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

Measure equivalence rigidity of the handlebody groups

Sebastian Hensel and Camille Horbez
Abstract

Let VV be a connected 33-dimensional handlebody of finite genus at least 33. We prove that the handlebody group Mod(V)\mathrm{Mod}(V) is superrigid for measure equivalence, i.e. every countable group which is measure equivalent to Mod(V)\mathrm{Mod}(V) is in fact virtually isomorphic to Mod(V)\mathrm{Mod}(V). Applications include a rigidity theorem for lattice embeddings of Mod(V)\mathrm{Mod}(V), an orbit equivalence rigidity theorem for free ergodic measure-preserving actions of Mod(V)\mathrm{Mod}(V) on standard probability spaces, and a WW^{*}-rigidity theorem among weakly compact group actions.

Introduction

A central quest in measured group theory is to classify countable groups up to measure equivalence, a notion coined by Gromov in [Gro93] as a measurable analogue to the geometric notion of quasi-isometry between finitely generated groups.

The definition is as follows: two infinite countable groups Γ1\Gamma_{1} and Γ2\Gamma_{2} are measure equivalent if there exists a standard infinite measure space Ω\Omega equipped with an action of Γ1×Γ2\Gamma_{1}\times\Gamma_{2} by measure-preserving Borel automorphisms, such that for every i{1,2}i\in\{1,2\}, the action of Γi\Gamma_{i} on Ω\Omega is free and has a fundamental domain of finite measure. The typical example is that any two (possibly non-uniform) lattices in the same locally compact second countable group GG are always measure equivalent, by considering the left and right multiplications on GG equipped with its Haar measure.

Dye proved in [Dye59, Dye63] that all countably infinite abelian groups are measure equivalent. This was famously generalized by Ornstein and Weiss to all countably infinite amenable groups [OW80], and in fact these form a class of the measure equivalence relation on the set of all countably infinite groups. At the other extreme of the picture, some groups satisfy very strong rigidity properties. A first striking example is the following: building on earlier work of Zimmer [Zim80, Zim91], Furman proved that every countable group which is measure equivalent to a lattice in a higher rank Lie group, is virtually a lattice in the same Lie group [Fur99a]. In [MS06], Monod and Shalom proved superrigidity type results for direct products of groups that satisfy an analytic form of negative curvature, phrased in terms of a bounded cohomology criterion. Later, Kida proved that, with the exception of some low-complexity cases, mapping class groups Mod(Σ)\mathrm{Mod}(\Sigma) of finite-type surfaces are ME-superrigid, i.e. every countable group that is measure equivalent to Mod(Σ)\mathrm{Mod}(\Sigma), is in fact commensurable to Mod(Σ)\mathrm{Mod}(\Sigma) up to a finite kernel [Kid10]. This led to further strong rigidity results, for certain amalgamated free products [Kid11], certain subgroups of Mod(Σ)\mathrm{Mod}(\Sigma) such as the Torelli group [CK15], some infinite classes of Artin groups of hyperbolic type [HH20a]. Very recently, Guirardel and the second named author established that Out(FN)\mathrm{Out}(F_{N}), the outer automorphism group of a finitely generated free group of rank N3N\geq 3, is also ME-superrigid [GH21].

In the present paper, we establish a superrigidity theorem for handlebody groups, defined as mapping class groups Mod(V)\mathrm{Mod}(V) of connected 33-dimensional handlebodies VV, i.e. VV is a disk-sum of finitely many copies of D2×S1D^{2}\times S^{1}. These groups are of particular importance in 33-dimensional topology, and most notably in the theory of Heegaard splittings, see e.g. the discussion in [Hen, Section 4]. They are also important in geometric group theory due to their direct connections to both mapping class groups of surfaces and outer automorphism groups of free groups. Notice indeed that V\partial V is a closed orientable surface of finite genus g0g\geq 0, and Mod(V)\mathrm{Mod}(V) embeds as a (highly distorted [HH12]) subgroup of Mod(V)\mathrm{Mod}(\partial V); it also surjects onto Out(Fg)\mathrm{Out}(F_{g}) via the action at the level of the fundamental group (with non-finitely generated kernel [McC85b]). Recently, the geometry of handlebody groups has been shown to share many features with outer automorphism groups of free groups rather than surface mapping class groups (e.g. concerning the growth of isoperimetric functions [HH21a] or the subgroup geometry of stabilisers [Hen21]).

Handlebody groups are known to satisfy some algebraic rigidity properties: Korkmaz and Schleimer proved in [KS09] that their outer automorphism group is trivial, and the first named author further proved in [Hen18] that the natural map from the handlebody group to its abstract commensurator is an isomorphism. To our knowledge, the question of the quasi-isometric rigidity of handlebody groups (which are finitely generated by work of Suzuki [Suz77], in fact finitely presented by work of Wajnryb [Waj98]) is still widely open. Our main theorem establishes their superrigidity from the viewpoint of measured group theory.

Theorem 1.

Let VV be a connected 33-dimensional handlebody of finite genus at least 33. Then Mod(V)\mathrm{Mod}(V) is ME-superrigid.

Consequences.

The techniques used in the proof of Theorem 1 have several other consequences. First, we recover (with a different argument) the commensurator rigidity statement established by the first named author in [Hen18], see Remark 3.3.

Second, using ideas of Furman [Fur11a] and Kida [Kid10], we can derive that handlebody groups cannot embed as lattices in second countable locally compact groups in any interesting way.

Corollary 2.

Let VV be a connected 33-dimensional handlebody of finite genus at least 33. Let GG be a locally compact second countable group, equipped with its Haar measure. Let Γ\Gamma be a finite index subgroup of Mod(V)\mathrm{Mod}(V), and let σ:ΓG\sigma:\Gamma\to G be an injective homomorphism whose image is a lattice.

Then there exists a homomorphism θ:GMod(V)\theta:G\to\mathrm{Mod}(V) with compact kernel such that for every fΓf\in\Gamma, one has θσ(f)=f\theta\circ\sigma(f)=f.

If SS is a finite generating set of Mod(V)\mathrm{Mod}(V), then Mod(V)\mathrm{Mod}(V) naturally embeds as a lattice in the automorphism group of the Cayley graph Cay(Mod(V),S)\mathrm{Cay}(\mathrm{Mod}(V),S), defined as the simplicial graph whose vertices are the elements of Mod(V)\mathrm{Mod}(V), with an edge between two distinct vertices g,hg,h whenever gh1SS1gh^{-1}\in S\cup S^{-1} (this convention excludes for instance loop-edges when SS contains the identity of Mod(V)\mathrm{Mod}(V), or multiple edges if SS contains an element and its inverse). The above rigidity statement about lattice embeddings has the following consequence (which can also be viewed as a very weak form of the conjectural quasi-isometry rigidity statement).

Corollary 3.

Let VV be a connected 33-dimensional handlebody of finite genus at least 33, and let SS be a finite generating set of Mod(V)\mathrm{Mod}(V). Then every graph automorphism of Cay(Mod(V),S)\mathrm{Cay}(\mathrm{Mod}(V),S) is at bounded distance from the left multiplication by an element of Mod(V)\mathrm{Mod}(V).

If Γ\Gamma is a torsion-free finite-index subgroup of Mod(V)\mathrm{Mod}(V), and if SS^{\prime} is a finite generating set of Γ\Gamma, then the automorphism group of Cay(Γ,S)\mathrm{Cay}(\Gamma,S^{\prime}) is countable, and in fact embeds as a subgroup of Mod(V)\mathrm{Mod}(V) containing Γ\Gamma.

The torsion-freeness assumption is crucial in the second part of the statement: for every finitely generated group GG containing a nontrivial torsion element, there exists a finite generating set SS of GG such that the automorphism group of Cay(G,S)\mathrm{Cay}(G,S) is uncountable, as was observed by de la Salle and Tessera in [dlST19, Lemma 6.1].

Thanks to work of Furman [Fur99b], the measure equivalence rigidity statement given in Theorem 1 can also be recast in the language of orbit equivalence rigidity of probability measure-preserving ergodic group actions. We reach the following corollary, analogous to a theorem of Kida [Kid11] for mapping class groups – see Section 4.2 for all definitions.

Corollary 4.

Let VV be a connected 33-dimensional handlebody of finite genus at least 33. Let Γ\Gamma be a countable group. Let Mod(V)X\mathrm{Mod}(V)\curvearrowright X and ΓY\Gamma\curvearrowright Y be two free ergodic measure-preserving group actions by Borel automorphisms on standard probability spaces.

If the actions Mod(V)X\mathrm{Mod}(V)\curvearrowright X and ΓY\Gamma\curvearrowright Y are stably orbit equivalent, then they are virtually isomorphic.

Finally, our work also yields strong rigidity statements for von Neumann algebras associated (via a celebrated construction of Murray and von Neumann [MvN36]) to probability measure-preserving ergodic group actions of handlebody groups. By combining Corollary 4 with the proper proximality of handlebody groups in the sense of Boutonnet, Ioana and Peterson [BIP21] (established in [HHL20]), we reach the following corollary – see Section 4.2 for definitions, and work of Ozawa and Popa [OP10, Definition 3.1] for the notion of a weakly compact group action (as an important example, the action of a residually finite group on its profinite completion is weakly compact).

Corollary 5.

Let VV be a connected 33-dimensional handlebody of finite genus at least 33. Let Γ\Gamma be a countable group. Let Mod(V)X\mathrm{Mod}(V)\curvearrowright X and ΓY\Gamma\curvearrowright Y be two free ergodic measure-preserving group actions by Borel automorphisms on standard probability spaces, and assume that ΓY\Gamma\curvearrowright Y is weakly compact.

If the von Neumann algebras L(X)Mod(V)L^{\infty}(X)\rtimes\mathrm{Mod}(V) and L(Y)ΓL^{\infty}(Y)\rtimes\Gamma are isomorphic, then the actions Mod(V)X\mathrm{Mod}(V)\curvearrowright X and ΓY\Gamma\curvearrowright Y are virtually conjugate.

Proof strategy.

The general strategy of our proof of Theorem 1 follows Kida’s approach for mapping class groups [Kid10]. General techniques from measured group theory, originating in the work of Furman [Fur99a], reduce the proof of Theorem 1 to a cocycle rigidity theorem (Theorem 3.2) for actions of Mod(V)\mathrm{Mod}(V) on standard probability spaces. In order to avoid some finite-order phenomena, it is in fact useful for us to work in a finite-index rotationless subgroup Mod0(V)\mathrm{Mod}^{0}(V) (see Section 1.2 for its precise definition). More precisely, we are given a measured groupoid 𝒢\mathcal{G}, which comes from restricting two actions of Mod0(V)\mathrm{Mod}^{0}(V) on standard finite measure spaces to a positive measure Borel subset YY on which their orbits coincide. The groupoid 𝒢\mathcal{G} is thus equipped with two cocycles ρ1,ρ2:𝒢Mod0(V)\rho_{1},\rho_{2}:\mathcal{G}\to\mathrm{Mod}^{0}(V), given by the two actions: whenever two points x,yYx,y\in Y are joined by an arrow g𝒢g\in\mathcal{G}, there is an element ρ1(g)\rho_{1}(g) sending xx to yy for the first action, and an element ρ2(g)\rho_{2}(g) sending xx to yy for the second action. Our goal is to build a canonical map φ:YMod(V)\varphi:Y\to\mathrm{Mod}(V) such that ρ1\rho_{1} and ρ2\rho_{2} are cohomologous through φ\varphi: this means that whenever x,yYx,y\in Y are joined by an arrow g𝒢g\in\mathcal{G}, then ρ2(g)=φ(y)ρ1(g)φ(x)1\rho_{2}(g)=\varphi(y)\rho_{1}(g)\varphi(x)^{-1}. In fact, using a theorem of Korkmaz and Schleimer which identifies Mod(V)\mathrm{Mod}(V) to the automorphism group of the disk graph 𝔻\mathbb{D} of VV, our goal is to build a (canonical) map YAut(𝔻)Y\to\mathrm{Aut}(\mathbb{D}). Recall that the disk graph is the graph whose vertices are the isotopy classes of meridians in V\partial V (i.e. essential simple closed curves that bound a properly embedded disk in VV), and two vertices are joined by an edge if the corresponding meridians have disjoint representatives in their respective isotopy classes.

In order to build the desired map YAut(𝔻)Y\to\mathrm{Aut}(\mathbb{D}), the main step is to characterize subgroupoids of 𝒢\mathcal{G} that arise as stabilizers of Borel maps Y𝔻Y\to\mathbb{D} in a purely groupoid-theoretic way, i.e. with no reference to the cocycles (so that a vertex stabilizer for ρ1\rho_{1} is also a vertex stabilizer for ρ2\rho_{2}).

