb{Z}/2\mathbb{Z})^{n+b-1}. The 1-1 in the exponent reflects the fact that the product of all the sphere twists about boundary components is trivial. Since we will need this fact later, we include a proof here.

\labellist
\hair

2pt \pinlabelzz at 310 680 \pinlabelS1S_{1} at 60 395 \pinlabelS2S_{2} at 350 490 \pinlabelS3S_{3} at 350 300 \pinlabelS4S_{4} at 350 110 \endlabellistRefer to caption

Figure 3: M0,4M_{0,4} embedded in 3\mathbb{R}^{3}.
Lemma 2.1.

If S1,,SbMn,bS_{1},\ldots,S_{b}\subset M_{n,b} be spheres parallel to the bb boundary components of Mn,bM_{n,b}, then the element TS1TSbT_{S_{1}}\cdots T_{S_{b}} is trivial in Mod(Mn,b)\operatorname{Mod}(M_{n,b}).

Proof.

We will prove this by induction on nn. As the base case, consider M0,bM_{0,b}. The argument in this case follows a proof of Hatcher and Wahl [15, Pg. 214-215], but we include the proof here as well for completeness. If b=0b=0, then the statement is trivial. If b>0b>0, then we can embed M0,bM_{0,b} in 3\mathbb{R}^{3} as the unit ball with b1b-1 smaller balls removed along the zz-axis (see Figure 3). We may then use the zz-axis as the axis of rotation for the sphere twists about all the boundary components. Taking S1S_{1} to be the unit sphere, we then see that the product T2TbT_{2}\cdots T_{b} is isotopic to T1T_{1}. Since sphere twists have order two, this gives the desired relation, and so we have completed the base case.

Next, consider Mn,bM_{n,b} for n>0n>0. Since n>0n>0, there exists a nonseparating sphere SMn,bS\subset M_{n,b} which is disjoint from S1,,SbS_{1},\ldots,S_{b}. Splitting Mn,bM_{n,b} along SS yields a submanifold diffeomorphic to Mn1,b+2M_{n-1,b+2}. Let ιM:Mod(Mn1,b+2)Mod(Mn,b)\iota_{M}:\operatorname{Mod}(M_{n-1,b+2})\to\operatorname{Mod}(M_{n,b}) be the map induced by inclusion. Let T1,,Tb+2T_{1},\ldots,T_{b+2} be the sphere twists about the boundary components of Mn1,b+2M_{n-1,b+2}, and order them such that ιM(Tj)=TSj\iota_{M}(T_{j})=T_{S_{j}} for 0jb0\leq j\leq b. With this ordering, notice that ιM(Tb+1)=ιM(Tb+2)=TS\iota_{M}(T_{b+1})=\iota_{M}(T_{b+2})=T_{S}. Since sphere twists have order two,

ιM(T1Tb+2)=TS1TSbTS2=TS1TSb.\iota_{M}(T_{1}\cdots T_{b+2})=T_{S_{1}}\cdots T_{S_{b}}\cdot T_{S}^{2}=T_{S_{1}}\cdots T_{S_{b}}.

By our induction hypothesis, T1Tb+2T_{1}\cdots T_{b+2} is trivial in Mod(Mn1,b+2)\operatorname{Mod}(M_{n-1,b+2}), and so we are done. ∎

If b=1b=1, this shows that the sphere twist about the boundary component is trivial. However, if b>1b>1, then the sphere twists about boundary components are nontrivial. We will also need this fact, so we prove it here.

Lemma 2.2.

Let b>1b>1, and let \partial be a boundary component of Mn,bM_{n,b}. Then TMod(Mn,b)T_{\partial}\in\operatorname{Mod}(M_{n,b}) is nontrivial.

Proof.

Let \partial^{\prime} be a boundary component of Mn,bM_{n,b} different from \partial. Then we get an embedding ι:Mn,bMn+1\iota:M_{n,b}\hookrightarrow M_{n+1} by attaching \partial and \partial^{\prime} with a copy of S2×IS^{2}\times I, and capping off all the remainding boundary components. Let ιM:Mod(Mn,b)Mod(Mn+1)\iota_{M}:\operatorname{Mod}(M_{n,b})\to\operatorname{Mod}(M_{n+1}) be the map induced by ι\iota. Then ιM(T)\iota_{M}(T_{\partial}) is a sphere twist about about a nonseparating sphere. Earlier in this section, we saw that such sphere twists are nontrivial, and so we conclude that TT_{\partial} is nontrivial as well. ∎

3 Restriction Theorem

Fix n,b0n,b\geq 0, and let PP be a partition of the boundary components of Mn,bM_{n,b}. Recall that we have defined H1P(Mn,b)H_{1}^{P}(M_{n,b}) to be the submodule of H1(Mn,b,Mn,b)H_{1}(M_{n,b},\partial M_{n,b}) generated by

{[h]H1(Mn,b,Mn,b)\displaystyle\{[h]\in H_{1}(M_{n,b},\partial M_{n,b})\mid either hh is a simple closed curve or
hh is a properly embedded arc with endpoints
in distinct P-adjacent boundary components},\displaystyle\text{ in distinct $P$-adjacent boundary components}\},

and IOn,bPIO_{n,b}^{P} is the kernel of the action of Out(Fn,b)\mathrm{Out}(F_{n,b}) on H1P(Mn,b)H_{1}^{P}(M_{n,b}) induced by the action of Mod(Mn,b)\operatorname{Mod}(M_{n,b}).

Remark.

This version of homology is simpler than the one used in [22]. There are two reasons for this.

  • In our case, we can take homology relative to the entire boundary, whereas in [22], homology is taken relative to a set consisting of a single point from each boundary component. This is because in surfaces, the boundary components give nontrivial elements of H1H_{1}, and the arcs considered in H1P(Σg,b)H_{1}^{P}(\Sigma_{g,b}) can get “wrapped around” those boundary components. This is not a problem in our setting because loops in boundary components of Mn,bM_{n,b} are trivial in H1H_{1}.

  • Next, suppose we have an embedding Σg,bΣg\Sigma_{g,b}\hookrightarrow\Sigma_{g^{\prime}} of surfaces. It is possible for a nontrivial element aH1(Σg,b)a\in H_{1}(\Sigma_{g,b}) to become trivial in H1(Σg)H_{1}(\Sigma_{g^{\prime}}) (for instance, if a boundary component is capped off). So, there could be elements of Mod(Σg,b)\operatorname{Mod}(\Sigma_{g,b}) which act trivially on H1(Σg)H_{1}(\Sigma_{g^{\prime}}), but not fix aa. In other words, the Torelli group would not be closed under restrictions. To fix this, the author in [22] must mod out by the submodules of H1(Σg,b)H_{1}(\Sigma_{g,b}) spanned by the pPp\in P (with proper orientation chosen). This is not a problem in the 3-dimensional case however, since an inclusion Mn,bMmM_{n,b}\hookrightarrow M_{m} induces an injection on homology.

We can now move on to the proof of Theorem A.

Proof of Theorem A.

Let ι:Mn,bMm\iota:M_{n,b}\hookrightarrow M_{m} be an embedding, and let ι:Out(Fn,b)Out(Fm)\iota_{*}:\mathrm{Out}(F_{n,b})\to\operatorname{Out}(F_{m}) be the induced map. Recall that we must show that ι1(IOm)=IOn,bP\iota_{*}^{-1}(IO_{m})=IO_{n,b}^{P}, where PP is the partition of the boundary components induced by ι\iota as described in the introduction.

This proof will follow the proof of [22, Theorem 3.3]. Define the following subsets of H1(Mm)H_{1}(M_{m}) (we use \cdot to denote concatenation of arcs):

Q1={[h]H1(Mm)\displaystyle Q_{1}=\{[h]\in H_{1}(M_{m})\mid h is a simple closed curve in MmMn,b}\displaystyle\text{ $h$ is a simple closed curve in $M_{m}\setminus M_{n,b}$}\}
Q2={[h]H1(Mm)\displaystyle Q_{2}=\{[h]\in H_{1}(M_{m})\mid h is a simple closed curve in Mn,b}\displaystyle\text{ $h$ is a simple closed curve in $M_{n,b}$}\}
Q3={[h1h2]H1(Mm)\displaystyle Q_{3}=\{[h_{1}\cdot h_{2}]\in H_{1}(M_{m})\mid h1h_{1} is a properly embedded arc in Mn,bM_{n,b} with
endpoints in distinct PP-adjacent boundary
components and h2h_{2} is a properly embedded arc
in MmMn,b with the same endpoints as h1}.\displaystyle\text{ in $M_{m}\setminus M_{n,b}$ with the same endpoints as $h_{1}$}\}.

We claim that the homology group H1(Mm)H_{1}(M_{m}) is spanned by Q1Q2Q3Q_{1}\cup Q_{2}\cup Q_{3}. To see why, let [α]H1(Mm)[\alpha]\in H_{1}(M_{m}) be the class of a loop α\alpha. If α\alpha can be homotoped to lie entirely inside Mn,bM_{n,b} or MmMn,bM_{m}\setminus M_{n,b}, then we are done. On the other hand, suppose that crosses the boundary of Mn,bM_{n,b}. Without loss of generality, we may assume that α\alpha crosses the boundary of Mn,bM_{n,b} exactly twice since any loop can be surgered into a collection of such loops (see Figure 4). It follows that α\alpha has the form α=γδ\alpha=\gamma\cdot\delta, where γMn,b\gamma\subset M_{n,b} is an arc connecting boundary components of Mn,bM_{n,b}, and δMmMn,b\delta\subset M_{m}\setminus M_{n,b} is a arc with the same endpoints as γ\gamma. Recall that under the partition PP induced by the inclusion ι\iota, two boundary components are PP-adjacent if they lie on the same component of MmMn,bM_{m}\setminus M_{n,b}. Therefore, the existence of δ\delta implies that the boundary components intersected by α\alpha are PP-adjacent, and thus [α]Q3[\alpha]\in Q_{3}. This completes the proof of the claim.

\labellist
\hair

2pt \pinlabelMn,bM_{n,b} at 88 70 \pinlabelMn,bM_{n,b} at 320 70 \pinlabelα\alpha at 88 240 \endlabellistRefer to caption

Figure 4: A loop can be surgered into a collection of loops which intersect Mn,b\partial M_{n,b} exactly twice.

Let fIOn,bPf\in IO_{n,b}^{P}. By the definition of IOn,bPIO_{n,b}^{P}, the element ι(f)\iota_{*}(f) acts trivially on Q2Q_{2}. Moreover, ι(f)\iota_{*}(f) acts trivially on Q1Q_{1} by the definition of ι\iota_{*}. Lastly, suppose that [h1h2]Q3[h_{1}\cdot h_{2}]\in Q_{3}. Then ι(f)\iota_{*}(f) fixes the homology class of h1h_{1} since fIOn,bPf\in IO_{n,b}^{P}, and fixes h2h_{2} pointwise by the definition of ι\iota_{*}. Therefore, fι1(IOm)f\in\iota_{*}^{-1}(IO_{m}).

Next, suppose that fι1(IOm)f\in\iota_{*}^{-1}(IO_{m}). By definition, ι(f)\iota_{*}(f) acts trivially on H1(Mm)H_{1}(M_{m}), and thus on Q2Q_{2} as well since the map H1(Mn,b)H1(Mm)H_{1}(M_{n,b})\to H_{1}(M_{m}) induced by ι\iota is injective. This implies that ff acts trivially on homology classes of simple closed curves in Mn,bM_{n,b}. So, we only need to check that ff preserves the homology classes of arcs in MmM_{m} connecting distinct PP-adjacent boundary components. Suppose there is a class of arcs [α]H1P(Mn,b)[\alpha]\in H_{1}^{P}(M_{n,b}). Since PP is the partition of the boundary components induced by ι\iota, [α][\alpha] can be completed to a homology class [αβ]H1(Mm)[\alpha\cdot\beta]\in H_{1}(M_{m}), where β\beta is an arc in MmMn,bM_{m}\setminus M_{n,b} connecting the endpoints of α\alpha. Then since ι(f)IOm\iota_{*}(f)\in IO_{m} and ι(f)\iota_{*}(f) fixes β\beta pointwise, we have

0=([αβ])ι(f)([αβ])=[α]f([α]).0=([\alpha\cdot\beta])-\iota_{*}(f)([\alpha\cdot\beta])=[\alpha]-f([\alpha]).

This shows that ff acts trivially on [α][\alpha]. Therefore, fIOn,bPf\in IO_{n,b}^{P}. ∎

4 Birman exact sequence

In this section, we give a version of the Birman exact sequence for the groups IOn,bPIO_{n,b}^{P}. We will start by giving a Birman exact sequence on the level of mapping class groups. We note that Banks has proved a version of the Birman exact sequence for 3-manifolds (see [4]). However, this version involves forgetting a puncture rather than capping a boundary component, so we will prove our own version here. Once we have the sequence for mapping class groups, we will mod out by sphere twists to get a corresponding sequence for Out(Fn,b)\mathrm{Out}(F_{n,b}), and finally restrict to get a sequence for IOn,bPIO_{n,b}^{P}.

Remark.

In the following theorems, we exclude the case (n,b)=(1,1)(n,b)=(1,1). This is because boundary drags in Mod(M1,1)\operatorname{Mod}(M_{1,1}) are trivial (see the proof of Theorem 4.1 for the definition of boundary drags). In this case, we have isomorphisms

  • Mod(M1,1)Mod(M1)/2×/2\operatorname{Mod}(M_{1,1})\cong\operatorname{Mod}(M_{1})\cong\mathbb{Z}/2\times\mathbb{Z}/2,

  • Out(F1,1)Out(F1)/2\mathrm{Out}(F_{1,1})\cong\operatorname{Out}(F_{1})\cong\mathbb{Z}/2,

  • IO1,1{}IO11IO_{1,1}^{\{\partial\}}\cong IO_{1}\cong 1,

where one of the generators of Mod(M1)=Mod(S1×S2)\operatorname{Mod}(M_{1})=\operatorname{Mod}(S^{1}\times S^{2}) is a sphere twist about the sphere ×S2*\times S^{2} and the other is the antipodal map in both coordinates.

Theorem 4.1.

Fix n,b>0n,b>0 such that (n,b)(1,1)(n,b)\neq(1,1). Glue a ball to a boundary component of Mn,bM_{n,b}, and let Mn,bMn,b1M_{n,b}\hookrightarrow M_{n,b-1} be the resulting embedding. Fix xMn,b1Mn,bx\in M_{n,b-1}\setminus M_{n,b}.

  1. (a)

    If b>1b>1, choose a lift x~Frx(Mn,b1)\tilde{x}\in\operatorname{Fr}_{x}(M_{n,b-1}) of xx, where Fr(Mn,b1)\operatorname{Fr}(M_{n,b-1}) is the oriented frame bundle of Mn,b1M_{n,b-1}. We then have an exact sequence

    1π1(Fr(Mn,b1),x~)Mod(Mn,b)Mod(Mn,b1)1.1\to\pi_{1}(\operatorname{Fr}(M_{n,b-1}),\tilde{x})\to\operatorname{Mod}(M_{n,b})\to\operatorname{Mod}(M_{n,b-1})\to 1.
  2. (b)

    If b=1b=1 and n>1n>1, then we have an exact sequence

    1π1(Mn,b1,x)Mod(Mn,b)Mod(Mn,b1)1.1\to\pi_{1}(M_{n,b-1},x)\to\operatorname{Mod}(M_{n,b})\to\operatorname{Mod}(M_{n,b-1})\to 1.
Proof.

Let Diff+(Mn,b1)\operatorname{Diff^{+}}(M_{n,b-1}) denote the space of orientation-preserving diffeomorphisms of Mn,b1M_{n,b-1} which restrict to the identity on Mn,b1\partial M_{n,b-1}, and let Diff+(Mn,b1,x~)\operatorname{Diff^{+}}(M_{n,b-1},\tilde{x}) be the subspace of Diff+(Mn,b1)\operatorname{Diff^{+}}(M_{n,b-1}) consisting of diffeomorphisms which fix the framing x~\tilde{x}. It is standard that there is a fiber bundle

Diff+(Mn,b1,x~)Diff+(Mn,b1)𝑝Fr(Mn,b1),\operatorname{Diff^{+}}(M_{n,b-1},\tilde{x})\to\operatorname{Diff^{+}}(M_{n,b-1})\overset{p}{\to}\operatorname{Fr}(M_{n,b-1}), (2)

where the map p:Diff+(Mn,b1)Fr(Mn,b1)p:\operatorname{Diff^{+}}(M_{n,b-1})\to\operatorname{Fr}(M_{n,b-1}) is given by φdφ(x~)\varphi\mapsto d\varphi(\tilde{x}). Passing to the long exact sequence of homotopy group associated to this fiber bundle, we find the segment

π1(Fr(Mn,b1))π0(Diff+(Mn,b1,x~))π0(Diff+(Mn,b1))π0(Fr(Mn,b1)).\pi_{1}(\operatorname{Fr}(M_{n,b-1}))\to\pi_{0}(\operatorname{Diff^{+}}(M_{n,b-1},\tilde{x}))\to\pi_{0}(\operatorname{Diff^{+}}(M_{n,b-1}))\to\pi_{0}(\operatorname{Fr}(M_{n,b-1})).

