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

On the finiteness of the classifying space of diffeomorphisms of reducible three manifolds

Sam Nariman Department of Mathematics
Purdue University
150 N. University Street
West Lafayette, IN 47907-2067
[email protected]
Abstract.

Kontsevich ([Kir95, Problem 3.48]) conjectured that BDiff(M,rel )\mathrm{BDiff}(M,\text{rel }\partial) has the homotopy type of a finite CW complex for all compact 33-manifolds with non-empty boundary. Hatcher-McCullough ([HM97]) proved this conjecture when MM is irreducible. We prove the homological version of Kontsevich’s conjecture to show that BDiff(M,rel )\mathrm{BDiff}(M,\text{rel }\partial) has finitely many nonzero homology groups each finitely generated when MM is a connected sum of irreducible 33-manifolds that each has a nontrivial boundary.

1. Introduction

For a closed surface Σg\Sigma_{g} of genus g>1g>1, it is well-known that the classifying space BDiff(Σg)\mathrm{BDiff}(\Sigma_{g}) is rationally equivalent to g\mathcal{M}_{g}, the moduli space of Riemann surfaces of genus gg. Therefore, in particular, the rational homology groups of BDiff(Σg)\mathrm{BDiff}(\Sigma_{g}) vanish above a certain degree, and in fact, more precisely they vanish above degree 4g54g-5, which is the virtual cohomological dimension of the mapping class group Mod(Σg)\text{Mod}(\Sigma_{g}). And for a surface Σg,k\Sigma_{g,k} with k>0k>0 boundary components, the classifying space BDiff(Σg,k,rel )\mathrm{BDiff}(\Sigma_{g,k},\text{rel }\partial) is in fact homotopy equivalent to the moduli space g,k\mathcal{M}_{g,k}. Therefore, BDiff(Σg,k,rel )\mathrm{BDiff}(\Sigma_{g,k},\text{rel }\partial) has the homotopy type of a finite-dimensional CW-complex.

Similarly, Kontsevich ([Kir95, Problem 3.48]) conjectured for compact 33-manifold MM with non-empty boundary, the classifying space BDiff(M,rel )\mathrm{BDiff}(M,\text{rel }\partial) has a finite-dimensional model. This conjecture is known to hold for irreducible 33-manifolds with non-empty boundary ([HM97]). In this paper, we shall prove the homological finiteness of these classifying spaces for reducible 33 manifolds with a condition on its boundary.

Throughout this paper, for brevity, we write Diff(M,rel )\mathrm{Diff}(M,\text{rel }\partial) and Homeo(M,rel )\text{{Homeo}}(M,\text{rel }\partial) to denote the smooth orientation preserving diffeomorphisms and orientation preserving homeomorphisms respectively whose supports are away from the boundary M\partial M and in general when we use rel X\text{rel }X in the diffeomorphism group, for some XMX\subset M, we mean those diffeomorphisms or homeomorphisms whose supports are away from XX.

Theorem 1.1.

Let MM be an orientable reducible 33-manifold that is a connected sum of irreducible 33-manifolds that each has a nontrivial boundary. Then the classifying space BDiff(M,rel )\mathrm{BDiff}(M,\text{rel }\partial) has finitely many nonzero homology groups which are each finitely generated.

In the irreducible case, the homotopy type of the group Diff(M)\mathrm{Diff}(M) is very well studied. When MM admits one of Thurston’s geometries, there has been an encompassing program known as the generalized Smale’s conjecture that relates the homotopy type of Diff(M)\mathrm{Diff}(M) with the isometry group of the corresponding geometry (for more details and history, see the discussions in Problem 3.47 in [Kir95] and Sections 1.2 and 1.3 in [HKMR12]). For 𝕊3\mathbb{S}^{3}, it was proved by Hatcher ([Hat83]), and for Haken 33-manifolds, it is a consequence of Hatcher’s work and also understanding the space of incompressible surfaces ([Wal68, Hat76, Iva76]) inside such manifolds. Recently Bamler and Kleiner ([BK19, BK21]) used Ricci flow techniques to settle the generalized Smale’s conjecture for all 33-manifolds admitting the spherical geometry and in the Nil geometry. Hence, this recent body of work using Ricci flow techniques addresses all cases of the generalized Smale’s conjecture.

Recall that a 33-manifold MM is called prime if it is not diffeomorphic to a connected sum of more than one 33-manifold so that none of which is diffeomorphic to the 33-sphere. The prime decomposition theorem says that every closed 33 manifold is diffeomorphic to the connected sum of prime manifolds. A prime closed 33-manifold is either diffeomorphic to 𝕊1×𝕊2\mathbb{S}^{1}\times\mathbb{S}^{2} or it is irreducible (i.e. every embedding 𝕊2\mathbb{S}^{2} bounds a ball). On the other hand, geometric manifolds are the building blocks for irreducible manifolds. Given the generalized Smale’s conjecture, we have a good understanding of the homotopy type of the diffeomorphism groups for these atomic pieces. And the JSJ and geometric decomposition theorems (see [Neu96, Chapter 2, section 6] for the statement of these theorems) give a way to cut an irreducible manifold along embedded tori into these building blocks. If the JSJ decomposition is non-trivial for an irreducible manifold, then it will be Haken whose diffeomorphism groups are well studied. Hence, given that we also know the homotopy type of the diffeomorphism group of 𝕊1×𝕊2\mathbb{S}^{1}\times\mathbb{S}^{2} by Hatcher’s theorem ([Hat81]), we have a complete understanding of the homotopy type of diffeomorphism group of prime manifolds. In the reducible case, the prime decomposition theorem cuts the manifold along separating spheres into its prime factors. The difficulty, however, in understanding the reducible case is to relate the diffeomorphism group of a reducible manifold to the diffeomorphisms of its prime factors.

César de Sá and Rourke ([CdSR79]) proposed to describe the homotopy type of Diff(M)\mathrm{Diff}(M) in terms of the homotopy type of diffeomorphisms of the prime factors and an extra factor of the loop space on “the space of prime decompositions". Hendriks-Laudenbach ([HL84]) and Hendriks-McCullough ([HM87]) found a model for this extra factor. Later Hatcher, in an interesting unpublished note, proposed a finite dimensional model for this “space of prime decompositions" and more interestingly, he proposed that there should be a “wrong-way map" between BDiff(M)\mathrm{BDiff}(M) and the classifying space of diffeomorphisms of prime factors.

Hatcher’s approach, if completed, would also solve Kontsevich’s conjecture in the special case of reducible 33 manifolds such that all the irreducible factors have non-empty boundaries. So our result is the homological version of what Hatcher intended to prove about Kontsevich’s conjecture. However, instead of trying to build this wrong way map, we take the geometric group theory approach by letting the abstract group of diffeomorphisms act on a “huge" simplicial complex inspired by the techniques that Kathryn Mann and the author ([MN20]) used to study the second homology of BDiff(M)\mathrm{BDiff}(M).

For technical simplicity, we work with the homeomorphism groups instead of diffeomorphism groups. The reason is that Cerf ([Cer61]) assumed Smale’s conjecture which was later proved by Hatcher ([Hat83]) to show that in these low dimensions, the inclusion Diff(M)Homeo(M)\mathrm{Diff}(M)\hookrightarrow\mathrm{Homeo}(M) is, in fact, a weak homotopy equivalence. On the other hand, in all dimensions, by Mather-Thurston’s theorem ([Thu74, Corollary (b) of theorem 5]) for homeomorphisms, we have the natural map

(1.1) BHomeoδ(M)BHomeo(M),\mathrm{B}\text{{Homeo}}^{\delta}(M)\to\mathrm{B}\text{{Homeo}}(M),

which is an acyclic map and in particular it induces a homology isomorphism in all degrees. The same statement also holds in the relative case in particular relative to the boundary when it is non-empty (see [McD80]).

Hence to prove the main theorem, we use a homological approach where we consider the action of Homeoδ(M,rel )\text{{Homeo}}^{\delta}(M,\text{rel }\partial) on a simplicial complex 𝒮(M)\mathcal{S}(M) given by the complex of essential spheres, to give a model for BHomeoδ(M,rel )\mathrm{B}\text{{Homeo}}^{\delta}(M,\text{rel }\partial) suitable for an inductive argument to prove the main theorem.

Acknowledgment

The author was partially supported by NSF grants DMS-2113828, NSF CAREER Grant DMS-2239106, a grant from the Simons Foundation (41000919, SN), and the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program (grant agreement No 682922). He thanks Sander Kupers and Andrea Bianchi for their comments on the first draft of this paper.

2. Proof strategy for an inductive argument

In this section we assume that MM is a compact reducible 33-manifold with a non-empty boundary. We assume that we do not have spherical boundary components in order to have a prime decomposition in the presence of the boundary ([Hem76, Chapter 3]). To induct on the number of its prime factors to study the homological finiteness of BHomeoδ(M,rel )\mathrm{B}\text{{Homeo}}^{\delta}(M,\text{rel }\partial), we shall first construct a simplicial complex 𝒮(M)\mathcal{S}(M) on which Homeoδ(M,rel )\text{{Homeo}}^{\delta}(M,\text{rel }\partial) acts simplicially.