In the surface mapping class group setting (where the disk graph is replaced by the curve graph of the surface Σ\Sigma), the important observation made by Kida is the following: curve stabilizers inside Mod(Σ)\mathrm{Mod}(\Sigma) are characterized as maximal nonamenable subgroups of Mod(Σ)\mathrm{Mod}(\Sigma) which contain an infinite amenable normal subgroup (namely, the cyclic subgroup generated by the twist about the curve). This has a groupoid-theoretic analogue, through notions of amenable and normal subgroupoids.

The situation is more complicated for handlebodies, and the above algebraic statement does not give a characterization of meridian stabilizers any longer, for several reasons that we will now explain; for simplicity we will sketch the group-theoretic version of our arguments, but in reality everything has to be phrased in the language of measured groupoids. Our most challenging task, which occupies a large part of Section 3, is in fact to characterize stabilizers of nonseparating meridians. Inspired by the surface setting, we want to start with a maximal nonamenable subgroup HH of Mod0(V)\mathrm{Mod}^{0}(V) which contains an infinite amenable normal subgroup AA. A first bad situation we encounter is the following: AA could be generated by a partial pseudo-Anosov, supported on a subsurface SVS\subseteq\partial V, and HH be its normalizer. In the surface setting considered by Kida [Kid10], such an HH is not maximal, as it is contained in the stabilizer HH^{\prime} of the boundary multicurve γ\gamma of SS. But for us, the group of multitwists about γ\gamma could intersect Mod0(V)\mathrm{Mod}^{0}(V) trivially; in this case HH^{\prime} may not contain any infinite normal amenable subgroup, so HH^{\prime} may not violate the maximality of HH. We resolve this first difficulty by further imposing that HH should not be contained in a subgroup containing two normal nonamenable subgroups that centralize each other (typically, the stabilizers of a subsurface and its complement); this is why we need to exclude separating meridians from our analysis at first. With a bit more work, we manage to reduce to the case where the pair (H,A)(H,A) is given by the following situation: there is a multicurve XX, together with a (possibly empty) collection 𝔄\mathfrak{A} of complementary components of XX labeled active, HH is the stabilizer of XX, and AA is exactly the active subgroup of (X,𝔄)(X,\mathfrak{A}), i.e. the subgroup of the stabilizer of XX acting trivially on all inactive subsurfaces, and it is amenable. This still includes several possibilities: XX could be a nonseparating meridian and 𝔄=\mathfrak{A}=\emptyset (in which case AA is the twist subgroup). But (still with 𝔄=\mathfrak{A}=\emptyset), the multicurve XX could also be of the form α1α2\alpha_{1}\cup\alpha_{2}, where α1\alpha_{1} and α2\alpha_{2} together bound an annulus in VV (see Figure 1): the cyclic subgroup generated by the product of twists Tα1Tα21T_{\alpha_{1}}T_{\alpha_{2}}^{-1} is then normal in the handlebody group stabilizer of the annulus. To exclude annuli (and in fact only retain nonseparating meridians), we use a combinatorial argument: roughly, we can always complete a nonseparating meridian to a collection of 3g33g-3 such, while doing this with annulus pairs will introduce redundancy, as the same curves will be used more than once. Combinatorially, in a collection of 3g33g-3 annuli, it is always possible to remove one without changing the link of the collection in an appropriate graph of disks and annuli.

Once we have characterized nonseparating meridians, we actually have enough information to also recover the separating ones, exploiting that these can be completed to a pair of pants decomposition by adding 3g43g-4 nonseparating meridians. Finally, a characterization of adjacency in the disk graph comes from observing that two meridians are disjoint up to isotopy if the corresponding twists commute, or in other words if these twists together they generate an amenable subgroup of Mod(V)\mathrm{Mod}(V).

Acknowledgments.

The first named author is partially supported by the DFG as part of the SPP 2026 “Geometry at Infinity”. The second named author acknowledges support from the Agence Nationale de la Recherche under Grant ANR-16-CE40-0006 DAGGER.

1 Handlebody and mapping class group facts

In this section, we collect a few facts about handlebody groups that will be useful in the paper. The reader is refered to [Joh95, Hen] for general information about handlebody groups.

1.1 General background

Handlebodies.

By a handlebody of (finite) genus g0g\geq 0, we mean a connected orientable 33-manifold which is a disk-sum of gg copies of D2×S1D^{2}\times S^{1}, where D2D^{2} is a closed disk and S1S^{1} is a circle. The boundary V\partial V of a handlebody VV of genus gg is a closed, connected, orientable surface of the same genus gg. The handlebody group Mod(V)\mathrm{Mod}(V) is the mapping class group of VV, i.e. the group of all isotopy classes of orientation-preserving homeomorphisms of VV. There is a restriction homomorphism Mod(V)Mod(V)\mathrm{Mod}(V)\to\mathrm{Mod}(\partial V), which is injective, thus allowing us to view Mod(V)\mathrm{Mod}(V) as a subgroup of Mod(V)\mathrm{Mod}(\partial V) (see e.g. [Hen, Lemma 3.1]).

Meridians and annuli.

Let VV be a handlebody. Recall that a simple closed curve on V\partial V is essential if it is homotopically nontrivial, i.e. it does not bound a disk on V\partial V. An essential simple closed curve on V\partial V is a meridian (represented in blue in Figure 1) if it bounds a properly embedded disk in VV.

Refer to caption
Figure 1: On the left: a meridian δ\delta, i.e. an essential curve bounding a disk in the handlebody. On the right: two curves α1,α2\alpha_{1},\alpha_{2} which individually do not bound disks in the handlebody, and which are not homotopic on the boundary surface, but bound a properly embedded annulus in the handlebody.

If cVc\subseteq\partial V is a meridian, then the Dehn twist TcT_{c} associated to cc belongs to Mod(V)\mathrm{Mod}(V), viewed as a subgroup of Mod(V)\mathrm{Mod}(\partial V) – and this is in fact a characterisation of meridians, as follows from [McC06, Theorem 1] or [Oer02, Theorem 1.11].

For multitwists, there is another possibility. Namely, a pair {α1,α2}\{\alpha_{1},\alpha_{2}\} of disjoint nonisotopic essential simple closed curves on V\partial V is an annulus pair (represented in red in Figure 1) if neither α1\alpha_{1} nor α2\alpha_{2} is a meridian, and there exists a properly embedded annulus AVA\subseteq V such that A=α1α2\partial A=\alpha_{1}\cup\alpha_{2}. An annulus twist is a mapping class of the form Tα1Tα21T_{\alpha_{1}}T_{\alpha_{2}}^{-1} for some annulus pair {α1,α2}\{\alpha_{1},\alpha_{2}\}. Annulus twists belong to Mod(V)\mathrm{Mod}(V) ([McC06, Theorem 1] or [Oer02, Theorem 1.11]).

Lemma 1.1.

Let cc be a meridian. Then every connected component of Vc\partial V\setminus c supports two handlebody group elements which both restrict to a pseudo-Anosov mapping class of Vc\partial V\setminus c and together generate a nonabelian free subgroup.

Proof.

Let XX be a connected component of Vc\partial V\setminus c which is not a once-holed torus, and denote by c1c_{1} a boundary component of XX (corresponding to one of the sides of cc). Then, for any essential simple closed curve αX\alpha\subset X which is not boundary parallel in XX, we can (and shall) choose an essential simple closed curve αX\alpha^{\prime}\subset X which is not isotopic to α\alpha and bounds a pair of pants on XX together with c1,αc_{1},\alpha (here, we are using that XX is not a once-holed torus). Since cc is a meridian, α,α\alpha,\alpha^{\prime} are either both meridians, or form an annulus pair. Thus, in either case, the multitwist fα=TαTα1f_{\alpha}=T_{\alpha^{\prime}}T_{\alpha}^{-1} is a handlebody group element supported in XX.

By choosing curves α,β\alpha,\beta which fill XX we can thus find fα,fβf_{\alpha},f_{\beta} so that no essential simple closed curve in XX is fixed by both up to isotopy. This implies that the group generated by fα,fβf_{\alpha},f_{\beta} contains a pseudo-Anosov ψ\psi ([Iva92], see also the discussion in [Man13, Section 2.4]). Conjugating ψ\psi by fαf_{\alpha} yields a second one, and sufficiently high powers of ψ\psi and fαψfα1f_{\alpha}\psi f_{\alpha}^{-1} generate a nonabelian free subgroup. ∎

Lemma 1.1 implies in particular that when VV has genus at least 33, the stabilizer in Mod(V)\mathrm{Mod}(V) of every meridian cc in V\partial V contains a nonabelian free subgroup (because at least one connected component of Vc\partial V\setminus c is not a one-holed torus). The requirement of having genus at least 33 is necessary, as the following shows.

Lemma 1.2.

Suppose that cc is a separating meridian, and suppose that XX is a component of Vc\partial V\setminus c which is a once-holed torus. Then XX contains a unique (nonseparating) meridian dXd_{X} which is not peripheral in XX up to isotopy, and therefore

StabMod(V)(c)StabMod(V)(dX).\mathrm{Stab}_{\mathrm{Mod}(V)}(c)\subsetneq\mathrm{Stab}_{\mathrm{Mod}(V)}(d_{X}).

If the genus of VV is at least 33, then dXd_{X} is the only other meridian whose stabiliser contains StabMod(V)(c)\mathrm{Stab}_{\mathrm{Mod}(V)}(c) (or even a finite-index subgroup of StabMod(V)(c)\mathrm{Stab}_{\mathrm{Mod}(V)}(c)).

Proof.

The subsurface XX is the boundary of a once-spotted genus 11 handlebody V11V_{1}^{1}. Hence, there is a nonseparating meridian dXd_{X} contained in XX. We claim that it is the only one up to isotopy. Namely, recall that in a once-holed torus any two isotopically distinct essential simple closed curves have nonzero algebraic intersection number. However, any two meridians have algebraic intersection number zero.

In particular StabMod(V)(c)StabMod(V)(dX)\mathrm{Stab}_{\mathrm{Mod}(V)}(c)\subseteq\mathrm{Stab}_{\mathrm{Mod}(V)}(d_{X}). This inclusion is strict, because there exists a handlebody group element φ\varphi which fixes dXd_{X} and restricts to a pseudo-Anosov homeomorphism on the complementary subsurface, in particular φ\varphi does not fix the isotopy class of cc.

To show the final claim, recall from Lemma 1.1 that there are elements in StabMod(V)(c)\mathrm{Stab}_{\mathrm{Mod}(V)}(c) restricting to pseudo-Anosov elements on any component of Vc\partial V\setminus c which is not a once-holed torus. If the genus of VV is at least 33, the complement of XX will be such a component. Hence, dXd_{X} is the unique other meridian fixed by StabMod(V)(c)\mathrm{Stab}_{\mathrm{Mod}(V)}(c) (or any finite-index subgroup). ∎

1.2 Rotationless mapping classes

In order to avoid finite-order phenomena, it will be useful to work in certain finite index subgroups. We say that a mapping class ff is rotationless (or pure) if the following holds: if a power of ff fixes the isotopy class of a simple closed curve cc, then ff actually fixes the oriented isotopy class of cc.

Let Σ\Sigma be a surface obtained from a closed, connected, orientable surface by removing at most finitely many points. We denote by Mod0(Σ)\mathrm{Mod}^{0}(\Sigma) the (finite index) subgroup of Mod(Σ)\mathrm{Mod}(\Sigma) consisting of mapping classes acting trivially on homology mod 33 of the surface. It is well-known [Iva92] that every fMod0(Σ)f\in\mathrm{Mod}^{0}(\Sigma) is rotationless.

We observe that if ff is rotationless, and SΣS\subset\Sigma is a subsurface which is preserved by ff, then the restriction f|Sf|_{S} is rotationless – however, if fMod0(Σ)f\in\mathrm{Mod}^{0}(\Sigma) then f|Sf|_{S} need not be contained in Mod0(S)\mathrm{Mod}^{0}(S).

To avoid this issue, we use the following construction (which is likely known to experts, but which we were unable to locate in the literature).

Lemma 1.3.

Denote by p:XΣp:X\to\Sigma the mod-22-homology cover of the surface Σ\Sigma. Let Mod1(Σ)\mathrm{Mod}^{1}(\Sigma) be the subgroup of those mapping classes which admit a lift to XX which acts trivially on H1(X;/3)H_{1}(X;\mathbb{Z}/3\mathbb{Z}).

Then Mod1(Σ)\mathrm{Mod}^{1}(\Sigma) is a finite index subgroup of Mod0(Σ)\mathrm{Mod}^{0}(\Sigma), and if hMod1(Σ)h\in\mathrm{Mod}^{1}(\Sigma) is any element preserving a connected subsurface SΣS\subset\Sigma, then the restriction h|Sh|_{S} is an element of Mod0(S)\mathrm{Mod}^{0}(S).