Since Fr(Mn,b1)\operatorname{Fr}(M_{n,b-1}) is the oriented frame bundle, it is connected, and so π0(Fr(Mn,b1))\pi_{0}(\operatorname{Fr}(M_{n,b-1})) is trivial. Moreover, π0(Diff+(Mn,b1,x~))\pi_{0}(\operatorname{Diff^{+}}(M_{n,b-1},\tilde{x})) is isomorphic to Mod(Mn,b)\operatorname{Mod}(M_{n,b}). For a proof of this fact in the surface case, see [12, p. 102]; the proof goes exactly the same way in our setting. Therefore, the above sequence becomes

π1(Fr(Mn,b1))Mod(Mn,b)Mod(Mn,b1)1.\pi_{1}(\operatorname{Fr}(M_{n,b-1}))\to\operatorname{Mod}(M_{n,b})\to\operatorname{Mod}(M_{n,b-1})\to 1. (3)

To get a short exact sequence, we must understand the kernel of the map Push~:π1(Fr(Mn,b1))Mod(Mn,b)\operatorname{\widetilde{\operatorname{Push}}}:\pi_{1}(\operatorname{Fr}(M_{n,b-1}))\to\operatorname{Mod}(M_{n,b}). We remark here that the map Push~\operatorname{\widetilde{\operatorname{Push}}} is given by pushing and rotating a small ball containing xx about a loop based at xx. This is in analogy with the “disk pushing maps” seen in the case of surfaces. Since Mn,b1M_{n,b-1} is parallelizable, we have

π1(Fr(Mn,b1))π1(Mn,b1)×π1(SO(3))=π1(Mn,b1)×/2.\pi_{1}(\operatorname{Fr}(M_{n,b-1}))\cong\pi_{1}(M_{n,b-1})\times\pi_{1}(SO(3))=\pi_{1}(M_{n,b-1})\times\mathbb{Z}/2.

Consider the map Mod(Mn,b)Aut(π1(Mn,b,y))\operatorname{Mod}(M_{n,b})\to\operatorname{Aut}(\pi_{1}(M_{n,b},y)), where the basepoint yy is on the boundary component \partial being capped off. As is shown in Figure 5, the composition

π1(Fr(Mn,b1))π1(Mn,b1)×/2Push~Mod(Mn,b)Aut(π1(Mn,b,y))\pi_{1}(\operatorname{Fr}(M_{n,b-1}))\cong\pi_{1}(M_{n,b-1})\times\mathbb{Z}/2\overset{\operatorname{\widetilde{\operatorname{Push}}}}{\longrightarrow}\operatorname{Mod}(M_{n,b})\to\operatorname{Aut}(\pi_{1}(M_{n,b},y))

is given by conjugation about the loop being pushed around. Since Aut(π1(Mn,b,y))Aut(Fn)\operatorname{Aut}(\pi_{1}(M_{n,b},y))\cong\operatorname{Aut}(F_{n}) is centerless for n>1n>1, the entire kernel of Push~\operatorname{\widetilde{\operatorname{Push}}} must be contained in 1×/2π1(Mn,b1)×/21\times\mathbb{Z}/2\subset\pi_{1}(M_{n,b-1})\times\mathbb{Z}/2. However, the image of the generator of this subgroup in Mod(Mn,b)\operatorname{Mod}(M_{n,b}) is the sphere twist TT_{\partial}. By Theorems 2.1 and 2.2, this sphere twist is nontrivial if and only if b>1b>1. If b>1b>1, this shows that Push\operatorname{Push} is injective, and (3) gives us the desired exact sequence. On the other hand, if b=1b=1, then ker(Push~)=1×/2\ker(\operatorname{\widetilde{\operatorname{Push}}})=1\times\mathbb{Z}/2. Therefore, the image of Push~\operatorname{\widetilde{\operatorname{Push}}} in Mod(Mn,b)\operatorname{Mod}(M_{n,b}) is isomorphic to

π1(Fr(Mn,b1))/Tπ1(Mn,b1)\pi_{1}(\operatorname{Fr}(M_{n,b-1}))/\langle T_{\partial}\rangle\cong\pi_{1}(M_{n,b-1})

as desired. ∎

\labellist
\hair

2pt \pinlabelα\alpha at 150 165 \pinlabelγ\gamma at 150 85 \pinlabelγ1αγ\gamma^{-1}\alpha\gamma at 480 145 \pinlabelyy at 245 142 \endlabellistRefer to caption

Figure 5: The image of α\alpha under Push~(γ,T)\operatorname{\widetilde{\operatorname{Push}}}(\gamma,T) is γ1αγ\gamma^{-1}\alpha\gamma. Here, TT can be either TT_{\partial} or trivial.

Modding out by sphere twists.

Now that we have a Birman exact sequence for Mod(Mn,b)\operatorname{Mod}(M_{n,b}), we can mod out by sphere twists to get a Birman exact sequence for Out(Fn,b)\mathrm{Out}(F_{n,b}). Consider the map iM:Mod(Mn,b)Mod(Mn,b1)i_{M}:\operatorname{Mod}(M_{n,b})\to\operatorname{Mod}(M_{n,b-1}) given by capping off a boundary component \partial. Since ιM\iota_{M} takes sphere twists to sphere twists, this map descends to a map ι:Out(Fn,b)Out(Fn,b1)\iota_{*}:\mathrm{Out}(F_{n,b})\to\mathrm{Out}(F_{n,b-1}). Since ιM\iota_{M} is surjective, ι\iota_{*} is as well. Let KK be the kernel of ι\iota_{*}, and let ψ:Mod(Mn,b)Out(Fn,b)\psi:\operatorname{Mod}(M_{n,b})\to\mathrm{Out}(F_{n,b}) be the quotient map. If b>1b>1, then the kernel of ιM\iota_{M} is π1(Fr(Mn,b1),x~)\pi_{1}(\operatorname{Fr}(M_{n,b-1}),\widetilde{x}) by Theorem 4.1. Let Push~:π1(Fr(Mn,b1),x~)Mod(Mn,b)\operatorname{\widetilde{\operatorname{Push}}}:\pi_{1}(\operatorname{Fr}(M_{n,b-1}),\widetilde{x})\to\operatorname{Mod}(M_{n,b}) be the map defined in the proof of Theorem 4.1, and fix an identification π1(Fr(Mn,b1),x~)=π1(Mn,b1,x)×/2\pi_{1}(\operatorname{Fr}(M_{n,b-1}),\widetilde{x})=\pi_{1}(M_{n,b-1},x)\times\mathbb{Z}/2. Since

ι(ψ(Push~(γ,T)))=ψ(ιM(Push~(γ,T)))=ψ(id)=id\iota_{*}(\psi(\operatorname{\widetilde{\operatorname{Push}}}(\gamma,T)))=\psi(\iota_{M}(\operatorname{\widetilde{\operatorname{Push}}}(\gamma,T)))=\psi(\operatorname{id})=\operatorname{id}

for all (γ,T)π1(Fr(Mn,b1),x~)(\gamma,T)\in\pi_{1}(\operatorname{Fr}(M_{n,b-1}),\widetilde{x}), the image of π1(Fr(Mn,b1),x~)\pi_{1}(\operatorname{Fr}(M_{n,b-1}),\widetilde{x}) under ψPush~\psi\circ\operatorname{\widetilde{\operatorname{Push}}} is contained in KK. In other words, we have the following commutative diagram:

1{1}π1(Fr(Mn,b1),x~){\pi_{1}(\operatorname{Fr}(M_{n,b-1}),\tilde{x})}Mod(Mn,b){\operatorname{Mod}(M_{n,b})}Mod(Mn,b1){\operatorname{Mod}(M_{n,b-1})}1{1}1{1}K{K}Out(Fn,b){\mathrm{Out}(F_{n,b})}Out(Fn,b1){\mathrm{Out}(F_{n,b-1})}1,{1,}Push~\scriptstyle{\operatorname{\widetilde{\operatorname{Push}}}}ψP\scriptstyle{\psi_{P}}ιM\scriptstyle{\iota_{M}}ψ\scriptstyle{\psi}ι\scriptstyle{\iota_{*}}

where ψP=ψPush~\psi_{P}=\psi\circ\operatorname{\widetilde{\operatorname{Push}}}. Next, we claim that the map ψP:π1(Fr(Mn,b1),x~)K\psi_{P}:\pi_{1}(\operatorname{Fr}(M_{n,b-1}),\widetilde{x})\to K is surjective. To see this, let fKf\in K, and choose a lift 𝔣Mod(Mn,b)\mathfrak{f}\in\operatorname{Mod}(M_{n,b}) of ff. Since ι(f)=id\iota_{*}(f)=\operatorname{id}, the image ιM(𝔣)\iota_{M}(\mathfrak{f}) is a product of sphere twists TS1TSjT_{S_{1}}\cdots T_{S_{j}}. For each TSiMod(Mn,b1)T_{S_{i}}\in\operatorname{Mod}(M_{n,b-1}), choose a preimage TSiMod(Mn,b)T_{S_{i}}^{\prime}\in\operatorname{Mod}(M_{n,b}) which is also a sphere twist. Then

ιM(TS1TSj𝔣)=id,\iota_{M}(T_{S_{1}}^{\prime}\cdots T_{S_{j}}^{\prime}\cdot\mathfrak{f})=\operatorname{id},

which implies that TS1TSj𝔣=Push~(γ,T)T_{S_{1}}^{\prime}\cdots T_{S_{j}}^{\prime}\cdot\mathfrak{f}=\operatorname{\widetilde{\operatorname{Push}}}(\gamma,T) for some (γ,T)π1(Mn,b1,x)×/2(\gamma,T)\in\pi_{1}(M_{n,b-1},x)\times\mathbb{Z}/2\mathbb{Z}. Moreover, ψ(TS1TSj𝔣)=f\psi(T_{S_{1}}^{\prime}\cdots T_{S_{j}}^{\prime}\cdot\mathfrak{f})=f, which verifies our claim that ψP:π1(Fr(Mn,b1),x~)K\psi_{P}:\pi_{1}(\operatorname{Fr}(M_{n,b-1}),\widetilde{x})\to K is surjective.

Now, we wish to identify the kernel of ψP\psi_{P}. Let (γ,T)π1(Mn,b1,x~)(\gamma,T)\in\pi_{1}(M_{n,b-1},\widetilde{x}), and fix a basepoint yy on the boundary component being capped off. At the end of the proof of Theorem 4.1, we saw that Push~(γ,T)\operatorname{\widetilde{\operatorname{Push}}}(\gamma,T) acts nontrivially on π1(Mn,b,y)\pi_{1}(M_{n,b},y) if and only if γ\gamma is trivial. Since sphere twists act trivially on homotopy classes of curves, it follows that ψP(γ,T)\psi_{P}(\gamma,T) is nontrivial if γ\gamma is nontrivial. Therefore, the kernel of ψP\psi_{P} must lie inside 1×/2π1(Mn,b1,x)×/21\times\mathbb{Z}/2\mathbb{Z}\subset\pi_{1}(M_{n,b-1},x)\times\mathbb{Z}/2\mathbb{Z}. However, the generator of 1×/21\times\mathbb{Z}/2\mathbb{Z} gets mapped to TT_{\partial} under Push~\operatorname{\widetilde{\operatorname{Push}}}, which is killed in Out(Fn,b)\mathrm{Out}(F_{n,b}). Therefore, ker(ψ)=1×/2\ker(\psi)=1\times\mathbb{Z}/2\mathbb{Z}, and so it follows that Kπ1(Mn,b1,x)K\cong\pi_{1}(M_{n,b-1},x).

On the other hand, if b=1b=1 and n>1n>1, then the kernel of the map ιM:Mod(Mn,b)Mod(Mn,b1)\iota_{M}:\operatorname{Mod}(M_{n,b})\to\operatorname{Mod}(M_{n,b-1}) is π1(Mn,b1,x)\pi_{1}(M_{n,b-1},x) by Theorem 4.1. Almost exactly the same argument used above shows that the quotient map restricts to a surjective map ψP:π1(Mn,b1,x)K\psi_{P}:\pi_{1}(M_{n,b-1},x)\to K. However, in this case, ψP\psi_{P} is injective since the sphere twist TT_{\partial} has already been killed off. Thus, we find that Kπ1(Mn,b1,x)K\cong\pi_{1}(M_{n,b-1},x) in this case as well.

From now on, we will identify the kernel of the map ι:Out(Fn,b)Out(Fn,b1)\iota_{*}:\mathrm{Out}(F_{n,b})\to\mathrm{Out}(F_{n,b-1}) with π1(Mn,b1,x)\pi_{1}(M_{n,b-1},x). The map π1(Mn,b1,x)Out(Fn,b)\pi_{1}(M_{n,b-1},x)\to\mathrm{Out}(F_{n,b}) will play a significant role throughout the remainder of the paper, and so we give a formal definition here.

Definition.

The map Push:π1(Mn,b1,x)Out(Fn,b)\operatorname{Push}:\pi_{1}(M_{n,b-1},x)\to\mathrm{Out}(F_{n,b}) is defined as Push(γ)=ψ(Push~(γ,T))\operatorname{Push}(\gamma)=\psi(\operatorname{\widetilde{\operatorname{Push}}}(\gamma,T)), where T/2T\in\mathbb{Z}/2\mathbb{Z} is arbitrary. Since sphere twists become trivial in Out(Fn,b)\mathrm{Out}(F_{n,b}), this element depends only on γ\gamma.

The upshot of this is that we have proven the Birman exact sequence for Out(Fn,b)\mathrm{Out}(F_{n,b}).

Theorem 4.2.

Fix n,b>0n,b>0 such that (n,b)(1,1)(n,b)\neq(1,1), and let Mn,bMn,b1M_{n,b}\hookrightarrow M_{n,b-1} be an embedding obtained by gluing a ball to a boundary component. Fix xMn,b1Mn,bx\in M_{n,b-1}\setminus M_{n,b}. Then the following sequence is exact:

1π1(Mn,b1,x)PushOut(Fn,b)ιOut(Fn,b1)1.1\to\pi_{1}(M_{n,b-1},x)\overset{\operatorname{Push}}{\to}\mathrm{Out}(F_{n,b})\overset{\iota_{*}}{\to}\mathrm{Out}(F_{n,b-1})\to 1.

Restrict to Torelli.

We now move on to proving Theorem B, which gives a Birman exact sequence for IOn,bPIO_{n,b}^{P}. We start by recalling its statement. Let PP be a partition of the boundary components of Mn,bM_{n,b}, and fix a boundary component \partial. Let pPp\in P be the set containing \partial, and let ι:Mn,bMn,b1\iota:M_{n,b}\hookrightarrow M_{n,b-1} be the inclusion obtained by capping off \partial. The partition PP induces a partition PP^{\prime} of the boundary components of Mn,b1M_{n,b-1} by removing \partial from pp. With this definition of PP^{\prime}, the map ι:Out(Fn,b)Out(Fn,b1)\iota_{*}:\mathrm{Out}(F_{n,b})\to\mathrm{Out}(F_{n,b-1}) restricts to a map IOn,bPIOn,b1PIO_{n,b}^{P}\to IO_{n,b-1}^{P^{\prime}}, which we will also call ι\iota_{*}. The sequence from Theorem 4.2 then restricts to

1π1(Mn,b1)IOn,bPIOn,bPιIOn,b1P.1\to\pi_{1}(M_{n,b-1})\cap IO_{n,b}^{P}\to IO_{n,b}^{P}\overset{\iota_{*}}{\to}IO_{n,b-1}^{P^{\prime}}.

Theorem B asserts that ι\iota_{*} is surjective, and identifies its kernel π1(Mn,b1)IOn,bP\pi_{1}(M_{n,b-1})\cap IO_{n,b}^{P}. We start with surjectivity.

Lemma 4.3.

The induced map ι:IOn,bPIOn,b1P\iota_{*}:IO_{n,b}^{P}\to IO_{n,b-1}^{P^{\prime}} is surjective for any embedding ι:Mn,bMn,b1\iota:M_{n,b}\hookrightarrow M_{n,b-1}.

Proof.

Consider an element gIOn,b1Pg\in IO_{n,b-1}^{P^{\prime}}. Our goal is to find some fIOn,bPf\in IO_{n,b}^{P} such that ι(f)=g\iota_{*}(f)=g. There are two cases.

First, suppose that p={}p=\{\partial\}. Then the inclusion ι\iota induces an isomorphism ιH:H1P(Mn,b)H1P(Mn,b1)\iota_{H}:H_{1}^{P}(M_{n,b})\to H_{1}^{P^{\prime}}(M_{n,b-1}) which is equivariant with respect to the actions of Out(Fn,b)\mathrm{Out}(F_{n,b}) and Out(Fn,b1)\mathrm{Out}(F_{n,b-1}). In other words, for any [h]H1P(Mn,b)[h]\in H_{1}^{P}(M_{n,b}) and fOut(Fn,b)f\in\mathrm{Out}(F_{n,b}), we have

ιH(f[h])=ι(f)ιH([h]).\iota_{H}(f\cdot[h])=\iota_{*}(f)\cdot\iota_{H}([h]). (4)

By Theorem 4.2, there exists some fOut(Fn,b)f\in\mathrm{Out}(F_{n,b}) such that ι(f)=g\iota_{*}(f)=g. We claim that fIOn,bPf\in IO_{n,b}^{P}. To see this, let [h]H1P(Mn,b)[h]\in H_{1}^{P}(M_{n,b}). Then, by Equation (4), we see that

ιH(f[h])=ι(f)ιH([h])=gιH([h])=ιH([h]).\iota_{H}(f\cdot[h])=\iota_{*}(f)\cdot\iota_{H}([h])=g\cdot\iota_{H}([h])=\iota_{H}([h]).

Since ιH\iota_{H} is an isomorphism, this implies that f[h]=[h]f\cdot[h]=[h], and so fIOn,bPf\in IO_{n,b}^{P}, as desired.

Next, suppose that p{}p\neq\{\partial\}. Again, choose fOut(Fn,b)f\in\mathrm{Out}(F_{n,b}) such that ι(f)=g\iota_{*}(f)=g. In this case, there is no longer a well-defined map H1P(Mn,b)H1P(Mn,b1)H_{1}^{P}(M_{n,b})\to H_{1}^{P^{\prime}}(M_{n,b-1}). However, there is a subgroup of H1P(Mn,b)H_{1}^{P}(M_{n,b}) which projects isomorphically onto H1P(Mn,b1)H_{1}^{P^{\prime}}(M_{n,b-1}). Let AH1P(Mn,b)A\subset H_{1}^{P}(M_{n,b}) be the subgroup generated by

{[a]H1(Mn,b,Mn,b)\displaystyle\{[a]\in H_{1}(M_{n,b},\partial M_{n,b})\mid either aa is a simple closed curve
or aa is a properly embedded arc with
neither endpoint on }.\displaystyle\text{ neither endpoint on $\partial$}\}.