Definition 2.1.

Let 𝒮(M)\mathcal{S}(M) be a simplicial complex whose vertices are given by locally flat embeddings ϕ:𝕊2M\phi\colon\mathbb{S}^{2}\hookrightarrow M whose images are essential spheres, and simplices are given by a collection of locally flat embeddings whose images are disjoint.

Remark 2.2.

Note that in a simplex there could be vertices that are given by isotopic spheres and since they are disjoint, they bound an embedded 𝕊2×[0,1]\mathbb{S}^{2}\times[0,1].

Note that the group Homeoδ(M,rel )\text{{Homeo}}^{\delta}(M,\text{rel }\partial) acts on 𝒮(M)\mathcal{S}(M) simplicially. We shall prove in 2.6 that 𝒮(M)\mathcal{S}(M) is contractible. Therefore, the homotopy quotient 𝒮(M)//Homeoδ(M,rel )\mathcal{S}(M)/\!\!/\text{{Homeo}}^{\delta}(M,\text{rel }\partial) is homotopy equivalent to BHomeoδ(M,rel )\mathrm{B}\text{{Homeo}}^{\delta}(M,\text{rel }\partial). The stabilizer of each simplex in 𝒮(M)\mathcal{S}(M) is the subgroup of Homeoδ(M,rel )\text{{Homeo}}^{\delta}(M,\text{rel }\partial) that fixes a set of essential spheres pointwise so it is isomorphic to diffeomorphism group of a 33-manifold whose connected components have fewer prime factors. But one issue is that 𝒮(M)\mathcal{S}(M) has simplices of arbitrary large dimensions since we allow parallel spheres. To account for this infinite dimensionality, we use the simplicial complex that Hatcher and McCullough defined in [HM90, Section 1].

Definition 2.3.

Let 𝒮nd(M)\mathcal{S}_{nd}(M) be the simplicial complex whose vertices are the isotopy classes of essential embedded spheres in MM and a set of vertices {[S0],[S1],,[Sn]}\{[S_{0}],[S_{1}],\dots,[S_{n}]\} constitutes an nn-simplex if there are pairwise disjoint embedded spheres SiS_{i}^{\prime} in MM such that for each ii, the sphere SiS_{i} is isotopic to SiS_{i}^{\prime}.

The mapping class group Mod(M,rel )=π0(Homeo(M,rel ))\text{Mod}(M,\text{rel }\partial)=\pi_{0}(\text{{Homeo}}(M,\text{rel }\partial)) acts on 𝒮nd(M)\mathcal{S}_{nd}(M) simplicially. The complex 𝒮nd(M)\mathcal{S}_{nd}(M) is finite-dimensional and also by Hatcher-McCullough’s theorem ([HM90, Proposition 2.2]) the set of the orbits of the action of Mod(M,rel )\text{Mod}(M,\text{rel }\partial) on simplices is also finite.

To briefly recall why this is the case, they use a theorem of Scharlemann (see [Bon83, Appendix A, Lemma A.1]) to find a “normal" representative of each orbit. Let the prime decomposition of MM be given by P1#P2##Pr#g𝕊1×𝕊2P_{1}\#P_{2}\#\cdots\#P_{r}\#^{g}\mathbb{S}^{1}\times\mathbb{S}^{2} where gg summands are diffeomorphic to 𝕊1×𝕊2\mathbb{S}^{1}\times\mathbb{S}^{2}. Let BB be a punctured 33-cell having ordered r+2gr+2g boundary components so that MM is obtained by gluing Pi\int(𝔻3)P_{i}\backslash\text{int}(\mathbb{D}^{3}) to ii-th sphere boundaries for 1ir1\leq i\leq r and gg copies of 𝕊2×[0,1]\mathbb{S}^{2}\times[0,1] are glues along the remaining 2g2g boundary components (see [Bon83, Appendix A, Lemma A.1] for more detail).

P1P_{1}P2P_{2}P3P_{3}BB
Figure 1. σ\sigma here is a 22-simplex consisting of 3 separating spheres that are drawn in one dimension lower.
Lemma 2.4 (Scharlemann).

For any simplex σ𝒮(M)\sigma\subset\mathcal{S}(M), there is a homeomorphism ff such that f(σ)Bf(\sigma)\subset B.

Now as Hatcher and McCullough observed in [HM90, Proposition 2.2], there are finitely many isotopy classes of essential spheres in BB since they are determined by the way they partition the boundary components of BB. And this observation implies the finiteness of the orbits of the action of Mod(M,rel )\text{Mod}(M,\text{rel }\partial) on simplices of 𝒮nd(M)\mathcal{S}_{nd}(M).

The skeletal filtration on 𝒮nd(M)\mathcal{S}_{nd}(M) induces a filtration on the quotient space

(2.1) 01n=𝒮nd(M)/Mod(M,rel ),\mathcal{F}_{0}\subset\mathcal{F}_{1}\subset\dots\subset\mathcal{F}_{n}=\mathcal{S}_{nd}(M)/\text{Mod}(M,\text{rel }\partial),

and by Hatcher and McCullough’s observation, the filtration quotients are given by the wedge of a finite number of spheres. Let OpO_{p} be the set of orbits of the action of Mod(M,rel )\text{Mod}(M,\text{rel }\partial) on pp-simplices of 𝒮nd(M)\mathcal{S}_{nd}(M). So pp1\mathcal{F}_{p}-\mathcal{F}_{p-1} is homeomorphic to the disjoint union of open pp-simplices indexed by OpO_{p}.

Now there is a natural simplicial map 𝒮(M)𝒮nd(M)\mathcal{S}(M)\to\mathcal{S}_{nd}(M) which is equivariant with respect to the map Homeoδ(M,rel )Mod(M,rel )\text{{Homeo}}^{\delta}(M,\text{rel }\partial)\to\text{Mod}(M,\text{rel }\partial). So we have a map 𝒮(M)/Homeoδ(M,rel )𝒮nd(M)/Mod(M,rel )\mathcal{S}(M)/\text{{Homeo}}^{\delta}(M,\text{rel }\partial)\to\mathcal{S}_{nd}(M)/\text{Mod}(M,\text{rel }\partial) which in turn induces a map

η:𝒮(M)//Homeoδ(M,rel )𝒮nd(M)/Mod(M,rel ).\eta\colon\mathcal{S}(M)/\!\!/\text{{Homeo}}^{\delta}(M,\text{rel }\partial)\to\mathcal{S}_{nd}(M)/\text{Mod}(M,\text{rel }\partial).

The preimage filtration 𝒢p=η1(i)\mathcal{G}_{p}=\eta^{-1}(\mathcal{F}_{i}) on 𝒮(M)//Homeoδ(M,rel )\mathcal{S}(M)/\!\!/\text{{Homeo}}^{\delta}(M,\text{rel }\partial) induces a spectral sequence

Ep,q1=Hp+q(𝒢p/𝒢p1;)Hp+q(BHomeoδ(M,rel );).E^{1}_{p,q}=H_{p+q}(\mathcal{G}_{p}/\mathcal{G}_{p-1};\mathbb{Z})\Longrightarrow H_{p+q}(\mathrm{B}\text{{Homeo}}^{\delta}(M,\text{rel }\partial);\mathbb{Z}).

To prove the classifying space BHomeoδ(M,rel )\mathrm{B}\text{{Homeo}}^{\delta}(M,\text{rel }\partial) has finitely many nonzero homology groups which are each finitely generated, it is enough to prove the same for the filtration quotients 𝒢p/𝒢p1\mathcal{G}_{p}/\mathcal{G}_{p-1}. And this is equivalent to the following theorem.

Theorem 2.5.

Let us identify pp1\mathcal{F}_{p}-\mathcal{F}_{p-1} with the finite disjoint union of open simplices σOpΔ˙σp\coprod_{\sigma\in O_{p}}\dot{\Delta}^{p}_{\sigma}. Then for each σOp\sigma\in O_{p}, as xx varies in Δ˙σp\dot{\Delta}^{p}_{\sigma}, the homotopy type of η1(x)\eta^{-1}(x) does not change and its homology groups are finitely generated and concentrated in finitely many degrees.