The proof uses the following covering argument.

Lemma 1.4.

Let SΣS\subset\Sigma be an essential connected subsurface. Denote by p:XΣp:X\to\Sigma the mod-22-homology cover, and let XSXX_{S}\subset X be a connected component of p1(S)p^{-1}(S). Then the map

H1(XS;)H1(X;)H_{1}(X_{S};\mathbb{Z})\to H_{1}(X;\mathbb{Z})

induced by the inclusion is injective, and the same is true with \mathbb{Z} replaced with /n\mathbb{Z}/n for any nn.

Proof.

Choose a subsurface YXSY\subset X_{S} with one boundary component, so that XSYX_{S}\setminus Y is a bordered sphere. Denote by δ0,,δk\delta_{0},\ldots,\delta_{k} the boundary components of XSX_{S} (which are all contained in YY). We then have that

H1(XS;)=H1(Y;)k,H_{1}(X_{S};\mathbb{Z})=H_{1}(Y;\mathbb{Z})\oplus\mathbb{Z}^{k},

where the latter summand is [δ1],,[δk]\langle[\delta_{1}],\ldots,[\delta_{k}]\rangle. The first summand injects into H1(X;)H_{1}(X;\mathbb{Z}), since YY is a subsurface of XX with one boundary component.

We now aim to show that for all i>0i>0 there is a curve βi\beta_{i} which is disjoint from YY, intersects δ0,δi\delta_{0},\delta_{i} each in a single point, and is disjoint from all other δj\delta_{j} (i.e. that the complement of XSX_{S} is connected). This will show that [δ1],,[δk][\delta_{1}],\ldots,[\delta_{k}] are linearly independent from each other and from H1(Y;)H_{1}(Y;\mathbb{Z}) in H1(X;)H_{1}(X;\mathbb{Z}) thus showing the lemma.

For simplicity of notation, we will perform the construction only for i=1i=1. Choose a basepoint q~\widetilde{q} in YY, and let q=p(q~)q=p(\widetilde{q}) be its image in Σ\Sigma. Since the mod-22 homology cover is normal (Galois), the preimage p1(q)p^{-1}(q) is exactly the orbit of q~\widetilde{q} under the deck group D=H1(Σ;/2)D=H_{1}(\Sigma;\mathbb{Z}/2).

To describe the intersection p1(q)XSp^{-1}(q)\cap X_{S}, first observe that since XSX_{S} is connected, a point q~p1(q)\widetilde{q}^{\prime}\in p^{-1}(q) is contained in XSX_{S} exactly if there is a path γ~\widetilde{\gamma} connecting q~\widetilde{q} to q~\widetilde{q}^{\prime} contained in XSX_{S}. Such paths are exactly the lifts of loops γ\gamma based at qq which are contained in SS. So q~\widetilde{q}^{\prime} is contained in XSX_{S} if and only if the deck group element gg mapping q~\widetilde{q} to q~\widetilde{q}^{\prime} is the image of some γπ1(S,q)π1(Σ,q)\gamma\in\pi_{1}(S,q)\subseteq\pi_{1}(\Sigma,q). The image of π1(S,q)\pi_{1}(S,q) in the deck group is exactly the subgroup DS=im(H1(S;/2)H1(Σ;/2))D_{S}=\mathrm{im}(H_{1}(S;\mathbb{Z}/2)\to H_{1}(\Sigma;\mathbb{Z}/2)). Together this shows that p1(q)XS=DSq~p^{-1}(q)\cap X_{S}=D_{S}\widetilde{q}.

Similarly, the components of p1(S)p^{-1}(S) can be identified with the cosets of the subgroup DSDD_{S}\subseteq D.

To describe the cover more precisely, we choose curves γi\gamma_{i} based at qq in the following way:

  1. (1)

    The homology classes [γi]=xi[\gamma_{i}]=x_{i} form a basis x1,,xNx_{1},\ldots,x_{N} of H1(Σ;)H_{1}(\Sigma;\mathbb{Z}),

  2. (2)

    x1,,xkx_{1},\ldots,x_{k} is a basis of im(H1(S;)H1(Σ;))\mathrm{im}(H_{1}(S;\mathbb{Z})\to H_{1}(\Sigma;\mathbb{Z})), and the curves γi\gamma_{i} are contained in SS.

  3. (3)

    The curves γi\gamma_{i} for i=k+1,,Ni=k+1,\ldots,N intersect S\partial S in exactly two points.

To see that these curves exist, we argue as follows. Denote by S1,,SrS_{1},\ldots,S_{r} the components of ΣS\Sigma\setminus S. Choose a curve αiSi\alpha_{i}\subset\partial S_{i}. The connectivity of SS implies that for every i{1,,r}i\in\{1,\dots,r\}, the curve αi\alpha_{i} is homologically nontrivial (in H1(Σ)H_{1}(\Sigma)) exactly if Si\partial S_{i} has more than one component. For each boundary curve βSiαi\beta\subset\partial S_{i}\setminus\alpha_{i} we can find a loop γβ\gamma_{\beta} based at qq which intersects S\partial S in two points, one on β\beta and one on αi\alpha_{i}. We can thus choose independent homology classes ziz_{i} defined by curves intersecting S\partial S in at most two points, so that for any xH1(Σ)x\in H_{1}(\Sigma) there is a linear combination zz of the ziz_{i}, so that x+zx+z has algebraic intersection number 0 with all curves in S\partial S. Any such class x+zx+z can be realised by a multicurve disjoint from S\partial S. Since every homology class defined by a curve (without specified basepoint) in SiS_{i} can be realised by a loop based at qq which intersects S\partial S in two points, and every curve in SS can be realised by a loop disjoint from S\partial S the desired existence follows.

Lifting a curve of the type in (2) at a point hq~h\widetilde{q} stays in the same connected component hXShX_{S}, while lifting a curve of the type in (3) joins hXShX_{S} to hXSh^{\prime}X_{S} and intersects hXS\partial hX_{S} in a single point. To see that last claim observe that a lift of a curve as in (3) cannot join two points of hXShX_{S}, as the image of that curve in H1(Σ;/2)H_{1}(\Sigma;\mathbb{Z}/2) would then be contained in im(H1(S;/2)H1(Σ;/2))\mathrm{im}(H_{1}(S;\mathbb{Z}/2)\to H_{1}(\Sigma;\mathbb{Z}/2)), contradicting (1) and (2).

For every i{0,1}i\in\{0,1\}, denote by ZiZ_{i} the component of Xp1(S)X\setminus p^{-1}(S) adjacent to δi\delta_{i}. There are hiim(H1(S;/2)H1(Σ;/2))h_{i}\notin\mathrm{im}(H_{1}(S;\mathbb{Z}/2)\to H_{1}(\Sigma;\mathbb{Z}/2)), so that hiXSh_{i}X_{S} are surfaces adjacent to ZiZ_{i}. Namely, either p(Zi)p(Z_{i}) has genus (which is automatically the case if p(δi)p(\delta_{i}) is separating), and contains a curve defining one of the xjx_{j} (of the second type), or p(Zi)p(Z_{i}) is a punctured sphere so that for the boundary component p(δi)p(\delta_{i}) there is some xj,j>kx_{j},j>k (of the third type) which intersects it once (namely, if all xix_{i} would interesect p(δi)p(\delta_{i}) in an even number of points, the xix_{i} could not be a basis of H1(Σ;)H_{1}(\Sigma;\mathbb{Z}), since p(δi)p(\delta_{i}) is nonseparating). In both cases the desired component is ±[xj]XS\pm[x_{j}]X_{S}. Choose paths ci,i=0,1c_{i},i=0,1 joining q~\widetilde{q} to hiq~h_{i}\widetilde{q} which intersect only δi\delta_{i} among the δj\delta_{j}.

Since im(H1(S;/2)H1(Σ;/2))=(/2)k\mathrm{im}(H_{1}(S;\mathbb{Z}/2)\to H_{1}(\Sigma;\mathbb{Z}/2))=(\mathbb{Z}/2)^{k} is a subgroup of H1(Σ;/2))=(/2)NH_{1}(\Sigma;\mathbb{Z}/2))=(\mathbb{Z}/2)^{N} generated by a subset of the generators, there is a path in the Cayley graph of H1(Σ;/2)H_{1}(\Sigma;\mathbb{Z}/2) from h0h_{0} to h1h_{1} which is disjoint from the Cayley graph of the subgroup im(H1(S;/2)H1(Σ;/2)\mathrm{im}(H_{1}(S;\mathbb{Z}/2)\to H_{1}(\Sigma;\mathbb{Z}/2). Each edge in such a path corresponds to a right multiplication hhxsh\mapsto hx_{s}, and we can choose a corresponding path joining hq~h\widetilde{q} to hxsq~hx_{s}\widetilde{q} which is disjoint from XSX_{S}. By concatenating these paths with c0,c1c_{0},c_{1} (in the right order) we then find the desired path β1\beta_{1}. ∎

Lemma 1.3 is now an immediate consequence of the following corollary.

Corollary 1.5.

Suppose that ff is a mapping class so that

  1. 1.

    ff admits a lift f~\widetilde{f} to the mod-22-homology-cover XX, which acts trivially on H1(X;/3)H_{1}(X;\mathbb{Z}/3)

  2. 2.

    ff preserves a subsurface SΣS\subset\Sigma

Then the restriction f|Sf|_{S} acts trivially on H1(S;/3)H_{1}(S;\mathbb{Z}/3).

Proof.

Let αS\alpha\subset S be a simple closed curve which is part of a basis for H1(S;/3)H_{1}(S;\mathbb{Z}/3). Then there is a power N=2nN=2^{n} so that αN\alpha^{N} lifts to a curve α~XS\widetilde{\alpha}\subset X_{S} (with notation as in the previous lemma).

Denote by pS:XSSp_{S}:X_{S}\to S the restriction of the covering map (which is then also a covering). We have (pS)[α~]=N[α](p_{S})_{\ast}[\widetilde{\alpha}]=N[\alpha]. Since NN is invertible mod 33, there is a multiple kk so that (pS)k[α~]=[α](p_{S})_{\ast}k[\widetilde{\alpha}]=[\alpha] mod 33.

By Lemma 1.4, H1(XS;/3)H_{1}(X_{S};\mathbb{Z}/3) is a subspace of H1(X;/3)H_{1}(X;\mathbb{Z}/3). Since f~\widetilde{f} acts trivially on H1(X;/3)H_{1}(X;\mathbb{Z}/3), this implies that the restriction f~XS\widetilde{f}_{X_{S}} acts trivially on H1(XS;/3)H_{1}(X_{S};\mathbb{Z}/3). Hence, we have (f~XS)k[α~]=k[α~](\widetilde{f}_{X_{S}})_{\ast}k[\widetilde{\alpha}]=k[\widetilde{\alpha}]. Since f~XS\widetilde{f}_{X_{S}} is a lift of fSf_{S} this implies (fS)[α]=[α](f_{S})_{\ast}[\alpha]=[\alpha]. ∎

In the sequel of the paper, we will always let Mod0(V)=Mod(V)Mod1(V)\mathrm{Mod}^{0}(V)=\mathrm{Mod}(V)\cap\mathrm{Mod}^{1}(\partial V), where Mod1(V)\mathrm{Mod}^{1}(\partial V) is as in Lemma 1.3.

1.3 Infinite conjugacy classes

A countable group GG is said to be ICC (standing for infinite conjugacy classes) if the conjugacy class of every nontrivial element of GG is infinite.

Lemma 1.6.

Let VV be a handlebody of genus at least 22, and let φMod(V)\varphi\in\mathrm{Mod}(V) be a handlebody group element. Then either the conjugacy class of φ\varphi is infinite, or φ\varphi fixes the isotopy class of every meridian.

In particular, when the genus of VV is at least 33, the group Mod(V)\mathrm{Mod}(V) is ICC.

We remark that in genus 22, the hyperelliptic involution fixes the isotopy class of every essential simple closed curve on V\partial V, and its conjugacy class is finite in Mod(V)\mathrm{Mod}(V).

Proof.

Suppose that φ\varphi is an element with finite conjugacy class. For any meridian cc, consider the elements TciφTciT_{c}^{i}\varphi T_{c}^{-i} for ii\in\mathbb{N}. By finiteness of the conjugacy class, two of these have to be equal, and thus there is some N>0N>0 so that

TcNφTcN=φ,T_{c}^{N}\varphi T_{c}^{-N}=\varphi,

or equivalently,

TcN=φTcNφ1=Tφ(c)N.T_{c}^{N}=\varphi T_{c}^{N}\varphi^{-1}=T_{\varphi(c)}^{N}.