It is clear that AH1P(Mn,b1)A\cong H_{1}^{P^{\prime}}(M_{n,b-1}).

Let [k]H1P(Mn,b)[k]\in H_{1}^{P}(M_{n,b}) be the class of an arc kk which has an endpoint on \partial. We claim that H1P(Mn,b)H_{1}^{P}(M_{n,b}) is generated by AA and [k][k]. To establish this claim, it suffices to show that []A,[k][\ell]\in\langle A,[k]\rangle, where []H1P(Mn,b)[\ell]\in H_{1}^{P}(M_{n,b}) is the class of any arc with an endpoint on \partial and the other elsewhere. Such an \ell exists since p{}p\neq\{\partial\}. Fix such a class [][\ell], and let α\alpha\subset\partial be an arc connecting the endpoints of \ell and kk on \partial. Orient \ell, α\alpha, and kk such that the curve αk\ell\cdot\alpha\cdot k is well-defined.

If the endpoints of \ell and kk which are not on \partial lie on distinct boundary components, then αk\ell\cdot\alpha\cdot k is an arc connecting PP^{\prime}-adjacent boundary components. Therefore, []+[α]+[k]A[\ell]+[\alpha]+[k]\in A. Since [α]=0[\alpha]=0 in H1P(Mn,b)H_{1}^{P}(M_{n,b}), it follows that []A,[k][\ell]\in\langle A,[k]\rangle. On the other hand, if the endpoints of \ell and kk which are not on \partial lie on the same boundary component \partial^{\prime}, then we can complete αk\ell\cdot\alpha\cdot k to a loop αkβ\ell\cdot\alpha\cdot k\cdot\beta, where β\beta\subset\partial^{\prime} is an arc connecting the endpoints of \ell and kk. Then

[]+[k]=[]+[α]+[k]+[β]=[αkβ]A,[\ell]+[k]=[\ell]+[\alpha]+[k]+[\beta]=[\ell\cdot\alpha\cdot k\cdot\beta]\in A,

and so []A,[k][\ell]\in\langle A,[k]\rangle. This completes the proof of the claim that H1P(Mn,b)H_{1}^{P}(M_{n,b}) is generated by AA and [k][k].

Since AA projects isomorphically onto H1P(Mn,b1)H_{1}^{P^{\prime}}(M_{n,b-1}), and this projection is equivariant with respect to the actions of Out(Fn,b)\mathrm{Out}(F_{n,b}) and Out(Fn,b1)\mathrm{Out}(F_{n,b-1}), we have f[a]=[a]f\cdot[a]=[a]. It follows that ff acts trivially on AA. Therefore, if ff fixes [k][k], then fIOn,bPf\in IO_{n,b}^{P} by the discussion in the preceding paragraph, and so we are done. On the other hand, if ff does not fix [k][k], then γ=kf(k)1\gamma=k\cdot f(k)^{-1} is a nontrivial loop based at a point on \partial. So, the element Push(γ)1fOut(Fn,b)\operatorname{Push}(\gamma)^{-1}\cdot f\in\mathrm{Out}(F_{n,b}) fixes [k][k]. Moreover, Push(γ)\operatorname{Push}(\gamma) acts trivially on AA, and so Push(γ)1f\operatorname{Push}(\gamma)^{-1}\cdot f does as well. Thus, Push(γ)1fIOn,bP\operatorname{Push}(\gamma)^{-1}\cdot f\in IO_{n,b}^{P}. Finally, since Push(γ)ker(i)\operatorname{Push}(\gamma)\in\ker(i_{*}), we have that ι(Push(γ)1f)=g\iota_{*}(\operatorname{Push}(\gamma)^{-1}\cdot f)=g, and so we are done. ∎

We now move on to the proof of Theorem B.

Proof of Theorem B.

Recall that we want to show that we have an exact sequence

1LPushIOn,bPιIOn,b1P1,1\to L\overset{\operatorname{Push}}{\to}IO_{n,b}^{P}\overset{\iota_{*}}{\to}IO_{n,b-1}^{P^{\prime}}\to 1,

where LL is equal to:

  1. (a)

    π1(Mn,b1,x)Fn\pi_{1}(M_{n,b-1},x)\cong F_{n} if p={}p=\{\partial\}.

  2. (b)

    [π1(Mn,b1,x),π1(Mn,b1,x)][Fn,Fn][\pi_{1}(M_{n,b-1},x),\pi_{1}(M_{n,b-1},x)]\cong[F_{n},F_{n}] if p{}p\neq\{\partial\}.

By Lemma 4.3 and the discussion preceding it, all that is left to show is that π1(Mn,b1)IOn,bP\pi_{1}(M_{n,b-1})\cap IO_{n,b}^{P} agrees with the subgroups LL given above.

We begin with the case p={}p=\{\partial\}. Recall that π1(Mn,b1)\pi_{1}(M_{n,b-1}) acts on Mn,bM_{n,b} by pushing the boundary component \partial about a given loop. Since \partial is not PP-adjacent to any other boundary components, it follows that π1(Mn,b1)\pi_{1}(M_{n,b-1}) acts trivially on H1P(Mn,b)H_{1}^{P}(M_{n,b}). Therefore, π1(Mn,b1)IOn,bP\pi_{1}(M_{n,b-1})\subset IO_{n,b}^{P}, and so π1(Mn,b1)IOn,bP=π1(Mn,b1)\pi_{1}(M_{n,b-1})\cap IO_{n,b}^{P}=\pi_{1}(M_{n,b-1}). This completes this case.

Next, suppose that p{}p\neq\{\partial\}. In this case, not all elements of π1(Mn,b)\pi_{1}(M_{n,b}) are contained in IOn,bPIO_{n,b}^{P}. This is because dragging \partial about loops may change the homology class of arcs connected to \partial. In particular, if γπ1(Mn,b1)\gamma\in\pi_{1}(M_{n,b-1}) and [h]H1P(Mn,b)[h]\in H_{1}^{P}(M_{n,b}) is the class of arc with an endpoint in \partial and the other elsewhere, then Push(γ)\operatorname{Push}(\gamma) acts on [h][h] via

Push(γ)[h]=[γ]+[h].\operatorname{Push}(\gamma)\cdot[h]=[\gamma]+[h].

See Figure 6 for an illustration. This implies that an element Push(γ)\operatorname{Push}(\gamma) is in IOn,bPIO_{n,b}^{P} if and only if [γ]=0[\gamma]=0 in H1(Mn,b1)H_{1}(M_{n,b-1}). Thus,

π1(Mn,b1)IOn,bP=[π1(Mn,b1),π1(Mn,b1)],\pi_{1}(M_{n,b-1})\cap IO_{n,b}^{P}=[\pi_{1}(M_{n,b-1}),\pi_{1}(M_{n,b-1})],

which is what we wanted to show. ∎

\labellist
\hair

2pt \pinlabelhh at 205 150 \pinlabelα\alpha at 145 190 \pinlabel\partial at 235 170 \endlabellistRefer to caption

Figure 6: Dragging \partial around α\alpha takes [h][h] to [α]+[h][\alpha]+[h].

5 Generators

In this section, we will define our generators of IOn,bPIO_{n,b}^{P}. The definition of these generators will involve splitting and dragging boundary components, so we will discuss these processes in more detail first, then move on to the definitions.

Splitting along spheres.

Let SMn,bS\subset M_{n,b} be an embedded 2-sphere. By splitting along SS, we mean removing an open tubular neighborhood NN of SS from Mn,bM_{n,b}. If SS is nonseparating, the resulting manifold will be diffeomorphic to Mn1,b+2M_{n-1,b+2} and if SS is separating, the result will be diffeomorphic to Mm1,c1Mm2,c2M_{m_{1},c_{1}}\sqcup M_{m_{2},c_{2}}, where m1+m2=nm_{1}+m_{2}=n and c1+c2=b+2c_{1}+c_{2}=b+2. Notice that the resulting manifold is a submanifold of Mn,bM_{n,b}, and so we get a corresponding map Mod(Mn1,b+2)Mod(Mn,b)\operatorname{Mod}(M_{n-1,b+2})\to\operatorname{Mod}(M_{n,b}) if SS is nonseparating, or Mod(Mm1,c1)×Mod(Mm2,c2)Mod(Mn,b)\operatorname{Mod}(M_{m_{1},c_{1}})\times\operatorname{Mod}(M_{m_{2},c_{2}})\to\operatorname{Mod}(M_{n,b}) if SS is separating. In either case, this map sends sphere twists to sphere twists, and thus induces a map ι:Out(Fn1,b+2)Out(Fn,b)\iota_{*}:\mathrm{Out}(F_{n-1,b+2})\to\mathrm{Out}(F_{n,b}) or ι:Out(Fm1,c1)×Out(Fm2,c2)Out(Fn,b)\iota_{*}:\mathrm{Out}(F_{m_{1},c_{1}})\times\mathrm{Out}(F_{m_{2},c_{2}})\to\mathrm{Out}(F_{n,b}), depending on whether or not SS separates Mn,bM_{n,b}.

Dragging boundary components.

Let \partial be a boundary component of Mn,bM_{n,b}, and let ι:Mn,bMn,b1\iota:M_{n,b}\hookrightarrow M_{n,b-1} be the embedding obtained by capping off \partial. By Theorem 4.2, we have an exact sequence

1π1(Mn,b1,x)PushOut(Fn,b)ιOut(Fn,b1)1,1\to\pi_{1}(M_{n,b-1},x)\overset{\operatorname{Push}}{\longrightarrow}\mathrm{Out}(F_{n,b})\overset{\iota_{*}}{\longrightarrow}\mathrm{Out}(F_{n,b-1})\to 1,

where xMn,b1Mn,bx\in M_{n,b-1}\setminus M_{n,b}. Given γπ1(Mn,b1,x)\gamma\in\pi_{1}(M_{n,b-1},x), recall that the element Push(γ)Out(Mn,b)\operatorname{Push}(\gamma)\in\operatorname{Out}(M_{n,b}) is given by pushing \partial about the loop γ\gamma. In the remainder of this section, we will be dragging multiple boundary components at a time. So, from now on we will write Push(γ)\operatorname{Push}_{\partial}(\gamma) in order to keep track of which boundary component is being pushed.

Magnus generators.

We now move on to defining our generators for IOn,bPIO_{n,b}^{P}. In the b=0b=0 case, we have that IOn,0P=IOnIO_{n,0}^{P}=IO_{n}, where IOnIO_{n} is the subgroup of Out(Fn)\operatorname{Out}(F_{n}) acting trivially on homology. In [21], Magnus found the following generating set for IOnIO_{n}.

Theorem 5.1 (Magnus).

Let Fn=x1,,xnF_{n}=\langle x_{1},\ldots,x_{n}\rangle. The group IOnIO_{n} is generated by the Out(Fn)\operatorname{Out}(F_{n})-classes of the automorphisms

Mij:xixjxixj1,Mijk:xixi[xj,xk],M_{ij}:x_{i}\mapsto x_{j}x_{i}x_{j}^{-1},\qquad M_{ijk}:x_{i}\mapsto x_{i}[x_{j},x_{k}],

for all distinct i,j,k{1,,n}i,j,k\in\{1,\ldots,n\} with j<kj<k. Here, the automorphisms are understood to fix xx_{\ell} for i\ell\neq i.

\labellist
\hair

2pt \pinlabelxx at 332 280 \pinlabelα1\alpha_{1} at 262 385 \pinlabelα2\alpha_{2} at 175 271 \pinlabelα3\alpha_{3} at 262 175 \pinlabelσ1+\sigma_{1}^{+} at 140 420 \pinlabelσ1\sigma_{1}^{-} at 311 481 \pinlabelσ2+\sigma_{2}^{+} at 66 345 \pinlabelσ2\sigma_{2}^{-} at 66 184 \pinlabelσ3\sigma_{3}^{-} at 122 87 \pinlabelσ3+\sigma_{3}^{+} at 296 49 \pinlabel11\partial_{1}^{1} at 446 420 \pinlabel21\partial_{2}^{1} at 490 324 \pinlabel12\partial_{1}^{2} at 490 213 \pinlabel13\partial_{1}^{3} at 437 131 \endlabellistRefer to caption

Figure 7: M3,4M_{3,4} split along S1S2S3S_{1}\cup S_{2}\cup S_{3} with the partition P={{11,21},{12},{13}}P=\{\{\partial_{1}^{1},\partial_{2}^{1}\},\{\partial_{1}^{2}\},\{\partial_{1}^{3}\}\}.

Our generating set will be inspired by Magnus’s, and will indeed reduce to it when b=0b=0. In order to choose a concrete collection of elements, we will need to make some choices. First, fix a basepoint int(Mn,b)*\in\operatorname{int}(M_{n,b}) and a set {α1,,αn}\{\alpha_{1},\ldots,\alpha_{n}\} of oriented simple closed curves intersecting only at * whose homotopy classes form a free basis for π1(Mn,b,)\pi_{1}(M_{n,b},*). We will call such a set {α1,,αn}\{\alpha_{1},\ldots,\alpha_{n}\} a geometric free basis for π1(Mn,b,)\pi_{1}(M_{n,b},*). In addition, choose a corresponding sphere basis; that is, a collection of nn disjointly embedded oriented 2-spheres S1,,SnMn,bS_{1},\ldots,S_{n}\subset M_{n,b} such that each SiS_{i} intersects αi\alpha_{i} exactly once with a positive orientation and is disjoint from the other αj\alpha_{j}. Notice that splitting Mn,bM_{n,b} along the SiS_{i} reduces it to a 3-sphere 𝒵Mn,b\mathcal{Z}\subset M_{n,b} with b+2nb+2n boundary components. The submanifold 𝒵\mathcal{Z} will play a significant role throughout the remainder of this section because it will allow all of our choices made in the definitions to be unique. For each SiS_{i}, let σi+\sigma_{i}^{+} and σi\sigma_{i}^{-} be the boundary components of 𝒵\mathcal{Z} arising from the split along SiS_{i}, where σi+\sigma_{i}^{+} (resp. σi\sigma_{i}^{-}) is the component lying on the positive (resp. negative) side of SiS_{i}. We will also choose an ordering P={p1,,p|P|}P=\{p_{1},\ldots,p_{|P|}\} and an ordering pr={1r,,brr}p_{r}=\{\partial_{1}^{r},\ldots,\partial_{b_{r}}^{r}\} for each r{1,,|P|}r\in\{1,\ldots,|P|\}. See Figure 7.

The following lemma will be helpful in showing that our generators lie in IOn,bPIO_{n,b}^{P}.

\labellist
\hair

2pt \pinlabelhh at 105 95 \pinlabelα\alpha at 435 95 \pinlabelγs\gamma_{s} at 520 105 \pinlabelγe\gamma_{e} at 500 180 \pinlabel* at 462 141 \endlabellistRefer to caption

Figure 8: The arc hh homotoped to be put in the form γsαγe\gamma_{s}\cdot\alpha\cdot\gamma_{e}.
Lemma 5.2.

Let 𝒵\mathcal{Z} be as above, and suppose that hMn,bh\subset M_{n,b} is a properly embedded oriented arc connecting PP-adjacent boundary components of Mn,bM_{n,b}. Then the homology class of [h]H1P(Mn,b)[h]\in H_{1}^{P}(M_{n,b}) has the form

[h]=[α]+[h0],[h]=[\alpha]+[h_{0}],

where α\alpha is a loop based at *, and h0h_{0} is the unique arc (up to isotopy) in 𝒵\mathcal{Z} which has the same endpoints as hh.

Proof.

We may homotope hh such that it has the form h=γsαγeh=\gamma_{s}\cdot\alpha\cdot\gamma_{e}, where (see Figure 8):

  • γs𝒵\gamma_{s}\subset\mathcal{Z} is the unique arc (up to isotopy) from the initial point of hh to the basepoint * of Mn,bM_{n,b},

  • γe𝒵\gamma_{e}\subset\mathcal{Z} is the unique arc from * to the endpoint of hh,

  • απ1(Mn,b,)\alpha\in\pi_{1}(M_{n,b},*).

Then,

[h]=[γsαγe]=[α]+[γsγe]=[α]+[h0],[h]=[\gamma_{s}\cdot\alpha\cdot\gamma_{e}]=[\alpha]+[\gamma_{s}\cdot\gamma_{e}]=[\alpha]+[h_{0}],

as desired. ∎

Handle drags.

Let i{1,,n}i\in\{1,\ldots,n\}, and let hih_{i} be the unique (up to isotopy) properly embedded arc in 𝒵\mathcal{Z} connecting σi+\sigma_{i}^{+} and σi\sigma_{i}^{-} which is disjoint from the αk\alpha_{k}. Choose a tubular neighborhood NiN_{i} of σi+hiσi\sigma_{i}^{+}\cup h_{i}\cup\sigma_{i}^{-} that does not intersect any αk\alpha_{k} for kik\neq i. Let Σi\Sigma_{i} be the boundary component of NiN_{i} which is not isotopic to σi+\sigma_{i}^{+} or σi\sigma_{i}^{-} (notice that Σi\Sigma_{i} is diffeomorphic to a 2-sphere). Splitting Mn,bM_{n,b} along Σi\Sigma_{i} yields Mn1,b+1M1,1M_{n-1,b+1}\sqcup M_{1,1}. Let ΣiMn1,b+1\Sigma_{i}^{\prime}\subset\partial M_{n-1,b+1} be the boundary component coming from this split, and fix a basepoint yiΣiy_{i}\in\Sigma_{i}^{\prime}. Fix an oriented arc δiZ\delta_{i}\subset Z from yiy_{i} to * which only intersects Σi\Sigma_{i}^{\prime} at yiy_{i}. Since 𝒵\mathcal{Z} is a 3-sphere with spherical boundary components, δi\delta_{i} is unique up to isotopy. The arc δi\delta_{i} induces an isomorphism π1(Mn1,b+1,)π1(Mn1,b+1,yi)\pi_{1}(M_{n-1,b+1},*)\to\pi_{1}(M_{n-1,b+1},y_{i}) given by γδiγδi1\gamma\mapsto\delta_{i}\gamma\delta_{i}^{-1}. Define βji=δiαjδi1\beta_{j}^{i}=\delta_{i}\alpha_{j}\delta_{i}^{-1}. Then we define the handle drag HDij:=ι(PushΣi(βji),id)Out(Fn,b)\operatorname{HD}_{ij}:=\iota_{*}(\operatorname{Push}_{\Sigma_{i}^{\prime}}(\beta_{j}^{i}),\operatorname{id})\in\mathrm{Out}(F_{n,b}) for jij\neq i, where ι\iota_{*} is the map Out(Fn1,b+1)×Out(F1,1)Out(Fn,b)\mathrm{Out}(F_{n-1,b+1})\times\mathrm{Out}(F_{1,1})\to\mathrm{Out}(F_{n,b}) induced by splitting along Σi\Sigma_{i}.