The filtration 𝒢i\mathcal{G}_{i}’s are sub-CW-complexes of realization of the semisimplicial set given by the bar construction that realizes to the homotopy quotient 𝒮(M)//Homeoδ(M,rel )\mathcal{S}(M)/\!\!/\text{{Homeo}}^{\delta}(M,\text{rel }\partial). Hence, the inclusions 𝒢p1𝒢p\mathcal{G}_{p-1}\hookrightarrow\mathcal{G}_{p} are cofibrations. On the other hand, given 2.5, the homotopy type of η1(x)\eta^{-1}(x) does not change as xx varies in each open simplex Δ˙σp\dot{\Delta}^{p}_{\sigma}. So Ep,q1E^{1}_{p,q} in the spectral sequence is isomorphic to

σOpHp+q((Δp/Δp)η1(bσ);),\bigoplus_{\sigma\in O_{p}}H_{p+q}((\Delta^{p}/\partial\Delta^{p})\wedge\eta^{-1}(b_{\sigma});\mathbb{Z}),

where bσb_{\sigma} is the barycenter of the open simplex Δ˙σp\dot{\Delta}^{p}_{\sigma}. Hence, given 2.5, the homology groups of the filtration quotients 𝒢p/𝒢p1\mathcal{G}_{p}/\mathcal{G}_{p-1} are finitely generated and concentrated in finitely many degrees which implies 1.1.

In the following sections, we find a model for η1(x)\eta^{-1}(x) to which we can apply the induction hypothesis (i.e. the homological finiteness of BHomeo(M,rel )\mathrm{B}\text{{Homeo}}(M,\text{rel }\partial) for MM with fewer prime factors) when MM is connected sum of irreducible factors that each has a non-trivial boundary. And we end this section with the proof of the contractibility of 𝒮(M)\mathcal{S}(M).

Proposition 2.6.

The simplicial complex 𝒮(M)\mathcal{S}(M) is contractible.

In fact, it is a weakly Cohen-Macaulay complex of dimension infinity. Recall that a simplicial complex XX is called weakly Cohen-Macaulay of dimension nn if it is (n1)(n-1)-connected and the link of any pp-simplex is (np2)(n-p-2)-connected. In this case, we denote this property by wCM(X)n\text{wCM}(X)\geq n (see [HW10, Definition 3.4]).

Proof.

It is enough to show that for each kk, any continuous map f:Sk𝒮(M)f\colon S^{k}\to\mathcal{S}(M) is nullhomotopic. Note that by the simplicial approximation theorem, we can assume that ff is a PL map concerning a triangulation KK of SkS^{k} and we shall change ff up to a simplicial homotopy such that there exists a vertex vv in 𝒮(M)\mathcal{S}(M), that cones off f(Sk)f(S^{k}) in 𝒮(M)\mathcal{S}(M).

It is easy to modify the map ff via a simplicial homotopy given by moving vertices to their parallel copies to make sure that the images of vertices of f(K)f(K) are pairwise transverse 111One can alternatively change ff up to simplicial homotopy to replace each vertex f(x)f(x) by a smooth nearby parallel copy and use transversality in the smooth category. (see [Nar20, Lemma 3.31 and Lemma 4.3] for treating transversality in the locally flat settings). Now we choose a vertex vv in 𝒮(M)\mathcal{S}(M) so that as an embedded sphere, it is transverse to all vertices in f(K)f(K). The intersection of spheres in f(K)f(K) with vv gives a collection of circles on the sphere given by vv. From this collection, choose a maximal family of disjoint circles, and let CC be the innermost circle in this family. Then CC is given by the intersection of the sphere vv and a sphere w=f(x)w=f(x) in f(K)f(K). Note that by the innermost circle, we mean that CC bounds a disk DD on vv whose interior is disjoint from all spheres in the maximal collection. We cut ww along cc and glue two nearby parallel copies of the disk DD to obtain two disjoint embedded spheres ww^{\prime} and w′′w^{\prime\prime} (see Figure 2).

wwww^{\prime}w′′w^{\prime\prime}vv
Figure 2. Surgery on spheres in one dimension lower

We can arrange this so that ww, ww^{\prime}, and w′′w^{\prime\prime} are disjoint spheres. Since ww is an essential sphere, ww^{\prime} and w′′w^{\prime\prime} both cannot bound a ball so let us assume that ww^{\prime} is essential and we give it an arbitrary parametrization to consider it as a vertex in 𝒮(M)\mathcal{S}(M). Let zz be a vertex in the Star(w)\text{Star}(w) where star here means as a subcomplex of f(K)f(K). The corresponding sphere for zz cannot intersect ww^{\prime}, since if it does intersect ww^{\prime}, it has to intersect DD, which contradicts that DD is innermost on vv. Hence, vertices in Star(w)\text{Star}(w) represent spheres that are disjoint from the representative sphere of ww^{\prime}. So we can define a simplicial homotopy G:K×[0,1]𝒮(M)G\colon K\times[0,1]\to\mathcal{S}(M) such that G(,1)G(-,1) is the same as f()f(-) on all vertices but xx and G(x,1)=wG(x,1)=w^{\prime}. Note that the vertices in the image of G(,1):K𝒮(M)G(-,1)\colon K\to\mathcal{S}(M) have fewer numbers of circles in their intersection with vv. By repeating this process to reduce the number of circles in the intersection of spheres in f(K)f(K) with vv, we could homotope the map ff to a map whose image lies in the star of vv. Therefore, ff is nullhomotopic. ∎

Remark 2.7.

To prove the weak Cohen-Macaulay condition, we need to prove that the links of simplices in 𝒮(M)\mathcal{S}(M) are contractible. To do so, it is easier to consider auxiliary complexes. Let RR be a collection of boundary components that is diffeomorphic to the union of spheres. Let 𝒮(M,R)\mathcal{S}(M,R) be a complex whose vertices are given by embedded essential separating spheres that could be parallel to components of RR and whose simplices are given by a set of such spheres that are disjoint. If RR is empty, this complex is the same as 𝒮(M)\mathcal{S}(M), and if RR is non-empty, then the complex 𝒮(M,R)\mathcal{S}(M,R) is still contractible. In fact, the proof, in this case, is easier since for any PL map f:K𝒮(M,R)f\colon K\to\mathcal{S}(M,R) where KK is a finite complex, we can find an embedded sphere vv that is sufficiently close to a component of RR which is disjoint from all vertices in f(K)f(K). Hence, the vertex vv cones off f(K)f(K) which implies that ff is nullhomotopic.

Now let σ\sigma be a simplex in 𝒮(M)\mathcal{S}(M) consisting of spheres {S0,S1,,Sp}\{S_{0},S_{1},\dots,S_{p}\}. Then the link of σ\sigma is the join of complexes 𝒮(Mi,Ri)\mathcal{S}(M_{i},R_{i}) where MiM_{i}’s are the components obtained by cutting MM along spheres in σ\sigma and RiR_{i}’s are non-empty. Therefore, the link of σ\sigma is contractible.

3. Parallel spheres and bar constructions

In this section, we shall use the hypothesis that the prime decomposition of MM consists of irreducible factors that each has non-empty boundary. The goal is for each pp and each xpp1x\in\mathcal{F}_{p}-\mathcal{F}_{p-1} to find a semi-simplicial space XX_{\bullet} whose realization is homology equivalent to η1(x)\eta^{-1}(x) and sits in a fibration sequence so that by induction on the number of prime factors we could argue that the fiber and the base have finite CW complex model.

The advantage of working with 33 manifolds that are connected sum of irreducible pieces such that each has a non-empty boundary is:

  • When we cut along essential separating spheres, the remaining pieces each has a non-spherical boundary component that is fixed and we shall use this for the inductive argument.

  • Since each irreducible factor has a non-trivial boundary that is fixed, homeomorphic irreducible factors cannot be permuted under the action of Homeo(M,rel )\text{{Homeo}}(M,\text{rel }\partial).

To start, let xx be in 0\mathcal{F}_{0} in the filtration 2.1 which is the image of a separatings sphere SMS\subset M. Let 𝒮(M,[S])\mathcal{S}(M,[S]) be the full subcomplex of 𝒮(M)\mathcal{S}(M) whose vertices are the orbits of SS under the action of Homeoδ(M,rel )\text{{Homeo}}^{\delta}(M,\text{rel }\partial). Note that the preimage of η1(x)\eta^{-1}(x) is homotopy equivalent to

𝒮(M,[S])//Homeoδ(M,rel ).\mathcal{S}(M,[S])/\!\!/\text{{Homeo}}^{\delta}(M,\text{rel }\partial).

Suppose that SS cuts the manifold MM into two pieces M1M_{1} and M2M_{2} where M1M_{1} contains M\partial_{*}M, the boundary component of MM with the base point. . And let Homeo(M1,S,rel M)\text{{Homeo}}(M_{1},S,\text{rel }\partial M) be the subgroup of Homeo(M1)\text{{Homeo}}(M_{1}) that fixes the boundary component SS set-wise and the rest of the boundary components pointwise. Towards our goal in this section, we prove that η1(x)\eta^{-1}(x) is homology equivalent to the realization of a semi-simplicial space (in fact a two-sided bar construction) XX_{\bullet} that fits in a fibration

(3.1) BHomeo(M2,rel )XBHomeo(M1,S,rel M).\mathrm{B}\text{{Homeo}}(M_{2},\text{rel }\partial)\to||X_{\bullet}||\to\mathrm{B}\text{{Homeo}}(M_{1},S,\text{rel }\partial M).

