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

Surface mapping class group actions on 3-manifolds

Alina al Beaini Lei Chen  and  Bena Tshishiku Alina Al Beaini
Department of Mathematics
Brown University
151 Thayer St.
Providence, RI, 02912, USA
alina_\_al_\_[email protected].
Lei Chen
Department of Mathematics
University of Maryland
4176 Campus Drive
College Park, MD 20742, USA
[email protected]
Bena Tshishiku
Department of Mathematics
Brown University
151 Thayer St.
Providence, RI, 02912, USA
bena_\_[email protected].
Abstract.

For each circle bundle S1XΣgS^{1}\to X\to\Sigma_{g} over a surface with genus g2g\geq 2, there is a natural surjection π:Homeo+(X)Mod(Σg)\pi:\operatorname{Homeo}^{+}(X)\to\operatorname{Mod}(\Sigma_{g}). When XX is the unit tangent bundle UΣgU\Sigma_{g}, it is well-known that π\pi splits. On the other hand π\pi does not split when the Euler number e(X)e(X) is not divisible by the Euler characteristic χ(Σg)\chi(\Sigma_{g}) by [CT23]. In this paper we show that this homomorphism does not split in many cases where χ(Σg)\chi(\Sigma_{g}) divides e(X)e(X).

1. Introduction

Let Σg\Sigma_{g} be a closed oriented surface of genus g2g\geq 2, and let Xg,eX_{g,e} denote the oriented S1S^{1}-bundle over Σg\Sigma_{g} with Euler number ee. Let Homeo+(Xg,e)\operatorname{Homeo}^{+}(X_{g,e}) be the group of orientation-preserving homeomorphisms of Xg,eX_{g,e} that act trivially on the center of π1(Xg,e)\pi_{1}(X_{g,e}), and let Mod(Xg,e):=π0(Homeo+(Xg,e))\operatorname{Mod}(X_{g,e}):=\pi_{0}\big{(}\operatorname{Homeo}^{+}(X_{g,e})\big{)} denote the mapping class group.

The (generalized) Nielsen realization problem for Xg,eX_{g,e} asks whether the surjective homomorphism

Homeo+(Xg,e)Mod(Xg,e)\operatorname{Homeo}^{+}(X_{g,e})\to\operatorname{Mod}(X_{g,e})

splits over subgroups of Mod(Xg,e)\operatorname{Mod}(X_{g,e}). In this paper we study a closely related problem. For each g,eg,e there is a surjection Mod(Xg,e)Mod(Σg)\operatorname{Mod}(X_{g,e})\to\operatorname{Mod}(\Sigma_{g}). Consider the composition

πg,e:Homeo+(Xg,e)Mod(Xg,e)Mod(Σg).\pi_{g,e}:\operatorname{Homeo}^{+}(X_{g,e})\twoheadrightarrow\operatorname{Mod}(X_{g,e})\twoheadrightarrow\operatorname{Mod}(\Sigma_{g}).
Problem 1.1.

Does πg,e:Homeo+(Xg,e)Mod(Σg)\pi_{g,e}:\operatorname{Homeo}^{+}(X_{g,e})\twoheadrightarrow\operatorname{Mod}(\Sigma_{g}) spilt?

If e=±(2g2)e=\pm(2g-2), then Xg,eX_{g,e} is the unit (co)tangent bundle, and πg,e\pi_{g,e} does split; see [Sou10, §1]. On the other hand, if ee is not divisible by 2g22g-2, then the surjection Mod(Xg,e)Mod(Σg)\operatorname{Mod}(X_{g,e})\to\operatorname{Mod}(\Sigma_{g}) does not split by work of the second two authors [CT23], so πg,e\pi_{g,e} also does not split in these cases. Given this, it remains to study the case when ee is divisible by 2g22g-2 and e±(2g2)e\neq\pm(2g-2). In these cases Mod(Xg,e)Mod(Σg)\operatorname{Mod}(X_{g,e})\to\operatorname{Mod}(\Sigma_{g}) does split [CT23], but we prove πg,e\pi_{g,e} does not split in many cases.

Theorem A.

Fix a surface Σg\Sigma_{g} of genus gg and ee\in\mathbb{Z}. Assume that g=4k1g=4k-1 where k3k\geq 3 and kk is not a power of 22, and assume that ee is divisible by (2g2)2p(2g-2)2p where pp is an odd prime dividing kk. Then the natural surjective homomorphism πg,e:Homeo+(Xg,e)Mod(Σg)\pi_{g,e}:\operatorname{Homeo}^{+}(X_{g,e})\to\operatorname{Mod}(\Sigma_{g}) does not split.

For example, if e=0e=0, we find that πg,e:Homeo+(Σg×S1)Mod(Σg)\pi_{g,e}:\operatorname{Homeo}^{+}(\Sigma_{g}\times S^{1})\to\operatorname{Mod}(\Sigma_{g}) does not split when g=11,19,23,27,35,39,43,47,g=11,19,23,27,35,39,43,47,\ldots.

Theorem A solves the Nielsen realization problem for Mod(Σg)\operatorname{Mod}(\Sigma_{g}) subgroups of Mod(Xg,e)\operatorname{Mod}(X_{g,e}) in the cases of the theorem. Specifically, if ee is divisible by 2g22g-2, then Mod(Xg,e)H1(Σg;)Mod(Σg)\operatorname{Mod}(X_{g,e})\cong H^{1}(\Sigma_{g};\mathbb{Z})\rtimes\operatorname{Mod}(\Sigma_{g}) [CT23], and every Mod(Σg)\operatorname{Mod}(\Sigma_{g}) subgroup of Mod(Xg,e)\operatorname{Mod}(X_{g,e}) is the image of a splitting of Mod(Xg,e)Mod(Σg)\operatorname{Mod}(X_{g,e})\to\operatorname{Mod}(\Sigma_{g}). By Theorem A, Homeo+(Xg,e)Mod(Xg,e)\operatorname{Homeo}^{+}(X_{g,e})\to\operatorname{Mod}(X_{g,e}) does not split over any of these Mod(Σg)\operatorname{Mod}(\Sigma_{g}) subgroups.

Theorem A has the following topological consequence. When 2g22g-2 divides ee, there is a “tautological” Xg,eX_{g,e}-bundle Eg,etautBHomeo(Σg)E_{g,e}^{\text{taut}}\to B\operatorname{Homeo}(\Sigma_{g}) whose monodromy

Mod(Σg)π1(BHomeo(Σg))Mod(Xg,e)\operatorname{Mod}(\Sigma_{g})\cong\pi_{1}\big{(}B\operatorname{Homeo}(\Sigma_{g})\big{)}\to\operatorname{Mod}(X_{g,e})

splits the surjection Mod(Xg,e)Mod(Σg)\operatorname{Mod}(X_{g,e})\to\operatorname{Mod}(\Sigma_{g}) (c.f. [CT23, §1]). One can ask whether or not the bundle Eg,etautBHomeo(Σg)E_{g,e}^{\text{taut}}\to B\operatorname{Homeo}(\Sigma_{g}) is flat. Recall that an XX-bundle EBE\to B is flat if there is a homomorphism ρ:π1(B)Homeo(X)\rho:\pi_{1}(B)\to\operatorname{Homeo}(X) and an XX-bundle isomorphism EXρBE\cong X\rtimes_{\rho}B. Such bundles are characterized by the existence of a horizontal foliation on EE, or, equivalently, by the property that their monodromy π1(B)Mod(X)\pi_{1}(B)\to\operatorname{Mod}(X) lifts to Homeo(X)\operatorname{Homeo}(X). When e=2g2e=2g-2, the bundle Eg,etautBHomeo(Σg)E_{g,e}^{\text{taut}}\to B\operatorname{Homeo}(\Sigma_{g}) is flat because of the splitting of πg,e\pi_{g,e} in this case. When πg,e\pi_{g,e} does not split, we deduce that Eg,etautBHomeo(Sg)E_{g,e}^{\text{taut}}\to B\operatorname{Homeo}(S_{g}) is not flat.

Corollary 1.2.

Fix g,eg,e as in the statement of Theorem A. Then the tautological Xg,eX_{g,e}-bundle Eg,etautBHomeo(Sg)E_{g,e}^{\text{taut}}\to B\operatorname{Homeo}(S_{g}) is not flat.