This implies cc is isotopic to φ(c)\varphi(c), see e.g. [FM12, Section 3.3]. The first part of the lemma follows since cc was arbitrary. The fact that Mod(V)\mathrm{Mod}(V) is ICC when the genus is at least 33 follows because every element fixing the isotopy class of every meridian is then trivial [KS09, Theorem 9.4]. ∎

2 Background on measured groupoids

The reader is refered to [AD13, Section 2.1], [Kid09] or [GH21, Section 3] for general background on measured groupoids.

Recall that a standard Borel space is a measurable space associated to a Polish space (i.e. separable and completely metrizable). A standard probability space is a standard Borel space equipped with a Borel probability measure.

A Borel groupoid is a standard Borel space 𝒢\mathcal{G} (whose elements are thought of as being arrows) equipped with two Borel maps s,r:𝒢Ys,r:\mathcal{G}\to Y towards a standard Borel space YY (giving the source and range of an arrow), and coming with a measurable composition law and inverse map and with a unit element eye_{y} per yYy\in Y. The Borel space YY is called the base space of the groupoid 𝒢\mathcal{G}. All Borel groupoids considered in the present paper are assumed to be discrete, i.e. there are countably many arrows in 𝒢\mathcal{G} with a given range (or source). It follows from a theorem of Lusin and Novikov (see e.g. [Kec95, Theorem 18.10]) that a discrete Borel groupoid 𝒢\mathcal{G} can always be written as a countable disjoint union of bisections, i.e. Borel subsets BB of 𝒢\mathcal{G} on which ss and rr are injective (in which case s(B)s(B) and r(B)r(B) are Borel subsets of YY, see [Kec95, Corollary 15.2]). A Borel groupoid 𝒢\mathcal{G} with base space YY is trivial if 𝒢={ey|yY}\mathcal{G}=\{e_{y}|y\in Y\}.

A finite Borel measure μ\mu on YY is quasi-invariant for the groupoid 𝒢\mathcal{G} if for every bisection B𝒢B\subseteq\mathcal{G}, one has μ(s(B))>0\mu(s(B))>0 if and only if μ(r(B))>0\mu(r(B))>0. A measured groupoid is a Borel groupoid together with a quasi-invariant finite Borel measure on its base space YY.

An important example of a measured groupoid to keep in mind is the following: when a countable group GG acts on a standard probability space YY by Borel automorphisms in a quasi-measure-preserving way, then G×YG\times Y has a natural structure of a measured groupoid over YY, denoted by GYG\ltimes Y: the source and range maps are given by s(g,y)=ys(g,y)=y and r(g,y)=gyr(g,y)=gy, the composition law is (g,hy)(h,y)=(gh,y)(g,hy)(h,y)=(gh,y), the inverse of (g,y)(g,y) is (g1,gy)(g^{-1},gy) and the units are ey=(e,y)e_{y}=(e,y).

A Borel subset 𝒢\mathcal{H}\subseteq\mathcal{G} which is stable under composition and inverse and contains all unit elements of 𝒢\mathcal{G} has the structure of a measured subgroupoid of 𝒢\mathcal{G} over the same base space YY. Given a Borel subset UYU\subseteq Y, the restriction 𝒢|U\mathcal{G}_{|U} is the measured groupoid over UU defined by only keeping the arrows whose source and range both belong to UU. Given two subgroupoids ,𝒢\mathcal{H},\mathcal{H}^{\prime}\subseteq\mathcal{G}, we denote by ,\langle\mathcal{H},\mathcal{H}^{\prime}\rangle the subgroupoid of 𝒢\mathcal{G} generated by \mathcal{H} and \mathcal{H}^{\prime}, i.e. the smallest subgroupoid of 𝒢\mathcal{G} containing \mathcal{H} and \mathcal{H}^{\prime} (it is made of all arrows obtained as finite compositions of arrows in \mathcal{H} and arrows in \mathcal{H}^{\prime}).

A measured groupoid 𝒢\mathcal{G} with base space YY is of infinite type if for every positive measure Borel subset UYU\subseteq Y and almost every yUy\in U, there are infinitely many elements of 𝒢|U\mathcal{G}_{|U} with source yy. Observe that if 𝒢\mathcal{G} is of infinite type, then for every Borel subset UYU\subseteq Y of positive measure, the restricted groupoid 𝒢|U\mathcal{G}_{|U} is again of infinite type.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, and let GG be a countable group. A strict cocycle ρ:𝒢G\rho:\mathcal{G}\to G is a Borel map such that for all g1,g2𝒢g_{1},g_{2}\in\mathcal{G}, if the source of g1g_{1} is equal to the range of g2g_{2} (so that the product g1g2g_{1}g_{2} is well-defined), then ρ(g1g2)=ρ(g1)ρ(g2)\rho(g_{1}g_{2})=\rho(g_{1})\rho(g_{2}). The kernel of a cocycle ρ\rho is the subgroupoid of 𝒢\mathcal{G} made of all g𝒢g\in\mathcal{G} such that ρ(g)=1\rho(g)=1. We say that ρ\rho has trivial kernel if its kernel is equal to the trivial subgroupoid of 𝒢\mathcal{G}, i.e. it only consists of the unit elements of 𝒢\mathcal{G}. We say that a strict cocycle 𝒢G\mathcal{G}\to G is action-type if ρ\rho has trivial kernel, and whenever HGH\subseteq G is an infinite subgroup, and UYU\subseteq Y is a Borel subset of positive measure, then ρ1(H)|U\rho^{-1}(H)_{|U} is a subgroupoid of 𝒢|U\mathcal{G}_{|U} of infinite type. An important example is that given a measure-preserving GG-action on a standard probability space YY, the natural cocycle ρ:GYG\rho:G\ltimes Y\to G is action-type [Kid09, Proposition 2.26]. We warn the reader that in the latter example, it is important that the GG-action on YY preserves the measure, as opposed to only quasi-preserving it.

Given a Polish space Δ\Delta equipped with a GG-action by Borel automorphisms, we say that a measurable map φ:YΔ\varphi:Y\to\Delta is (𝒢,ρ)(\mathcal{G},\rho)-equivariant if there exists a conull Borel subset YYY^{*}\subseteq Y such that for every g𝒢|Yg\in\mathcal{G}_{|Y^{*}}, one has φ(r(g))=ρ(g)φ(s(g))\varphi(r(g))=\rho(g)\varphi(s(g)). We say that an element δΔ\delta\in\Delta is (𝒢,ρ)(\mathcal{G},\rho)-invariant if the constant map with value δ\delta is (𝒢,ρ)(\mathcal{G},\rho)-equivariant (equivalently, there exists a conull Borel subset YYY^{*}\subseteq Y such that ρ(𝒢|Y)StabG(δ)\rho(\mathcal{G}_{|Y})\subseteq\mathrm{Stab}_{G}(\delta)). The (𝒢,ρ)(\mathcal{G},\rho)-stabilizer of δ\delta is the subgroupoid of 𝒢\mathcal{G} made of all elements gg such that ρ(g)StabG(δ)\rho(g)\in\mathrm{Stab}_{G}(\delta). A measurable map φ:YΔ\varphi:Y\to\Delta is stably (𝒢,ρ)(\mathcal{G},\rho)-equivariant if one can partition YY into at most countably many Borel subsets YiY_{i} such that for every ii, the map φ|Yi\varphi_{|Y_{i}} is (𝒢|Yi,ρ)(\mathcal{G}_{|Y_{i}},\rho)-equivariant.

Given two measured subgroupoids ,𝒢\mathcal{H},\mathcal{H}^{\prime}\subseteq\mathcal{G}, we say that \mathcal{H} is stably contained in \mathcal{H}^{\prime} if there exist a conull Borel subset YYY^{*}\subseteq Y and a partition Y=iIYiY^{*}=\sqcup_{i\in I}Y_{i} into at most countably many Borel subsets such that for every iIi\in I, one has |Yi|Yi\mathcal{H}_{|Y_{i}}\subseteq\mathcal{H}^{\prime}_{|Y_{i}}. We say that \mathcal{H} and \mathcal{H}^{\prime} are stably equal if there exist a conull Borel subset and a partition as above such that for every iIi\in I, one has |Yi=|Yi\mathcal{H}_{|Y_{i}}=\mathcal{H}^{\prime}_{|Y_{i}}. We say that \mathcal{H} is stably trivial if it is stably equal to the trivial subgroupoid of 𝒢\mathcal{G}.

Let \mathcal{H} be a measured subgroupoid of 𝒢\mathcal{G}, and B𝒢B\subseteq\mathcal{G} be a bisection. We say that \mathcal{H} is BB-invariant if there exists a conull Borel subset YYY^{*}\subseteq Y such that for every g1,g2B𝒢|Yg_{1},g_{2}\in B\cap\mathcal{G}_{|Y^{*}} and every h𝒢|Yh\in\mathcal{G}_{|Y^{*}} such that the composition g2hg11g_{2}hg_{1}^{-1} is well-defined, we have h|Yh\in\mathcal{H}_{|Y^{*}} if and only if g2hg11|Yg_{2}hg_{1}^{-1}\in\mathcal{H}_{|Y^{*}}. Let now \mathcal{H}^{\prime} be another measured subgroupoid of 𝒢\mathcal{G}. The groupoid \mathcal{H} is normalized by \mathcal{H}^{\prime} if \mathcal{H}^{\prime} can be covered by countably many bisections Bn𝒢B_{n}\subseteq\mathcal{G} in such a way that \mathcal{H} is BnB_{n}-invariant for every nn\in\mathbb{N}. The subgroupoid \mathcal{H} is stably normalized by \mathcal{H}^{\prime} if one can partition YY into at most countably many Borel subsets YiY_{i} in such a way that for every ii, the groupoid |Yi\mathcal{H}_{|Y_{i}} is normalized by |Yi\mathcal{H}^{\prime}_{|Y_{i}}. When \mathcal{H}\subseteq\mathcal{H}^{\prime}, we will simply say that \mathcal{H} is stably normal in \mathcal{H}^{\prime}.

There is a notion of amenability of a measured groupoid, generalizing Zimmer’s notion of amenability of a group action, for which we refer to [Kid09]; here we only list the properties of amenable groupoids we will need. First, if 𝒢\mathcal{G} is amenable and comes with a cocycle ρ:𝒢G\rho:\mathcal{G}\to G towards a countable group GG, and if GG acts by homeomorphisms on a compact metrizable space KK, then there exists a (𝒢,ρ)(\mathcal{G},\rho)-equivariant Borel map YProb(K)Y\to\mathrm{Prob}(K), see [Kid09, Proposition 4.14]. Here Prob(K)\mathrm{Prob}(K) denotes the set of Borel probability measures on KK, equipped with the weak-\ast topology coming from the duality with the space of real-valued continuous functions on KK given by the Riesz–Markov–Kakutani theorem. Second, whenever ρ:𝒢G\rho:\mathcal{G}\to G is a cocycle with trivial kernel, and AGA\subseteq G is an amenable subgroup of GG, then ρ1(A)\rho^{-1}(A) is an amenable subgroupoid of 𝒢\mathcal{G} (see e.g. [GH21, Corollary 3.39]). Amenability is stable under subgroupoids and restrictions. Furthermore, if there exists a conull Borel subset YYY^{*}\subseteq Y and a partition Y=iIYiY^{*}=\sqcup_{i\in I}Y_{i} into at most countably many Borel subsets such that for every iIi\in I, the groupoid 𝒢|Yi\mathcal{G}_{|Y_{i}} is amenable, then 𝒢\mathcal{G} is amenable (this is immediate with the definition of amenability given in [GH21, Definition 3.33], see also [GH21, Remark 3.34] for the comparison to equivalent definitions).

A groupoid 𝒢\mathcal{G} over a standard probability space YY is everywhere nonamenable if for every Borel subset UYU\subseteq Y of positive measure, the groupoid 𝒢|U\mathcal{G}_{|U} is nonamenable.

3 Measure equivalence rigidity of the handlebody group

In this section, we prove the main theorem of the present paper.

Theorem 3.1.

Let VV be a handlebody of genus at least 33. Then Mod(V)\mathrm{Mod}(V) is ME-superrigid.

Using the fact that Mod(V)\mathrm{Mod}(V) is ICC (Lemma 1.6), Theorem 3.1 is a consequence of the following statement combined with [GH21, Theorem 4.5] (which builds on earlier works of Furman [Fur99a, Fur99b] and Kida [Kid10]).

Theorem 3.2.