\labellist
\hair

2pt \pinlabelΣ1\Sigma_{1}^{\prime} at 200 275 \pinlabelδ1\delta_{1} at 170 215 \pinlabely1y_{1} at 135 230 \pinlabelα2\alpha_{2} at 100 150 \pinlabelα2α1α21\alpha_{2}\alpha_{1}\alpha_{2}^{-1} at 560 210 \endlabellistRefer to caption

Figure 9: Setup of the handle drag HD12\operatorname{HD}_{12} and the image of α1\alpha_{1} under HD12\operatorname{HD}_{12}

To see that HDijIOn,bP\operatorname{HD}_{ij}\in IO_{n,b}^{P}, notice that HDij\operatorname{HD}_{ij} acts trivially on αk\alpha_{k} for kik\neq i, and acts on αi\alpha_{i} via αiαjαiαj1\alpha_{i}\mapsto\alpha_{j}\alpha_{i}\alpha_{j}^{-1}. See Figure 9. This shows that HDij\operatorname{HD}_{ij} acts trivially on homology classes of simple closed curves. Additionally, this shows that HDij\operatorname{HD}_{ij} reduces to MijM_{ij} of the Magnus generators if b=0b=0.

Next, suppose that hh is an arc connecting PP-adjacent boundary components. By Lemma 5.2, we may write [h]=[α]+[h0][h]=[\alpha]+[h_{0}], where α\alpha is a loop based at *, and h0h_{0} is the unique arc (up to isotopy) in 𝒵\mathcal{Z} which has the same endpoints as hh. We have seen that HDij\operatorname{HD}_{ij} fixes the homology class of α\alpha. Moreover, we may homotope HDij\operatorname{HD}_{ij} such that it fixes the arc h0h_{0}. Thus, HDij\operatorname{HD}_{ij} fixes the homology class of hh, and we conclude that HDijIOn,bP\operatorname{HD}_{ij}\in IO_{n,b}^{P}.

Commutator drags.

Let i,j,k{1,,n}i,j,k\in\{1,\ldots,n\} be distinct with j<kj<k. Split Mn,bM_{n,b} along SiS_{i} to get Mn,b+2M_{n,b+2}, where 𝒵Mn,b+2Mn,b\mathcal{Z}\subset M_{n,b+2}\subset M_{n,b}. Fix basepoint yiσi+y_{i}\in\sigma_{i}^{+} and ziσiz_{i}\in\sigma_{i}^{-}, and choose oriented arcs δi,εi𝒵\delta_{i},\varepsilon_{i}\subset\mathcal{Z} connecting yiy_{i} and ziz_{i} to *, respectively. Just as in the construction of handle drags, δi\delta_{i} and εi\varepsilon_{i} are unique up to isotopy. Let βi=δiαδi1\beta_{\ell}^{i}=\delta_{i}\alpha_{\ell}\delta_{i}^{-1} and γi=εiαεi1\gamma_{\ell}^{i}=\varepsilon_{i}\alpha_{\ell}\varepsilon_{i}^{-1} for {j,m}\ell\in\{j,m\}. Then, we define the commutator drags CDijk+,CDijkOut(Fn,b)\operatorname{CD}_{ijk}^{+},\operatorname{CD}_{ijk}^{-}\in\mathrm{Out}(F_{n,b}) as ι(Pushσi+([βji,βki]))\iota_{*}(\operatorname{Push}_{\sigma_{i}^{+}}([\beta_{j}^{i},\beta_{k}^{i}])) and ι(Pushσi([γji,γki]))\iota_{*}(\operatorname{Push}_{\sigma_{i}^{-}}([\gamma_{j}^{i},\gamma_{k}^{i}])), respectively, where ι:Out(Fn,b+2)Out(Fn,b)\iota_{*}:\mathrm{Out}(F_{n,b+2})\to\mathrm{Out}(F_{n,b}) is the map induced by splitting along SiS_{i}. See Figure 10.

Again, we see that CDijk±\operatorname{CD}_{ijk}^{\pm} acts trivially on α\alpha_{\ell} for i\ell\neq i, the commutator drag CDijk+\operatorname{CD}_{ijk}^{+} sends αi\alpha_{i} to αi[αj,αk]1\alpha_{i}[\alpha_{j},\alpha_{k}]^{-1}, and CDijk\operatorname{CD}_{ijk}^{-} sends αi\alpha_{i} to [αj,αk]αi[\alpha_{j},\alpha_{k}]\alpha_{i}. This shows that CDijk+\operatorname{CD}_{ijk}^{+} reduces to Mijk1M_{ijk}^{-1} of the Magnus generators when b=0b=0.

Now, suppose that hh is an arc connecting PP-adjacent boundary components of Mn,bM_{n,b}. By Lemma 5.2, we may express [h][h] in the form [h]=[α]+[h0][h]=[\alpha]+[h_{0}]. We just saw that CDijk±\operatorname{CD}_{ijk}^{\pm} fixes [α][\alpha]. We may also homotope CDijk±\operatorname{CD}_{ijk}^{\pm} such that it fixes h0h_{0}. Thus, CDijk±\operatorname{CD}_{ijk}^{\pm} fixes [h][h], and so CDijk±IOn,bP\operatorname{CD}_{ijk}^{\pm}\in IO_{n,b}^{P}.

\labellist
\hair

2pt \pinlabelz1z_{1} at 109 225 \pinlabelε1\varepsilon_{1} at 155 205 \pinlabelα2\alpha_{2} at 100 150 \pinlabelα3\alpha_{3} at 150 100 \pinlabelσ1\sigma_{1}^{-} at 60 240 \endlabellistRefer to caption

Figure 10: Setup of the commutator drag CD123\operatorname{CD}_{123}^{-}.

Boundary commutator drags.

Let prPp_{r}\in P and srpr\partial_{s}^{r}\in p_{r}. Fix i,j{1,,n}i,j\in\{1,\ldots,n\} such that i<ji<j. Choose a basepoint ysrsry_{s}^{r}\in\partial_{s}^{r}. Let γsr𝒵\gamma_{s}^{r}\subset\mathcal{Z} be the unique arc (up to isotopy) from ysry_{s}^{r} to *. Let βkrs=γsrαk(γsr)1\beta_{k}^{rs}=\gamma_{s}^{r}\alpha_{k}(\gamma_{s}^{r})^{-1} for k{i,j}k\in\{i,j\}. Then, we define the boundary commutator drags BCDijrs=Pushsr([βirs,βjrs])Out(Fn,b)\operatorname{BCD}_{ij}^{rs}=\operatorname{Push}_{\partial_{s}^{r}}([\beta_{i}^{rs},\beta_{j}^{rs}])\in\mathrm{Out}(F_{n,b}).

It is clear from the definition that BCDijrs\operatorname{BCD}_{ij}^{rs} acts trivially on α1,,αn\alpha_{1},\ldots,\alpha_{n} and arcs that do not have an endpoint on sr\partial_{s}^{r}. Suppose that hh is an oriented arc with an endpoint on sr\partial_{s}^{r}. Without loss of generality, suppose the terminal endpoint of hh lies on sr\partial_{s}^{r}. Applying lemma 5.2, we may write [h]=[α]+[h0][h]=[\alpha]+[h_{0}], where α\alpha is a loop based at * and h0𝒵h_{0}\subset\mathcal{Z} is the unique arc (up to isotopy) which shares endpoints with hh. We just saw that BCDijrs\operatorname{BCD}_{ij}^{rs} fixes the αk\alpha_{k}, and thus fixes the homology class [α][\alpha]. Therefore,

BCDjm([h])\displaystyle\operatorname{BCD}_{j\ell m}([h]) =BCDijrs([α]+[h0])\displaystyle=\operatorname{BCD}_{ij}^{rs}([\alpha]+[h_{0}])
=[α]+BCDijrs([h0])\displaystyle=[\alpha]+\operatorname{BCD}_{ij}^{rs}([h_{0}])
=[α]+[h0]+[αiαjαi1αj1]\displaystyle=[\alpha]+[h_{0}]+[\alpha_{i}\cdot\alpha_{j}\cdot\alpha_{i}^{-1}\cdot\alpha_{j}^{-1}]
=[α]+[h0]\displaystyle=[\alpha]+[h_{0}]
=[h].\displaystyle=[h].

So, it follows that BCDijrsIOn,bP\operatorname{BCD}_{ij}^{rs}\in IO_{n,b}^{P} as well.

PP-drags.

The final type of elements we will define are called PP-drags, where PP is a partition of the boundary components of Mn,bM_{n,b}. Let prPp_{r}\in P and i{1,,n}i\in\{1,\ldots,n\}. Let Σr𝒵\Sigma_{r}\subset\mathcal{Z} be the unique 2-sphere (up to isotopy) which separates the boundary components of prp_{r} from the remaining boundary components and the σj±\sigma_{j}^{\pm}. Splitting Mn,bM_{n,b} along Σr\Sigma_{r} gives Mn,bc+1M0,c+1M_{n,b-c+1}\sqcup M_{0,c+1}, where cc is the number of boundary components in pp. Let ΣrMn,bc+1\Sigma_{r}^{\prime}\subset\partial M_{n,b-c+1} be the boundary component coming from this splitting. Just as in the construction of the other drags, fix a basepoint yrΣry_{r}\in\Sigma_{r}^{\prime} and an oriented arc γr\gamma_{r} from yry_{r} to * to get a basis {β1r,,βnr}\{\beta_{1}^{r},\ldots,\beta_{n}^{r}\} of π1(Mn,bc+1,yr)\pi_{1}(M_{n,b-c+1},y_{r}). See Figure 11. Then, we define the PP-drag PDir:=ι(PushΣr(βi),id)Out(Fn,b)\operatorname{PD}_{i}^{r}:=\iota_{*}(\operatorname{Push}_{\Sigma_{r}^{\prime}}(\beta_{i}),\operatorname{id})\in\mathrm{Out}(F_{n,b}), where ι:Out(Fn,bc+1)×Out(F0,c+1)Out(Fn,b)\iota_{*}:\mathrm{Out}(F_{n,b-c+1})\times\mathrm{Out}(F_{0,c+1})\to\mathrm{Out}(F_{n,b}) is the map induced by splitting along Σr\Sigma_{r}.

To see why PDirIOn,bP\operatorname{PD}_{i}^{r}\in IO_{n,b}^{P}, first notice that we can isotope PDir\operatorname{PD}_{i}^{r} to fix all the αj\alpha_{j}. Next, if hh is an arc connecting PP-adjacent boundary components, we write [h]=[α]+[h0][h]=[\alpha]+[h_{0}] as in Lemma 5.2. As we just noted, PDir\operatorname{PD}_{i}^{r} fixes [α][\alpha], so it suffices to show that PDir\operatorname{PD}_{i}^{r} fixes the homology class of h0h_{0}. If the endpoints of hh lie on boundary components in prp_{r}, then we may homotope h0h_{0} such that it never crosses Σr\Sigma_{r}. Then, PDir\operatorname{PD}_{i}^{r} fixes h0h_{0}. On the other hand, if the endpoints of hh lie on boundary components which are not in prp_{r}, then again we can homotope h0h_{0} such that it does not cross Σr\Sigma_{r}, and then homotope PDir\operatorname{PD}_{i}^{r} such that it fixes h0h_{0}. In either case, PDir\operatorname{PD}_{i}^{r} fixes the homology class of of h0h_{0}, and so we conclude that PDirIOn,bP\operatorname{PD}_{i}^{r}\in IO_{n,b}^{P}.

\labellist
\hair

2pt \pinlabelγp\gamma_{p} at 210 192 \pinlabelα2\alpha_{2} at 100 150 \pinlabel21\partial_{2}^{1} at 282 197 \pinlabel11\partial_{1}^{1} at 298 222 \pinlabelΣp\Sigma_{p}^{\prime} at 235 240 \endlabellistRefer to caption

Figure 11: Setup of the PP-drag PD2p\operatorname{PD}_{2}^{p}, where p={11,21}Pp=\{\partial_{1}^{1},\partial_{2}^{1}\}\in P.

Images under capping.

Suppose we have an embedding ι:Mn,bMn,b1\iota:M_{n,b}\hookrightarrow M_{n,b-1} given by capping off the boundary component \partial. Let ι:IOn,bPIOn,b1P\iota_{*}:IO_{n,b}^{P}\to IO_{n,b-1}^{P^{\prime}} be the induced map, where PP^{\prime} is the partition of the boundary components of Mn,b1M_{n,b-1} induced by PP. Using the geometric free basis {α1,,αn}\{\alpha_{1},\ldots,\alpha_{n}\} and corresponding sphere basis {S1,,Sn}\{S_{1},\ldots,S_{n}\} for Mn,bM_{n,b}, we get a corresponding geometric free basis {ι(α1),,ι(αn)}\{\iota(\alpha_{1}),\ldots,\iota(\alpha_{n})\} and sphere basis {ι(S1),,ι(Sn)}\{\iota(S_{1}),\ldots,\iota(S_{n})\} for Mn,b1M_{n,b-1}. Moreover, the ordering on PP (and each prPp_{r}\in P) induces an ordering on PP^{\prime}. We can repeat the process described throughout this section to define handle drags, commutator drags, boundary commutator drag, and PP^{\prime}-drags in IOn,b1PIO_{n,b-1}^{P^{\prime}}, which we will denote by HD¯ij\overline{\operatorname{HD}}_{ij}, CD¯ijk±\overline{\operatorname{CD}}_{ijk}^{\pm}, BCD¯ijrs\overline{\operatorname{BCD}}_{ij}^{rs}, and PD¯ir\overline{\operatorname{PD}}_{i}^{r^{\prime}}, respectively. With this setup, we find that:

  • ι(HDij)=HD¯ij\iota_{*}(\operatorname{HD}_{ij})=\overline{\operatorname{HD}}_{ij}

  • ι(CDijk±)=CD¯ijk±\iota_{*}(\operatorname{CD}_{ijk}^{\pm})=\overline{\operatorname{CD}}_{ijk}^{\pm}

  • ι(BCDijrs)=id\iota_{*}(\operatorname{BCD}_{ij}^{rs})=\operatorname{id} if =sr\partial=\partial_{s}^{r} and ι(BCDijrs)=BCD¯ijrs\iota_{*}(\operatorname{BCD}_{ij}^{rs})=\overline{\operatorname{BCD}}_{ij}^{rs} if sr\partial\neq\partial_{s}^{r}

  • ι(PDir)=id\iota_{*}(\operatorname{PD}_{i}^{r})=\operatorname{id} if pr={1r}p_{r}=\{\partial_{1}^{r}\} and ι(PDir)=PD¯ir\iota_{*}(\operatorname{PD}_{i}^{r})=\overline{\operatorname{PD}}_{i}^{r} if pr{1r}p_{r}\neq\{\partial_{1}^{r}\}.

6 Finite Generation

Now that we have defined our collection of candidate generators for IOn,bPIO_{n,b}^{P}, we now move on to proving that they do in fact generate. The first step in this proof will be an induction on bb to reduce to the case of b=0b=0. This induction will rely on the following theorem of Tomaszewski [25] (see [24] for a geometric proof).

Theorem 6.1 (Tomaszewski).

Let FnF_{n} be the free group on nn letters {x1,,xn}\{x_{1},\ldots,x_{n}\}. The commutator subgroup [Fn,Fn][F_{n},F_{n}] of FnF_{n} is freely generated by the set

{[xi,xj]xidixndn,1i<jn,d,in},\left\{[x_{i},x_{j}]^{x_{i}^{d_{i}}\cdots x_{n}^{d_{n}}},1\leq i<j\leq n,d_{\ell}\in\mathbb{Z},i\leq\ell\leq n\right\},

where the superscript denotes conjugation.

We will also need the following lemma from group theory.

Lemma 6.2.

Consider an exact sequence of groups

1KGQ1.1\to K\to G\to Q\to 1.

Let SQS_{Q} be a generating set for QQ. Moreover, assume that there are sets SKKS_{K}\subset K and SGGS_{G}\subset G such that KK is contained in the subgroup of GG generated by SKS_{K} and SGS_{G}. Then GG is generated by the set SKSGS~QS_{K}\cup S_{G}\cup\widetilde{S}_{Q}, where S~Q\widetilde{S}_{Q} is a set consisting of one lift q~G\widetilde{q}\in G for every element qSqq\in S_{q}.

Proof of lemma.

Let GGG^{\prime}\subset G be the subgroup generated by SKSGS~QS_{K}\cup S_{G}\cup\widetilde{S}_{Q}, and let K=GKK^{\prime}=G^{\prime}\cap K. Then the following diagram commutes and has exact rows:

1{1}K{K^{\prime}}G{G^{\prime}}Q{Q}1{1}1{1}K{K}G{G}Q{Q}1.{1.}φ\scriptstyle{\varphi}=\scriptstyle{=}

The vertical maps are all inclusions, and hence injective. Also, by assumption, the map φ\varphi is surjective. Therefore, by the five lemma, all of the vertical maps are isomorphisms, and so we are done. ∎

We now prove Theorem C by proving the following stronger result.

Theorem 6.3.

The group IOn,bPIO_{n,b}^{P} is generated by handle, commutator, boundary commutator, and PP-drags for b0b\geq 0, n>0n>0.