By induction the fiber BHomeo(M2,rel )\mathrm{B}\text{{Homeo}}(M_{2},\text{rel }\partial) has a finite CW complex model and in 3.9, we show that the base also has a finite CW complex model. First, we note that there is a natural order on vertices of each simplex in 𝒮(M,[S])\mathcal{S}(M,[S]).

Lemma 3.1.

When MM is a compact 33-manifold with a non-empty boundary and its prime decomposition of MM consists of irreducible factors that each have a non-empty boundary, then vertices in each simplex in 𝒮(M,[S])\mathcal{S}(M,[S]) have a natural order.

Definition 3.2.

Let 𝒮(M,[S])\mathcal{S}_{\bullet}(M,[S]) denote the semisimplicial set given by this ordering of vertices.

Proof of 3.1.

Since the prime decomposition of MM has irreducible factors with non-empty boundaries, an edge in 𝒮(M,[S])\mathcal{S}(M,[S]) consists of two disjoint isotopic spheres in the orbit of SS. This is because if we had two disjoint non-isotopic spheres in the orbit of SS, given that SS is separating, these two spheres cut out homeomorphic pieces that are permuted by an element in Homeo(M,rel )\text{{Homeo}}(M,\text{rel }\partial) which contradicts the hypothesis. Hence each simplex in 𝒮(M,[S])\mathcal{S}(M,[S]) consists of disjoint isotopic spheres. We call them parallel spheres.

Now we show that there is an induced order on parallel separating spheres when the manifold MM has a non-empty boundary. To describe this a priori order on parallel spheres, we choose once and for all a base point on one of the boundary components. We denote this boundary component by M\partial_{*}M. Each separating sphere SS separates MM into connected components and one of them PSP_{S} contains the base point. If we have isotopic disjoint separating spheres SiS_{i}’s, we order them by the inclusion of the components PSiP_{S_{i}}’s. In other words, we can put a metric on MM and order SiS_{i}’s by their distance to the base point. We call this order from inside to outside direction. ∎

Now we define a semi-simplicial space whose underlying semi-simplicial set is 𝒮(M,[S])\mathcal{S}_{\bullet}(M,[S]).

Definition 3.3.

𝒮τ(M,[S])\mathcal{S}^{\tau}_{\bullet}(M,[S]) is a semi-simplicial space whose 0-simplices as a set is the same as 𝒮0(M,[S])\mathcal{S}_{0}(M,[S]) but it is topologized as the subspace of locally flat embeddings Emb𝗅𝖿(𝕊2,M)\text{Emb}^{\sf{lf}}(\mathbb{S}^{2},M). And for each p>0p>0 the space 𝒮pτ(M,[S])\mathcal{S}^{\tau}_{p}(M,[S]) as a set is the same as 𝒮p(M,[S])\mathcal{S}_{p}(M,[S]) but it is topologized as a subspace of 𝒮0τ(M,[S])p+1\mathcal{S}^{\tau}_{0}(M,[S])^{p+1}.

Since the action of Homeoδ(M,rel )\text{{Homeo}}^{\delta}(M,\text{rel }\partial) on the set of pp-simplices 𝒮p(M,[S])\mathcal{S}_{p}(M,[S]) is transitive, it is easy to use Thurston’s homology isomorphism 1.1 to obtain that for each pp, the natural map

𝒮p(M,[S])//Homeoδ(M,rel )𝒮pτ(M,[S])//Homeo(M,rel ),\mathcal{S}_{p}(M,[S])/\!\!/\text{{Homeo}}^{\delta}(M,\text{rel }\partial)\to\mathcal{S}^{\tau}_{p}(M,[S])/\!\!/\text{{Homeo}}(M,\text{rel }\partial),

induces a homology isomorphism. Therefore, by the spectral sequence that calculates the homology of the realization (see [ERW19, Section 1.4]), we have a homology isomorphism between fat realizations

||𝒮(M,[S])//Homeoδ(M,rel )||||𝒮τ(M,[S])//Homeo(M,rel )||||\mathcal{S}_{\bullet}(M,[S])/\!\!/\text{{Homeo}}^{\delta}(M,\text{rel }\partial)||\to||\mathcal{S}^{\tau}_{\bullet}(M,[S])/\!\!/\text{{Homeo}}(M,\text{rel }\partial)||

Hence, to prove homological finiteness for η1(x)\eta^{-1}(x), we find a model for

||𝒮τ(M,[S])//Homeo(M,rel )||,||\mathcal{S}^{\tau}_{\bullet}(M,[S])/\!\!/\text{{Homeo}}(M,\text{rel }\partial)||,

that sits in a fibration 3.1.

We can define the smooth version of 𝒮τ(M,[S])\mathcal{S}^{\tau}_{\bullet}(M,[S]) and work with diffeomorphism groups. But in this dimension and for codimension 11 embeddings, the corresponding objects in the C0C^{0} and CC^{\infty}-category are weakly homotopy equivalent. So we stick to C0C^{0}-category.

3.A. Moduli spaces of manifold models

Let e:M{0}×e\colon\partial_{*}M\hookrightarrow\{0\}\times\mathbb{R}^{\infty} be a fixed locally flat embedding of the boundary component that contains the base point and let Emb𝗅𝖿(M,[0,)×)\text{Emb}^{\sf{lf}}_{\partial}(M,[0,\infty)\times\mathbb{R}^{\infty}) be the space of locally flat embeddings of MM whose intersection with {0}×\{0\}\times\mathbb{R}^{\infty} is e(M)e(\partial_{*}M). By [Las76, Appendix, theorem 1] and [Kup15, Lemma 2.2], the space Emb𝗅𝖿(M,[0,)×)\text{Emb}^{\sf{lf}}_{\partial}(M,[0,\infty)\times\mathbb{R}^{\infty}) is weakly contractible. Hence, the semi-simplicial space

(M,[S])\coloneqq𝒮τ(M,[S])×Emb𝗅𝖿(M,[0,)×)Homeo(M,rel ),\mathcal{M}_{\bullet}(M,[S])\coloneqq\frac{\mathcal{S}^{\tau}_{\bullet}(M,[S])\times\text{Emb}^{\sf{lf}}_{\partial}(M,[0,\infty)\times\mathbb{R}^{\infty})}{\text{{Homeo}}(M,\text{rel }\partial)},

is level-wise weakly equivalent to 𝒮τ(M,[S])//Homeo(M,rel )\mathcal{S}^{\tau}_{\bullet}(M,[S])/\!\!/\text{{Homeo}}(M,\text{rel }\partial) and we think of (M,[S])\mathcal{M}_{\bullet}(M,[S]) as a configuration space of the manifolds in [0,)×[0,\infty)\times\mathbb{R}^{\infty} that are homeomorphic to MM satisfying the boundary condition and with a choice of parallel spheres in the orbit of SS.

Now we shall define a two-sided bar construction model for (M,[S])\mathcal{M}_{\bullet}(M,[S]). Let ι0:𝕊2{0}×\iota_{0}\colon\mathbb{S}^{2}\hookrightarrow\{0\}\times\mathbb{R}^{\infty} be a fixed embedding and we denote the embedding ι0+te1\iota_{0}+t\cdot e_{1} in {t}×\{t\}\times\mathbb{R}^{\infty} by ιt\iota_{t}.

Definition 3.4.

Let BD\mathrm{B}\mathrm{D} be the topological monoid given by space of pairs (t,f)[0,)×Emb𝗅𝖿(𝕊2×[0,1],[0,)×)(t,f)\in[0,\infty)\times\text{Emb}^{\sf{lf}}_{\partial}(\mathbb{S}^{2}\times[0,1],[0,\infty)\times\mathbb{R}^{\infty}) where f(𝕊2×[0,1])[0,t]×f(\mathbb{S}^{2}\times[0,1])\subset[0,t]\times\mathbb{R}^{\infty} and the restriction of ff to 𝕊2×{0}\mathbb{S}^{2}\times\{0\} and 𝕊2×{1}\mathbb{S}^{2}\times\{1\} are given by embeddings ι0\iota_{0} and ιt\iota_{t} respectively. The monoid structure is given by adding the tt-coordinates and stacking the embeddings next to each other.

It is standard to see that the topological monoid BD\mathrm{B}\mathrm{D} is homotopy equivalent to BHomeo(𝕊2×[0,1],rel )\mathrm{B}\text{{Homeo}}(\mathbb{S}^{2}\times[0,1],\text{rel }\partial). But since the homotopy type of Homeo(𝕊2×[0,1],rel )\text{{Homeo}}(\mathbb{S}^{2}\times[0,1],\text{rel }\partial) is known ([Hat83, Appendix]) to be the loop space Ω(SO(3))\Omega(\text{SO}(3)), we can also determine the homotopy type of the delooping BBD\mathrm{B}\mathrm{B}\mathrm{D}.