Short proof sketch of Theorem A. The proof strategy is similar to an argument of Chen–Salter [CS22] that shows that Homeo+(Σg)Mod(Σg)\operatorname{Homeo}^{+}(\Sigma_{g})\to\operatorname{Mod}(\Sigma_{g}) does not split when g2g\geq 2. Theorem A is proved by contradiction: assuming the existence of a splitting Mod(Σg)Homeo+(Xg,e)\operatorname{Mod}(\Sigma_{g})\to\operatorname{Homeo}^{+}(X_{g,e}), first we obtain, by lifting, an action of the based mapping class group Mod(Σg,)\operatorname{Mod}(\Sigma_{g},*) on the cover X^g,e2×S1\widehat{X}_{g,e}\cong\mathbb{R}^{2}\times S^{1} corresponding to the center of π1(Xg,e)\pi_{1}(X_{g,e}). The conditions on gg and ee in Theorem A guarantee the existence of a /2p\mathbb{Z}/2p\mathbb{Z} subgroup of Mod(Σg,)\operatorname{Mod}(\Sigma_{g},*) for which we can show the action on X^g,e\widehat{X}_{g,e} has a fixed circle. Denoting a generator of /2p\mathbb{Z}/2p\mathbb{Z} by α\alpha, we show that Mod(Σg,)\operatorname{Mod}(\Sigma_{g},*) is generated by the centralizers of α2\alpha^{2} and αp\alpha^{p}. This shows that the entire group Mod(Σg,)\operatorname{Mod}(\Sigma_{g},*) acts on X^g,e\widehat{X}_{g,e} with a fixed circle, which contradicts the fact that the point-pushing subgroup π1(Σg)<Mod(Σg,)\pi_{1}(\Sigma_{g})<\operatorname{Mod}(\Sigma_{g},*) acts freely (by deck transformations) on X^g,e\widehat{X}_{g,e}.


Other questions. Related to the Mod(Σg)\operatorname{Mod}(\Sigma_{g}) action on the unit tangent bundle UΣgU\Sigma_{g}, we pose the following question.

Question 1.3.

Do either of the following surjections split?

Diff+(UΣg)Mod(Σg) or Homeo(UΣg)Mod(UΣg)\operatorname{Diff}^{+}(U\Sigma_{g})\to\operatorname{Mod}(\Sigma_{g})\>\>\>\text{ or }\>\>\>\operatorname{Homeo}(U\Sigma_{g})\to\operatorname{Mod}(U\Sigma_{g})

If one includes orientation-reversing diffeomorphisms and mapping classes, then if g12g\geq 12, then Diff(UΣg)Mod±(Σg)\operatorname{Diff}(U\Sigma_{g})\to\operatorname{Mod}^{\pm}(\Sigma_{g}) does not split by Souto [Sou10, Thm. 1].


Acknowledgement. The authors LC and BT are supported by NSF grants DMS2203178, DMS-2104346, and DMS-2005409.

2. Proof of Theorem A

Fix g=4k1g=4k-1 and ee as in the theorem statement, and set Σ=Σg\Sigma=\Sigma_{g} and X=Xg,eX=X_{g,e}. Suppose for a contradiction that there is a homomorphism

σ:Mod(Σ)Homeo(X)\sigma:\operatorname{Mod}(\Sigma)\to\operatorname{Homeo}(X)

whose composition with π=πg,e:Homeo(X)Mod(Σ)\pi=\pi_{g,e}:\operatorname{Homeo}(X)\to\operatorname{Mod}(\Sigma) is the identity.

2.1. Step 1: lifting argument

Consider the covering space X^=Σ~×S1\widehat{X}=\widetilde{\Sigma}\times S^{1} of XX, where Σ~2\widetilde{\Sigma}\cong\mathbb{R}^{2} is the universal cover. This is the covering corresponding to the center ζ\zeta of π1(X)\pi_{1}(X). Given the action of Mod(Σ)\operatorname{Mod}(\Sigma) on XX, we consider the set of all lifts of homeomorphisms in this action to X^\widehat{X}. This is an action of the pointed mapping class group Mod(Σ,)\operatorname{Mod}(\Sigma,*) on X^\widehat{X}. To explain this, we start with the following general proposition.

Proposition 2.1.

Let YY be a closed manifold. Let ζ<π1(Y)\zeta<\pi_{1}(Y) be the center of the fundamental group, and denote Δ=π1(Y)/ζ\Delta=\pi_{1}(Y)/\zeta. Let Y^Y\widehat{Y}\to Y be the covering space with π1(Y^)=ζ\pi_{1}(\widehat{Y})=\zeta. Fix a basepoint Y*\in Y. Assume that the evaluation map

Homeo(Y)Y,ff()\operatorname{Homeo}(Y)\to Y,\>\>\>\>f\mapsto f(*)

induces a surjection π1(Homeo(Y))ζ<π1(Y)\pi_{1}\big{(}\operatorname{Homeo}(Y)\big{)}\twoheadrightarrow\zeta<\pi_{1}(Y). Then there is a commutative diagram

(1) 1\textstyle{1}Δ\textstyle{\Delta}Homeo(Y^)Δ\textstyle{\operatorname{Homeo}(\widehat{Y})^{\Delta}}Homeo(Y)\textstyle{\operatorname{Homeo}(Y)}1\textstyle{1}1\textstyle{1}Δ\textstyle{\Delta}Mod(Y,)\textstyle{\operatorname{Mod}(Y,*)}Mod(Y)\textstyle{\operatorname{Mod}(Y)}1\textstyle{1}q\textstyle{q}p^\textstyle{\widehat{p}}p\textstyle{p}

whose rows are exact, where the bottom row is the (generalized) Birman exact sequence. Furthermore, this diagram is a pullback diagram.

A version of Proposition 2.1 when Y=ΣgY=\Sigma_{g} (whose center is trivial) is used in [CS22].

We prove Proposition 2.1 after explaining how it gives the desired lifting. In our situation, the center of π1(X)\pi_{1}(X) is the kernel of π1(X)π1(Σ)\pi_{1}(X)\to\pi_{1}(\Sigma) since π1(Σ)\pi_{1}(\Sigma) has trivial center. Thus Proposition 2.1 gives us the following diagram.

1\textstyle{1}π1(Σ)\textstyle{\pi_{1}(\Sigma)}Homeo(X^)Δ\textstyle{\operatorname{Homeo}(\widehat{X})^{\Delta}}Homeo(X)\textstyle{\operatorname{Homeo}(X)}1\textstyle{1}1\textstyle{1}π1(Σ)\textstyle{\pi_{1}(\Sigma)}Mod(X,)\textstyle{\operatorname{Mod}(X,*)}Mod(X)\textstyle{\operatorname{Mod}(X)}1\textstyle{1}1\textstyle{1}π1(Σ)\textstyle{\pi_{1}(\Sigma)}Mod(Σ,)\textstyle{\operatorname{Mod}(\Sigma,*)}Mod(Σ)\textstyle{\operatorname{Mod}(\Sigma)}1\textstyle{1}q\textstyle{q}p^\textstyle{\widehat{p}}p\textstyle{p}

The splitting σ\sigma defines a subgroup Mod(Σ)<Mod(X)\operatorname{Mod}(\Sigma)<\operatorname{Mod}(X) and a splitting of pp over this subgroup. Since the top row is a pullback of the middle row, it follows that p^\widehat{p} splits over Mod(Σ,)\operatorname{Mod}(\Sigma,*) (this uses only general facts about pullbacks). Denote this splitting by

σ^:Mod(Σ,)Homeo(X^)Δ.\widehat{\sigma}:\operatorname{Mod}(\Sigma,*)\to\operatorname{Homeo}(\widehat{X})^{\Delta}.

Under this splitting the point-pushing subgroup π1(Σ)\pi_{1}(\Sigma) acts by deck transformations.

Remark 2.2.

If G<Mod(Σ)G<\operatorname{Mod}(\Sigma) and σ(G)\sigma(G) has a fixed point *, then after choosing a lift ^\widehat{*} of *, one can lift canonically elements of σ(G)\sigma(G) to X^\widehat{X} by choosing the unique lift that fixes ^\widehat{*}. This implies that G<Mod(Σ)G<\operatorname{Mod}(\Sigma) can be lifted to G<Mod(Σ,)G<\operatorname{Mod}(\Sigma,*) so that σ^(G)\widehat{\sigma}(G) has a fixed point.

Proof of Proposition 2.1.

First we recall the construction of the bottom row of diagram (1). Evaluation at Y*\in Y defines a fibration

Homeo(Y,)Homeo(Y)ϵY.\operatorname{Homeo}(Y,*)\to\operatorname{Homeo}(Y)\xrightarrow{\epsilon}Y.

The long exact sequence of homotopy groups gives an exact sequence

π1(Homeo(Y))ϵπ1(Y)Mod(Y,)Mod(Y)1.\pi_{1}\big{(}\operatorname{Homeo}(Y))\xrightarrow{\epsilon_{*}}\pi_{1}(Y)\to\operatorname{Mod}(Y,*)\to\operatorname{Mod}(Y)\to 1.

In general the image of ϵ\epsilon_{*} is contained in the center of π1(Y)\pi_{1}(Y); see e.g. [Hat02, §1.1, Exer. 20]. By assumption, ϵ\epsilon_{*} surjects onto the center, so we obtain the short exact sequence in the bottom row of (1). The homomorphism π1(Y)Mod(Y,)\pi_{1}(Y)\to\operatorname{Mod}(Y,*) is the so-called “point-pushing” homomorphism. It sends ηπ1(Y)\eta\in\pi_{1}(Y) (basepoint= *) to the time-1 map of an isotopy that pushes * around η\eta in reverse (this follows directly from the definition of the connecting homomorphism in the long exact sequence; note that it makes sense for the reverse of η\eta to appear in defining this homomorphism since concatenation of paths is left-to-right, while composition of functions is right-to-left).