Let VV be a handlebody of genus at least 33. Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY (with source map ss and range map rr), and let ρ1,ρ2:𝒢Mod0(V)\rho_{1},\rho_{2}:\mathcal{G}\to\mathrm{Mod}^{0}(V) be two strict action-type cocycles.

Then there exist a Borel map θ:YMod(V)\theta:Y\to\mathrm{Mod}(V) and a conull Borel subset YYY^{*}\subseteq Y such that for all g𝒢|Yg\in\mathcal{G}_{|Y^{*}}, one has ρ1(g)=θ(r(g))1ρ2(g)θ(s(g))\rho_{1}(g)=\theta(r(g))^{-1}\rho_{2}(g)\theta(s(g)).

Remark 3.3.

The case where YY is reduced to a point is actually already relevant: if ff is an automorphism of Mod0(V)\mathrm{Mod}^{0}(V), then the group Mod0(V)\mathrm{Mod}^{0}(V), viewed as a groupoid over a point, comes equipped with two (action-type) cocycles towards Mod0(V)\mathrm{Mod}^{0}(V), given by the identity and ff. The conclusion in this case is that every automorphism of Mod0(V)\mathrm{Mod}^{0}(V) is a conjugation. More generally, a consequence of Theorem 3.2 is that the natural map from Mod(V)\mathrm{Mod}(V) to its abstract commensurator is an isomorphism (using that Mod(V)\mathrm{Mod}(V) is ICC for its injectivity). Our work therefore recovers the commensurator rigidity statement from [Hen18].

The rest of the section is devoted to the proof of Theorem 3.2. Starting from a measured groupoid 𝒢\mathcal{G} with two action-type cocycles ρ1,ρ2\rho_{1},\rho_{2} towards Mod0(V)\mathrm{Mod}^{0}(V), we ultimately aim to show that subgroupoids of 𝒢\mathcal{G} corresponding to meridian stabilizers for ρ1\rho_{1} - in the precise sense that they are of meridian type as in Definition 3.4 below - are also of meridian type with respect to ρ2\rho_{2}. Additionally, we will prove that the property that two subgroupoids stabilize disjoint meridians is also independent of the action-type cocycle we choose. This will be used to build a canonical map θ\theta from the base space YY of the groupoid 𝒢\mathcal{G} to the group of all automorphisms of the disk graph. We will finally appeal to the theorem of Korkmaz and Schleimer [KS09] saying that the automorphism group of the disk graph is precisely Mod(V)\mathrm{Mod}(V) to conclude. We make the following definition.

Definition 3.4 (Subgroupoids of meridian type).

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, and let ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V) be a strict cocycle. A measured subgroupoid \mathcal{H} of 𝒢\mathcal{G} is of meridian type with respect to ρ\rho if there exists a conull Borel subset YYY^{*}\subseteq Y and a partition Y=iIYiY^{*}=\sqcup_{i\in I}Y_{i} into at most countably many Borel subsets such that for every iIi\in I, the groupoid |Yi\mathcal{H}_{|Y_{i}} is equal to the (𝒢|Yi,ρ)(\mathcal{G}_{|Y_{i}},\rho)-stabilizer of the isotopy class of a meridian cic_{i}.

When \mathcal{H} can be written as in Definition 3.4, we say that the map φ\varphi sending every yYiy\in Y_{i} to the isotopy class of the meridian cic_{i} is a meridian map for (,ρ)(\mathcal{H},\rho). The essential uniqueness of this map (i.e. the fact that, up to measure 0, it does not depend on the choice of a partition and meridians cic_{i} as above) will follow from Lemmas 3.13 and 3.14.

Likewise, we define the notions of subgroupoids of nonseparating meridian type, and of separating meridian type, by respectively requiring cic_{i} to be nonseparating, or separating. Before completing our characterisation of subgroupoids of meridian type in Proposition 3.36, we will go through successive characterisations of subgroupoids of nonseparating-meridian type (Section 3.9) and of separating-meridian type (Section 3.10).

3.1 Groupoids with cocycles to a free group, after Adams [Ada94], Kida [Kid10]

Throughout the paper, we will work with the following definition.

Definition 3.5 (Strongly Schottky pairs of subgroupoids).

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY. A strongly Schottky pair of subgroupoids of 𝒢\mathcal{G} is a pair (𝒜1,𝒜2)(\mathcal{A}^{1},\mathcal{A}^{2}) of amenable subgroupoids of 𝒢\mathcal{G} of infinite type such that for every Borel subset UYU\subseteq Y of positive measure, there exists a Borel subset VUV\subseteq U of positive measure such that every normal amenable subgroupoid of 𝒜|V1,𝒜|V2\langle\mathcal{A}^{1}_{|V},\mathcal{A}^{2}_{|V}\rangle is stably trivial.

We observe that this notion is stable under restrictions: if (𝒜1,𝒜2)(\mathcal{A}^{1},\mathcal{A}^{2}) is a strongly Schottky pair of subgroupoids of 𝒢\mathcal{G}, then for every Borel subset UYU\subseteq Y of positive measure, the pair (𝒜|U1,𝒜|U2)(\mathcal{A}^{1}_{|U},\mathcal{A}^{2}_{|U}) is a strongly Schottky pair of subgroupoids of 𝒢|U\mathcal{G}_{|U}. In addition, this notion is stable under stabilization: given a pair (𝒜1,𝒜2)(\mathcal{A}^{1},\mathcal{A}^{2}) of subgroupoids of 𝒢\mathcal{G}, and a partition Y=iIYiY=\sqcup_{i\in I}Y_{i} into at most countably many Borel subsets, if (𝒜|Yi1,𝒜|Yi2)(\mathcal{A}^{1}_{|Y_{i}},\mathcal{A}^{2}_{|Y_{i}}) is a strongly Schottky pair of subgroupoids of 𝒢|Yi\mathcal{G}_{|Y_{i}} for every iIi\in I, then (𝒜1,𝒜2)(\mathcal{A}^{1},\mathcal{A}^{2}) is a strongly Schottky pair of subgroupoids of 𝒢\mathcal{G}.

Notice that the last conclusion implies in particular that 𝒜|V1,𝒜|V2\langle\mathcal{A}^{1}_{|V},\mathcal{A}^{2}_{|V}\rangle is nonamenable. So the existence of a strongly Schottky pair of subgroupoids of 𝒢\mathcal{G} forces 𝒢\mathcal{G} to be everywhere nonamenable.

Definition 3.5 is a strengthening of the notion of a Schottky pair of subgroupoids from [GH21, Definition 13.1], which only required the groupoid 𝒜|U1,𝒜|U2\langle\mathcal{A}^{1}_{|U},\mathcal{A}^{2}_{|U}\rangle to be nonamenable. The following lemma is a variation over arguments of Adams [Ada94, Section 3] and Kida [Kid10, Lemma 3.20], and gives the main example of a strongly Schottky pair of subgroupoids.

Lemma 3.6.

Let GG be a countable group, and let g,hGg,h\in G be two elements that generate a nonabelian free subgroup of GG. Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict action-type cocycle ρ:𝒢G\rho:\mathcal{G}\to G.

Then (ρ1(g),ρ1(h))(\rho^{-1}(\langle g\rangle),\rho^{-1}(\langle h\rangle)) is a strongly Schottky pair of subgroupoids of 𝒢\mathcal{G} (in particular 𝒢\mathcal{G} is everywhere nonamenable).

In the following proof, whenever Δ\Delta is a Polish space, the set Prob(Δ)\mathrm{Prob}(\Delta) of all Borel probability measures on Δ\Delta is equipped with the topology generated by the maps μXf𝑑μ\mu\mapsto\int_{X}fd\mu, where ff varies over the set of all real-valued bounded continuous functions. When Δ\Delta is compact, this is nothing but the weak-\ast topology coming from the duality given by the Riesz–Markov–Kakutani theorem. When Δ\Delta is a countable discrete space, this is nothing but the topology of pointwise convergence. The reader is refered to [Kec95, Section 17.E] for more information and basic facts regarding the Borel structure on Prob(Δ)\mathrm{Prob}(\Delta) which justify the measurability of all maps in the following proof.

Proof of Lemma 3.6.

As g\langle g\rangle and h\langle h\rangle are amenable subgroups of GG and ρ\rho has trivial kernel, the subgroupoids ρ1(g)\rho^{-1}(\langle g\rangle) and ρ1(h)\rho^{-1}(\langle h\rangle) are amenable. As g\langle g\rangle and h\langle h\rangle are infinite and ρ\rho is action-type, the subgroupoids ρ1(g)\rho^{-1}(\langle g\rangle) and ρ1(h)\rho^{-1}(\langle h\rangle) are of infinite type.

Now it is enough to prove that if UYU\subseteq Y is a Borel subset of positive measure, and 𝒜\mathcal{A} is a normal amenable subgroupoid of ρ1(g)|U,ρ1(h)|U\langle\rho^{-1}(\langle g\rangle)_{|U},\rho^{-1}(\langle h\rangle)_{|U}\rangle, then 𝒜\mathcal{A} is stably trivial.

Let TT be the Cayley tree of the free group F=g,hF=\langle g,h\rangle, with respect to the generating set {g,h}\{g,h\}. The FF-action on TT by isometries extends to an FF-action on T\partial_{\infty}T by homeomorphisms. As 𝒜\mathcal{A} is amenable and T\partial_{\infty}T is compact and metrizable, there exists an (𝒜,ρ)(\mathcal{A},\rho)-equivariant Borel map UProb(T)U\to\mathrm{Prob}(\partial_{\infty}T), see [Kid09, Proposition 4.14].

We claim that we can find a (possibly null) Borel subset U1UU_{1}\subseteq U of maximal measure such that there exists a Borel map μ:U1Prob(T)\mu:U_{1}\to\mathrm{Prob}(\partial_{\infty}T) which is stably (𝒜|U1,ρ)(\mathcal{A}_{|U_{1}},\rho)-equivariant and such that for every yU1y\in U_{1}, the support of the measure μ(y)\mu(y) has cardinality at least 33. Indeed, let α\alpha be the supremum of the measures of all Borel subsets of UU having the above property, and let (U1,n)n(U_{1,n})_{n\in\mathbb{N}} be a measure-maximizing sequence of such sets – in particular, for every nn\in\mathbb{N}, we have a Borel map μn:U1,nProb(T)\mu_{n}:U_{1,n}\to\mathrm{Prob}(\partial_{\infty}T) as above. Then the countable union U1,U_{1,\infty} of all subsets U1,nU_{1,n} has measure at least α\alpha, and we claim that it also satisfies the above property. To see this, inductively define V1,0=U1,0V_{1,0}=U_{1,0} and V1,n=U1,n(U1,nV1,n1)V_{1,n}=U_{1,n}\setminus(U_{1,n}\cap V_{1,n-1}). Then the map μ:U1,Prob(T)\mu_{\infty}:U_{1,\infty}\to\mathrm{Prob}(\partial_{\infty}T), defined to coincide with μn\mu_{n} when restricted to V1,nV_{1,n}, is stably (𝒜|U1,,ρ)(\mathcal{A}_{|U_{1,\infty}},\rho)-equivariant, and for every yU1,y\in U_{1,\infty}, the support of the measure μ(y)\mu(y) has cardinality at least 33. This concludes the proof of our claim.

From now on, we choose a Borel subset U1UU_{1}\subseteq U as in the above claim. We first prove that 𝒜|U1\mathcal{A}_{|U_{1}} is stably trivial. Up to partitioning U1U_{1} into at most countably many Borel subsets, we can assume that the map μ1\mu_{1} is (𝒜|U1,ρ)(\mathcal{A}_{|U_{1}},\rho)-equivariant (and not just stably equivariant). For every yU1y\in U_{1}, the probability measure μ(y)μ(y)μ(y)\mu(y)\otimes\mu(y)\otimes\mu(y) on (T)3(\partial_{\infty}T)^{3} gives positive measure to the FF-invariant subset (T)(3)(\partial_{\infty}T)^{(3)} made of pairwise distinct triples. Thus, after restricting this measure to (T)(3)(\partial_{\infty}T)^{(3)} and renormalizing to turn this restricted measure into a probability measure, we get an (𝒜|U1,ρ)(\mathcal{A}_{|U_{1}},\rho)-equivariant Borel map U1Prob((T)(3))U_{1}\to\mathrm{Prob}((\partial_{\infty}T)^{(3)}). Now, denoting by V(T)V(T) the vertex set of TT, there is a natural FF-equivariant barycenter map (T)(3)V(T)(\partial_{\infty}T)^{(3)}\to V(T). By pushing the probability measures through this map, we get an (𝒜|U1,ρ)(\mathcal{A}_{|U_{1}},\rho)-equivariant Borel map U1Prob(V(T))U_{1}\to\mathrm{Prob}(V(T)). Let 𝒫<(V(T))\mathcal{P}_{<\infty}(V(T)) be the set of all nonempty finite subsets of V(T)V(T). As V(T)V(T) is countable, there is also a natural FF-equivariant Borel map Prob(V(T))𝒫<(V(T))\mathrm{Prob}(V(T))\to\mathcal{P}_{<\infty}(V(T)), sending a probability measure ν\nu to the finite subset of V(T)V(T) made of all vertices that have maximal ν\nu-measure. We thus derive an (𝒜|U1,ρ)(\mathcal{A}_{|U_{1}},\rho)-equivariant Borel map ϕ:U1𝒫<(V(T))\phi:U_{1}\to\mathcal{P}_{<\infty}(V(T)). As 𝒫<(V(T))\mathcal{P}_{<\infty}(V(T)) is countable, we can then find a Borel partition U1=iIU1,iU_{1}=\sqcup_{i\in I}U_{1,i} into at most countably many Borel subsets such that for every iIi\in I, the map ϕ|U1,i\phi_{|U_{1,i}} is constant, with value a nonempty finite set i\mathcal{F}_{i} of vertices of TT. In other words, there exists a conull Borel subset U1,iU1,iU_{1,i}^{*}\subseteq U_{1,i} such that ρ(𝒜|U1,i)\rho(\mathcal{A}_{|U_{1,i}^{*}}) is contained in the FF-stabilizer of i\mathcal{F}_{i}. As this stabilizer is trivial and ρ\rho has trivial kernel, it follows that 𝒜|U1\mathcal{A}_{|U_{1}} is stably trivial.