Lemma 3.5.

The space BBD\mathrm{B}\mathrm{B}\mathrm{D} which is the classifying space of the topological monoid BD\mathrm{B}\mathrm{D} is homotopy equivalent to BSO(3)\mathrm{B}\text{SO}(3).

Proof.

It is easier to see this by describing BBD\mathrm{B}\mathrm{B}\mathrm{D} as the realization of a bi-semi-simplicial space and then realize it in two different simplicial directions. Consider the standard embedding of [0,)×𝕊2[0,)×[0,\infty)\times\mathbb{S}^{2}\hookrightarrow[0,\infty)\times\mathbb{R}^{\infty}. Let D\mathrm{D} be the subspace of pairs

(t,f)[0,)×Homeoc([0,)×𝕊2),(t,f)\in[0,\infty)\times\text{{Homeo}}_{c}([0,\infty)\times\mathbb{S}^{2}),

such that supp(f)[0,t]×𝕊2\text{supp}(f)\subset[0,t]\times\mathbb{S}^{2}. The multiplication of (t,f)(t,f)=(t+t,ff)(t,f)\cdot(t^{\prime},f^{\prime})=(t+t^{\prime},f\sqcup f^{\prime}) where fff\sqcup f^{\prime} is in Homeo([0,t+t]×𝕊2,rel )\text{{Homeo}}([0,t+t^{\prime}]\times\mathbb{S}^{2},\text{rel }\partial) given by concatenating ff and ff^{\prime}. Let D\mathrm{D}_{\bullet} be a simplicial monoid where the space of kk-simplices is given by the tuples (t,f1,,fk)(t,f_{1},\dots,f_{k}) where supp(fi)[0,t]×𝕊2\text{supp}(f_{i})\subset[0,t]\times\mathbb{S}^{2} for all ii. The face maps are given by the composition of homeomorphisms and the degeneracies are given by inserting the identity. It is easy to see that the realization D||\mathrm{D}_{\bullet}|| is a topological monoid homotopy equivalent to BD\mathrm{B}\mathrm{D}.

Let ΩtSO(3)\Omega_{t}\text{SO}(3) be the Moore monoid model for the space of loops on SO(3)\text{SO}(3). This is a submonoid of D\mathrm{D} by sending a loop f:[0,t]SO(3)f\colon[0,t]\to\text{SO}(3) to the homeomorphism of [0,t]×𝕊2[0,t]\times\mathbb{S}^{2} that sends (s,x)(s,x) to (s,f(s)(x))(s,f(s)(x)). This inclusion is a weak equivalence and also respects the composition of homeomorphism groups. Hence, if we define the simplicial monoid ΩtSO(3)\Omega_{t}\text{SO}(3)_{\bullet} similar to D\mathrm{D}_{\bullet}, then its nerve is a bi-simplicial space that can be realized in both directions. The homotopy type of the realization of this bi-simplicial space is the same as BBD\mathrm{B}\mathrm{B}\mathrm{D}. So if we realize ΩtSO(3)\Omega_{t}\text{SO}(3)_{\bullet} in the monoid direction first and then in the simplicial direction, we obtain the homotopy type of BSO(3)\mathrm{B}\text{SO}(3). ∎

Recall that when we cut MM along SS, we obtain two pieces M1M_{1} and M2M_{2} where M1M_{1} contains M\partial_{*}M, the boundary component of MM with the base point. Now we define moduli space models for BHomeo(M1,rel )\mathrm{B}\text{{Homeo}}(M_{1},\text{rel }\partial) and BHomeo(M2,rel )\mathrm{B}\text{{Homeo}}(M_{2},\text{rel }\partial) that are modules over the topological monoid BD\mathrm{B}\mathrm{D}.

Definition 3.6.

Let BL\mathrm{B}\mathrm{L} be the space of pairs (t,f)[0,)×Emb𝗅𝖿(M1,[0,)×)(t,f)\in[0,\infty)\times\text{Emb}^{\sf{lf}}_{\partial}(M_{1},[0,\infty)\times\mathbb{R}^{\infty}) such that

  • The image f(M1)f(M_{1}) lies in the strip [0,t]×[0,t]\times\mathbb{R}^{\infty}.

  • The intersection f(M1){0}×f(M_{1})\cap\{0\}\times\mathbb{R}^{\infty} is given by the embedding ee and f(M1){t}×f(M_{1})\cap\{t\}\times\mathbb{R}^{\infty} is given by ιt\iota_{t}.

And similarly, let BR\mathrm{B}\mathrm{R} to be the space of pairs (t,f)[0,)×Emb𝗅𝖿(M2,[0,)×)(t,f)\in[0,\infty)\times\text{Emb}^{\sf{lf}}_{\partial}(M_{2},[0,\infty)\times\mathbb{R}^{\infty}) such that

  • The image f(M2)f(M_{2}) lies in [t,)×[t,\infty)\times\mathbb{R}^{\infty}.

  • The intersection f(M2){t}×f(M_{2})\cap\{t\}\times\mathbb{R}^{\infty} is given by ιt\iota_{t}.

It is easy to see that BL\mathrm{B}\mathrm{L} and BR\mathrm{B}\mathrm{R} are weakly equivalent to BHomeo(M1,rel )\mathrm{B}\text{{Homeo}}(M_{1},\text{rel }\partial) and BHomeo(M2,rel )\mathrm{B}\text{{Homeo}}(M_{2},\text{rel }\partial) respectively.

Note that there is a right BD\mathrm{B}\mathrm{D}-module structure on BL\mathrm{B}\mathrm{L} such that the action of (t,f)BD(t,f)\in\mathrm{B}\mathrm{D} on (t,f)BL(t^{\prime},f^{\prime})\in\mathrm{B}\mathrm{L} is the pair (t+t,f(f+te1))(t+t^{\prime},f^{\prime}\sqcup(f+t^{\prime}\cdot e_{1})) where f+te1f+t^{\prime}\cdot e_{1} is the embedding ff shifted in the first coordinate to the right by tt^{\prime}. Similarly, there is a left BD\mathrm{B}\mathrm{D}-module structure on BR\mathrm{B}\mathrm{R}.

{a}×\{a\}\times\mathbb{R}^{\infty}{b}×\{b\}\times\mathbb{R}^{\infty}{0}×\{0\}\times\mathbb{R}^{\infty}[0,1]×𝕊2[0,1]\times\mathbb{S}^{2}M1M_{1}

\cong

\cong

\cong

M2M_{2}
Figure 3. Schematic picture in one dimension lower on how BD\mathrm{BD} acts on BR\mathrm{BR} and BL\mathrm{BL}

We consider the two-sided bar resolution given by the semi-simplicial space

Bp(BL,BD,BR)=BL×BDp×BRB_{p}(\mathrm{B}\mathrm{L},\mathrm{B}\mathrm{D},\mathrm{B}\mathrm{R})=\mathrm{B}\mathrm{L}\times\mathrm{B}\mathrm{D}^{p}\times\mathrm{B}\mathrm{R}

where the face map d0d_{0} and dpd_{p} are given by the actions of BD\mathrm{B}\mathrm{D} on BL\mathrm{B}\mathrm{L} and BR\mathrm{B}\mathrm{R} respectively and other face maps are induced by the monoid structure of BD\mathrm{B}\mathrm{D}.

Note that there is a natural semi-simplicial map

hp:Bp(BL,BD,BR)p(M,[S])h_{p}\colon B_{p}(\mathrm{B}\mathrm{L},\mathrm{B}\mathrm{D},\mathrm{B}\mathrm{R})\to\mathcal{M}_{p}(M,[S])

by gluing the embeddings in order and choosing the spheres along which the embeddings are glued as a choice of parallel spheres in the orbit of SS. On the other hand, recall that the action of Homeo(M,rel )\text{{Homeo}}(M,\text{rel }\partial) on 𝒮pτ(M,[S])\mathcal{S}^{\tau}_{p}(M,[S]) is transitive for each pp. So for a pp-simplex σ𝒮pτ(M,[S])\sigma\in\mathcal{S}^{\tau}_{p}(M,[S]), the homotopy quotient 𝒮pτ(M,[S])//Homeo(M,rel )\mathcal{S}^{\tau}_{p}(M,[S])/\!\!/\text{{Homeo}}(M,\text{rel }\partial) is weakly equivalent to BStab(σ)\mathrm{B}\text{Stab}(\sigma). Hence, the semi-simplicial map hh_{\bullet} is level-wise a weak equivalence, and we have weak equivalences between the (fat) realizations

||B(BL,BD,BR)||||(M,[S])||||𝒮τ(M,[S])//Homeo(M,rel )||.||B_{\bullet}(\mathrm{B}\mathrm{L},\mathrm{B}\mathrm{D},\mathrm{B}\mathrm{R})||\xrightarrow{\simeq}||\mathcal{M}_{\bullet}(M,[S])||\simeq||\mathcal{S}^{\tau}_{\bullet}(M,[S])/\!\!/\text{{Homeo}}(M,\text{rel }\partial)||.