Next we define p^:Homeo(Y^)ΔMod(Y,)\widehat{p}:\operatorname{Homeo}(\widehat{Y})^{\Delta}\to\operatorname{Mod}(Y,*). Fix a point ^Y^\widehat{*}\in\widehat{Y} that covers the basepoint Y*\in Y. Given fHomeo(Y^)Δf\in\operatorname{Homeo}(\widehat{Y})^{\Delta}. Choose a path [0,1]Y^[0,1]\to\widehat{Y} from ^\widehat{*} to f(^)f(\widehat{*}) and let γf\gamma_{f} denote the composition [0,1]Y^Y[0,1]\to\widehat{Y}\to Y. By isotopy extension, there exists an isotopy ht:YYh_{t}:Y\to Y where h0=idYh_{0}=\operatorname{id}_{Y} and ht()=γf(t)h_{t}(*)=\gamma_{f}(t) for each t[0,1]t\in[0,1]. Define

p^(f)=[h1q(f)].\widehat{p}(f)=[h_{1}\circ q(f)].

The map p^\widehat{p} is well-defined. The choice of γf\gamma_{f} is unique only up to an element of π1(Y^)=ζ\pi_{1}(\widehat{Y})=\zeta. This implies that the isotopy class [h1f][h_{1}\circ f] is only well-defined up to composition by a point-pushing mapping class by an element of ζ\zeta, but such a point-push is trivial by assumption.

It is a straightforward exercise to check that p^\widehat{p} is a homomorphism. The right square in diagram (1) commutes because q(f)q(f) and h1q(f)h_{1}\circ q(f) are isotopic by construction. It is easy to see that the left square in the diagram commutes by applying the definition of p^\widehat{p} to deck transformations.

Finally, regarding the claim that the diagram is a pullback, we show that the map to the fibered product

p^×q:Homeo(Y^)ΔMod(Y,)×Mod(Y)Homeo(Y)\widehat{p}\times q:\operatorname{Homeo}(\widehat{Y})^{\Delta}\to\operatorname{Mod}(Y,*)\times_{\operatorname{Mod}(Y)}\operatorname{Homeo}(Y)

is an isomorphism. The codomain consists of pairs (ϕ,g)Mod(Y,)×Homeo(Y)(\phi,g)\in\operatorname{Mod}(Y,*)\times\operatorname{Homeo}(Y) such that gg is isotopic to a representative of the isotopy class ϕ\phi.

We define an inverse ι\iota to p^×q\widehat{p}\times q. Given (ϕ,g)(\phi,g) in the fibered product, choose an isotopy gtg_{t} from gg to a homeomorphism representing ϕ\phi. Lift gtg_{t} to an isotopy g~t\widetilde{g}_{t} such that g~1\widetilde{g}_{1} fixes ^\widehat{*}, and define ι(ϕ,g)=g~0\iota(\phi,g)=\widetilde{g}_{0}. The reader can check that the maps ι\iota and p^×q\widehat{p}\times q are inverses. ∎

2.2. Step 2: finite group action rigidity

Recall that g=4k1g=4k-1 and k3k\geq 3 is not a power of 2; let pp be an odd prime dividing kk. From Step 1, we have homomorphism σ^:Mod(Σ,)Homeo(X^)\widehat{\sigma}:\operatorname{Mod}(\Sigma,*)\to\operatorname{Homeo}(\widehat{X}) that descends to a splitting σ:Mod(Σ)Homeo(X)\sigma:\operatorname{Mod}(\Sigma)\to\operatorname{Homeo}(X). In this section we describe the action of a particular finite subgroup of Mod(Σ,)\operatorname{Mod}(\Sigma,*) on X^\widehat{X}.

Proposition 2.3.

There exists an element αMod(Σ,)\alpha\in\operatorname{Mod}(\Sigma,*) of order 2p2p such that the fixed sets of σ^(α)\widehat{\sigma}(\alpha), σ^(α)2\widehat{\sigma}(\alpha)^{2}, and σ^(α)p\widehat{\sigma}(\alpha)^{p} coincide and are equal to an embedded circle cX^c\subset\widehat{X}.

It is worth noting that the fixed set of a finite-order, orientation-preserving homeomorphism of a 3-manifold can be wildly embedded [MZ54].

In order to prove Proposition 2.3 we first construct the specific element α\alpha. Then we prove (Proposition 2.5) a weaker version of Proposition 2.3 with the additional assumption that the action is smooth. Finally, we combine this with a result of Pardon [Par21] and Smith theory to prove Proposition 2.3.


Construction of α\alpha. We obtain α\alpha as an element in a dihedral subgroup D4kD_{4k} of Mod(Σ)\operatorname{Mod}(\Sigma), where D4kD_{4k} denotes the dihedral group of order 8k8k. The dihedral action D4kΣD_{4k}\curvearrowright\Sigma we use has quotient Σ/D4k\Sigma/D_{4k} homeomorphic to T2T^{2} and the quotient ΣΣ/D4k\Sigma\to\Sigma/D_{4k} has a single branch point; it is determined by the homomorphism

x,y=F2π1(T2pt)D4k=a,ba4k=b2=1,bab=a1\langle x,y\rangle=F_{2}\cong\pi_{1}(T^{2}\setminus\operatorname{pt})\to D_{4k}=\langle a,b\mid a^{4k}=b^{2}=1,bab=a^{-1}\rangle
xa,yb.x\mapsto a,\>\>\>\>y\mapsto b.

By Riemann–Hurwitz, the genus of Σ\Sigma is 4k14k-1. The orbifold O=Σ/D4kO=\Sigma/D_{4k} has fundamental group

π1orb(O)=x,y,hh2k=1,h=[x,y],\pi_{1}^{orb}(O)=\langle x,y,h\mid h^{2k}=1,h=[x,y]\rangle,

and there is a short exact sequence

(2) 1π1(Σ)π1orb(O)D4k1.1\to\pi_{1}(\Sigma)\to\pi_{1}^{orb}(O)\to D_{4k}\to 1.

This sequence induces a homomorphism π1orb(O)Mod(Σ,)\pi_{1}^{orb}(O)\to\operatorname{Mod}(\Sigma,*). We take α=hk/p\alpha=h^{k/p}, where pp, as defined above, is an odd prime dividing kk, which exists by assumption. Then α\alpha is an element of order 2p2p in the subgroup h/2k\langle h\rangle\cong\mathbb{Z}/2k\mathbb{Z} of π1orb(O)<Mod(Σ,)\pi_{1}^{orb}(O)<\operatorname{Mod}(\Sigma,*).

Remark 2.4.

The argument that follows works equally well when Σ/D4k\Sigma/D_{4k} is a genus-gg surface and ΣΣ/D4k\Sigma\to\Sigma/D_{4k} has a single branched point. This provides more values of g,eg,e for which the conclusion of Theorem A holds.


Smooth case. Here we prove the following proposition.

Proposition 2.5.

Fix D4k<Mod(Σ)D_{4k}<\operatorname{Mod}(\Sigma) as above. Suppose that σ:D4kDiff+(X)\sigma:D_{4k}\to\operatorname{Diff}^{+}(X) and is a splitting of π:Homeo+(X)Mod(Σ)\pi:\operatorname{Homeo}^{+}(X)\to\operatorname{Mod}(\Sigma) over D4kD_{4k}. Then σ^(α)\widehat{\sigma}(\alpha) fixes a unique circle on X^=2×S1\widehat{X}=\mathbb{H}^{2}\times S^{1}. Consequently, the fixed set of σ(a2k/p)\sigma(a^{2k/p}) is nonempty.

The last part of the statement of Proposition 2.5 follows from the preceding statement because the image of α\alpha under π1orb(O)D4k\pi_{1}^{orb}(O)\to D_{4k} is a2k/pa^{2k/p}.

Proof of Proposition 2.5.

First we reduce to a more geometric setting. By Meeks–Scott [MS86, Thm. 2.1], the smooth(!) action σ(D4k)X\sigma(D_{4k})\curvearrowright X preserves some geometric metric on XX. There are two possibilities for the geometry: if e(X)=0e(X)=0, then XX has 2×\mathbb{H}^{2}\times\mathbb{R}-geometry, and if e(X)0e(X)\neq 0, then XX has PSL2()~\widetilde{\operatorname{PSL}_{2}(\mathbb{R})}-geometry. We treat these cases in parallel.

The universal cover X~\widetilde{X} (with the induced geometric structure) is either 2×\mathbb{H}^{2}\times\mathbb{R} or PSL2()~\widetilde{\operatorname{PSL}_{2}(\mathbb{R})}. In either case, X~\widetilde{X} has an isometric foliation by lines whose leaf space is isometric to 2\mathbb{H}^{2}, and this foliation is preserved by Isom(X~)\operatorname{Isom}(\widetilde{X}), so there is a homomorphism Isom(X~)Isom(2)\operatorname{Isom}(\widetilde{X})\to\operatorname{Isom}(\mathbb{H}^{2}). Let Isom+(X~)<Isom(X~)\operatorname{Isom}^{+}(\widetilde{X})<\operatorname{Isom}(\widetilde{X}) be the group whose action on the leaves and on the leaf space are both orientation preserving. There is an exact sequence

(3) 1Isom+(X~)𝐹Isom+(2)1.1\to\mathbb{R}\to\operatorname{Isom}^{+}(\widetilde{X})\xrightarrow{F}\operatorname{Isom}^{+}(\mathbb{H}^{2})\to 1.