We will now prove that U2=UU1U_{2}=U\setminus U_{1} is a null set, which will conclude the proof of the lemma. So assume towards a contradiction that U2U_{2} has positive measure. We know that there exists an (𝒜|U2,ρ)(\mathcal{A}_{|U_{2}},\rho)-equivariant Borel map μ:U2Prob(T)\mu:U_{2}\to\mathrm{Prob}(\partial_{\infty}T), and that for every such map and almost every yU2y\in U_{2}, the support of μ(y)\mu(y) has cardinality at most 22. Let 𝒫2(T)\mathcal{P}_{\leq 2}(\partial_{\infty}T) be the set of all nonempty subsets of T\partial_{\infty}T of cardinality at most 22. As in [Ada94, Lemma 3.2], we can thus find an (𝒜|U2,ρ)(\mathcal{A}_{|U_{2}},\rho)-equivariant Borel map θmax:U2𝒫2(T)\theta_{\max}:U_{2}\to\mathcal{P}_{\leq 2}(\partial_{\infty}T) which is maximal in the sense that for every other (𝒜|U2,ρ)(\mathcal{A}_{|U_{2}},\rho)-equivariant Borel map θ:U2𝒫2(T)\theta:U_{2}\to\mathcal{P}_{\leq 2}(\partial_{\infty}T) and a.e. yYy\in Y, one has θ(y)θmax(y)\theta(y)\subseteq\theta_{\max}(y). Being canonical, the map θmax\theta_{\max} is then equivariant under the groupoid ρ1(g)|U2,ρ1(h)|U2\langle\rho^{-1}(\langle g\rangle)_{|U_{2}},\rho^{-1}(\langle h\rangle)_{|U_{2}}\rangle which normalizes 𝒜|U2\mathcal{A}_{|U_{2}}. Recall that the groupoid ρ1(g)|U2\rho^{-1}(\langle g\rangle)_{|U_{2}} is amenable and of infinite type. Therefore, repeating the argument from the present proof shows that there exists a maximal (ρ1(g)|U2,ρ)(\rho^{-1}(\langle g\rangle)_{|U_{2}},\rho)-equivariant Borel map U2𝒫2(T)U_{2}\to\mathcal{P}_{\leq 2}(\partial_{\infty}T), and this must then be the constant map with value {g,g+}\{g^{-\infty},g^{+\infty}\}. Likewise, the constant map with value {h,h+}\{h^{-\infty},h^{+\infty}\} is the maximal (ρ1(h)|U2,ρ)(\rho^{-1}(\langle h\rangle)_{|U_{2}},\rho)-equivariant Borel map U2𝒫2(T)U_{2}\to\mathcal{P}_{\leq 2}(\partial_{\infty}T). As {g,g+}{h,h+}=\{g^{-\infty},g^{+\infty}\}\cap\{h^{-\infty},h^{+\infty}\}=\emptyset, we have reached a contradiction. This completes our proof. ∎

3.2 Canonical reduction sets, after Kida [Kid08a]

In this section, we review work of Kida [Kid08a] regarding groupoids with cocycles towards a surface mapping class group. Since our terminology slightly differs from Kida’s, we recall proofs for the convenience of the reader. We also mention that the results in this section can also be viewed as a special case of those in [HH20b, Section 3.6], applied by taking for \mathbb{P} the set of all elementwise stabilizers of collections of curves on the surface, but we believe it is useful to have the arguments specified in our context. In the whole section, we let Σ\Sigma be a (possibly disconnected) orientable surface of finite type, i.e. Σ\Sigma is obtained from the disjoint union of finitely many closed connected orientable surfaces by removing at most finitely many points. We define Mod0(Σ)\mathrm{Mod}^{0}(\Sigma) as the group of all isotopy classes of orientation-preserving diffeomorphisms of Σ\Sigma that do not permute the connected components of Σ\Sigma, and act trivially on the homology mod 33 of each connected component; in other words Mod0(Σ)=Mod0(Σ1)××Mod0(Σk)\mathrm{Mod}^{0}(\Sigma)=\mathrm{Mod}^{0}(\Sigma_{1})\times\dots\times\mathrm{Mod}^{0}(\Sigma_{k}), where Σ1,,Σk\Sigma_{1},\dots,\Sigma_{k} are the connected components of Σ\Sigma.

Definition 3.7 (Irreducibility).

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict cocycle ρ:𝒢Mod0(Σ)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(\Sigma).

We say that (𝒢,ρ)(\mathcal{G},\rho) is reducible if there exist a Borel subset UYU\subseteq Y of positive measure and an essential simple closed curve cc on Σ\Sigma such that the isotopy class of cc is (𝒢|U,ρ)(\mathcal{G}_{|U},\rho)-invariant.

Otherwise, we say that (𝒢,ρ)(\mathcal{G},\rho) is irreducible.

Definition 3.8 (Canonical reduction set).

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict cocycle ρ:𝒢Mod0(Σ)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(\Sigma). A (possibly infinite) set 𝒞\mathcal{C} of isotopy classes of essential simple closed curves on Σ\Sigma is a canonical reduction set for (𝒢,ρ)(\mathcal{G},\rho) if

  1. 1.

    every c𝒞c\in\mathcal{C} is (𝒢,ρ)(\mathcal{G},\rho)-invariant, and

  2. 2.

    for every Borel subset UYU\subseteq Y of positive measure, every isotopy class cc^{\prime} of essential simple closed curves which is (𝒢|U,ρ)(\mathcal{G}_{|U},\rho)-invariant belongs to 𝒞\mathcal{C}.

Note that (𝒢,ρ)(\mathcal{G},\rho) is irreducible if and only if \emptyset is a canonical reduction set for 𝒢\mathcal{G}. Notice also that if a canonical reduction set for (𝒢,ρ)(\mathcal{G},\rho) exists, then it is unique. The following statement shows that up to a countable Borel partition of the base space, canonical reduction sets always exist.

Lemma 3.9.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict cocycle ρ:𝒢Mod0(Σ)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(\Sigma).

Then there exist a partition Y=iIYiY=\sqcup_{i\in I}Y_{i} into at most countably many Borel subsets such that for every iIi\in I, (𝒢|Yi,ρ)(\mathcal{G}_{|Y_{i}},\rho) has a canonical reduction set.

Proof.

Let Y0YY^{\prime}_{0}\subseteq Y be a Borel subset of maximal measure such that there exists a partition Y0=iI0YiY^{\prime}_{0}=\sqcup_{i\in I_{0}}Y_{i} into at most countably many Borel subsets such that for every iI0i\in I_{0}, the set 𝒞i\mathcal{C}_{i} of all isotopy classes of essential simple closed curves that are (𝒢|Yi,ρ)(\mathcal{G}_{|Y_{i}},\rho)-invariant is nonempty. We emphasise that 𝒞i\mathcal{C}_{i} is possibly infinite, and may contain non-disjoint curves. Notice that such a Borel subset Y0Y^{\prime}_{0} exists, because if (Y0,n)n(Y^{\prime}_{0,n})_{n\in\mathbb{N}} is a measure-maximizing sequence of such sets, then their countable union also satisfies the same property. Let now Y0=YY0Y_{0}=Y\setminus Y^{\prime}_{0}. The maximality of the measure of Y0Y^{\prime}_{0} ensures that (𝒢|Y0,ρ)(\mathcal{G}_{|Y_{0}},\rho) is irreducible.

For every iI0i\in I_{0}, we then let Γ𝒞i\Gamma_{\mathcal{C}_{i}} be the elementwise stabilizer of 𝒞i\mathcal{C}_{i} in Mod0(Σ)\mathrm{Mod}^{0}(\Sigma): this is a proper subgroup of Mod0(Σ)\mathrm{Mod}^{0}(\Sigma) (because 𝒞i\mathcal{C}_{i}\neq\emptyset). Repeating the above argument, for every iI0i\in I_{0}, there exists a Borel partition Yi=Yi,0Yi,0Y_{i}=Y_{i,0}\sqcup Y_{i,0}^{\prime} such that

  1. 1.

    for every Borel subset UYi,0U\subseteq Y_{i,0} of positive measure, every (𝒢|U,ρ)(\mathcal{G}_{|U},\rho)-invariant isotopy class of essential simple closed curve belongs to 𝒞i\mathcal{C}_{i},

  2. 2.

    there exists a partition Yi,0=jJiYi,jY^{\prime}_{i,0}=\sqcup_{j\in J_{i}}Y_{i,j} into at most countably many Borel subsets such that for every jJij\in J_{i}, the set 𝒞i,j\mathcal{C}_{i,j} of all isotopy classes of essential simple closed curves that are (𝒢|Yi,j,ρ)(\mathcal{G}_{|Y_{i,j}},\rho)-invariant contains 𝒞i\mathcal{C}_{i} properly.

For every jJij\in J_{i}, we then let Γ𝒞i,j\Gamma_{\mathcal{C}_{i,j}} be the elementwise stabilizer of 𝒞i,j\mathcal{C}_{i,j} in Mod0(Σ)\mathrm{Mod}^{0}(\Sigma). We observe that Γ𝒞i,j\Gamma_{\mathcal{C}_{i,j}} is a proper subgroup of Γ𝒞i\Gamma_{\mathcal{C}_{i}}. Indeed, there exists a conull Borel subset YiYiY_{i}^{*}\subseteq Y_{i} such that ρ(𝒢|Yi)Γ𝒞i\rho(\mathcal{G}_{|Y_{i}^{*}})\subseteq\Gamma_{\mathcal{C}_{i}}. If Γ𝒞i,j=Γ𝒞i\Gamma_{\mathcal{C}_{i,j}}=\Gamma_{\mathcal{C}_{i}}, then every curve in 𝒞i,j𝒞i\mathcal{C}_{i,j}\setminus\mathcal{C}_{i} is Γ𝒞i\Gamma_{\mathcal{C}_{i}}-invariant, and therefore (𝒢|Yi,ρ)(\mathcal{G}_{|Y_{i}},\rho)-invariant, contradicting the definition of 𝒞i\mathcal{C}_{i}. This contradiction shows that Γ𝒞i,jΓ𝒞i\Gamma_{\mathcal{C}_{i,j}}\subsetneq\Gamma_{\mathcal{C}_{i}}.

We now repeat the above procedure inductively. Since there is a bound, only depending on the topology of Σ\Sigma, on a chain (for inclusion) of collections of curves on Σ\Sigma with pairwise distinct elementwise stabilizers in Mod0(Σ)\mathrm{Mod}^{0}(\Sigma), we attain a partition of YY with the required properties after finitely many iterations of the above procedure. This completes the proof. ∎

The following lemma justifies the canonicity of a canonical reduction set.

Lemma 3.10.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict cocycle ρ:𝒢Mod0(Σ)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(\Sigma), and let \mathcal{H} be a measured subgroupoid of 𝒢\mathcal{G}. Assume that (,ρ)(\mathcal{H},\rho) has a canonical reduction set 𝒞\mathcal{C}.