Note that we have a fibration

(3.2) BRB(BL,BD,BR)B(BL,BD,).\mathrm{B}\mathrm{R}\to||B_{\bullet}(\mathrm{B}\mathrm{L},\mathrm{B}\mathrm{D},\mathrm{B}\mathrm{R})||\to||B_{\bullet}(\mathrm{B}\mathrm{L},\mathrm{B}\mathrm{D},*)||.

We shall use the technique of the Weiss fibration as was explained in [Kup19] to show that this is the desired fibration 3.1.

3.B. Kupers’ bar resolution for self-embeddings

We shall use Kupers’ theorem ([Kup19, Section 4]) to determine the homotopy type of B(BL,BD,)||B_{\bullet}(\mathrm{B}\mathrm{L},\mathrm{B}\mathrm{D},*)|| to be able to say that has finite CW complex model.

The manifold M1M_{1} has a sphere boundary component 0M1=S\partial_{0}M_{1}=S which we call the free boundary component and we denote the union of the rest of the boundary components by 1M1\partial_{1}M_{1} which we call the fixed boundary components. Let Homeo(M1,S,rel 1)\text{{Homeo}}(M_{1},S,\text{rel }\partial_{1}) be the group homeomorphisms of M1M_{1} that fix SS set-wise and fix 1M1\partial_{1}M_{1} point-wise.

Theorem 3.7.

There is a zig-zag of weak equivalences between the bar resolution B(BL,BD,)||B_{\bullet}(\mathrm{B}\mathrm{L},\mathrm{B}\mathrm{D},*)|| and the classifying space BHomeo(M1,S,rel 1)\mathrm{B}\text{{Homeo}}(M_{1},S,\text{rel }\partial_{1}).

Kupers in [Kup19] gives a model for Weiss fiber sequence where the set-up is we have an nn-dimensional manifold MM with a non-empty boundary and we fix an embedded 𝔻n1M\mathbb{D}^{n-1}\hookrightarrow\partial M. Let Emb1/2(M)\text{Emb}^{\cong}_{1/2\partial}(M) be the space of self-embeddings of MM that are identity on M\int(𝔻n1)\partial M\backslash\text{int}(\mathbb{D}^{n-1}) and are isotopic to a diffeomorphism that fixes the boundary through isotopies fixing M\int(𝔻n1)\partial M\backslash\text{int}(\mathbb{D}^{n-1}). There exists a fiber sequence named after Michael Weiss

BDiff(M,rel )BEmb1/2(M)BBDiff(𝔻n,rel ),\mathrm{BDiff}(M,\text{rel }\partial)\to\mathrm{B}\text{Emb}^{\cong}_{1/2\partial}(M)\to\mathrm{B}\mathrm{BDiff}(\mathbb{D}^{n},\text{rel }\partial),

where the delooping BBDiff(𝔻n,rel )\mathrm{B}\mathrm{BDiff}(\mathbb{D}^{n},\text{rel }\partial) is defined by considering BDiff(𝔻n,rel )\mathrm{BDiff}(\mathbb{D}^{n},\text{rel }\partial) as a topological monoid similar to 3.4 and the E1E_{1}-structure on this topological monoid is given by stacking along the first coordinate when we consider the interior of the cube as a model for the interior of the disk. We want to use a similar fiber sequence for a compact 33-manifold MM with a non-empty boundary whose one of its boundary components is homeomorphic to 𝕊2\mathbb{S}^{2}.

Proof of 3.7.

Let Emb1(M)\text{Emb}^{\cong}_{\partial_{1}}(M) the space of self locally flat embeddings of M1M_{1} that are the identity on the fixed boundary components 1M1\partial_{1}M_{1} and are isotopic to a homeomorphism that fixes the boundary through isotopies fixing 1M1\partial_{1}M_{1}.

Given that in dimension 33, the corresponding objects in C0C^{0} and CC^{\infty} category are weakly equivalent, we may apply the proof of [Kup19, Theorem 4.17] mutatis mutandis to conclude that there is a fiber sequence

(3.3) BHomeo(M1,rel )BEmb1(M1)BBD,\mathrm{B}\text{{Homeo}}(M_{1},\text{rel }\partial)\to\mathrm{B}\text{Emb}^{\cong}_{\partial_{1}}(M_{1})\to\mathrm{B}\mathrm{B}\mathrm{D},

and a weak equivalence B(BL,BD,)BEmb1(M1)||B_{\bullet}(\mathrm{B}\mathrm{L},\mathrm{B}\mathrm{D},*)||\simeq\mathrm{B}\text{Emb}^{\cong}_{\partial_{1}}(M_{1}). On the other hand, we have a fiber sequence

(3.4) BHomeo(M1,rel )BHomeo(M1,S,rel 1)BHomeo0(𝕊2),\mathrm{B}\text{{Homeo}}(M_{1},\text{rel }\partial)\to\mathrm{B}\text{{Homeo}}(M_{1},S,\text{rel }\partial_{1})\to\mathrm{B}\text{{Homeo}}_{0}(\mathbb{S}^{2}),

where the last map is the restriction to SS. Since homeomorphisms in Homeo(M1,S,rel 1)\text{{Homeo}}(M_{1},S,\text{rel }\partial_{1}) fix at least one boundary component, they are orientation preserving so they restrict to Homeo0(𝕊2)\text{{Homeo}}_{0}(\mathbb{S}^{2}).

Recall from the proof of 3.5 that there is a map from BSO(3)\mathrm{B}\text{SO}(3) to BBD\mathrm{B}\mathrm{B}\mathrm{D} which is a weak equivalence. And also the inclusion SO(3)Homeo0(𝕊2)\text{SO}(3)\hookrightarrow\text{{Homeo}}_{0}(\mathbb{S}^{2}) is a weak equivalence ([Ham74]). Hence, there is a map from the fiber sequence 3.4 to the fiber sequence 3.3 that induces weak equivalences between bases and the fibers. Therefore, their total spaces are also weakly equivalent. ∎

In the next section, we prove in 3.9 that BHomeo(M1,S,rel 1)\mathrm{B}\text{{Homeo}}(M_{1},S,\text{rel }\partial_{1}) has a finite CW complex model by induction on the number of prime factors. The space BR\mathrm{B}\mathrm{R} in the fibration 3.2 also by induction has a finite CW complex model. So in the fibration 3.2 the base and the fiber have a finite CW complex model which implies the same for the total space. Hence, this implies that η1(x)\eta^{-1}(x) when x0x\in\mathcal{F}_{0} is homology isomorphic to a finite CW complex.

Remark 3.8.

For MM being connected sum of two irreducible 33-manifolds with non-empty boundaries, Hatcher’s theorem ([Hat81]) about 33-manifolds also implies Kontsevich’s finiteness. Let SS be a separating sphere in MM, then his theorem implies that ι:BDiff(M,S)BDiff(M)\iota\colon\mathrm{BDiff}(M,S)\to\mathrm{BDiff}(M) is a homotopy equivalence where Diff(M,S)\mathrm{Diff}(M,S) is the subgroup of Diff(M)\mathrm{Diff}(M) that fixes SS set wise. And we have a fibration

BDiff(M2,rel )BDiff(M,S)BDiff(M1,S,rel M),\mathrm{BDiff}(M_{2},\text{rel }\partial)\to\mathrm{BDiff}(M,S)\to\mathrm{BDiff}(M_{1},S,\text{rel }\partial M),

where M1M_{1} and M2M_{2} are obtained by cutting MM along SS. This fibration is similar to our fibration 3.1.

3.C. Higher filtrations and finishing the proof of 2.5

For p>0p>0 suppose xpp1x\in\mathcal{F}_{p}-\mathcal{F}_{p-1}. We want to generalize the above bar resolution model by iterating the same construction (p+1)(p+1) times for each separating sphere in different orbits. And then write this iterated bar construction in a fiber sequence whose base and fiber, by induction, have finite CW complex models.

Let 𝐒={S0,S1,,Sp}{\bf S}=\{S_{0},S_{1},\dots,S_{p}\} be a set of p+1p+1 separating spheres where SiS_{i}’s are pairwise in different orbit classes under the action of Homeoδ(M,rel )\text{{Homeo}}^{\delta}(M,\text{rel }\partial). We pick an order on these spheres and note that they cannot be permuted via the action of Homeoδ(M,rel )\text{{Homeo}}^{\delta}(M,\text{rel }\partial). We similarly define 𝒮(M,[𝐒])\mathcal{S}(M,[{\bf S}]) to be the full subcomplex of 𝒮(M)\mathcal{S}(M) whose vertices are in the orbits of spheres in 𝐒{\bf S} under the action of Homeoδ(M,rel )\text{{Homeo}}^{\delta}(M,\text{rel }\partial). Note that the preimage of η1(x)\eta^{-1}(x) is homotopy equivalent to

𝒮(M,[𝐒])//Homeoδ(M,rel ).\mathcal{S}(M,[{\bf S}])/\!\!/\text{{Homeo}}^{\delta}(M,\text{rel }\partial).