See also [Sco83, §4].

Next consider the group Λ\Lambda of all lifts of elements of σ(D4k)<Isom(X)\sigma(D_{4k})<\operatorname{Isom}(X) to Isom+(X~)\operatorname{Isom}^{+}(\widetilde{X}). This yields an exact sequence

1π1(X)ΛD4k1.1\to\pi_{1}(X)\to\Lambda\to D_{4k}\to 1.

The action of Λ\Lambda on X~\widetilde{X} induces an action of Λ/ζ\Lambda/\zeta on X~/ζ=X^2×S1\widetilde{X}/\zeta=\widehat{X}\cong\mathbb{H}^{2}\times S^{1}, where ζ\zeta is the center of π1(X)\pi_{1}(X). This action extends to an action of Isom+(X~)/ζ\operatorname{Isom}^{+}(\widetilde{X})/\zeta, and there is a homomorphism

ρ:Λ/ζIsom+(X~)/ζIsom(X^).\rho:\Lambda/\zeta\to\operatorname{Isom}^{+}(\widetilde{X})/\zeta\xrightarrow{\cong}\operatorname{Isom}(\widehat{X}).

The last map is an isomorphism by the general formula Isom(X~/Λ)=NIsom(X~)(Λ)/Λ\operatorname{Isom}(\widetilde{X}/\Lambda)=N_{\operatorname{Isom}(\widetilde{X})}(\Lambda)/\Lambda for discrete subgroups Λ<Isom(X~)\Lambda<\operatorname{Isom}(\widetilde{X}).

To prove the proposition, we first identify Λ/ζ\Lambda/\zeta with π1orb(O)\pi_{1}^{orb}(O) (Claim 2.6). Then it is a formal consequence of our setup that ρ(hk/p)=σ^(α)\rho(h^{k/p})=\widehat{\sigma}(\alpha), and after showing Isom+(X~)/ζIsom+(2)×SO(2)\operatorname{Isom}^{+}(\widetilde{X})/\zeta\cong\operatorname{Isom}^{+}(\mathbb{H}^{2})\times\operatorname{SO}(2) (Claim 2.7), we show that ρ(hk/p)\rho(h^{k/p}) fixes a unique circle in X^\widehat{X} (Claim 2.8).

Claim 2.6.

The restriction of the sequence (3) to Λ\Lambda is a short exact sequence

1ζΛπ1orb(O)11\to\zeta\to\Lambda\to\pi_{1}^{orb}(O)\to 1

where ζ\zeta is the center of π1(X)\pi_{1}(X).

Proof of Claim 2.6.

Recall the map F:Isom+(X~)Isom+(2)F:\operatorname{Isom}^{+}(\widetilde{X})\to\operatorname{Isom}^{+}(\mathbb{H}^{2}) from (3). First we identify F(Λ)<Isom+(2)F(\Lambda)<\operatorname{Isom}^{+}(\mathbb{H}^{2}) with π1orb(O)\pi_{1}^{orb}(O). For this, it suffices to show that F(Λ)F(\Lambda) fits into a short exact sequence

(4) 1π1(Σ)F(Λ)D4k1,1\to\pi_{1}(\Sigma)\to F(\Lambda)\to D_{4k}\to 1,

where the “monodromy” D4kOut+(π1(Σ))Mod(Σ)D_{4k}\to\operatorname{Out}^{+}\big{(}\pi_{1}(\Sigma)\big{)}\cong\operatorname{Mod}(\Sigma) has image the given subgroup D4k<Mod(Σ)D_{4k}<\operatorname{Mod}(\Sigma). This implies that F(Λ)π1orb(O)F(\Lambda)\cong\pi_{1}^{orb}(O) because π1orb(O)\pi_{1}^{orb}(O) is an extension of the same form (see (2)), and extensions of π1(Σ)\pi_{1}(\Sigma) are determined by their monodromy [Bro82, §IV.3].

To construct the extension (4), first note that the restriction of (3) to π1(X)\pi_{1}(X) is the short exact sequence

1ζπ1(X)π1(Σ)1.1\to\zeta\to\pi_{1}(X)\to\pi_{1}(\Sigma)\to 1.

The group π1(Σ)=F(π1(X))\pi_{1}(\Sigma)=F(\pi_{1}(X)) is normal in F(Λ)F(\Lambda) because π1(X)\pi_{1}(X) is normal in Λ\Lambda. Furthermore, the surjection ΛF(Λ)\Lambda\to F(\Lambda) induces a surjection D4k=Λ/π1(X)F(Λ)/π1(Σ)D_{4k}=\Lambda/\pi_{1}(X)\to F(\Lambda)/\pi_{1}(\Sigma).

The quotient map X~2\widetilde{X}\to\mathbb{H}^{2}, which is equivariant with respect to ΛF(Λ)\Lambda\to F(\Lambda) descends to a map X=X~/π1(X)2/π1(Σ)=ΣX=\widetilde{X}/\pi_{1}(X)\to\mathbb{H}^{2}/\pi_{1}(\Sigma)=\Sigma that’s equivariant with respect to D4k=Λ/π1(X)F(Λ)/π1(Σ)D_{4k}=\Lambda/\pi_{1}(X)\to F(\Lambda)/\pi_{1}(\Sigma).

Since σ\sigma is a realization, the induced action of σ(D4k)\sigma(D_{4k}) on Σ\Sigma is a realization of the D4k<Mod(Σ)D_{4k}<\operatorname{Mod}(\Sigma), and in particular the D4kD_{4k} action on Σ\Sigma is faithful. Therefore, F(Λ)/π1(Σ)D4kF(\Lambda)/\pi_{1}(\Sigma)\cong D_{4k}, and the monodromy of the associated extension

1π1(Σ)F(Λ)D4k1,1\to\pi_{1}(\Sigma)\to F(\Lambda)\to D_{4k}\to 1,

is the given inclusion D4k<Mod(Σ)D_{4k}<\operatorname{Mod}(\Sigma). This concludes the proof that F(Λ)F(\Lambda) is isomorphic to π1orb(O)\pi_{1}^{orb}(O).

To finish the proof of Claim 2.6, it remains to show that the intersection of Λ\Lambda with =ker(F)\mathbb{R}=\ker(F) is ζ\zeta. We do this by showing (i) Λ\Lambda\cap\mathbb{R} is the center of Λ\Lambda, and (ii) the center of Λ\Lambda is contained in π1(X)\pi_{1}(X). Together with the obvious containment ζ<Λ\zeta<\Lambda\cap\mathbb{R}, (i) and (ii) imply Λ=ζ\Lambda\cap\mathbb{R}=\zeta.

(i): First note that Λ\Lambda\cap\mathbb{R} is central because \mathbb{R} is central in Isom(X~)\operatorname{Isom}(\widetilde{X}). On the other hand, the center of Λ\Lambda is contained in Λ\Lambda\cap\mathbb{R} because the center of Λ/(Λ)π1orb(O)\Lambda/(\Lambda\cap\mathbb{R})\cong\pi_{1}^{orb}(O) has trivial center.

(ii): To show the center of Λ\Lambda is contained in π1(X)\pi_{1}(X), we show that the center of Λ\Lambda projects trivially to D4k=Λ/π1(X)D_{4k}=\Lambda/\pi_{1}(X). This is true because ΛD4k\Lambda\to D_{4k} factors through π1orb(O)\pi_{1}^{orb}(O), which has trivial center.∎

We summarize the relation between the relevant groups in Diagram (5).

(5) 1\textstyle{1}π1(X)\textstyle{\pi_{1}(X)}Λ\textstyle{\Lambda}D4k\textstyle{D_{4k}}1\textstyle{1}1\textstyle{1}π1(Σ)\textstyle{\pi_{1}(\Sigma)}π1orb(O)\textstyle{\pi_{1}^{orb}(O)}D4k\textstyle{D_{4k}}1\textstyle{1}ζ\textstyle{\zeta}ζ\textstyle{\zeta}1\textstyle{1}1\textstyle{1}1\textstyle{1}1\textstyle{1}

By Claim 2.6, Λ/ζπ1orb(O)\Lambda/\zeta\cong\pi_{1}^{orb}(O), so ρ\rho takes the form

ρ:π1orb(O)Isom+(X~)/ζIsom(X^)\rho:\pi_{1}^{orb}(O)\to\operatorname{Isom}^{+}(\widetilde{X})/\zeta\cong\operatorname{Isom}(\widehat{X})

By construction, this homomorphism is the restriction of σ^:Mod(Σ,)Homeo(X^)\widehat{\sigma}:\operatorname{Mod}(\Sigma,*)\to\operatorname{Homeo}(\widehat{X}) to π1orb(O)\pi_{1}^{orb}(O). Since α=hk/p\alpha=h^{k/p}, to show the fixed set of σ^(α)\widehat{\sigma}(\alpha) is a circle, it suffices to show the same statement for ρ(hk/p)\rho(h^{k/p}). To prove this, we first compute Isom(X^)Isom+(X~)/ζ\operatorname{Isom}(\widehat{X})\cong\operatorname{Isom}^{+}(\widetilde{X})/\zeta.