Then for every measured subgroupoid \mathcal{H}^{\prime} of 𝒢\mathcal{G} that normalizes \mathcal{H}, the set 𝒞\mathcal{C} is (,ρ)(\mathcal{H}^{\prime},\rho)-invariant. In other words, denoting by Stab(𝒞)\mathrm{Stab}(\mathcal{C}) the global stabilizer of 𝒞\mathcal{C} in Mod0(Σ)\mathrm{Mod}^{0}(\Sigma), there exists a conull Borel subset YYY^{*}\subseteq Y such that ρ(|Y)Stab(𝒞)\rho(\mathcal{H}^{\prime}_{|Y^{*}})\subseteq\mathrm{Stab}(\mathcal{C}).

Proof.

Since \mathcal{H}^{\prime} normalizes \mathcal{H}, there exists a covering of \mathcal{H}^{\prime} by countably many bisections BnB_{n} that all leave \mathcal{H} invariant. Up to subdividing the bisections BnB_{n}, we will assume that for every nn\in\mathbb{N}, the ρ\rho-image of BnB_{n} is a single element γnMod0(Σ)\gamma_{n}\in\mathrm{Mod}^{0}(\Sigma). For every nn\in\mathbb{N}, we let UnU_{n} and VnV_{n} be the source and range of BnB_{n}.

Every isotopy class c𝒞c\in\mathcal{C} is (|Un,ρ)(\mathcal{H}_{|U_{n}},\rho)-invariant, so γnc\gamma_{n}c is (|Vn,ρ)(\mathcal{H}_{|V_{n}},\rho)-invariant. If VnV_{n} has positive measure, the maximality condition in the definition of a canonical reduction set ensures that γnc𝒞\gamma_{n}c\in\mathcal{C}. By reversing the arrows in the bisection BnB_{n}, we also derive that γnc𝒞\gamma_{n}c\notin\mathcal{C} if c𝒞c\notin\mathcal{C}. Let YYY^{*}\subseteq Y be a conull Borel subset which avoids each of the countably many subsets UnU_{n} and VnV_{n} of zero measure. Then ρ(|Y)Stab(𝒞)\rho(\mathcal{H}^{\prime}_{|Y^{*}})\subseteq\mathrm{Stab}(\mathcal{C}). This concludes our proof. ∎

Given a (possibly infinite) set 𝒞\mathcal{C} of isotopy classes of essential simple closed curves on Σ\Sigma, there is up to isotopy a unique essential subsurface SΣS\subseteq\Sigma such that every curve in 𝒞\mathcal{C} is isotopic into SS, and if SS^{\prime} is another subsurface with this property, then up to isotopy SSS\subseteq S^{\prime}. We call SS the subsurface of Σ\Sigma filled by 𝒞\mathcal{C}. The multicurve XX, obtained from S\partial S by only keeping one curve in each isotopy class, is called the boundary multicurve of 𝒞\mathcal{C}.

Corollary 3.11.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict cocycle ρ:𝒢Mod0(Σ)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(\Sigma). Let ,𝒢\mathcal{H},\mathcal{H}^{\prime}\subseteq\mathcal{G} be measured subgroupoids. Assume that \mathcal{H} is stably normalized by \mathcal{H}^{\prime}, and that for every Borel subset UYU\subseteq Y of positive measure, one has ρ(|U){1}\rho(\mathcal{H}_{|U})\neq\{1\}.

If (,ρ)(\mathcal{H},\rho) is reducible, then so is (,ρ)(\mathcal{H}^{\prime},\rho).

Proof.

Since (,ρ)(\mathcal{H},\rho) is reducible, we can find a Borel subset UYU\subseteq Y of positive measure such that (|U,ρ)(\mathcal{H}_{|U},\rho) has a nonempty canonical reduction set 𝒞\mathcal{C}. As ρ(|U){1}\rho(\mathcal{H}_{|U})\neq\{1\}, the set 𝒞\mathcal{C} does not fill Σ\Sigma, so the boundary multicurve XX of 𝒞\mathcal{C} is nonempty. Up to restricting to a Borel subset of UU of positive measure, we can assume that |U\mathcal{H}_{|U} is normalized by |U\mathcal{H}^{\prime}_{|U}. Then Lemma 3.10 ensures that 𝒞\mathcal{C} is (|U,ρ)(\mathcal{H}^{\prime}_{|U},\rho)-invariant. In particular XX is (|U,ρ)(\mathcal{H}^{\prime}_{|U},\rho)-invariant, showing that (,ρ)(\mathcal{H}^{\prime},\rho) is reducible. ∎

When 𝒞\mathcal{C} is the canonical reduction set for (,ρ)(\mathcal{H},\rho), the boundary multicurve XX of 𝒞\mathcal{C} will be called the canonical reduction multicurve of (,ρ)(\mathcal{H},\rho). A connected component SS of ΣX\Sigma\setminus X is then called active for (,ρ)(\mathcal{H},\rho) if it contains an essential simple closed curve whose isotopy class does not belong to 𝒞\mathcal{C}, and inactive for (,ρ)(\mathcal{H},\rho) otherwise (because in the latter case, every element in the elementwise stabilizer of 𝒞\mathcal{C} acts trivially on SS).

We give a few examples of active and inactive subsurfaces in the case that the essential image of ρ\rho is a cyclic subgroup generated by φ\varphi.

  1. i)

    If φ\varphi is a partial pseudo-Anosov supported on a connected subsurface ZΣZ\subset\Sigma, possibly composed with Dehn twists about curves contained in Z\partial Z, then the canonical reduction multicurve is Z\partial Z, and ZZ is the only active complementary component.

  2. ii)

    If φ\varphi is a Dehn twist about a curve α\alpha, then the canonical reduction multicurve is α\alpha, and all complementary components are inactive.

3.3 Exploiting amenable normalized subgroupoids, after Kida [Kid08a]

The following result of Kida will be used extensively in the remainder of this section, applied either to V\partial V or to subsurfaces of V\partial V. We include a proof to explain how to deal with disconnected subsurfaces.

Lemma 3.12 (Kida [Kid08a]).

Let Σ\Sigma be a (possibly disconnected) surface of finite type, so that every connected component has negative Euler characteristic. Let 𝒢\mathcal{G} be a measured groupoid, equipped with a strict cocycle ρ:𝒢Mod0(Σ)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(\Sigma). Let \mathcal{H} be a measured subgroupoid of 𝒢\mathcal{G} such that ρ|\rho_{|\mathcal{H}} has trivial kernel.

If \mathcal{H} stably normalizes an amenable subgroupoid 𝒜\mathcal{A} of 𝒢\mathcal{G}, with (𝒜,ρ)(\mathcal{A},\rho) irreducible, then \mathcal{H} is amenable.

Proof.

Up to a countable Borel partition of the base space YY of 𝒢\mathcal{G} (which does not affect the conclusion), we will assume that \mathcal{H} normalizes 𝒜\mathcal{A}.

Let Σ1,,Σk\Sigma_{1},\dots,\Sigma_{k} be the connected components of Σ\Sigma. Then Mod0(Σ)\mathrm{Mod}^{0}(\Sigma) decomposes as Mod0(Σ)=Mod0(Σ1)××Mod0(Σk)\mathrm{Mod}^{0}(\Sigma)=\mathrm{Mod}^{0}(\Sigma_{1})\times\dots\times\mathrm{Mod}^{0}(\Sigma_{k}). For i{1,,k}i\in\{1,\dots,k\}, let ρi:𝒢Mod0(Σi)\rho_{i}:\mathcal{G}\to\mathrm{Mod}^{0}(\Sigma_{i}) be the cocycle obtained by post-composing ρ\rho with the ithi^{\text{th}} projection.

Let i{1,,k}i\in\{1,\dots,k\}. Then Mod0(Σi)\mathrm{Mod}^{0}(\Sigma_{i}) acts on the compact metrizable space PML(Σi)\mathrm{PML}(\Sigma_{i}) of projective measured laminations on Σi\Sigma_{i}. As 𝒜\mathcal{A} is amenable, there exists an (𝒜,ρi)(\mathcal{A},\rho_{i})-equivariant Borel map μ:YProb(PML(Σi))\mu:Y\to\mathrm{Prob}(\mathrm{PML}(\Sigma_{i})). The space PML(Σi)\mathrm{PML}(\Sigma_{i}) has a Mod(Σi)\mathrm{Mod}(\Sigma_{i})-invariant Borel partition into the subspace ALi\mathrm{AL}_{i} made of arational laminations, and the subspace NALi\mathrm{NAL}_{i} made of non-arational laminations.

Let us first assume towards a contradiction that there exists a Borel subset UYU\subseteq Y of positive measure such that for all yUy\in U, the measure μ(y)\mu(y) gives positive measure to NALi\mathrm{NAL}_{i}. After restricting μ(y)\mu(y) to NALi\mathrm{NAL}_{i} and renormalizing it to get a probability measure, we obtain an (𝒜|U,ρi)(\mathcal{A}_{|U},\rho_{i})-equivariant Borel map UProb(NALi)U\to\mathrm{Prob}(\mathrm{NAL}_{i}). Let 𝒫<(𝒞(Σi))\mathcal{P}_{<\infty}(\mathcal{C}(\Sigma_{i})) be the countable set of all nonempty finite sets of isotopy classes of essential simple closed curves on Σi\Sigma_{i}. There is a Mod(Σi)\mathrm{Mod}(\Sigma_{i})-equivariant map NALi𝒫<(𝒞(Σi))\mathrm{NAL}_{i}\to\mathcal{P}_{<\infty}(\mathcal{C}(\Sigma_{i})), sending a lamination to the union of all simple closed curves it contains together with all boundaries of the subsurfaces it fills. We thus get an (𝒜|U,ρi)(\mathcal{A}_{|U},\rho_{i})-equivariant Borel map UProb(𝒫<(𝒞(Σi)))U\to\mathrm{Prob}(\mathcal{P}_{<\infty}(\mathcal{C}(\Sigma_{i}))). As 𝒫<(𝒞(Σi))\mathcal{P}_{<\infty}(\mathcal{C}(\Sigma_{i})) is countable, there is also a Mod(Σi)\mathrm{Mod}(\Sigma_{i})-equivariant map Prob(𝒫<(𝒞(Σi)))𝒫<(𝒞(Σi))\mathrm{Prob}(\mathcal{P}_{<\infty}(\mathcal{C}(\Sigma_{i})))\to\mathcal{P}_{<\infty}(\mathcal{C}(\Sigma_{i})), sending a probability measure ν\nu to the union of all finite sets with maximal ν\nu-measure. In summary, we have found an (𝒜|U,ρi)(\mathcal{A}_{|U},\rho_{i})-equivariant Borel map U𝒫<(𝒞(Σi))U\to\mathcal{P}_{<\infty}(\mathcal{C}(\Sigma_{i})). Let VUV\subseteq U be a Borel subset of positive measure where this map is constant, with value a finite set \mathcal{F}. As we are working in the finite-index subgroup Mod0(Σi)\mathrm{Mod}^{0}(\Sigma_{i}), every curve in \mathcal{F} is (𝒜|V,ρi)(\mathcal{A}_{|V},\rho_{i})-invariant, contradicting the irreducibility of (𝒜,ρ)(\mathcal{A},\rho).

Therefore μ\mu determines an (𝒜,ρi)(\mathcal{A},\rho_{i})-equivariant Borel map YProb(ALi)Y\to\mathrm{Prob}(\mathrm{AL}_{i}). Klarreich’s description [Kla18] of the boundary 𝒞i\partial_{\infty}\mathcal{C}_{i} of the curve graph of Σi\Sigma_{i} yields a continuous Mod(Σi)\mathrm{Mod}(\Sigma_{i})-equivariant map ALi𝒞i\mathrm{AL}_{i}\to\partial_{\infty}\mathcal{C}_{i}, so we get an (𝒜,ρi)(\mathcal{A},\rho_{i})-equivariant Borel map YProb(𝒞i)Y\to\mathrm{Prob}(\partial_{\infty}\mathcal{C}_{i}). Denoting by (𝒞i)(3)(\partial_{\infty}\mathcal{C}_{i})^{(3)} the space of pairwise distinct triples, Kida proved in [Kid08a, Section 4.1] the existence of a Mod(Σi)\mathrm{Mod}(\Sigma_{i})-equivariant Borel map (𝒞i)(3)𝒫<(𝒞(Σi))(\partial_{\infty}\mathcal{C}_{i})^{(3)}\to\mathcal{P}_{<\infty}(\mathcal{C}(\Sigma_{i})). Using again the irreducibility of (𝒜,ρ)(\mathcal{A},\rho), together with an Adams-type argument as in the proof of Lemma 3.6, we deduce that there exists a Borel map Y𝒫2(𝒞i)Y\to\mathcal{P}_{\leq 2}(\partial_{\infty}\mathcal{C}_{i}) which is both (𝒜,ρi)(\mathcal{A},\rho_{i})-equivariant and (,ρi)(\mathcal{H},\rho_{i})-equivariant.