Proof.

As mentioned above, we will prove this by induction on bb. The base case b=0b=0 follows directly from Magnus’s Theorem 1.1.

If b>0b>0, fix a boundary component \partial of Mn,bM_{n,b} and let pPp\in P be the partition containing \partial. Let ι:Mn,bMn,b1\iota:M_{n,b}\hookrightarrow M_{n,b-1} be an embedding obtained by capping off \partial, and choose a basepoint Mn,b1Mn,b*\in M_{n,b-1}\setminus M_{n,b}. By Theorem B, there is an exact sequence

1LPushIOn,bPιIOn,b1P1,1\to L\overset{\operatorname{Push}}{\longrightarrow}IO_{n,b}^{P}\overset{\iota_{*}}{\longrightarrow}IO_{n,b-1}^{P^{\prime}}\to 1,

where L=π1(Mn,b,)L=\pi_{1}(M_{n,b},*) if p={}p=\{\partial\} and L=[π1(Mn,b,),π1(Mn,b,)]L=[\pi_{1}(M_{n,b},*),\pi_{1}(M_{n,b},*)] otherwise. As we saw in the discussion at the end of Section 5, we can define the drags of IOn,bPIO_{n,b}^{P} and IOn,b1PIO_{n,b-1}^{P^{\prime}} in a consistent way; that is, we can define our drags in such a way that ι\iota_{*} takes handle drags to handle drags, commutator drags to commutator drags, and so on. By induction, IOn,b1PIO_{n,b-1}^{P^{\prime}} is generated by the desired drags. Therefore, it suffices to show that Push(L)\operatorname{Push}(L) is generated by our drags as well. If p={}p=\{\partial\}, then Push(L)\operatorname{Push}(L) is precisely the subgroup of IOn,bPIO_{n,b}^{P} generated by the PP-drags, and so we are done in this case.

The case of p{}p\neq\{\partial\} is less straightforward since the commutator subgroup of a free group is not finitely generated when n2n\geq 2. However, this is not necessary for IOn,bPIO_{n,b}^{P} to be finitely generated by our collection of drags. We will appeal to Lemma 6.2. Suppose that p{}p\neq\{\partial\}. Then, by Theorem 6.1, the kernel L=[π1(Mn,b,),π1(Mn,b,)]L=[\pi_{1}(M_{n,b},*),\pi_{1}(M_{n,b},*)] of the Birman exact sequence is generated by elements of the form [xi,xj]xidixndn[x_{i},x_{j}]^{x_{i}^{d_{i}}\cdots x_{n}^{d_{n}}}. First, notice that Push([xi,xj])\operatorname{Push}([x_{i},x_{j}]) is the boundary commutator drag BCDijrs\operatorname{BCD}_{ij}^{rs}, where sr=\partial_{s}^{r}=\partial is the boundary component of Mn,bM_{n,b} being capped off. Moreover, we have seen that the handle drag HDk\operatorname{HD}_{k\ell} acts on xkx_{k} by xkxxkx1x_{k}\mapsto x_{\ell}x_{k}x_{\ell}^{-1}. It follows that HDikHDjk([xi,xj])=[xi,xj]xk\operatorname{HD}_{ik}\cdot\operatorname{HD}_{jk}([x_{i},x_{j}])=[x_{i},x_{j}]^{x_{k}}. Continuing this pattern, we see that

[xi,xj]xidixndn=(HDinHDjn)dn(HDiiHDji)di([xi,xj]),[x_{i},x_{j}]^{x_{i}^{d_{i}}\cdots x_{n}^{d_{n}}}=(\operatorname{HD}_{in}\cdot\operatorname{HD}_{jn})^{d_{n}}\cdots(\operatorname{HD}_{ii}\cdot\operatorname{HD}_{ji})^{d_{i}}([x_{i},x_{j}]),

where HDii\operatorname{HD}_{ii} is taken to be trivial. Therefore,

Push([xi,xk]xidixndn)=(HDinHDkn)dn(HDiiHDki)diBCDijrs.\operatorname{Push}([x_{i},x_{k}]^{x_{i}^{d_{i}}\cdots x_{n}^{d_{n}}})=(\operatorname{HD}_{in}\cdot\operatorname{HD}_{kn})^{d_{n}}\cdots(\operatorname{HD}_{ii}\cdot\operatorname{HD}_{ki})^{d_{i}}\cdot\operatorname{BCD}_{ij}^{rs}.

This shows that Push(L)\operatorname{Push}(L) is contained in the subgroup of IOn,bPIO_{n,b}^{P} generated by boundary commutator and handle drags. Applying Lemma 6.2 (taking SG={handle drags}S_{G}=\{\text{handle drags}\} and SK={boundary commutator drags}S_{K}=\{\text{boundary commutator drags}\}), we conclude that IOn,bPIO_{n,b}^{P} is generated by the desired drags. ∎

7 Partial Proof of Magnus’s Theorem

In this section, we will give a partial proof of Magnus’s Theorem 1.1, which constituted the base case in the proof of Theorem 6.3. As stated in the introduction, the original proof of Magnus’s Theorem involved two steps: showing that the elements MijM_{ij} and MijkM_{ijk} normally generate IOnIO_{n}, and then showing that the subgroup generated by these elements is normal. We will give a proof of the first step here (Theorem D).

In order to establish this fact, we will examine the action of IOn,0{}=IOnIO_{n,0}^{\{\}}=IO_{n} on a certain simplicial complex, and apply the following theorem of Armstrong [2]. We say that a group GG acts on a simplicial complex XX without rotations if every simplex ss is fixed pointwise by every element of its stabilizer, which we will denote by GsG_{s}.

Theorem 7.1 (Armstrong).

Suppose the group GG acts on a simply-connected simplicial complex XX without rotations. If X/GX/G is simply-connected, then GG is generated by the set

vX(0)Gv.\bigcup_{v\in X^{(0)}}G_{v}.

Here X(0)X^{(0)} is the 0-skeleton of XX.

Remark.

In [2], Armstrong proves the converse of this theorem as well. For a modern discussion of the proof of Theorem 7.1, along with some generalizations, we refer the reader to [23, Section 3].

The nonseparating sphere complex.

The complex to which we will apply Theorem 7.1 will be the nonseparating sphere complex 𝕊nns\operatorname{\mathbb{S}_{n}^{\text{ns}}}. Vertices of 𝕊nns\operatorname{\mathbb{S}_{n}^{\text{ns}}} are isotopy classes of smoothly embedded non-nullhomotopic 2-spheres in MnM_{n}, and 𝕊nns\operatorname{\mathbb{S}_{n}^{\text{ns}}} has a kk-simplex {S0,,Sk}\{S_{0},\ldots,S_{k}\} if the spheres S0,,SkS_{0},\ldots,S_{k} can be realized pairwise disjointly and their union does not separate MnM_{n}. This is a subcomplex of the more ubiquitous sphere complex, which was introduced by Hatcher in [13] as a tool to explore the homological stability of Out(Fn)\operatorname{Out}(F_{n}) and Aut(Fn)\operatorname{Aut}(F_{n}). In [13, Proposition 3.1], Hatcher proves the following connectivity result about 𝕊nns\operatorname{\mathbb{S}_{n}^{\text{ns}}}.

Proposition 7.2 (Hatcher).

The complex 𝕊nns\operatorname{\mathbb{S}_{n}^{\text{ns}}} is (n2)(n-2)-connected.

In particular, 𝕊nns\operatorname{\mathbb{S}_{n}^{\text{ns}}} is simply connected for n3n\geq 3. Recall that sphere twists act trivially on isotopy classes of embedded surfaces, and so we get an action of IOnIO_{n} on 𝕊nns\operatorname{\mathbb{S}_{n}^{\text{ns}}}. Notice that spheres in a simplex of 𝕊nns\operatorname{\mathbb{S}_{n}^{\text{ns}}} necessarily represent distinct H2H_{2}-classes. By Poincaré duality, elements of IOnIO_{n} act trivially on H2(Mn)H_{2}(M_{n}), and so this implies that IOnIO_{n} acts on 𝕊nns\operatorname{\mathbb{S}_{n}^{\text{ns}}} without rotations. Thus, in order to apply Theorem 7.1, we must show that 𝕊nns/IOn\operatorname{\mathbb{S}_{n}^{\text{ns}}}/IO_{n} is simply-connected.

To do this, we will give a description of 𝕊nns/IOn\operatorname{\mathbb{S}_{n}^{\text{ns}}}/IO_{n} in terms of linear algebra. Fix an identification H2(Mn)=nH_{2}(M_{n})=\mathbb{Z}^{n}. Let FS(n)\operatorname{FS(\mathbb{Z}^{n})} be the simplicial complex whose vertices are rank 1 summands of n\mathbb{Z}^{n}, and there is a \ell-simplex {A0,,A}\{A_{0},\ldots,A_{\ell}\} if A0AA_{0}\oplus\cdots\oplus A_{\ell} is a rank +1\ell+1 summand of n\mathbb{Z}^{n}. There is a map φ:𝕊nns/IOnFS(n)\varphi:\operatorname{\mathbb{S}_{n}^{\text{ns}}}/IO_{n}\to\operatorname{FS(\mathbb{Z}^{n})} defined as follows. Let s𝕊nns/IOns\in\operatorname{\mathbb{S}_{n}^{\text{ns}}}/IO_{n} be a vertex, and choose a sphere SMnS\subset M_{n} which represents ss. As noted above, elements of IOnIO_{n} act trivially on H2(Mn)H_{2}(M_{n}). Therefore, the homology class [S]H2(Mn)[S]\in H_{2}(M_{n}) does not depend on the choice of representative SS. We then define φ(s)\varphi(s) to be the span of [S][S] in H2(Mn)H_{2}(M_{n}). It is clear that φ\varphi extends to simplices.

Lemma 7.3.

The map φ:𝕊nns/IOnFS(n)\varphi:\operatorname{\mathbb{S}_{n}^{\text{ns}}}/IO_{n}\to\operatorname{FS(\mathbb{Z}^{n})} is an isomorphism of simplicial complexes.

Proof.

Let σ={A0,,A}\sigma=\{A_{0},\ldots,A_{\ell}\} be an \ell-simplex of FS(n)\operatorname{FS(\mathbb{Z}^{n})}. We must show that, up to the action of IOnIO_{n}, there exists a unique \ell-simplex σ~\tilde{\sigma} of 𝕊nns\operatorname{\mathbb{S}_{n}^{\text{ns}}} which projects to σ\sigma.

Let vjH2(Mn)v_{j}\in H_{2}(M_{n}) be a primitive element generating AjA_{j} for 0j0\leq j\leq\ell, and extend this to a basis {v0,,vn1}\{v_{0},\ldots,v_{n-1}\} for H2(Mn,b)=nH_{2}(M_{n,b})=\mathbb{Z}^{n}. In Appendix B, we will prove Lemma B.2, which says that there exists a collection {S0,,Sn1}\{S_{0},\ldots,S_{n-1}\} of disjoint embedded 2-spheres such that [Sj]=vj[S_{j}]=v_{j} for 0jn10\leq j\leq n-1. Then the simplex σ~={S0,,S}\tilde{\sigma}=\{S_{0},\ldots,S_{\ell}\} of 𝕊nns\operatorname{\mathbb{S}_{n}^{\text{ns}}} maps to the σ\sigma under the composition

𝕊nns𝕊nns/IOn𝜑FS(n).\operatorname{\mathbb{S}_{n}^{\text{ns}}}\to\operatorname{\mathbb{S}_{n}^{\text{ns}}}/IO_{n}\overset{\varphi}{\to}\operatorname{FS(\mathbb{Z}^{n})}.

We will now show that σ~\tilde{\sigma} is unique up to the action of IOnIO_{n}. Suppose that σ~={S0,,S}\tilde{\sigma}^{\prime}=\{S_{0}^{\prime},\ldots,S_{\ell}^{\prime}\} is another simplex of 𝕊nns\operatorname{\mathbb{S}_{n}^{\text{ns}}} which projects to σ\sigma. Since σ~\tilde{\sigma} and σ~\tilde{\sigma}^{\prime} bother project to σ\sigma, we may order and orient the spheres such that [Sj]=[Sj][S_{j}]=[S_{j}^{\prime}] for 0j0\leq j\leq\ell. Again by Lemma B.2, we can extend {S1,,S}\{S_{1},\ldots,S_{\ell}\} and {S1,,S}\{S_{1}^{\prime},\ldots,S_{\ell}^{\prime}\} to collections of spheres {S1,,Sn}\{S_{1},\ldots,S_{n}\} and {S1,,Sn}\{S_{1}^{\prime},\ldots,S_{n}^{\prime}\} such that [Sj]=[Sj]=vj[S_{j}]=[S_{j}^{\prime}]=v_{j} for 0jn10\leq j\leq n-1. Notice that splitting MnM_{n} along either of these collections reduces MnM_{n} to a sphere with 2n2n boundary components. Therefore, there exists some 𝔣Mod(Mn)\mathfrak{f}\in\operatorname{Mod}(M_{n}) such that 𝔣(Sj)=Sj\mathfrak{f}(S_{j})=S_{j}^{\prime} for all jj. Let fOut(Fn)f\in\operatorname{Out}(F_{n}) be the image of ff. By construction, f(σ~)=σ~f(\tilde{\sigma})=\tilde{\sigma}^{\prime}. Furthermore, ff fixes a basis for homology, and so fIOnf\in IO_{n}. This completes the proof. ∎

This description of 𝕊nns/IOn\operatorname{\mathbb{S}_{n}^{\text{ns}}}/IO_{n} is advantageous because FS(n)\operatorname{FS(\mathbb{Z}^{n})} is known to be (n2)(n-2)-connected, and hence simply connected for n3n\geq 3. The first proof of this fact is due to Maazen [20] in his unpublished thesis (see [8, Theorem E] for a published proof). Thus, we have shown that 𝕊nns/IOn\operatorname{\mathbb{S}_{n}^{\text{ns}}}/IO_{n} is sufficiently connected.

Corollary 7.4 (Maazen).

The complex 𝕊nns/IOn\operatorname{\mathbb{S}_{n}^{\text{ns}}}/IO_{n} is simply connected for n3n\geq 3.

As indicated in Theorem 7.1, the stabilizers of spheres play an important role in the proof of Theorem D, and so we introduce notation for them here. If SS is an isotopy class of embedded sphere in MnM_{n}, we denote by Out(Fn,S)\operatorname{Out}(F_{n},S) the stabilizer of SS in Out(Fn)\operatorname{Out}(F_{n}), and define IOn(S)=Out(Fn,S)IOnIO_{n}(S)=\operatorname{Out}(F_{n},S)\cap IO_{n}. We now move on to the proof of Theorem D.

Proof of Theorem D.

We will induct on nn. The base cases are easy; IO1IO_{1} and IO2IO_{2} are both trivial. Suppose now that IOn1IO_{n-1} is Out(Fn1)\operatorname{Out}(F_{n-1})-normally generated by handle and commutator drags. We must now show that IOnIO_{n} is Out(Fn)\operatorname{Out}(F_{n})-normally generated by handle and commutator drags as well. By Theorem 7.1, Proposition 7.2, and Corollary 7.4, it suffices to show that IOn(S)IO_{n}(S) is generated by Out(Fn)\operatorname{Out}(F_{n})-conjugates of these drags for all SS. Let {α1,,αn}\{\alpha_{1},\ldots,\alpha_{n}\} be the geometric free basis of π1(Mn)\pi_{1}(M_{n}) identified with our fixed generating set {x1,,xn}\{x_{1},\ldots,x_{n}\} of FnF_{n}, and let {S1,,Sn}\{S_{1},\ldots,S_{n}\} be a corresponding sphere basis. Use these bases to construct the handle and commutator drags as in Section 5. Recall that handle drags correspond to the automorphisms MijM_{ij} of Magnus’s generators, and commutator drags correspond to MijkM_{ijk}. We will first show that IOn(S1)IO_{n}(S_{1}) is Out(Fn,S1)\operatorname{Out}(F_{n},S_{1})-normally generated by handle and commutator drags.

Splitting MnM_{n} along S1S_{1} yields a copy of Mn1,2M_{n-1,2}. Let NN be the tubular neighborhood of S1S_{1} removed in this splitting, and let 1\partial_{1} and 2\partial_{2} be the boundary components of Mn1,2M_{n-1,2}. Then this splitting induces a surjective map Out(Fn1,2)Out(Fn,S1)\mathrm{Out}(F_{n-1,2})\to\operatorname{Out}(F_{n},S_{1}), which restricts to a map ι:IOn1,2PIOn(S1)\iota_{*}:IO_{n-1,2}^{P}\to IO_{n}(S_{1}), where P={p1}={{1,2}}P=\{p_{1}\}=\{\{\partial_{1},\partial_{2}\}\}. This map is also surjective.

Use the bases {α2,,αn}\{\alpha_{2},\ldots,\alpha_{n}\} and {S2,,Sn}\{S_{2},\ldots,S_{n}\} to construct the handle, commutator, boundary commutator, and PP-drags in IOn1,2PIO_{n-1,2}^{P}. By our induction hypothesis combined with the proof of Theorem 6.3, these drags Out(Fn1,2)\mathrm{Out}(F_{n-1,2})-normally generate IOn1,2PIO_{n-1,2}^{P}. Notice that with these choices of drags, the map ι\iota_{*} takes handle and commutator drags to handle and commutator drags. Moreover, ι\iota_{*} takes boundary commutator drags in IOn1,2PIO_{n-1,2}^{P} to commutator drags in IOn(S)IO_{n}(S), and takes the PP-drag PDiP\operatorname{PD}_{i}^{P} to the handle drag HD1i\operatorname{HD}_{1i}. Thus, IOn(S1)IO_{n}(S_{1}) is Out(Fn,S1)\operatorname{Out}(F_{n},S_{1})-normally generated by handle and commutator drags.