Claim 2.7.

The group Isom+(X~)/ζ\operatorname{Isom}^{+}(\widetilde{X})/\zeta is isomorphic to Isom+(2)×SO(2)\operatorname{Isom}^{+}(\mathbb{H}^{2})\times\operatorname{SO}(2).

Proof of Claim 2.7.

First note that there is an extension

1SO(2)Isom+(X~)/ζIsom+(2)11\to\operatorname{SO}(2)\to\operatorname{Isom}^{+}(\widetilde{X})/\zeta\to\operatorname{Isom}^{+}(\mathbb{H}^{2})\to 1

induced from (3). This sequence is obviously split when X~=2×\widetilde{X}=\mathbb{H}^{2}\times\mathbb{R} since Isom+(X~)Isom+(2)×\operatorname{Isom}^{+}(\widetilde{X})\cong\operatorname{Isom}^{+}(\mathbb{H}^{2})\times\mathbb{R} is a product.

Assume now that X~=PSL2()~\widetilde{X}=\widetilde{\operatorname{PSL}_{2}(\mathbb{R})}, and write e=(2g2)ne=(2g-2)n where nn is a nonzero integer. Let KK denote the kernel of the universal cover homomorphism PSL2()~PSL2()\widetilde{\operatorname{PSL}_{2}(\mathbb{R})}\to\operatorname{PSL}_{2}(\mathbb{R}).

We claim that ζ=1nK\zeta=\frac{1}{n}K. To see this, note that the extension

1KPSL2()~PSL2()11\to K\to\widetilde{\operatorname{PSL}_{2}(\mathbb{R})}\to\operatorname{PSL}_{2}(\mathbb{R})\to 1

pulled back under a Fuchsian representation π1(Σ)PSL2()\pi_{1}(\Sigma)\to\operatorname{PSL}_{2}(\mathbb{R}) induces the extension of the unit tangent bundle group π1(UΣ)\pi_{1}(U\Sigma), which has Euler number 22g2-2g, and there is an nn-fold fiberwise cover UΣXU\Sigma\to X, so the center of π1(UΣ)<π1(X)\pi_{1}(U\Sigma)<\pi_{1}(X) is generated by the nn-the power of the generator of the center of π1(X)\pi_{1}(X), i.e. ζ=1nK\zeta=\frac{1}{n}K.

The inclusion of PSL2()~\widetilde{\operatorname{PSL}_{2}(\mathbb{R})} in Isom(PSL2()~)\operatorname{Isom}\big{(}\widetilde{\operatorname{PSL}_{2}(\mathbb{R})}\big{)} (given by left-multiplication) descends to a homomorphism

PSL2()=PSL2()~/KIsom(PSL2()~)/KIsom(PSL2()~)/ζ\operatorname{PSL}_{2}(\mathbb{R})=\widetilde{\operatorname{PSL}_{2}(\mathbb{R})}/K\to\operatorname{Isom}\big{(}\widetilde{\operatorname{PSL}_{2}(\mathbb{R})}\big{)}/K\twoheadrightarrow\operatorname{Isom}\big{(}\widetilde{\operatorname{PSL}_{2}(\mathbb{R})}\big{)}/\zeta

that defines a splitting of the sequence

1SO(2)Isom(PSL2()~)/ζPSL2()1.1\to\operatorname{SO}(2)\to\operatorname{Isom}(\widetilde{\operatorname{PSL}_{2}(\mathbb{R})})/\zeta\to\operatorname{PSL}_{2}(\mathbb{R})\to 1.\qed

The following Claim 2.8 is the last step in the proof of Proposition 2.5.

Claim 2.8.

Let pp be an odd prime dividing kk. If ee is divisible by (2g2)2p(2g-2)2p, then the fixed set of ρ(hk/p)\rho(h^{k/p}) is a circle.

Before proving the claim, we explain how the factors of Isom+(2)×SO(2)Isom+(X^)\operatorname{Isom}^{+}(\mathbb{H}^{2})\times\operatorname{SO}(2)\cong\operatorname{Isom}^{+}(\widehat{X}) act on X^=X~/ζ\widehat{X}=\widetilde{X}/\zeta.

Remark 2.9.

Consider the isomorphism Isom(X^)Isom+(2)×SO(2)\operatorname{Isom}(\widehat{X})\cong\operatorname{Isom}^{+}(\mathbb{H}^{2})\times\operatorname{SO}(2) from Claim 2.7. In each case (e=0e=0 or e0e\neq 0) the action of SO(2)\operatorname{SO}(2) on X^\widehat{X} covers the identity of 2\mathbb{H}^{2} and acts freely by rotation on the circle fibers of X2X\to\mathbb{H}^{2}. For the Isom+(2)\operatorname{Isom}^{+}(\mathbb{H}^{2}) action, when e=0e=0, then X^2×S1\widehat{X}\cong\mathbb{H}^{2}\times S^{1} is a metric product, and the action of Isom+(2)\operatorname{Isom}^{+}(\mathbb{H}^{2}) is trivial on the S1S^{1} factor and is the natural action on 2\mathbb{H}^{2}. If e=(2g2)ne=(2g-2)n is nonzero, then

X^PSL2()~/ζPSL2()/(/n),\widehat{X}\cong\widetilde{\operatorname{PSL}_{2}(\mathbb{R})}/\zeta\cong\operatorname{PSL}_{2}(\mathbb{R})/(\mathbb{Z}/n\mathbb{Z}),

and with respect to this isomorphism, the action of Isom+()PSL2()\operatorname{Isom}^{+}(\mathbb{H})\cong\operatorname{PSL}_{2}(\mathbb{R}) on X^\widehat{X} is induced from left multiplication of PSL2()\operatorname{PSL}_{2}(\mathbb{R}) on PSL2()\operatorname{PSL}_{2}(\mathbb{R}). Identifying PSL2()\operatorname{PSL}_{2}(\mathbb{R}) with the unit tangent U2U\mathbb{H}^{2}, we can also view PSL2()/(/n)\operatorname{PSL}_{2}(\mathbb{R})/(\mathbb{Z}/n\mathbb{Z}) as the quotient of U2U\mathbb{H}^{2} by the /n\mathbb{Z}/n\mathbb{Z} action that covers the identity of 2\mathbb{H}^{2} and rotates each fiber.

Proof of Claim 2.8.

Write e=(2g2)2pme=(2g-2)2pm for some integer mm.

First note that since ρ(h)\rho(h) has finite order, the induced isometry of 2\mathbb{H}^{2} has a unique fixed point, so ρ(h)\rho(h) preserves a unique circle CC of the fibering X^2\widehat{X}\to\mathbb{H}^{2}. The same is true for ρ(hk/p)\rho(h^{k/p}), and we will show that ρ(hk/p)\rho(h^{k/p}) acts trivially on CC.

Since h=[x,y]h=[x,y] is a commutator in π1orb(O)\pi_{1}^{orb}(O) and SO(2)\operatorname{SO}(2) is abelian, we find that the projection

ρ(h)Isom+(X^)Isom+(2)×SO(2)SO(2)\rho(h)\in\operatorname{Isom}^{+}(\widehat{X})\cong\operatorname{Isom}^{+}(\mathbb{H}^{2})\times\operatorname{SO}(2)\to\operatorname{SO}(2)

is trivial. Therefore, the action of ρ(h)\rho(h) on X^\widehat{X} factors through Isom+(2)\operatorname{Isom}^{+}(\mathbb{H}^{2}) acting on X^\widehat{X}. This action is described in Remark 2.9. If e=0e=0, since Isom+(2)\operatorname{Isom}^{+}(\mathbb{H}^{2}) acts trivially on the S1S^{1} factor of X^2×S1\widehat{X}\cong\mathbb{H}^{2}\times S^{1}, we conclude that ρ(h)\rho(h) acts trivially on CC. If e0e\neq 0, then ρ(h)\rho(h) acts as a a rotation by 2π(pm/k)2\pi(pm/k) on CC, so ρ(hk/p)\rho(h^{k/p}) acts as a rotation by 2πm2\pi m, which is trivial. ∎

This completes the proof of Proposition 2.5. ∎


Homeomorphism case. Here prove Proposition 2.3.

Proof of Proposition 2.3.

By Pardon [Par21, Thm. 1.1], there is a sequence of smooth D4kD_{4k} actions converging in Hom(D4k,Homeo(X))\operatorname{Hom}\big{(}D_{4k},\operatorname{Homeo}(X)\big{)} to the given action of σ(D4k)\sigma(D_{4k}) on XX. Sufficiently close approximates also give a splitting of π\pi over D4k<Mod(Σ)D_{4k}<\operatorname{Mod}(\Sigma) because Homeo(X)\operatorname{Homeo}(X) is locally path connected [EK71].