In each simplex in 𝒮(M,[𝐒])\mathcal{S}(M,[{\bf S}]), the spheres parallel to SiS_{i} for each ii has a natural inside to outside order (see 3.1). So there is a natural multi-semi-simplicial set structure on 𝒮(M,[𝐒])\mathcal{S}(M,[{\bf S}]) that we denote by 𝒮,,(M,[𝐒])\mathcal{S}_{\bullet,\dots,\bullet}(M,[{\bf S}]) where the number of simplicial directions is p+1p+1. Similar to 3.3, we have a multi-semi-simplicial space 𝒮,,τ(M,[𝐒])\mathcal{S}^{\tau}_{\bullet,\dots,\bullet}(M,[{\bf S}]) and a homology isomorphism

||𝒮,,(M,[𝐒])//Homeoδ(M,rel )||||𝒮,,τ(M,[𝐒])//Homeo(M,rel )||.||\mathcal{S}_{\bullet,\dots,\bullet}(M,[{\bf S}])/\!\!/\text{{Homeo}}^{\delta}(M,\text{rel }\partial)||\to||\mathcal{S}^{\tau}_{\bullet,\dots,\bullet}(M,[{\bf S}])/\!\!/\text{{Homeo}}(M,\text{rel }\partial)||.

To finish the proof of 2.5, it is enough to show that ||𝒮,,τ(M,[𝐒])//Homeo(M,rel )||||\mathcal{S}^{\tau}_{\bullet,\dots,\bullet}(M,[{\bf S}])/\!\!/\text{{Homeo}}(M,\text{rel }\partial)|| has a finite CW complex model by putting it in a fibration whose base and the fiber have finite CW complex model by induction.

By doing the bar construction model in each simplicial direction, we have fibrations similar to the fibration 3.2. And by applying 3.7 we obtain a fibration similar to the fibration 3.1 whose fiber and the base are realizations of multi-semi-simplicial spaces with fewer simplicial directions to which we can apply induction.

Let M1M_{1} and M2M_{2} be the submanifolds obtained by cutting MM along S0S_{0} and M1M_{1} containing the boundary component M\partial_{*}M. Suppose that kk of spheres in {S1,,Sp}\{S_{1},\dots,S_{p}\} lies in M1M_{1} and the rest are in M2M_{2}. By doing the bar construction model and applying 3.7, we obtain a fibration

||R,,(M2)||||𝒮,,τ(M,[𝐒])//Homeo(M,rel )||||L,,(M1)||,||\mathrm{R}_{\bullet,\dots,\bullet}(M_{2})||\to||\mathcal{S}^{\tau}_{\bullet,\dots,\bullet}(M,[{\bf S}])/\!\!/\text{{Homeo}}(M,\text{rel }\partial)||\to||\mathrm{L}_{\bullet,\dots,\bullet}(M_{1})||,

where the number of simplicial directions in R,,(M2)\mathrm{R}_{\bullet,\dots,\bullet}(M_{2}) and L,,(M1)\mathrm{L}_{\bullet,\dots,\bullet}(M_{1}) are respectively pkp-k and kk. Hence, it is easy to see that we can exhaust simplicial directions by considering fibrations and using 3.7. So to finish the proof of 2.5, it is enough to prove the following base case.

Let PP be an irreducible 33 manifold with a non-empty boundary. Let ei:𝔻3Pe_{i}\colon\mathbb{D}^{3}\hookrightarrow P for 1ik+l1\leq i\leq k+l be disjoint embeddings and let NN be the 33 manifold obtained from PP by removing ei(int(𝔻3))e_{i}(\text{int}(\mathbb{D}^{3})) for all 1ik+l1\leq i\leq k+l. So the boundary of NN is the union of P\partial P with sphere boundary components SiS_{i}’s. We denote the union of the sphere boundary components {Si}i=1k\{S_{i}\}_{i=1}^{k} by SfreeS_{\text{free}} and union of the rest of sphere boundary components by SfixedS_{\text{fixed}}. Let Homeo(N,Sfree,rel PSfixed)\text{{Homeo}}(N,S_{\text{free}},\text{rel }\partial P\cup S_{\text{fixed}}) be the subgroup of Homeo(N)\text{{Homeo}}(N) whose elements fix each sphere in SfreeS_{\text{free}} set-wise and fix PSfixed\partial P\cup S_{\text{fixed}} pointwise.

Theorem 3.9.

Then BHomeo(N,Sfree,rel PSfixed)\mathrm{B}\text{{Homeo}}(N,S_{\text{free}},\text{rel }\partial P\cup S_{\text{fixed}}) has a finite CW complex model.

Since the corresponding diffeomorphism groups and homeomorphism groups are weakly equivalent, we shall instead prove that BDiff(N,Sfree,rel PSfixed)\mathrm{BDiff}(N,S_{\text{free}},\text{rel }\partial P\cup S_{\text{fixed}}) has a finite CW complex model. We already know the homotopical finiteness for an irreducible 33-manifold with a non-empty boundary ([HM97]).

Theorem 3.10 (Hatcher-McCullough).

If MM is an irreducible 33-manifold with a non-empty boundary, then BDiff(M,rel )\mathrm{BDiff}(M,\text{rel }\partial) has the homotopy type of a finite CW-complex.

So we know that BDiff(P,rel )\mathrm{BDiff}(P,\text{rel }\partial) has a finite CW complex model. And we want to inductively fix ei(𝔻3)e_{i}(\mathbb{D}^{3}) either set-wise or pointwise and still get a finite CW complex model.

Lemma 3.11.

Suppose MM is a 33-manifold with possibly nonempty boundary. Let 1\partial_{1} be a subset of boundary components containing the non-spherical components and let SfreeS_{\text{free}} be the union of remaining spherical components. Let e:D3Me\colon D^{3}\hookrightarrow M be an embedding of a ball inside MM. If BDiff(M,Sfree,rel 1)\mathrm{BDiff}(M,S_{\text{free}},\text{rel }\partial_{1}) has the homotopy type of a finite CW-complex, so does BDiff(M,Sfree,rel 1e(D3))\mathrm{BDiff}(M,S_{\text{free}},\text{rel }\partial_{1}\cup e(D^{3})).

Proof.

We give the argument when MM has a non-trivial boundary. The case of the closed manifold is similar. The fibration

Diff(M,Sfree,rel 1e(D3))Diff(M,Sfree,rel 1)Emb+(D3,M)Fr+(M),\mathrm{Diff}(M,S_{\text{free}},\text{rel }\partial_{1}\cup e(D^{3}))\to\mathrm{Diff}(M,S_{\text{free}},\text{rel }\partial_{1})\to\text{Emb}^{+}(D^{3},M)\simeq\text{Fr}^{+}(M),

where Emb+(D3,M)\text{Emb}^{+}(D^{3},M) is the space of orientation preserving embeddings and Fr+(M)\text{Fr}^{+}(M) is the oriented frame bundle of MM, induces the fibration

Fr+(M)BDiff(M,Sfree,rel 1e(D3))BDiff(M,Sfree,rel 1).\text{Fr}^{+}(M)\to\mathrm{BDiff}(M,S_{\text{free}},\text{rel }\partial_{1}\cup e(D^{3}))\to\mathrm{BDiff}(M,S_{\text{free}},\text{rel }\partial_{1}).

The base and the fiber of this fibration have the homotopy type of a finite CW-complex. Therefore, the total space also has a finite-dimensional model. ∎

Proof of 3.9.

We shall prove that BDiff(N,Sfree,rel PSfixed)\mathrm{BDiff}(N,S_{\text{free}},\text{rel }\partial P\cup S_{\text{fixed}}) has a finite CW complex model. Let MM be the manifold obtained from PP by removing ei(int(𝔻3))e_{i}(\text{int}(\mathbb{D}^{3})) for all 1ik1\leq i\leq k. Given 3.11, it is enough to prove that BDiff(M,Sfree,rel P)\mathrm{BDiff}(M,S_{\text{free}},\text{rel }\partial P) has a finite CW complex model.

Let xix_{i} be a point in PP given by the image of the center of the ball ei(int(𝔻3))e_{i}(\text{int}(\mathbb{D}^{3})). And let Diff(P,{x1,,xk},rel P)\mathrm{Diff}(P,\{x_{1},\dots,x_{k}\},\text{rel }\partial P) be the subgroup of Diff(P,rel P)\mathrm{Diff}(P,\text{rel }\partial P) consisting of those elements that fix each xix_{i}.

Claim.