Finally, let SS be an arbitrary vertex of 𝕊nns\operatorname{\mathbb{S}_{n}^{\text{ns}}}. Since SS is nonseparating, there exists some fOut(Fn)f\in\operatorname{Out}(F_{n}) such that f(S1)=Sf(S_{1})=S. It follows that

IOn(S)=fIOn(S1)f1.IO_{n}(S)=f\cdot IO_{n}(S_{1})\cdot f^{-1}.

Since IOn(S1)IO_{n}(S_{1}) is Out(Fn,S1)\operatorname{Out}(F_{n},S_{1})-normally generated by handle and commutator drags, it follows that IOn(S)IO_{n}(S) is generated by Out(Fn)\operatorname{Out}(F_{n})-conjugates of handle and commutator drags, which is what we wanted to show. ∎

8 Computing the abelianization

In this section, we compute the abelianization of the group IOn,bPIO_{n,b}^{P}, proving Theorem E. For the Torelli subgroup of the mapping class group of a surface, this was done by Johnson [17]. Some key tools used in this computation are the Johnson homomorphisms

τΣg,1:(Σg,1)3HandτΣg:(Σg)(3H)/H,\tau_{\Sigma_{g,1}}:\operatorname{\mathcal{I}}(\Sigma_{g,1})\to\wedge^{3}H\qquad\text{and}\qquad\tau_{\Sigma_{g}}:\operatorname{\mathcal{I}}(\Sigma_{g})\to(\wedge^{3}H)/H,

where H=H1(Σg,b)H=H_{1}(\Sigma_{g,b}). Johnson showed that these homomorphisms are the abelianization maps modulo torsion if g3g\geq 3. For IAn=IOn,1IA_{n}=IO_{n,1}, Andreadakis [1] and Bachmuth [3] used an analogous homomorphism τ:IAnHom(H,2H)\tau:IA_{n}\to\operatorname{Hom}(H,\wedge^{2}H) (where now H=H1(Fn)=nH=H_{1}(F_{n})=\mathbb{Z}^{n}) to show that

H1(IAn)Hom(H,2H)n(n2).H_{1}(IA_{n})\cong\operatorname{Hom}(H,\wedge^{2}H)\cong\mathbb{Z}^{n\cdot\binom{n}{2}}.

We will begin by recalling the definition of τ\tau, along with the computation of the ranks of H1(IAn)H_{1}(IA_{n}) and H1(IOn)H_{1}(IO_{n}), and then proceed to the case of multiple boundary components.

The Johnson homomorphism.

Recall that Out(Fn,1)Aut(Fn)\mathrm{Out}(F_{n,1})\cong\operatorname{Aut}(F_{n}), and the subgroup IAnIA_{n} is precisely those automorphisms which act trivially on H1(Fn)=nH_{1}(F_{n})=\mathbb{Z}^{n}. The goal is to construct a homomorphism τ:IAnHom(H,2H)\tau:IA_{n}\to\operatorname{Hom}(H,\wedge^{2}H), where H=H1(Fn)=nH=H_{1}(F_{n})=\mathbb{Z}^{n}.

First, we claim that the group [Fn,Fn]/[Fn,[Fn,Fn]][F_{n},F_{n}]/[F_{n},[F_{n},F_{n}]] is isomorphic to 2H\wedge^{2}H, where [Fn,Fn][F_{n},F_{n}] denotes the commutator subgroup of FnF_{n}. To see this, consider the short exact sequence

1[Fn,Fn]Fnn1.1\to[F_{n},F_{n}]\to F_{n}\to\mathbb{Z}^{n}\to 1.

Passing to the five-term exact sequence in homology, we get the sequence

0H2(n)H1([Fn,Fn])nH1(Fn)H1(n)0,0\to H_{2}(\mathbb{Z}^{n})\to H_{1}([F_{n},F_{n}])_{\mathbb{Z}^{n}}\to H_{1}(F_{n})\to H_{1}(\mathbb{Z}^{n})\to 0,

where H1([Fn,Fn])n=[Fn,Fn]/[Fn,[Fn,Fn]]H_{1}([F_{n},F_{n}])_{\mathbb{Z}^{n}}=[F_{n},F_{n}]/[F_{n},[F_{n},F_{n}]] denotes the group of co-invariants of H1([Fn,Fn])H_{1}([F_{n},F_{n}]) with respect to the action of n\mathbb{Z}^{n} (induced by the conjugation action of FnF_{n} on [Fn,Fn][F_{n},F_{n}]). The map H1(Fn)H1(n)H_{1}(F_{n})\to H_{1}(\mathbb{Z}^{n}) is clearly an isomorphism, and so it follows that the map H2(n)[Fn,Fn]/[Fn,[Fn,Fn]]H_{2}(\mathbb{Z}^{n})\to[F_{n},F_{n}]/[F_{n},[F_{n},F_{n}]] is an isomorphism as well. This proves our claim because H2(n)2nH_{2}(\mathbb{Z}^{n})\cong\wedge^{2}\mathbb{Z}^{n}. Let ρ:[Fn,Fn]2n\rho:[F_{n},F_{n}]\to\wedge^{2}\mathbb{Z}^{n} be the projection. Following the definitions above, we see that ρ\rho is defined by

ρ([x,y])=[x][y],\rho([x,y])=[x]\wedge[y],

where [x][x] and [y][y] denote the classes of xx and yy in HH, respectively.

Next, let fIAnf\in IA_{n}. Then f(x)x1f(x)x^{-1} is nullhomologous for all xFnx\in F_{n}, and therefore lies in [Fn,Fn][F_{n},F_{n}]. We define the map τ^f:Fn2H\hat{\tau}_{f}:F_{n}\to\wedge^{2}H via

τ^f(x)=ρ(f(x)x1).\hat{\tau}_{f}(x)=\rho(f(x)x^{-1}).

We now check that τ^f\hat{\tau}_{f} is a homomorphism. Let x,yFnx,y\in F_{n}. Applying the relation ab=[a,b]baab=[a,b]ba, we get

τ^f(xy)\displaystyle\hat{\tau}_{f}(xy) =ρ(f(x)f(y)y1x1)\displaystyle=\rho(f(x)f(y)y^{-1}x^{-1})
=ρ([f(x),f(y)y1](f(y)y1)(f(x)x1))\displaystyle=\rho\left([f(x),f(y)y^{-1}]\cdot(f(y)y^{-1})\cdot(f(x)x^{-1})\right)
=[f(x)][f(y)y1]+τ^f(y)+τ^f(y)\displaystyle=[f(x)]\wedge[f(y)y^{-}1]+\hat{\tau}_{f}(y)+\hat{\tau}_{f}(y)
=τ^f(y)+τ^f(y),\displaystyle=\hat{\tau}_{f}(y)+\hat{\tau}_{f}(y),

since [f(y)y1]=0[f(y)y^{-1}]=0. This shows that τ^\hat{\tau} is indeed a homomorphism. Furthermore, since 2H\wedge^{2}H is abelian, the map τ^:Fn2H\hat{\tau}:F_{n}\to\wedge^{2}H factors through the abelianization, inducing a map τf:H2H\tau_{f}:H\to\wedge^{2}H. Therefore, we have a map

τ:IAnHom(H,2H)\tau:IA_{n}\to\operatorname{Hom}(H,\wedge^{2}H)

sending ff to τf\tau_{f}. We now check that τ\tau is a homomorphism. Let f,gIAnf,g\in IA_{n}. Then

τfg([x])\displaystyle\tau_{fg}([x]) =ρ(f(g(x))x1)\displaystyle=\rho(f(g(x))x^{-1})
=ρ(f(g(x))(g(x))1g(x)x1)\displaystyle=\rho(f(g(x))(g(x))^{-1}g(x)x^{-1})
=τf([g(x)])+τg([x])\displaystyle=\tau_{f}([g(x)])+\tau_{g}([x])
=τf([x])+τg([x])\displaystyle=\tau_{f}([x])+\tau_{g}([x])

since gg fixes [x][x]. Thus, τ\tau is the desired homomorphism.

We now move on to computing the image of our generators under τ\tau. Since we are dealing with the case of one boundary component, boundary commutator drags are unnecessary since they are products of PP-drags. Fix a basepoint Mn,1*\in\partial M_{n,1}, and choose a basis {x1,,xn}\{x_{1},\ldots,x_{n}\} of π1(Mn,1,)\pi_{1}(M_{n,1},*). Construct the handle, commutator, and PP-drags with respect to this basis.

Handle drags.

Recall that the handle drag HDij\operatorname{HD}_{ij} acts on π1(Mn,1)\pi_{1}(M_{n,1}) by sending xix_{i} to xjxixj1x_{j}x_{i}x_{j}^{-1}, and fixing the remaining basis elements. Therefore,

τ(HDij)([x])=ρ(HDij(x)x1)={0if iρ(xjxixj1xi1)if =i.\displaystyle\tau(\operatorname{HD}_{ij})([x_{\ell}])=\rho(\operatorname{HD}_{ij}(x_{\ell})x_{\ell}^{-1})=\begin{cases}0&\text{if }\ell\neq i\\ \rho(x_{j}x_{i}x_{j}^{-1}x_{i}^{-1})&\text{if }\ell=i.\end{cases}

Thus, τ(HDij)\tau(\operatorname{HD}_{ij}) is the homomorphism [xi][xj][xi][x_{i}]\mapsto[x_{j}]\wedge[x_{i}] (and all other generators are sent to 0).

Commutator drags.

Notice that the product of commutator drags CDijk+CDijk\operatorname{CD}_{ijk}^{+}\cdot\operatorname{CD}_{ijk}^{-} is equal to a commutator of handle drags. Therefore, only the CDijk\operatorname{CD}_{ijk}^{-} are needed in our generating set, and we can disregard the CDijk+\operatorname{CD}_{ijk}^{+} from now on. Recall that CDijk\operatorname{CD}_{ijk}^{-} acts on π1(Mn,1)\pi_{1}(M_{n,1}) by sending xix_{i} to [xj,xk]xi[x_{j},x_{k}]x_{i}. Therefore,

τ(CDijk)([x])=ρ(CDijk(x)x1)={0if iρ([xj,xk])if =i.\displaystyle\tau(\operatorname{CD}_{ijk}^{-})([x_{\ell}])=\rho(\operatorname{CD}_{ijk}^{-}(x_{\ell})x_{\ell}^{-1})=\begin{cases}0&\text{if }\ell\neq i\\ \rho([x_{j},x_{k}])&\text{if }\ell=i.\end{cases}

It follows that τ(CDijk)\tau(\operatorname{CD}_{ijk}^{-}) is the map [xi][xj][xk][x_{i}]\mapsto[x_{j}]\wedge[x_{k}].

PP-drags.

Next, we note that the product

PDjHD1jHDnj\operatorname{PD}_{j}\cdot\operatorname{HD}_{1j}\cdots\operatorname{HD}_{nj} (5)

is trivial in IAnIA_{n}. For a justification of this fact, see the proof of the claim at the end of Theorem A.2. It follows that the PP-drags are also redundant in our generating set for IAnIA_{n}, and can be removed.

Abelianization of IAnIA_{n}.

To compute the rank of the abelianization of IAnIA_{n}, we use the following lemma.

Lemma 8.1.

Let GG be a group and SS a finite generating set for GG. Suppose that φ:G|S|\varphi:G\to\mathbb{Z}^{|S|} is a surjective homomorphism. Then H1(G)|S|H_{1}(G)\cong\mathbb{Z}^{|S|}.

Proof.

Let F(S)F(S) denote the free group on the set SS. Since |S|\mathbb{Z}^{|S|} is abelian, the homomorphism φ\varphi factors through the abelianization to give a map φ¯:H1(G)n\overline{\varphi}:H_{1}(G)\to\mathbb{Z}^{n}, which is also surjective. Additionally, by the universal property of free groups, we have a map ψ:F(S)G\psi:F(S)\to G. Passing to the abelianizations induces a map ψ¯:H1(F(S))H1(G)\overline{\psi}:H_{1}(F(S))\to H_{1}(G). Since SS is a generating set for GG, this map is also surjective. It follows that φ¯ψ¯\overline{\varphi}\circ\overline{\psi} is a surjective map between free abelian groups of equal rank, and is hence an isomorphism. Thus, φ¯\overline{\varphi} is an isomorphism as well. ∎

From the discussion in the preceeding paragraphs, we have a generating set for IAnIA_{n} of size

#(Handle Drags)+#(Commutator Drags)=n(n1)+n(n12)=n(n2)\#(\text{Handle Drags})+\#(\text{Commutator Drags})=n(n-1)+n\cdot\binom{n-1}{2}=n\cdot\binom{n}{2}

(since PP-drags can be written as a product of handle drags), and the image of this generating set spans Hom(H,2H)\operatorname{Hom}(H,\wedge^{2}H), which also has dimension n(n2)n\cdot\binom{n}{2}. Therefore, by Lemma 8.1, the group H1(IAn)H_{1}(IA_{n}) has rank n(n2)n\cdot\binom{n}{2}.

To compute the rank of H1(IOn)H_{1}(IO_{n}), consider the quotient map IAnIOnIA_{n}\to IO_{n}, whose kernel is the subgroup of inner automorphisms (or PP-drags under our geometric interpretation of IAnIA_{n}). We compute the image of a PP-drag under τ\tau:

τ(PDi)([x])=ρ(PDi(x)x1)=ρ(xi1xxix1)=[x][xi].\tau(\operatorname{PD}_{i})([x_{\ell}])=\rho(\operatorname{PD}_{i}(x_{\ell})x_{\ell}^{-1})=\rho(x_{i}^{-1}x_{\ell}x_{i}x_{\ell}^{-1})=[x_{\ell}]\wedge[x_{i}].

Since τ(PDi)\tau(\operatorname{PD}_{i}) is nontrivial, τ\tau does not descend to a map IOnHom(H,2H)IO_{n}\to\operatorname{Hom}(H,\wedge^{2}H). However, the images {τ(PDi)}\{\tau(\operatorname{PD}_{i})\} span a subgroup of Hom(H,2H)\operatorname{Hom}(H,\wedge^{2}H) isomorphic to HH (where the isomorphism is given by [h]([x][x][h])[h]\mapsto([x_{\ell}]\mapsto[x_{\ell}]\wedge[h])). So, τ\tau induces a map IOnHom(H,2H)/HIO_{n}\to\operatorname{Hom}(H,\wedge^{2}H)/H. Just as the element given in Equation (5) is trivial in IAnIA_{n}, the product HD1jHDnj\operatorname{HD}_{1j}\cdots\operatorname{HD}_{nj} is trivial in IOnIO_{n} for all j{1,,n}j\in\{1,\ldots,n\}. Thus, we may throw out nn handle drags from our generating set to obtain a generating set for IOnIO_{n} of size n(n2)nn\cdot\binom{n}{2}-n. Since Hom(H,2H)\operatorname{Hom}(H,\wedge^{2}H) has rank n(n2)nn\cdot\binom{n}{2}-n, Lemma 8.1 implies that H1(IOn)H_{1}(IO_{n}) has rank n(n2)nn\cdot\binom{n}{2}-n as well. This verifies Theorem 1.2 from the introduction.

Multiple boundary components.

We now move on to the case of multiple boundary components. Just as we did when constructing our drags in Section 5, fix an ordering P={p1,,p|P|}P=\{p_{1},\ldots,p_{|P|}\} and an ordering pr={1r,,brr}p_{r}=\{\partial_{1}^{r},\ldots,\partial_{b_{r}}^{r}\} for each prPp_{r}\in P. We cap off the boundary components of each pPp\in P as follows:

  • If |p|=1|p|=1, we attach a copy of M1,1M_{1,1} to the single boundary component of pp.

  • If |p|=k>1|p|=k>1, we attach a copy of M0,kM_{0,k} to these kk boundary components.

  • If p=p1p=p_{1}, we follow the same rules as above, except we introduce an additional boundary component in the piece glued to pp.

Capping off the boundary components in this way gives an embedding

ι:Mn,bMm,1,\iota:M_{n,b}\hookrightarrow M_{m,1},

where the boundary component of Mm,1M_{m,1} lies in the piece attached to p1p_{1}. This embedding induces a map ι:IOn,bPIAm\iota_{*}:IO_{n,b}^{P}\to IA_{m}. We obtain a similar map IAmIOmIA_{m}\to IO_{m} by attaching a disk to the boundary component of Mm,1M_{m,1}. In Appendix A, we will prove Theorem A.2, which says that the composition

IOn,bPιIAmIOmIO_{n,b}^{P}\overset{\iota_{*}}{\to}IA_{m}\to IO_{m}

is injective. It follows that ι\iota_{*} is injective as well. Therefore, to compute the rank of the abelianization of IOn,bPIO_{n,b}^{P}, it suffices to compute the rank of the abelianization of its image in IAmIA_{m}. Let H=H1(Mm,1)H=H_{1}(M_{m,1}), and let τ:IOn,bPHom(H,2H)\tau_{*}:IO_{n,b}^{P}\to\operatorname{Hom}(H,\wedge^{2}H) denote the composition

IOn,bPιIAm𝜏Hom(H,2H).IO_{n,b}^{P}\overset{\iota_{*}}{\to}IA_{m}\overset{\tau}{\to}\operatorname{Hom}(H,\wedge^{2}H).

Our goal now becomes computing the images of handle, commutator, boundary commutator, and PP-drags under τ\tau_{*}.

Choosing a basis.

To carry out this computation, it will be helpful to choose bases for π1(Mn,b)\pi_{1}(M_{n,b}) and π1(Mm,1)\pi_{1}(M_{m,1}) carefully. For simplicity, we will assume that |p1|>1|p_{1}|>1. The case of |p1|=1|p_{1}|=1 is more straightforward. Fix a basepoint z1111z_{1}^{1}\in\partial_{1}^{1} and a basis {x1,,xn}\{x_{1},\ldots,x_{n}\} for π1(Mn,b,z11)\pi_{1}(M_{n,b},z_{1}^{1}). Choose a corresponding sphere basis {S1,,Sn}\{S_{1},\ldots,S_{n}\}. We define our drags in IOn,bPIO_{n,b}^{P} with respect to these bases.