For each of the smooth approximations of σ(D4k)\sigma(D_{4k}), the fixed set of a2k/pa^{2k/p} is nonempty by Proposition 2.5. This implies that σ(a2k/p)\sigma(a^{2k/p}) has a fixed point (a sequence of fixed points, one for each smooth action, sub-converges to a fixed point of the σ(a2k/p)\sigma(a^{2k/p}) action). By Remark 2.2, there exists a lift of a2k/pD4k<Mod(Σ)a^{2k/p}\in D_{4k}<\operatorname{Mod}(\Sigma) to a finite order element απ1orb(O)<Mod(Σ,)\alpha^{\prime}\in\pi_{1}^{orb}(O)<\operatorname{Mod}(\Sigma,*) so that σ^(α)\widehat{\sigma}(\alpha^{\prime}) has a fixed point. Since π1orb(O)\pi_{1}^{orb}(O) has a unique conjugacy class of finite subgroup of order 2p2p, the subgroups α\langle\alpha^{\prime}\rangle and α\langle\alpha\rangle are conjugate, so the fixed set of σ^(α)\widehat{\sigma}(\alpha) is nonempty.

It remains to show the fixed set of σ^(α)\widehat{\sigma}(\alpha) is a circle, and that this circle is the same as the fixed sets of σ^(α)2\widehat{\sigma}(\alpha)^{2} and σ^(α)p\widehat{\sigma}(\alpha)^{p}.

First we show (using Smith theory) that both σ^(α)2\widehat{\sigma}(\alpha)^{2} and σ^(α)p\widehat{\sigma}(\alpha)^{p} have fixed set a single circle (we are not yet claiming/arguing that the fixed sets of σ^(α)2\widehat{\sigma}(\alpha)^{2} and σ^(α)p\widehat{\sigma}(\alpha)^{p} are the same). To see this, we focus on σ^(α)2\widehat{\sigma}(\alpha)^{2} for concreteness. Consider the group Λ0\Lambda_{0} of all lifts of powers of σ^(α)2\widehat{\sigma}(\alpha)^{2} to the universal cover X~\widetilde{X}. This group is an extension

1Λ0/p1,1\to\mathbb{Z}\to\Lambda_{0}\to\mathbb{Z}/p\mathbb{Z}\to 1,

which is central and split; hence Λ0×/p\Lambda_{0}\cong\mathbb{Z}\times\mathbb{Z}/p\mathbb{Z}. It is central because α\alpha acts orientation-preservingly on fibers of XΣX\to\Sigma (otherwise, the action of α\alpha on Σ\Sigma would reverse orientation, contrary to the construction); it splits because σ^(α)\widehat{\sigma}(\alpha) has a fixed point. The (lifted) action of σ^(α)2\widehat{\sigma}(\alpha)^{2} on X~\widetilde{X} has fixed set a line (i.e. embedded copy of \mathbb{R}) by Smith theory and local Smith theory [Bre12, Theorem 20.1], and this line is preserved and acted properly by <Λ0\mathbb{Z}<\Lambda_{0}; thus σ^(α)2\widehat{\sigma}(\alpha)^{2} acts on X^\widehat{X} with a circle in its fixed set. Furthermore, each component of the fixed set of σ^(α)2\widehat{\sigma}(\alpha)^{2} acting on X^\widehat{X} corresponds to a distinct conjugacy class of order-pp subgroup of Λ0\Lambda_{0}. Since there is only one /p\mathbb{Z}/p\mathbb{Z} subgroup of Λ0\Lambda_{0}, the fixed set of σ^(α)2\widehat{\sigma}(\alpha)^{2} is connected, i.e. a single circle. The same argument111Smith theory applies to prime-order finite cyclic group actions, so we cannot apply this argument directly to σ^(α)\widehat{\sigma}(\alpha). works for σ^(α)p\widehat{\sigma}(\alpha)^{p}.

Now we determine the fixed set of σ^(α)\widehat{\sigma}(\alpha). First observe that σ^(α)\widehat{\sigma}(\alpha) preserves the fixed set of σ^(α)2\widehat{\sigma}(\alpha)^{2} and has a fixed point there (the fixed set of σ^(α)\widehat{\sigma}(\alpha) is nonempty and contained in the fixed set of σ^(α)2\widehat{\sigma}(\alpha)^{2}). The only /2p\mathbb{Z}/2p\mathbb{Z} action on the circle with a fixed point is the trivial action, so in fact the fixed sets of σ^(α)\widehat{\sigma}(\alpha) and σ^(α)2\widehat{\sigma}(\alpha)^{2} are the same. The same argument applies to σ^(α)\widehat{\sigma}(\alpha) and σ^(α)p\widehat{\sigma}(\alpha)^{p}. This proves Proposition 2.3. ∎

2.3. Step 3: centralizer argument

Recall that we have defined α\alpha as an element of order 2p2p in π1orb(O)<Mod(Σ,)\pi_{1}^{orb}(O)<\operatorname{Mod}(\Sigma,*). In this step we prove that Mod(Σ,)\operatorname{Mod}(\Sigma,*) is generated by the centralizers of α2\alpha^{2} and αp\alpha^{p}.

Proposition 2.10 (centralizer property).

Let αMod(Σ,)\alpha\in\operatorname{Mod}(\Sigma,*) be the element of order 2p2p constructed above. Then

Mod(Σ,)=C(α2),C(αp),\operatorname{Mod}(\Sigma,*)=\langle C(\alpha^{2}),C(\alpha^{p})\rangle,

where C()C(-) denotes the centralizer in Mod(Σ,)\operatorname{Mod}(\Sigma,*).


Strategy for proving Proposition 2.10. Set Γ=C(α2),C(αp)\Gamma=\langle C(\alpha^{2}),C(\alpha^{p})\rangle. Our method for showing Γ=Mod(Σ,)\Gamma=\operatorname{Mod}(\Sigma,*), which is similar to the proof of [CS22, Thm. 1.1], is to inductively build subsurfaces

(6) S0S1SNΣ{}S_{0}\subset S_{1}\subset\cdots\subset S_{N}\subset\Sigma\setminus\{*\}

such that Mod(Sn)Γ\operatorname{Mod}(S_{n})\subset\Gamma for each nn and SNS_{N} fills Σ{}\Sigma\setminus\{*\} (i.e. each boundary component of SNS_{N} is inessential in Σ{}\Sigma\setminus\{*\}). The fact that SNS_{N} fills implies that Mod(SN)=Mod(Σ,)\operatorname{Mod}(S_{N})=\operatorname{Mod}(\Sigma,*), so then Mod(Σ,)Γ\operatorname{Mod}(\Sigma,*)\subset\Gamma by the last step in the inductive argument.

In order to ensure that Mod(Sn)Γ\operatorname{Mod}(S_{n})\subset\Gamma, the subsurface SnS_{n} is obtained from Sn1S_{n-1} by an operation known as subsurface stabilization. If SΣS\subset\Sigma is a subsurface and cΣc\subset\Sigma is a simple closed curve that intersects SS in a single arc, then the stabilization of SS along cc is the subsurface SN(c)S\cup N(c), where N(c)N(c) is a regular neighborhood of cc. It is easy to show that Mod(SN(c))\operatorname{Mod}(S\cup N(c)) is generated by Mod(S)\operatorname{Mod}(S) and the Dehn twist τc\tau_{c} [CS22, Lem. 4.2], so if Mod(S)Γ\operatorname{Mod}(S)\subset\Gamma and τcΓ\tau_{c}\in\Gamma, then Mod(SN(c))Γ\operatorname{Mod}(S\cup N(c))\subset\Gamma. Therefore, for the proof, it suffices to find a sequence of subsurface stabilizations along curves whose Dehn twist belongs to Γ=C(α2),C(αp)\Gamma=\langle C(\alpha^{2}),C(\alpha^{p})\rangle.


Model for the α\alpha action. Our proof of Proposition 2.10 makes use of an explicit model for Σ\Sigma with its α\alpha action, which is pictured below in the case k=6k=6 and p=3p=3 (recall that g=4k1g=4k-1 and pp is an odd prime dividing kk).