The classifying space BDiff(M,Sfree,rel P)\mathrm{BDiff}(M,S_{\text{free}},\text{rel }\partial P) is homotopy equivalent to BDiff(P,{x1,,xk},rel P)\mathrm{BDiff}(P,\{x_{1},\dots,x_{k}\},\text{rel }\partial P).

Proof of the claim: Consider the fibration

BDiff(M,rel P)B(Diff0(S2)k),\mathrm{BDiff}(M,\text{rel }\partial P)\to\mathrm{B}(\mathrm{Diff}_{0}(S^{2})^{k}),

given by the restriction of diffeomorphisms of MM to sphere boundary components. The fiber of this map is homotopy equivalent to BDiff(M,rel M)\mathrm{BDiff}(M,\text{rel }\partial M) i.e. the classifying space of diffeomorphisms of MM that fix all boundary components pointwise.

Now recall that tangent bundles for orientable 33-manifolds are trivial, so the derivative map at the marked points xix_{i}’s gives

BDiff(P,{x1,,xk},rel M)B(GL3+()k).\mathrm{BDiff}(P,\{x_{1},\dots,x_{k}\},\text{rel }\partial M)\to\mathrm{B}(\text{GL}^{+}_{3}(\mathbb{R})^{k}).

Recall that Smale’s theorem ([Sma59]) implies that Diff0(S2)GL3+()SO(3)\mathrm{Diff}_{0}(S^{2})\simeq\text{GL}^{+}_{3}(\mathbb{R})\simeq\text{SO}(3). and the fiber of this fibration is also homotopy equivalent to BDiff(M,rel M)\mathrm{BDiff}(M,\text{rel }\partial M). Hence, the natural map

BDiff(M,rel P)BDiff(P,{x1,,xk},rel P),\mathrm{BDiff}(M,\text{rel }\partial P)\to\mathrm{BDiff}(P,\{x_{1},\dots,x_{k}\},\text{rel }\partial P),

that is induced by capping off the sphere boundary components and extending the diffeomorphisms by the identity to the bounding balls, is a homotopy equivalence. \blacksquare

Let PConfk(P)\text{PConf}_{k}(P) be the space of ordered configuration space of kk points in the interior of PP. We have a fiber sequence

PConfk(P)BDiff(P,{x1,,xk},rel P)BDiff(P,rel P).\text{PConf}_{k}(P)\to\mathrm{BDiff}(P,\{x_{1},\dots,x_{k}\},\text{rel }\partial P)\to\mathrm{BDiff}(P,\text{rel }\partial P).

Since both BDiff(P,rel P)\mathrm{BDiff}(P,\text{rel }\partial P) and PConfk(P)\text{PConf}_{k}(P) have a finite CW complex model so does BDiff(P,{x1,,xk},rel P)\mathrm{BDiff}(P,\{x_{1},\dots,x_{k}\},\text{rel }\partial P). ∎

Remark 3.12.

To consider the general case of a 33-manifold with a non-empty boundary, this method runs into two difficulties. One is that in the prime decomposition, not all the prime factors necessarily contain a boundary component. So diffeomorphic prime factors may be permuted so the homological model for η1(x)\eta^{-1}(x) for xpp1x\in\mathcal{F}_{p}-\mathcal{F}_{p-1} should be modified in this case. But the author thinks the more difficult issue is to deal with the case where we only have sphere boundary components. More concretely, let PP be a closed prime 33 manifold and let 𝔻3P\mathbb{D}^{3}\hookrightarrow P be an embedded ball inside PP. Given the resolution of the generalized Smale’s conjecture ([BK19, BK21]), the author thinks one could show that BDiff(P,rel 𝔻3)\mathrm{BDiff}(P,\text{rel }\mathbb{D}^{3}) has a finite CW complex model. But at the moment he does not know how to show that BDiff(M,rel 𝔻3)\mathrm{BDiff}(M,\text{rel }\mathbb{D}^{3}) has a finite CW complex model when MM is a closed reducible 33 manifold.

References

  • [BK19] Richard H Bamler and Bruce Kleiner. Ricci flow and contractibility of spaces of metrics. arXiv preprint arXiv:1909.08710, 2019.
  • [BK21] Richard H Bamler and Bruce Kleiner. Diffeomorphism groups of prime 3-manifolds. arXiv preprint arXiv:2108.03302, 2021.
  • [Bon83] Francis Bonahon. Cobordism of automorphisms of surfaces. Ann. Sci. École Norm. Sup. (4), 16(2):237–270, 1983.
  • [CdSR79] Eugénia César de Sá and Colin Rourke. The homotopy type of homeomorphisms of 33-manifolds. Bull. Amer. Math. Soc. (N.S.), 1(1):251–254, 1979.
  • [Cer61] Jean Cerf. Topologie de certains espaces de plongements. Bulletin de la Société Mathématique de France, 89:227–380, 1961.
  • [ERW19] Johannes Ebert and Oscar Randal-Williams. Semisimplicial spaces. Algebr. Geom. Topol., 19(4):2099–2150, 2019.
  • [Ham74] Mary-Elizabeth Hamstrom. Homotopy in homeomorphism spaces, TOPTOP and PLPL. Bulletin of the American Mathematical Society, 80(2):207–230, 1974.
  • [Hat76] Allen Hatcher. Homeomorphisms of sufficiently large P2P^{2}-irreducible 33-manifolds. Topology, 15(4):343–347, 1976.
  • [Hat81] A. Hatcher. On the diffeomorphism group of S1×S2S^{1}\times S^{2}. Proc. Amer. Math. Soc., 83(2):427–430, 1981.
  • [Hat83] Allen E Hatcher. A proof of the Smale conjecture, Diff(S3)O(4)\text{Diff}(S^{3})\simeq O(4). Annals of Mathematics, pages 553–607, 1983.
  • [Hem76] John Hempel. 33-Manifolds. Annals of Mathematics Studies, No. 86. Princeton University Press, Princeton, N. J.; University of Tokyo Press, Tokyo, 1976.
  • [HKMR12] Sungbok Hong, John Kalliongis, Darryl McCullough, and J. Hyam Rubinstein. Diffeomorphisms of elliptic 3-manifolds, volume 2055 of Lecture Notes in Mathematics. Springer, Heidelberg, 2012.
  • [HL84] Harrie Hendriks and François Laudenbach. Difféomorphismes des sommes connexes en dimension trois. Topology, 23(4):423–443, 1984.
  • [HM87] Harrie Hendriks and Darryl McCullough. On the diffeomorphism group of a reducible 33-manifold. Topology Appl., 26(1):25–31, 1987.
  • [HM90] Allen Hatcher and Darryl McCullough. Finite presentation of 3-manifold mapping class groups. In Groups of Self-Equivalences and Related Topics, pages 48–57. Springer, 1990.
  • [HM97] Allen Hatcher and Darryl McCullough. Finiteness of classifying spaces of relative diffeomorphism groups of 33-manifolds. Geom. Topol., 1:91–109, 1997.
  • [HW10] Allen Hatcher and Nathalie Wahl. Stabilization for mapping class groups of 3-manifolds. Duke Mathematical Journal, 155(2):205–269, 2010.
  • [Iva76] N. V. Ivanov. Groups of diffeomorphisms of Waldhausen manifolds. Zap. Naučn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI), 66:172–176, 209, 1976. Studies in topology, II.
  • [Kir95] Rob Kirby. Problems in low-dimensional topology. In Proceedings of Georgia Topology Conference, Part 2. Citeseer, 1995.
  • [Kup15] Alexander Kupers. Proving homological stability for homeomorphisms of manifolds. arXiv preprint arXiv:1510.02456, 2015.
  • [Kup19] Alexander Kupers. Some finiteness results for groups of automorphisms of manifolds. Geometry & Topology, 23(5):2277–2333, 2019.
  • [Las76] R. Lashof. Embedding spaces. Illinois J. Math., 20(1):144–154, 03 1976.
  • [McD80] Dusa McDuff. The homology of some groups of diffeomorphisms. Commentarii Mathematici Helvetici, 55(1):97–129, 1980.
  • [MN20] Kathryn Mann and Sam Nariman. Dynamical and cohomological obstructions to extending group actions. Mathematische Annalen, 377(3):1313–1338, 2020.
  • [Nar20] Sam Nariman. A local to global argument on low dimensional manifolds. Transactions of the American Mathematical Society, 373(2):1307–1342, 2020.
  • [Neu96] Walter D Neumann. Notes on geometry and 3-manifolds. Topology Atlas, 1996.
  • [Sma59] Stephen Smale. Diffeomorphisms of the 22-sphere. Proc. Amer. Math. Soc., 10:621–626, 1959.
  • [Thu74] William Thurston. Foliations and groups of diffeomorphisms. Bulletin of the American Mathematical Society, 80(2):304–307, 1974.
  • [Wal68] Friedhelm Waldhausen. On irreducible 33-manifolds which are sufficiently large. Ann. of Math. (2), 87:56–88, 1968.