Next, choose a basepoint zMm,1z\in\partial M_{m,1}, and an oriented arc α1Mm,1int(Mn,b)\alpha_{1}\subset M_{m,1}\setminus\operatorname{int}(M_{n,b}) from zz to z11z_{1}^{1} (this is possible since the boundary component of Mm,1M_{m,1} lies on the piece attached to p1p_{1}). For i{1,,n}i\in\{1,\ldots,n\}, let yi=α1xiα11y_{i}=\alpha_{1}x_{i}\alpha_{1}^{-1}. Then {y1,,yn}\{y_{1},\ldots,y_{n}\} is a partial basis for π1(Mm,1,z)\pi_{1}(M_{m,1},z). We wish to extend this to a full basis. Throughout the definition of this extended basis, we encourage the reader to follow along in Figure 12.

For each boundary component sr\partial_{s}^{r} of Mn,bM_{n,b}, fix a point zsrsrz_{s}^{r}\in\partial_{s}^{r} (leaving z11z_{1}^{1} as before). For each zsrz11z_{s}^{r}\neq z_{1}^{1}, let βsr\beta_{s}^{r} be the unique oriented arc (up to isotopy) in Mn,bSiM_{n,b}\setminus\bigcup S_{i} from z11z_{1}^{1} to zsrz_{s}^{r}. For s{2,,b1}s\in\{2,\ldots,b_{1}\}, define ys1=α1βs1αs1y_{s}^{1}=\alpha_{1}\beta_{s}^{1}\alpha_{s}^{-1}. In Figure 12, the loops y11y_{1}^{1} and y21y_{2}^{1} are of this form.

Next, let r>1r>1. If |pr|=1|p_{r}|=1, let γ1r\gamma_{1}^{r} be an oriented loop based at z1rz_{1}^{r} which generates the fundamental group of the copy of M1,1M_{1,1} attached to prp_{r}. Then we define y1r=α1β1rγ11(β1r)1(α1)1y_{1}^{r}=\alpha_{1}\beta_{1}^{r}\gamma_{1}^{1}(\beta_{1}^{r})^{-1}(\alpha_{1})^{-1}. In Figure 12, the curve y13y_{1}^{3} is an example of such a loop.

On the other hand, suppose |pr|>1|p_{r}|>1. For s{2,,br}s\in\{2,\ldots,b_{r}\}, let γsr\gamma_{s}^{r} be the unique (up to isotopy) oriented curve in Mm,1int(Mn,b)M_{m,1}\setminus\operatorname{int}(M_{n,b}) from z1rz_{1}^{r} to zsrz_{s}^{r}. Then, define

ysr=α1β1rγsr(βsr)1(α1)1.y_{s}^{r}=\alpha_{1}\beta_{1}^{r}\gamma_{s}^{r}(\beta_{s}^{r})^{-1}(\alpha_{1})^{-1}.

The curve y22y_{2}^{2} is an example of this type of loop in Figure 12.

\labellist
\hair

2pt \pinlabelα1\alpha_{1} at 95 280

\pinlabel

y1y_{1} at 280 105 \pinlabely2y_{2} at 370 100 \pinlabely3y_{3} at 430 143

\pinlabel

y11y_{1}^{1} at 85 200 \pinlabely21y_{2}^{1} at 65 160

\pinlabel

y22y_{2}^{2} at 320 393

\pinlabel

y13y_{1}^{3} at 550 360

\pinlabel

M3,6M_{3,6} at 530 70 \endlabellistRefer to caption

Figure 12: M3,6M_{3,6} with the partition P={{11,21,31},{12,22},{13}}P=\{\{\partial_{1}^{1},\partial_{2}^{1},\partial_{3}^{1}\},\{\partial_{1}^{2},\partial_{2}^{2}\},\{\partial_{1}^{3}\}\} embedded into M7,1M_{7,1}. The loops {y1,y2,y3}\{y_{1},y_{2},y_{3}\} are freely homotopic to a basis for π1(M3,6)\pi_{1}(M_{3,6}), and this basis has been extended to a basis {y1,y2,y3,y11,y21,y22,y13}\{y_{1},y_{2},y_{3},y_{1}^{1},y_{2}^{1},y_{2}^{2},y_{1}^{3}\} of π1(M7,1,z)\pi_{1}(M_{7,1},z).

Let Y={y1,yn}Y=\{y_{1}\ldots,y_{n}\}. For r{1,,|P|}r\in\{1,\ldots,|P|\}, let Yr={y1r}Y_{r}=\{y_{1}^{r}\} if |pr|=1|p_{r}|=1, and Yr={y2r,,ybrr}Y_{r}=\{y_{2}^{r},\ldots,y_{b_{r}}^{r}\} otherwise. Then the collection

YY1Y|P|Y\cup Y_{1}\cup\cdots\cup Y_{|P|}

forms a free basis for π1(Mm,1,z)\pi_{1}(M_{m,1},z).

Computations and relations.

We now move on to the computation of the images of our collection of drags under τ\tau_{*}. These computations are straightforward, and are summarized in Figure 13. We see from these computations that there is a relation between the images of PP-drags and Handle Drags. Namely,

r=1|P|τ(PDjr)=i=1nτ(HDij)\sum_{r=1}^{|P|}\tau_{*}(\operatorname{PD}_{j}^{r})=-\sum_{i=1}^{n}\tau_{*}(\operatorname{HD}_{ij}) (6)

for all j{1,,n}j\in\{1,\ldots,n\}. As we saw in the case of one boundary component, this is because

PDj1PDj|P|HD1jHDnj=1\operatorname{PD}_{j}^{1}\cdots\operatorname{PD}_{j}^{|P|}\cdot\operatorname{HD}_{1j}\cdot\operatorname{HD}_{nj}=1 (7)

in IOn,bPIO_{n,b}^{P}. Additionally, we see a relation between the image of boundary commutator drags:

s=1brτ(BCDijrs)=0\sum_{s=1}^{b_{r}}\tau_{*}(\operatorname{BCD}_{ij}^{rs})=0 (8)

for all r{1,,|P|}r\in\{1,\ldots,|P|\} and i,j{1,,n}i,j\in\{1,\ldots,n\} with i<ji<j. This relation holds because

BCDijr1BCDijrbr=[PDir,PDjr]\operatorname{BCD}_{ij}^{r1}\cdots\operatorname{BCD}_{ij}^{rb_{r}}=[\operatorname{PD}_{i}^{r},\operatorname{PD}_{j}^{r}] (9)

in IOn,bPIO_{n,b}^{P}.

Drag Action on π1\pi_{1} Image under τ\tau_{*}
HDij\operatorname{HD}_{ij} yiyjyiyj1y_{i}\mapsto y_{j}y_{i}y_{j}^{-1} [yi][yj][yi][y_{i}]\mapsto[y_{j}]\wedge[y_{i}]
CDijk\operatorname{CD}_{ijk}^{-} yi[yj,yk]yiy_{i}\mapsto[y_{j},y_{k}]y_{i} [yi][yj][yk][y_{i}]\mapsto[y_{j}]\wedge[y_{k}]
BCDjkrs\operatorname{BCD}_{jk}^{rs} (r,s>1r,s>1) ysrysr[yj,yk]1y_{s}^{r}\mapsto y_{s}^{r}[y_{j},y_{k}]^{-1} [ysr][yk][yj][y_{s}^{r}]\mapsto[y_{k}]\wedge[y_{j}]
BCDjkr1\operatorname{BCD}_{jk}^{r1} (r>1r>1) ysr[yj,yk]ysry_{s}^{r}\mapsto[y_{j},y_{k}]y_{s}^{r} (s>1s>1) [ysr][yj][yk][y_{s}^{r}]\mapsto[y_{j}]\wedge[y_{k}] (s>1s>1)
BCDjk1s\operatorname{BCD}_{jk}^{1s} (s>1s>1) ys1ys1[yj,yk]y_{s}^{1}\mapsto y_{s}^{1}[y_{j},y_{k}] [ys1][yj][yk][y_{s}^{1}]\mapsto[y_{j}]\wedge[y_{k}]
BCDjk11\operatorname{BCD}_{jk}^{11} y[yj,yk]1y[yj,yk]y\mapsto[y_{j},y_{k}]^{-1}y[y_{j},y_{k}] (yY1y\not\in Y_{1}) [y]0[y]\mapsto 0 (yY1y\not\in Y_{1})
ys1[yj,yk]1ysry_{s}^{1}\mapsto[y_{j},y_{k}]^{-1}y_{s}^{r} (s>1s>1) [ys1][yk][yj][y_{s}^{1}]\mapsto[y_{k}]\wedge[y_{j}] (s>1s>1)
PDjr\operatorname{PD}_{j}^{r} (r>1r>1) ysryjysryj1y_{s}^{r}\mapsto y_{j}y_{s}^{r}y_{j}^{-1} (s>1s>1) [ysr][yj][ysr][y_{s}^{r}]\mapsto[y_{j}]\wedge[y_{s}^{r}] (s>1s>1)
PDj1\operatorname{PD}_{j}^{1} yyj1yyjy\mapsto y_{j}^{-1}yy_{j} (yY1y\not\in Y_{1}) [y][y][yj][y]\mapsto[y]\wedge[y_{j}] (yY1y\not\in Y_{1})
Figure 13: Computing the image of drags under τ\tau_{*}.

Contributions to abelianization.

From the computations and relations above, we see that the handle drags and commutator drags together still contribute n(n2)n\cdot\binom{n}{2} dimensions to the abelianization of IOn,bPIO_{n,b}^{P}. There are b(n2)b\cdot\binom{n}{2} boundary commutator drags, but the relations given in Equation (8) kill off |P|(n2)|P|\cdot\binom{n}{2} of these in the abelianization (though we can also remove this many elements from our generating set by using Equation (9)). Finally, the number of PP-drags is |P|n|P|\cdot n, but nn of them are killed in the abelianization by Equation (6) (and again, we may remove nn elements from our generating set by Equation (7)). Adding these all together, we find that the image of τ:IOn,bPHom(H,2H)\tau_{*}:IO_{n,b}^{P}\to\operatorname{Hom}(H,\wedge^{2}H) has rank

R=n(n2)+(b(n2)|P|(n2))+(|P|nn).R=n\cdot\binom{n}{2}+\left(b\cdot\binom{n}{2}-|P|\cdot\binom{n}{2}\right)+(|P|\cdot n-n).

Moreover, we can reduce our generating set (using Equations (7) and (9)) to a set of size RR as well. Thus, by Lemma 8.1, the group H1(IOn,bP)H_{1}(IO_{n,b}^{P}) has rank RR, which proves Theorem E.

Appendix A Injectivity of the inclusion map

We end this paper with a proof of the following facts, which are surely known to experts, but for which we do not know a reference. They are significant because they allow us to realize the groups Out(Fn,b)\mathrm{Out}(F_{n,b}) (and hence IOn,bPIO_{n,b}^{P}) as subgroups of Out(Fm)\operatorname{Out}(F_{m}). We will begin with a low-genus case.

Lemma A.1.

The induced map ι:Out(F1,1)Out(Fm)\iota_{*}:\mathrm{Out}(F_{1,1})\to\operatorname{Out}(F_{m}) is injective for any embedding ι:M1,1Mm\iota:M_{1,1}\hookrightarrow M_{m}.

Proof.

By Laudenbach [18], the group Out(F1,1)Aut(F1)/2\mathrm{Out}(F_{1,1})\cong\operatorname{Aut}(F_{1})\cong\mathbb{Z}/2, where the nontrivial element fOut(F1,1)f\in\mathrm{Out}(F_{1,1}) acts on π1(M1,1,x)\pi_{1}(M_{1,1},x)\cong\mathbb{Z} by inverting the generator. Therefore, ι(f)Out(Fm)\iota_{*}(f)\in\operatorname{Out}(F_{m}) is the class of the automorphism

{x1x11xjxjif j>1.\begin{cases}x_{1}\mapsto x_{1}^{-1}\\ x_{j}\mapsto x_{j}&\text{if }j>1.\end{cases}

This automorphism is not an inner automorphism for any m1m\geq 1, so ii_{*} is injective. ∎

Theorem A.2.

Fix n,b1n,b\geq 1 such that (n,b)(1,1)(n,b)\neq(1,1), and let ι:Mn,bMm\iota:M_{n,b}\hookrightarrow M_{m} be an embedding. The induced map ι:Out(Fn,b)Out(Fm)\iota_{*}:\mathrm{Out}(F_{n,b})\to\operatorname{Out}(F_{m}) is injective if and only if no component of Mmint(Mn,b)M_{m}\setminus\operatorname{int}(M_{n,b}) is diffeomorphic to a 3-disk.

Proof.

Suppose first that some component of Mmint(Mn,b)M_{m}\setminus\operatorname{int}(M_{n,b}) is diffeomorphic to a disk, and let \partial be the boundary component of Mn,bM_{n,b} capped off by this disk. By the Birman exact sequence (Theorem 4.2), dragging this boundary component along any nontrivial loop will give a nontrivial element in the kernel of ι\iota_{*}.

Suppose now that no component of Mmint(Mn,b)M_{m}\setminus\operatorname{int}(M_{n,b}) is a disk. We will first prove the theorem in the case b=1b=1, and then move on to the general result.

Case 1: Suppose we have an embedding ι:Mn,1Mm\iota:M_{n,1}\hookrightarrow M_{m}. Since no component of Mmint(Mn,b)M_{m}\setminus\operatorname{int}(M_{n,b}) is a disk, m>nm>n. If n=1n=1, then we are done by Lemma A.1, so we may assume that n>1n>1. Fix a basepoint xx on the boundary of Mn,1M_{n,1}, and choose a free basis {x1,,xn}\{x_{1},\ldots,x_{n}\} of π1(Mn,1,x)\pi_{1}(M_{n,1},x). The embedding ι\iota induces an injection π1(Mn,b,x)π1(Mm,x)\pi_{1}(M_{n,b},x)\hookrightarrow\pi_{1}(M_{m},x) which identifies π1(Mn,1,x)\pi_{1}(M_{n,1},x) with a free summand of π1(Mm,x)\pi_{1}(M_{m},x). This allows us to extend {x1,,xn}\{x_{1},\ldots,x_{n}\} to a free basis {x1,,xm}\{x_{1},\ldots,x_{m}\} of π1(Mm,x)\pi_{1}(M_{m},x). Given fOut(Fn,1)Aut(Fn)f\in\mathrm{Out}(F_{n,1})\cong\operatorname{Aut}(F_{n}), the image ι(f)Out(Fm)\iota_{*}(f)\in\operatorname{Out}(F_{m}) is the class of the automorphism φAut(Fm)\varphi\in\operatorname{Aut}(F_{m}) generated by

φ:{xif(xi) if 1inxixi if n<im.\varphi:\begin{cases}x_{i}\mapsto f(x_{i})&\text{ if }1\leq i\leq n\\ x_{i}\mapsto x_{i}&\text{ if }n<i\leq m.\end{cases}

Suppose that φ\varphi is an inner automorphism. If m>n+1m>n+1, then φ\varphi fixes at least two generators of FmF_{m}, and thus must be trivial. It follows that ff is trivial as well. On the other hand, if m=n+1m=n+1, then φ\varphi fixes xmx_{m}. Since φ\varphi is inner, φ\varphi must conjugate by a power of xmx_{m}. However, if φ\varphi conjugates by a nontrivial power of xmx_{m}, then ff would not act as an automorphism on x1,,xnFm\langle x_{1},\ldots,x_{n}\rangle\subset F_{m}, which is a contradiction. Thus, φ\varphi is trivial, and so ff is trivial as well.

In summary, we have shown that φ\varphi is an inner automorphism if and only if ff is trivial, which implies that ι\iota_{*} is injective.

\labellist
\hair

2pt \pinlabelxx at 214 240

\pinlabel

1\partial_{1} at 330 155 \pinlabel2\partial_{2} at 345 205 \pinlabel3\partial_{3} at 330 300

\pinlabel

x3x_{3} at 490 112 \pinlabelx4x_{4} at 467 320 \pinlabelx5x_{5} at 480 408

\pinlabel

Σ\Sigma at 245 100 \pinlabelM2,3M_{2,3} at 140 230 \endlabellistRefer to caption

Figure 14: M2,3M_{2,3} embedded inside M5M_{5}. For clarity, x1x_{1} and x2x_{2} are not shown, but they lie entirely on the opposite side of Σ\Sigma from x3x_{3}, x4x_{4}, and x5x_{5}.

Case 2: Next, suppose that ι:Mn,bMm\iota:M_{n,b}\hookrightarrow M_{m} is an embedding, where b>1b>1. Let 1,,b\partial_{1},\ldots,\partial_{b} be the boundary components of Mn,bM_{n,b}. Let ΣMn,b\Sigma\subset M_{n,b} be a 2-sphere which separates Mn,bM_{n,b} into Mn,1M_{n,1} and M0,b+1M_{0,b+1} (see Figure 14). Then we have a composition of inclusions

Mn,1Mn,bMm.M_{n,1}\hookrightarrow M_{n,b}\hookrightarrow M_{m}.