\labellist
\hair

2pt \pinlabel11 at 87 154 \pinlabel22 at 87 208 \pinlabel33 at 87 262 \pinlabel44 at 87 320 \pinlabel55 at 87 373 \pinlabel66 at 87 429

\pinlabel

77 at 173 154 \pinlabel88 at 173 208 \pinlabel99 at 173 262 \pinlabel1010 at 173 320 \pinlabel1111 at 173 373 \pinlabel1212 at 173 429

\pinlabel

11 at 326 154 \pinlabel22 at 326 208 \pinlabel33 at 326 262 \pinlabel44 at 326 320 \pinlabel55 at 326 373 \pinlabel66 at 326 429

\pinlabel

77 at 373 154 \pinlabel88 at 373 208 \pinlabel99 at 373 262 \pinlabel1010 at 373 320 \pinlabel1111 at 373 373 \pinlabel1212 at 373 429

\pinlabel

77 at 433 154 \pinlabel88 at 433 208 \pinlabel99 at 433 262 \pinlabel1010 at 433 320 \pinlabel1111 at 433 373 \pinlabel1212 at 433 429

\pinlabel

11 at 480 154 \pinlabel22 at 480 208 \pinlabel33 at 480 262 \pinlabel44 at 480 320 \pinlabel55 at 480 373 \pinlabel66 at 480 429

\pinlabel

77 at 637 154 \pinlabel88 at 637 208 \pinlabel99 at 637 262 \pinlabel1010 at 637 320 \pinlabel1111 at 637 373 \pinlabel1212 at 637 429

\pinlabel

11 at 722 154 \pinlabel22 at 722 208 \pinlabel33 at 722 262 \pinlabel44 at 722 320 \pinlabel55 at 722 373 \pinlabel66 at 722 429

\pinlabel

α\alpha at 820 218 \pinlabelT2T^{2} at 400 485 \pinlabel\rightarrow at 400 460 \pinlabel\rightarrow at 400 123 \pinlabel\blacktriangle at 287 291 \pinlabel\blacktriangle at 520 291 \pinlabelS2S^{2} at 130 485 \pinlabel\rightarrow at 130 460 \pinlabel\rightarrow at 130 123 \pinlabelS2S^{2} at 680 485 \pinlabel\rightarrow at 680 460 \pinlabel\rightarrow at 680 123 \endlabellistRefer to caption

Figure 1. The model of the surface Σ23\Sigma_{23} where p=3p=3.

The surface Σ\Sigma is built out of two copies of the standard action of /2p\mathbb{Z}/2p\mathbb{Z} on S2S^{2} and one copy of a free action of /2p\mathbb{Z}/2p\mathbb{Z} on T2T^{2}. We glue each copy of S2S^{2} to T2T^{2} along k/pk/p free orbits by an equivariant connected sum. In the figure, α\alpha acts by vertical translation. Note that the fixed points of α\alpha on S2S^{2} are not pictured in the figure – they are at ±\pm\infty along the xx-axis.

To derive this model, recall that D4k=a,ba4k=1=b2,bab=a1D_{4k}=\langle a,b\mid a^{4k}=1=b^{2},bab=a^{-1}\rangle has abelianization D4k(/2)2D_{4k}\to(\mathbb{Z}/2\mathbb{Z})^{2} with kernel a2/2k\langle a^{2}\rangle\cong\mathbb{Z}/2k\mathbb{Z}. Then there is a sequence of regular covers

Σa2Σ/a2(/2)2Σ/D4k.\Sigma\xrightarrow{\langle a^{2}\rangle}\Sigma/\langle a^{2}\rangle\xrightarrow{(\mathbb{Z}/2\mathbb{Z})^{2}}\Sigma/D_{4k}.

The cover Σ/a2Σ/D4k\Sigma/\langle a^{2}\rangle\to\Sigma/D_{4k} is unbranched and is the /2\mathbb{Z}/2\mathbb{Z}-homology cover of T2T^{2} (in particular, Σ/a2\Sigma/\langle a^{2}\rangle is also a torus). The cover ΣΣ/a2\Sigma\to\Sigma/\langle a^{2}\rangle is branched over four points; the local monodromy around the branched points is a2a^{2} at two of the branched points and a2a^{-2} at the other two. Choosing branched cuts joining branched points in a±2a^{\pm 2} pairs gives a model for Σ\Sigma, and one can check that this model is equivalent to the one described above. (The spheres in Figure 1 arise from pre-images under ΣΣ/a2\Sigma\to\Sigma/\langle a^{2}\rangle of neighborhoods of the branch cuts.)

By Remark 2.2, since σ(a2k/p)\sigma(a^{2k/p}) has a fixed point, the subgroup αMod(Σ,)\langle\alpha\rangle\subset\operatorname{Mod}(\Sigma,*), which lifts a2k/p\langle a^{2k/p}\rangle, has a fixed point. The different lifts of a2k/p\langle a^{2k/p}\rangle to a finite subgroup of Mod(Σ,)\operatorname{Mod}(\Sigma,*) are in one-to-one correspondence to fixed points of a2k/pa^{2k/p}. Since these fixed points are permuted transitively by the action of D4kD_{4k}, the different lifts of a2k/p\langle a^{2k/p}\rangle are conjugate. Consequently, for the purpose of our argument, we can choose * to be any one of the four fixed points of α\alpha and prove Proposition 2.10 for this choice, without loss of generality. (It will also be evident from the argument that a similar argument applies if * is changed to another fixed point.)

Remark 2.11.

We do not known how generally the relation Mod(Σ,)=C(α2),C(αp)\operatorname{Mod}(\Sigma,*)=\langle C(\alpha^{2}),C(\alpha^{p})\rangle holds. For example, it may hold for every /2p\mathbb{Z}/2p\mathbb{Z} subgroup of Mod(Σ,)\operatorname{Mod}(\Sigma,*). We do not know a general (abstract) approach to this problem.

Proof of Proposition 2.10.

Symmetry breaking. In preparation for constructing a sequence of subsurface stabilizations, in this paragraph we find a suitable collection of Dehn twists that belong to Γ\Gamma. The obvious way for τc\tau_{c} to belong to Γ\Gamma is if cc is preserved by either α2\alpha^{2} or αp\alpha^{p}. More generally, we use a process that we call symmetry breaking to show τcΓ\tau_{c}\in\Gamma for certain cc. We formulate this in the following lemma, which is similar to [CS22, Lem. 3.2].

Lemma 2.12 (Symmetry breaking).

Assume that c,dΣc,d\subset\Sigma are simple closed curves that intersect once and τdΓ\tau_{d}\in\Gamma. Suppose that either (i) αp(c)\alpha^{p}(c) is disjoint from cc and dd or (ii) the curves α2(c),α4(c),α2p2(c)\alpha^{2}(c),\alpha^{4}(c)\ldots,\alpha^{2p-2}(c) are disjoint from dd and the curves c,α2(c),α4(c),α2p2(c)c,\alpha^{2}(c),\alpha^{4}(c)\ldots,\alpha^{2p-2}(c) are pairwise disjoint. Then τcΓ\tau_{c}\in\Gamma.

Proof of Lemma 2.12.

We prove case (i) of the statement; case (ii) is similar. Since Dehn twists about disjoint curves commute,

(τcταp(c))τd(τcταp(c))1=τcτdτc1.\big{(}\tau_{c}\tau_{\alpha^{p}(c)}\big{)}\tau_{d}\big{(}\tau_{c}\tau_{\alpha^{p}(c)}\big{)}^{-1}=\tau_{c}\tau_{d}\tau_{c}^{-1}.

The left hand side of the equation is in Γ\Gamma because τdΓ\tau_{d}\in\Gamma by assumption and τcταp(c)Γ\tau_{c}\tau_{\alpha^{p}(c)}\in\Gamma because the curves c,αp(c)c,\alpha^{p}(c) are permuted by αp\alpha^{p} and are disjoint (so their twists commute), and thus τcταp(c)C(αp)Γ\tau_{c}\tau_{\alpha^{p}(c)}\in C(\alpha^{p})\subset\Gamma. Since cc and dd intersect once, the braid relation implies that τdτcτd1=τcτdτc1\tau_{d}\tau_{c}\tau_{d}^{-1}=\tau_{c}\tau_{d}\tau_{c}^{-1} also belongs to Γ\Gamma. Since τdΓ\tau_{d}\in\Gamma this implies that τcΓ\tau_{c}\in\Gamma, as desired. ∎

Remark 2.13.

When applying Lemma 2.12(i) or (ii) we refer to it as the αp\alpha^{p}- or α2\alpha^{2}-symmetry breaking, respectively.

Lemma 2.14.

Dehn twists about the curves in Figure 2 are in Γ\Gamma.

\labellist
\hair

2pt \pinlabel11 at 87 154 \pinlabel22 at 87 208 \pinlabel33 at 87 262 \pinlabel44 at 87 320 \pinlabel55 at 87 373 \pinlabel66 at 87 429

\pinlabel

77 at 173 154 \pinlabel88 at 173 208 \pinlabel99 at 173 262 \pinlabel1010 at 173 320 \pinlabel1111 at 173 373 \pinlabel1212 at 173 429

\pinlabel

11 at 326 154 \pinlabel22 at 326 208 \pinlabel33 at 326 262 \pinlabel44 at 326 320 \pinlabel55 at 326 373 \pinlabel66 at 326 429

\pinlabel

77 at 373 154 \pinlabel88 at 373 208 \pinlabel99 at 373 262 \pinlabel1010 at 373 320 \pinlabel1111 at 373 373 \pinlabel1212 at 373 429

\pinlabel

77 at 433 154 \pinlabel88 at 433 208 \pinlabel99 at 433 262 \pinlabel1010 at 433 320 \pinlabel1111 at 433 373 \pinlabel1212 at 433 429

\pinlabel

11 at 480 154 \pinlabel22 at 480 208 \pinlabel33 at 480 262 \pinlabel44 at 480 320 \pinlabel55 at 480 373 \pinlabel66 at 480 429

\pinlabel

77 at 637 154 \pinlabel88 at 637 208 \pinlabel99 at 637 262 \pinlabel1010 at 637 320 \pinlabel1111 at 637 373 \pinlabel1212 at 637 429

\pinlabel

11 at 722 154 \pinlabel22 at 722 208 \pinlabel33 at 722 262 \pinlabel44 at 722 320 \pinlabel55 at 722 373 \pinlabel66 at 722 429

\pinlabel

α\alpha at 820 218

\pinlabel

\rightarrow at 400 459 \pinlabel\rightarrow at 400 123 \pinlabel\blacktriangle at 287 291 \pinlabel\blacktriangle at 520 291

\pinlabel

\rightarrow at 130 459 \pinlabel\rightarrow at 130 123

\pinlabel

\rightarrow at 680 459 \pinlabel\rightarrow at 680 123

\pinlabel

c1c_{1} at 410 308 \pinlabelc2c_{2} at 354 133 \pinlabelc3c_{3} at 385 186 \pinlabelc4c_{4} at 130 140 \pinlabelc5c_{5} at 507 189 \pinlabelc6c_{6} at 409 405 \pinlabelc7c_{7} at 130 210

\endlabellist
Refer to caption
Figure 2. The curves used in the proof of Lemma 2.14. Here we are using the model for Σ23\Sigma_{23}, but the proof follows in the same way for similar types of curves on any Σ\Sigma.

In Figure 2, we illustrate the case k=6,p=3k=6,p=3. The corresponding curves in the general case belong to Γ\Gamma by the exact same argument.

Proof of Lemma 2.14.

First observe that τc1\tau_{c_{1}} and τc6\tau_{c_{6}} are in Γ\Gamma because each is invariant under αp\alpha^{p}. We deduce that τc2Γ\tau_{c_{2}}\in\Gamma using α2\alpha^{2}-symmetry breaking with d=c1d=c_{1}. Each of τc3\tau_{c_{3}} and τc4\tau_{c_{4}} are in Γ\Gamma by αp\alpha^{p}-symmetry breaking with d=c2d=c_{2}. Finally, both τc5\tau_{c_{5}} and τc7\tau_{c_{7}} are in Γ\Gamma by αp\alpha^{p}-symmetry breaking with d=c3d=c_{3}. ∎


Surface stabilization sequence. We stabilize with the sequence of curves represented in Figure 3. To get the initial subsurface S0S_{0} we can take the subsurface spanned by the chain of curves c0,c1,,c4c_{0},c_{1},\ldots,c_{4}. This subsurface has genus 2 and one boundary component, and these curves are Humphries generators for S0S_{0} [FM12, Fig. 4.10]. Next we extend this chain with the curves c5,,c4k1c_{5},\ldots,c_{4k-1}; at this point the left genus-0 subsurface has been filled. Next we stabilize with c4kc_{4k} and the curves (c4k+1,c4k+2,,c8k2c_{4k+1},c_{4k+2},\cdots,c_{8k-2}) that fill the right genus-0 subsurface; there is some choice in the order of curves we stabilize, but this is not important. Finally we stabilize with the curves c8k1c_{8k-1} and c8kc_{8k} that generate π1(T2)\pi_{1}(T^{2}).

All of the twists about the curves used are in Γ\Gamma. In each case, this can be seen either directly from the statement of Lemma 2.14 or by an argument that is a small variation of its proof. Since this collection of curves fills Σ\Sigma, we have shown that Γ=Mod(Σ,)\Gamma=\operatorname{Mod}(\Sigma,*). This proves Proposition 2.10. ∎

\labellist
\hair

2pt \pinlabelα\alpha at 820 218

\pinlabel

\rightarrow at 400 460 \pinlabel\rightarrow at 400 123 \pinlabel\blacktriangle at 287 291 \pinlabel\blacktriangle at 520 291

\pinlabel

\rightarrow at 130 460 \pinlabel\rightarrow at 130 123

\pinlabel

\rightarrow at 680 460 \pinlabel\rightarrow at 680 123

\hair

2pt

\pinlabel

c0c_{0} at 45 430 \pinlabelc1c_{1} at 88 430 \pinlabelc2c_{2} at 100 403 \pinlabelc3c_{3} at 88 375 \pinlabelc4c_{4} at 100 340 \pinlabelc5c_{5} at 88 320 \pinlabelc6c_{6} at 100 293 \pinlabelc7c_{7} at 88 265 \pinlabelc8c_{8} at 100 235 \pinlabelc9c_{9} at 88 205 \pinlabelc10c_{10} at 102 180 \pinlabelc11c_{11} at 88 155 \pinlabelc12c_{12} at 130 141 \pinlabelc13c_{13} at 173 155 \pinlabelc14c_{14} at 188 180 \pinlabelc15c_{15} at 173 205 \pinlabelc16c_{16} at 188 235 \pinlabelc17c_{17} at 173 265 \pinlabelc18c_{18} at 188 293 \pinlabelc19c_{19} at 173 320 \pinlabelc20c_{20} at 188 347 \pinlabelc21c_{21} at 173 375 \pinlabelc22c_{22} at 188 400 \pinlabelc23c_{23} at 173 430

\pinlabel

c24c_{24} at 230 310

\pinlabel

c25c_{25} at 637 320 \pinlabelc26c_{26} at 622 349 \pinlabelc27c_{27} at 637 375 \pinlabelc28c_{28} at 622 402 \pinlabelc29c_{29} at 637 430

\pinlabel

c30c_{30} at 678 439

\pinlabel

c31c_{31} at 723 430 \pinlabelc32c_{32} at 740 400 \pinlabelc33c_{33} at 723 375 \pinlabelc34c_{34} at 740 347 \pinlabelc35c_{35} at 723 320

\pinlabel

c36c_{36} at 637 155 \pinlabelc37c_{37} at 622 180 \pinlabelc38c_{38} at 637 208 \pinlabelc39c_{39} at 622 237 \pinlabelc40c_{40} at 637 265

\pinlabel

c41c_{41} at 678 141

\pinlabel

c42c_{42} at 723 155 \pinlabelc43c_{43} at 740 180 \pinlabelc44c_{44} at 723 208 \pinlabelc45c_{45} at 740 237 \pinlabelc46c_{46} at 723 265

\pinlabel

c47c_{47} at 302 190 \pinlabelc48c_{48} at 418 400

\endlabellistRefer to caption
Figure 3. The stabilization sequence we use for the case k=6k=6 and p=3p=3.

2.4. Step 4: conclusion

By Proposition 2.3, σ^(α)\widehat{\sigma}(\alpha), σ^(α2)\widehat{\sigma}(\alpha^{2}) and σ^(αp)\widehat{\sigma}(\alpha^{p}) all have the same fixed set, which is a circle cX^c\subset\widehat{X}. The centralizers C(α2)C(\alpha^{2}) and C(αp)C(\alpha^{p}) preserve cc. By Proposition 2.10, Mod(Σ,)=C(α2),C(αp)\operatorname{Mod}(\Sigma,*)=\langle C(\alpha^{2}),C(\alpha^{p})\rangle, so σ^(Mod(Σ,))\widehat{\sigma}\big{(}\operatorname{Mod}(\Sigma,*)\big{)} preserves cc. This contradicts the fact that σ^(π1(Σg))\widehat{\sigma}(\pi_{1}(\Sigma_{g})) acts as the deck group, which as a properly discontinuous action does not preserve any compact set.

References

  • [Bre12] Glen E Bredon. Sheaf theory, volume 170. Springer Science & Business Media, 2012.
  • [Bro82] K. S. Brown. Cohomology of groups, volume 87 of Graduate Texts in Mathematics. Springer-Verlag, New York-Berlin, 1982.
  • [CS22] L. Chen and N. Salter. Global fixed points of mapping class group actions and a theorem of Markovic. J. Topol., 15(3):1311–1324, 2022.
  • [CT23] Lei Chen and Bena Tshishiku. Mapping class groups of circle bundles over a surface, 2023.
  • [EK71] R. Edwards and R. Kirby. Deformations of spaces of imbeddings. Ann. of Math. (2), 93:63–88, 1971.
  • [FM12] B. Farb and D. Margalit. A primer on mapping class groups, volume 49 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 2012.
  • [Hat02] A. Hatcher. Algebraic topology. Cambridge University Press, Cambridge, 2002.
  • [MS86] W. H. Meeks, III and P. Scott. Finite group actions on 33-manifolds. Invent. Math., 86(2):287–346, 1986.
  • [MZ54] D. Montgomery and L. Zippin. Examples of transformation groups. Proc. Amer. Math. Soc., 5:460–465, 1954.
  • [Par21] J. Pardon. Smoothing finite group actions on three-manifolds. Duke Math. J., 170(6):1043–1084, 2021.
  • [Sco83] P. Scott. The geometries of 33-manifolds. Bull. London Math. Soc., 15(5):401–487, 1983.
  • [Sou10] J. Souto. A remark on the action of the mapping class group on the unit tangent bundle. Ann. Fac. Sci. Toulouse Math. (6), 19(3-4):589–601, 2010.