Let κ:Out(Fn,1)Out(Fn,b)\kappa_{*}:\mathrm{Out}(F_{n,1})\to\mathrm{Out}(F_{n,b}) be the map induced by inclusion. By the preceding case, ικ\iota_{*}\circ\kappa_{*} is injective. Let fOut(Fn,b)f\in\mathrm{Out}(F_{n,b}), and suppose that ι(f)=id\iota_{*}(f)=\operatorname{id}. By repeated applications of the Birman exact sequence (Theorem 4.2), ff has the form f=p1p2pbκ(g)f=p_{1}p_{2}\cdots p_{b}\cdot\kappa_{*}(g), where gOut(Fn,1)Aut(Fn)g\in\mathrm{Out}(F_{n,1})\cong\operatorname{Aut}(F_{n}) and pjOut(Fn,1)p_{j}\in\mathrm{Out}(F_{n,1}) is a boundary drag of j\partial_{j} along a loop βj\beta_{j}. Fix a basepoint xΣx\in\Sigma, and let γjπ1(Mn,b,x)\gamma_{j}\in\pi_{1}(M_{n,b},x) be representative of the free homotopy class of βj\beta_{j}. Choose a free basis {x1,,xn}\{x_{1},\ldots,x_{n}\} for π1(Mn,1,x)\pi_{1}(M_{n,1},x). Extend this to a free basis {x1,,xm}\{x_{1},\ldots,x_{m}\} for π1(Mm,x)\pi_{1}(M_{m},x) such that for each i>ni>n, the loop xix_{i} intersects the set j=1bj\bigcup_{j=1}^{b}\partial_{j} exactly twice: once when exiting Mn,bM_{n,b}, and once when re-entering (see Figure 14). For i>ni>n, let (i)\partial_{\ell(i)} be the boundary component through which αi\alpha_{i} leaves Mn,bM_{n,b}, and let r(i)\partial_{r(i)} be the boundary component through which it returns. Then ι(f)\iota_{*}(f) is the class of the automorphism φAut(Fm)\varphi\in\operatorname{Aut}(F_{m}) given by

φ:{xig(xi) for 1inxiγ(i)xiγr(i)1 for n<im.\varphi:\begin{cases}x_{i}\mapsto g(x_{i})&\text{ for }1\leq i\leq n\\ x_{i}\mapsto\gamma_{\ell(i)}x_{i}\gamma_{r(i)}^{-1}&\text{ for }n<i\leq m.\end{cases}

By assumption, this automorphism is an inner automorphism. Suppose that φ\varphi conjugates by a reduced word ww in the xix_{i}. Since gg is an automorphism of x1,,xnFm\langle x_{1},\ldots,x_{n}\rangle\subset F_{m}, it follows that wx1,xnw\in\langle x_{1},\ldots x_{n}\rangle. We will show that this implies that ff is trivial by induction on the reduced word length of ww.

For the base case, suppose that the word length of ww is 0. Then ww and φ\varphi are both trivial. Since ικ\iota_{*}\circ\kappa_{*} is injective, gg is trivial as well. Suppose now that some γj\gamma_{j} is non-nullhomotopic. Since no component of Mmint(Mn,b)M_{m}\setminus\operatorname{int}(M_{n,b}) is a disk, there exists some xix_{i} which passes through j\partial_{j}, where i>ni>n. In other words, either (i)=j\ell(i)=j or r(i)=jr(i)=j. This is a contradiction because then φ(xi)=γ(i)xiγr(i)1xi\varphi(x_{i})=\gamma_{\ell(i)}x_{i}\gamma_{r(i)}^{-1}\neq x_{i}. Thus, all γj\gamma_{j} are nullhomotopic, and so ff is trivial. This completes the base case.

Next, suppose that ww has positive word length, and let xi±1x_{i}^{\pm 1} be the last letter in the reduced form of ww. Then, w=wxi±1w=w^{\prime}x_{i}^{\pm 1}, where the length of ww^{\prime} is less than that of ww. To avoid notational complexity, we will assume that xi±1=x1x_{i}^{\pm 1}=x_{1}, but the same argument works for any other xix_{i}. Consider the element

h:=HD21HD31HDn1q1qbOut(Fn,b),h:=\operatorname{HD}_{21}\operatorname{HD}_{31}\cdots\operatorname{HD}_{n1}\cdot q_{1}\cdots q_{b}\in\mathrm{Out}(F_{n,b}),

where HDi1\operatorname{HD}_{i1} is the handle drag of the ii-th handle about the first handle (see Section 5) and qjq_{j} is obtained by dragging j\partial_{j} about a loop in the free homotopy class of x1x_{1}. By construction, ι(h)Out(Fm)\iota_{*}(h)\in\operatorname{Out}(F_{m}) is the class of the automorphism which conjugates by x1x_{1}. Therefore, ι(h1f)\iota_{*}(h^{-1}f) is the class of the automorphism which conjugates by ww^{\prime}. By our induction hypothesis, this implies that h1fh^{-1}f is trivial.

Claim.

The element hh is trivial.

Proof.

Let ΣMn,b\Sigma^{\prime}\subset M_{n,b} be a 2-sphere which separates Mn,bM_{n,b} into M1,1M_{1,1} and Mn1,b+1M_{n-1,b+1}, where the M1,1M_{1,1} is the handle containing x1x_{1}. Let λ:Out(F1,1)Out(Fn,b)\lambda_{*}:\mathrm{Out}(F_{1,1})\to\mathrm{Out}(F_{n,b}) be the map induced by this inclusion. Notice that h=λ(q)h=\lambda_{*}(q), where qOut(F1,1)q\in\mathrm{Out}(F_{1,1}) drags the boundary component of M1,1M_{1,1} about the nontrivial loop in the positive direction. We saw in the proof of Lemma A.1 that Out(F1,1)/2\mathrm{Out}(F_{1,1})\cong\mathbb{Z}/2, and the nontrivial element acts on π1(M1,1)\pi_{1}(M_{1,1}) by inversion. However, the element qq acts trivially on π1(M1,1)\pi_{1}(M_{1,1}), and is thus trivial itself. It follows that hh is trivial as well. ∎

Combining the claim with the fact that h1fh^{-1}f is trivial, we find that ff is trivial. This completes the induction, and so we conclude that ι\iota_{*} is injective.

Appendix B Realizing homology classes as spheres

In this section, we prove a result used in the proof of Lemma 7.3 which involves realizing bases of H2(Mn)H_{2}(M_{n}) as collections of 2-spheres. Recall that H2(Mn)=nH_{2}(M_{n})=\mathbb{Z}^{n}. This identification induces a homomorphism η:Mod(Mn)GLn()\eta:\operatorname{Mod}(M_{n})\to\operatorname{GL}_{n}(\mathbb{Z}) which takes a mapping class to its action on homology.

Lemma B.1.

The map η:Mod(Mn)GLn()\eta:\operatorname{Mod}(M_{n})\to\operatorname{GL}_{n}(\mathbb{Z}) is surjective.

Proof.

First, notice that H1(Mn)=nH^{1}(M_{n})=\mathbb{Z}^{n}. This identification also induces a homomorphism η:Mod(Mn)GLn()\eta^{\prime}:\operatorname{Mod}(M_{n})\to\operatorname{GL}_{n}(\mathbb{Z}) which is well-known to be surjective. Indeed, this map factors as

Mod(Mn)𝑞Out(Fn)𝜑GLn(),\operatorname{Mod}(M_{n})\overset{q}{\to}\operatorname{Out}(F_{n})\overset{\varphi}{\to}\operatorname{GL}_{n}(\mathbb{Z}),

where qq is the quotient map, and φ\varphi sends an automorphism class to its action on H1H^{1}. Therefore, if we choose our identifications H1(Mn)=nH^{1}(M_{n})=\mathbb{Z}^{n} and H2(Mn)=nH_{2}(M_{n})=\mathbb{Z}^{n} to agree with Poincaré duality, then η\eta and η\eta^{\prime} are the same map. Thus, η\eta is surjective. ∎

Lemma B.2.

Let {v1,,vn}\{v_{1},\ldots,v_{n}\} be a basis for H2(Mn)=nH_{2}(M_{n})=\mathbb{Z}^{n}, and let A={S1,,S}A=\{S_{1},\ldots,S_{\ell}\} be a collection of disjoint embedded oriented 2-spheres in MnM_{n} which satisfy [Sj]=vj[S_{j}]=v_{j} for 1j1\leq j\leq\ell. Then AA can be extended to a collection A¯={S1,,Sn}\overline{A}=\{S_{1},\ldots,S_{n}\} of disjoint embedded oriented 2-spheres such that [Sj]=vj[S_{j}]=v_{j} for 1jn1\leq j\leq n.

Proof.

We will induct on nn. The base case n=0n=0 is trivial. So assume n>0n>0, and let {v1,,vn}\{v_{1},\ldots,v_{n}\} and A={S1,,S}A=\{S_{1},\ldots,S_{\ell}\} be as stated. There are two cases.

First, suppose that =0\ell=0. If we identity H2(Mn)H_{2}(M_{n}) with n\mathbb{Z}^{n}, then by Lemma B.1 the resulting map η:Mod(Mn)GLn()\eta:\operatorname{Mod}(M_{n})\to\operatorname{GL}_{n}(\mathbb{Z}) is surjective. Choose any collection Σ1,,ΣnMn\Sigma_{1},\ldots,\Sigma_{n}\subset M_{n} of disjoint non-nullhomotopic embedded 2-spheres. Then {[Σ1],,[Σn]}\{[\Sigma_{1}],\ldots,[\Sigma_{n}]\} is a basis for H2(Mn)H_{2}(M_{n}). Since GLn()\operatorname{GL}_{n}(\mathbb{Z}) acts transitively on ordered bases of n\mathbb{Z}^{n} and the map η\eta is surjectve (Lemma B.1), there exists some 𝔣Mod(Mn)\mathfrak{f}\in\operatorname{Mod}(M_{n}) such that η(𝔣)[Σj]=vj\eta(\mathfrak{f})\cdot[\Sigma_{j}]=v_{j} for all 1jn1\leq j\leq n. In other words, [𝔣(Σj)]=vj[\mathfrak{f}(\Sigma_{j})]=v_{j}, and so {𝔣(Σ1),,𝔣(Σn)}\{\mathfrak{f}(\Sigma_{1}),\ldots,\mathfrak{f}(\Sigma_{n})\} is the desired collection of spheres.

Next, suppose that >0\ell>0. Splitting MnM_{n} along S1S_{1} gives an embedding ι:Mn1,2Mn\iota:M_{n-1,2}\hookrightarrow M_{n}. Notice that the induced map ιH:H2(Mn1,2)H2(Mn)\iota_{H}:H_{2}(M_{n-1,2})\to H_{2}(M_{n}) is an isomorphism. Let wj=ιH1(vj)w_{j}=\iota_{H}^{-1}(v_{j}) for 1jn1\leq j\leq n, and let \partial and \partial^{\prime} be the boundary components of Mn1,2M_{n-1,2}. Capping the two boundary components of Mn1,2M_{n-1,2} with disks DD and DD^{\prime}, we get another embedding ι:Mn1,2Mn1\iota^{\prime}:M_{n-1,2}\hookrightarrow M_{n-1}. This embedding induces a surjective map ιH:H2(Mn1,2)H2(Mn1)\iota_{H}^{\prime}:H_{2}(M_{n-1,2})\to H_{2}(M_{n-1}) whose kernel is generated by [][\partial]. Let wj=ιH(wj)w_{j}^{\prime}=\iota_{H}^{\prime}(w_{j}) for 2jn2\leq j\leq n, and let Sk=ι(Sk)S_{k}^{\prime}=\iota^{\prime}(S_{k}) for 2k2\leq k\leq\ell. By our induction hypothesis, we can extend the collection {S2,,S}\{S_{2}^{\prime},\ldots,S_{\ell}^{\prime}\} to a collection {S2,,Sn1}\{S_{2}^{\prime},\ldots,S_{n-1}^{\prime}\} of disjoint embedded oriented 2-spheres in Mn1M_{n-1} such that [Sj]=wj[S_{j}^{\prime}]=w_{j}^{\prime} for 2jn2\leq j\leq n. Moreover, since the disks DD and DD^{\prime} used to cap the boundary components of Mn1,2M_{n-1,2} are contractible, we may isotope S+1,,Sn1S_{\ell+1}^{\prime},\ldots,S_{n-1}^{\prime} such that they are disjoint from DD and DD^{\prime}. Let Sj=(ι)1(Sj)S_{j}=(\iota^{\prime})^{-1}(S_{j}^{\prime}) for +1jn\ell+1\leq j\leq n. If [Sk]=wk[S_{k}]=w_{k} for all kk, then {S1,,Sn}\{S_{1},\ldots,S_{n}\} is the desired collection, and we are done. However, since the kernel of ιH\iota_{H}^{\prime} is generated by [][\partial], we have

[Sk]=wk+ck[],[S_{k}]=w_{k}+c_{k}[\partial],

where ckc_{k}\in\mathbb{Z}. Note that ck=0c_{k}=0 for 2k2\leq k\leq\ell. To fix this, we may surger parallel copies of \partial or \partial^{\prime} onto SkS_{k} such that it has the correct homology class. The process is as follows (see Figure 15):

  1. (i)

    If ck>0c_{k}>0, take ckc_{k} parallel copies of \partial^{\prime}, which we denote by 1,,ck\partial_{1},\ldots,\partial_{c_{k}}. If instead ck<0c_{k}<0, take 1,,ck\partial_{1},\ldots,\partial_{c_{k}} to be parallel copies of \partial. Order the j\partial_{j} such that 1\partial_{1} is furthest from its respective boundary component, then 2\partial_{2}, and so on.

  2. (ii)

    Let γ1\gamma_{1} be a properly embedded arc connecting the positive side of SkS_{k} to 1\partial_{1} which does not intersect any of the other SjS_{j} or j\partial_{j}.

  3. (iii)

    Surger SkS_{k} and 1\partial_{1} together via a tube running along γ1\gamma_{1}.

  4. (iv)

    Repeat steps (ii) and (iii) for the remaining j\partial_{j}.

Once we have carried out this process for all the SkS_{k}, we will have obtained a collection collection {S2,,Sn}\{S_{2},\ldots,S_{n}\} of spheres whose homology classes are exactly w2,,wnw_{2},\ldots,w_{n}. Thus, {S1,,Sn}\{S_{1},\ldots,S_{n}\} is the desired collection of 2-spheres. ∎

\labellist\hair

2pt \pinlabelSkS_{k} at 135 110 \pinlabelγ1\gamma_{1} at 125 70 \pinlabel1\partial_{1} at 100 45

\pinlabel

γ2\gamma_{2} at 330 120 \pinlabelSkS_{k} at 330 80

\pinlabel

SkS_{k} at 530 80 \endlabellistRefer to caption

Figure 15: Surgering boundary spheres onto SkS_{k}.

References

  • [1] S. Andreadakis. On the automorphisms of free groups and free nilpotent groups. Proc. London Math. Soc. (3), 15:239–268, 1965.
  • [2] M. A. Armstrong. On the fundamental group of an orbit space. Proc. Cambridge Philos. Soc., 61:639–646, 1965.
  • [3] S. Bachmuth. Induced automorphisms of free groups and free metabelian groups. Trans. Amer. Math. Soc., 122:1–17, 1966.
  • [4] Jessica E. Banks. The Birman exact sequence for 3-manifolds. Bull. Lond. Math. Soc., 49(4):604–629, 2017.
  • [5] Mladen Bestvina, Kai-Uwe Bux, and Dan Margalit. Dimension of the Torelli group for Out(Fn){\rm Out}(F_{n}). Invent. Math., 170(1):1–32, 2007.
  • [6] Tara Brendle, Nathan Broaddus, and Andrew Putman. The mapping class group of connect sums of S2×S1S^{2}\times S^{1}. Preprint, 2020.
  • [7] Martin R. Bridson and Karen Vogtmann. Automorphism groups of free groups, surface groups and free abelian groups. In Problems on mapping class groups and related topics, volume 74 of Proc. Sympos. Pure Math., pages 301–316. Amer. Math. Soc., Providence, RI, 2006.
  • [8] Thomas Church, Benson Farb, and Andrew Putman. Integrality in the Steinberg module and the top-dimensional cohomology of SLn𝒪K{\rm SL}_{n}\mathcal{O}_{K}. Amer. J. Math., 141(5):1375–1419, 2019.
  • [9] Matthew Day and Andrew Putman. A Birman exact sequence for Aut(Fn){\rm Aut}(F_{n}). Adv. Math., 231(1):243–275, 2012.
  • [10] Matthew Day and Andrew Putman. The complex of partial bases for FnF_{n} and finite generation of the Torelli subgroup of Aut(Fn)\rm{Aut}(F_{n}). Geom. Dedicata, 164:139–153, 2013.
  • [11] Matthew Day and Andrew Putman. A Birman exact sequence for the Torelli subgroup of Aut(Fn)\rm{Aut}(F_{n}). Internat. J. Algebra Comput., 26(3):585–617, 2016.
  • [12] Benson Farb and Dan Margalit. A primer on mapping class groups, volume 49 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 2012.
  • [13] Allen Hatcher. Homological stability for automorphism groups of free groups. Comment. Math. Helv., 70(1):39–62, 1995.
  • [14] Allen Hatcher and Nathalie Wahl. Stabilization for the automorphisms of free groups with boundaries. Geom. Topol., 9:1295–1336, 2005.
  • [15] Allen Hatcher and Nathalie Wahl. Stabilization for mapping class groups of 3-manifolds. Duke Math. J., 155(2):205–269, 2010.
  • [16] Craig A. Jensen and Nathalie Wahl. Automorphisms of free groups with boundaries. Algebr. Geom. Topol., 4:543–569, 2004.
  • [17] Dennis Johnson. The structure of the Torelli group. III. The abelianization of 𝒯\mathcal{T}. Topology, 24(2):127–144, 1985.
  • [18] F. Laudenbach. Sur les 22-sphères d’une variété de dimension 33. Ann. of Math. (2), 97:57–81, 1973.
  • [19] François Laudenbach. Topologie de la dimension trois: homotopie et isotopie. Société Mathématique de France, Paris, 1974. With an English summary and table of contents, Astérisque, No. 12.
  • [20] Hendrik Maazen. Homology stability for the general linear group. thesis, 1951.
  • [21] Wilhelm Magnus. Über nn-dimensionale Gittertransformationen. Acta Math., 64(1):353–367, 1935.
  • [22] Andrew Putman. Cutting and pasting in the Torelli group. Geom. Topol., 11:829–865, 2007.
  • [23] Andrew Putman. Obtaining presentations from group actions without making choices. Algebr. Geom. Topol., 11(3):1737–1766, 2011.
  • [24] Andrew Putman. The commutator subgroups of free groups and surface groups. Preprint, 2021.
  • [25] Witold Tomaszewski. A basis of Bachmuth type in the commutator subgroup of a free group. Canad. Math. Bull., 46(2):299–303, 2003.
  • [26] Nathalie Wahl. From mapping class groups to automorphism groups of free groups. J. London Math. Soc. (2), 72(2):510–524, 2005.