Combining all these maps as ii varies in {1,,k}\{1,\dots,k\} yields an (,ρ)(\mathcal{H},\rho)-equivariant Borel map

Y𝒫2(𝒞1)××𝒫2(𝒞k).Y\to\mathcal{P}_{\leq 2}(\partial_{\infty}\mathcal{C}_{1})\times\dots\times\mathcal{P}_{\leq 2}(\partial_{\infty}\mathcal{C}_{k}).

For every i{1,,k}i\in\{1,\dots,k\}, the action of Mod(Σi)\mathrm{Mod}(\Sigma_{i}) on 𝒞i\partial_{\infty}\mathcal{C}_{i} is Borel amenable [Kid08a, Ham09], and therefore so is the action of Mod0(Σ)\mathrm{Mod}^{0}(\Sigma) on 𝒫2(𝒞1)××𝒫2(𝒞k)\mathcal{P}_{\leq 2}(\partial_{\infty}\mathcal{C}_{1})\times\dots\times\mathcal{P}_{\leq 2}(\partial_{\infty}\mathcal{C}_{k}) (see e.g. [HH20b, Section 3.4.1] for the relevant background). As ρ|\rho_{|\mathcal{H}} has trivial kernel, it then follows from [GH21, Proposition 3.38] (originially due to Kida [Kid08a, Proposition 4.33]) that \mathcal{H} is amenable. ∎

3.4 Uniqueness statements

Lemma 3.13.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V). Let \mathcal{H} be a measured subgroupoid of 𝒢\mathcal{G}.

Let c,cc,c^{\prime} be two meridians, with cc nonseparating. Assume that there exists a Borel subset UYU\subseteq Y of positive measure such that |U\mathcal{H}_{|U} is equal to the (𝒢|U,ρ)(\mathcal{G}_{|U},\rho)-stabilizer of the isotopy class of cc, and the isotopy class of cc^{\prime} is (|U,ρ)(\mathcal{H}_{|U},\rho)-invariant.

Then c=cc^{\prime}=c (up to isotopy).

Proof.

The stabilizer of cc in Mod0(V)\mathrm{Mod}^{0}(V) contains an element gg which restricts to a pseudo-Anosov element on Vc\partial V\setminus c (Lemma 1.1). The groupoid ρ1(g)|U\rho^{-1}(\langle g\rangle)_{|U} is contained in |U\mathcal{H}_{|U}, and it is of infinite type since ρ\rho is action-type. Therefore cc^{\prime} is fixed by some positive power of gg, which implies that c=cc^{\prime}=c up to isotopy. ∎

The following is a version of Lemma 3.13 for separating meridians, whose proof is similar and left to the reader (recall that although the stabiliser HH of a separating meridian cc may fix other nonseparating meridians, cc is the unique separating meridian it fixes by Lemma 1.2).

Lemma 3.14.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V). Let \mathcal{H} be a measured subgroupoid of 𝒢\mathcal{G}.

Let c,cc,c^{\prime} be two separating meridians. Assume that there exists a Borel subset UYU\subseteq Y of positive measure such that |U\mathcal{H}_{|U} is equal to the (𝒢|U,ρ)(\mathcal{G}_{|U},\rho)-stabilizer of the isotopy class of cc, and the isotopy class of cc^{\prime} is (|U,ρ)(\mathcal{H}_{|U},\rho)-invariant.

Then c=cc=c^{\prime} (up to isotopy). ∎

3.5 Property (Pnsep)(\mathrm{P}_{\mathrm{nsep}}) and subgroupoids of non-separating meridian type

We make the following definition.

Definition 3.15 (Product-like subgroupoid).

A measured groupoid 𝒫\mathcal{P} is product-like if there exist two subgroupoids 𝒫1,𝒫2𝒫\mathcal{P}_{1},\mathcal{P}_{2}\subseteq\mathcal{P} which are both stably normal in 𝒫\mathcal{P}, such that for every i{1,2}i\in\{1,2\}, the groupoid 𝒫i\mathcal{P}_{i} contains a strongly Schottky pair of subgroupoids (𝒜i1,𝒜i2)(\mathcal{A}_{i}^{1},\mathcal{A}_{i}^{2}), with 𝒜i1\mathcal{A}_{i}^{1} and 𝒜i2\mathcal{A}_{i}^{2} both stably normalized by 𝒫3i\mathcal{P}_{3-i}.

Notice that this notion is stable under restrictions and stabilization. In the terminology from [GH21, Definition 13.5], the subgroupoids 𝒫1\mathcal{P}_{1} and 𝒫2\mathcal{P}_{2} form a pseudo-product. One difference between our definition and [GH21, Definition 13.5] is that we are working with strongly Schottky pairs of subgroupoids, while [GH21, Definition 13.5] is phrased using the weaker notion of Schottky pairs of subgroupoids. Also, we are further imposing that 𝒫1\mathcal{P}_{1} and 𝒫2\mathcal{P}_{2} are stably normal in an ambient groupoid 𝒫\mathcal{P}.

We now introduce the following properties, which will be useful in order to detect subgroupoids of nonseparating-meridian type.

Definition 3.16.

Let 𝒢\mathcal{G} be a measured groupoid, and let 𝒜,\mathcal{A},\mathcal{H} be measured subgroupoids of 𝒢\mathcal{G}, with 𝒜\mathcal{A}\subseteq\mathcal{H}.

  1. 1.

    We say that the pair (,𝒜)(\mathcal{H},\mathcal{A}) satisfies Property (Qnsep)(\mathrm{Q}_{\mathrm{nsep}}) if the following conditions hold:

    1. (a)

      \mathcal{H} is everywhere nonamenable;

    2. (b)

      𝒜\mathcal{A} is amenable, of infinite type, and stably normal in \mathcal{H};

    3. (c)

      if \mathcal{B} is a stably normal amenable subgroupoid of \mathcal{H}, then \mathcal{B} is stably contained in 𝒜\mathcal{A};

    4. (d)

      if \mathcal{H}^{\prime} is another subgroupoid of 𝒢\mathcal{G} which is everywhere nonamenable and contains a stably normal amenable subgroupoid of infinite type, and if \mathcal{H} is stably contained in \mathcal{H}^{\prime}, then \mathcal{H} is stably equal to \mathcal{H}^{\prime};

    5. (e)

      for every Borel subset UYU\subseteq Y of positive measure, the groupoid |U\mathcal{H}_{|U} is not contained in any product-like subgroupoid of 𝒢|U\mathcal{G}_{|U}.

  2. 2.

    We say that \mathcal{H} satisfies Property (Pnsep)(\mathrm{P}_{\mathrm{nsep}}) if there exists a measured subgroupoid 𝒜\mathcal{A}\subseteq\mathcal{H} such that (,𝒜)(\mathcal{H},\mathcal{A}) satisfies Property (Qnsep)(\mathrm{Q}_{\mathrm{nsep}}).

Remark 3.17.

These properties are stable under restrictions and stabilization. Also, if \mathcal{H} satisfies Property (Pnsep)(\mathrm{P}_{\mathrm{nsep}}), then a subgroupoid 𝒜\mathcal{A}\subseteq\mathcal{H} such that (,𝒜)(\mathcal{H},\mathcal{A}) satisfies Property (Qnsep)(\mathrm{Q}_{\mathrm{nsep}}) is “stably unique” in the following sense: if 𝒜\mathcal{A} and 𝒜\mathcal{A}^{\prime} are two such subgroupoids, there exist a conull Borel subset YYY^{*}\subseteq Y and a partition Y=iIYiY^{*}=\sqcup_{i\in I}Y_{i} into at most countably many Borel subsets such that for every iIi\in I, one has 𝒜|Yi=𝒜|Yi\mathcal{A}_{|Y_{i}}=\mathcal{A}^{\prime}_{|Y_{i}}. Indeed, this is a consequence of Assumptions (b) and (c) from the definition.

The goal of the present section is to prove that subgroupoids of nonseparating meridian type with respect to an action-type cocycle 𝒢Mod0(V)\mathcal{G}\to\mathrm{Mod}^{0}(V) satisfy Property (Pnsep)(\mathrm{P}_{\mathrm{nsep}}).

Proposition 3.18.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V). Let cc be a nonseparating meridian, let \mathcal{H} be the (𝒢,ρ)(\mathcal{G},\rho)-stabilizer of the isotopy class of cc, and let 𝒜=ρ1(Tc)\mathcal{A}=\rho^{-1}(\langle T_{c}\rangle).

Then (,𝒜)(\mathcal{H},\mathcal{A}) satisfies Property (Qnsep)(\mathrm{Q}_{\mathrm{nsep}}).

Proposition 3.18 is the combination of our three next lemmas. The first checks Assertions (a),(b) and (c) from Definition 3.16. For later convenience, in this lemma, we also allow for separating meridians in the statement.

Lemma 3.19.

Let 𝒢\mathcal{G} be a measured groupoid, equipped with a strict action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V). Let cc be a meridian, and let \mathcal{H} be the (𝒢,ρ)(\mathcal{G},\rho)-stabilizer of the isotopy class of cc. Let ΣV\Sigma\subseteq\partial V be the union of all components of Vc\partial V\setminus c which are not once-holed tori. Let AStabMod0(V)(c)A\subseteq\mathrm{Stab}_{\mathrm{Mod}^{0}(V)}(c) be the kernel of the restriction homomorphism to Mod0(Σ)\mathrm{Mod}^{0}(\Sigma)111notice that if cc is nonseparating, or if cc is separating and both complementary components have genus at least 22, then A=TcMod0(V)A=\langle T_{c}\rangle\cap\mathrm{Mod}^{0}(V), and let 𝒜=ρ1(A)\mathcal{A}=\rho^{-1}(A).

Then \mathcal{H} is everywhere nonamenable, 𝒜\mathcal{A} is a normal amenable subgroupoid of \mathcal{H} of infinite type, and every stably normal amenable subgroupoid of \mathcal{H} is stably contained in 𝒜\mathcal{A}.

Proof.

Notice that the subsurface Σ\Sigma is nonempty because the genus of VV is at least 33. Lemma 1.1 ensures that StabMod0(V)(c)\mathrm{Stab}_{\mathrm{Mod}^{0}(V)}(c) contains a nonabelian free subgroup, so Lemma 3.6 shows that \mathcal{H} is everywhere nonamenable.

Normality of 𝒜\mathcal{A} in \mathcal{H} follows from the normality of AA in StabMod0(V)(c)\mathrm{Stab}_{\mathrm{Mod}^{0}(V)}(c). Notice that AA is amenable (using Lemma 1.2 in the case where one of the complementary components of cc is a once-holed torus). As ρ\rho has trivial kernel, it follows that 𝒜\mathcal{A} is amenable. And 𝒜\mathcal{A} is of infinite type because AA is infinite (it always contains a power of TcT_{c}) and ρ\rho is action-type.

Let now \mathcal{B}\subseteq\mathcal{H} be a stably normal amenable subgroupoid of \mathcal{H}. Let SΣS\subseteq\Sigma be a connected component of Σ\Sigma. Let ρS:Mod0(S)\rho_{S}:\mathcal{H}\to\mathrm{Mod}^{0}(S) be the cocycle obtained by post-composing ρ\rho with the restriction homomorphism. Let also FStabMod0(V)(c)F\subseteq\mathrm{Stab}_{\mathrm{Mod}^{0}(V)}(c) be a nonabelian free subgroup which embeds into Mod0(S)\mathrm{Mod}^{0}(S) under the restriction homomorphism, and whose image in Mod0(S)\mathrm{Mod}^{0}(S) contains a pseudo-Anosov mapping class (this exists because SS is not a once-holed torus, see Lemma 1.1). Let =ρ1(F)\mathcal{H}^{\prime}=\rho^{-1}(F).

By Lemma 3.9, we can find a partition Y=iIYiY=\sqcup_{i\in I}Y_{i} into at most countably many Borel subsets such that for every iIi\in I, the pair (|Yi,ρS)(\mathcal{B}_{|Y_{i}},\rho_{S}) has a canonical reduction set 𝒞i\mathcal{C}_{i}. As \mathcal{B} is stably normal in \mathcal{H}, up to refining the above partition, we can assume that for every iIi\in I, the groupoid |Yi\mathcal{B}_{|Y_{i}} is normal in |Yi\mathcal{H}_{|Y_{i}}. Lemma 3.10 thus ensures that 𝒞i\mathcal{C}_{i} is (|Yi,ρS)(\mathcal{H}_{|Y_{i}},\rho_{S})-invariant, so either 𝒞i=\mathcal{C}_{i}=\emptyset or 𝒞i\mathcal{C}_{i} fills SS.

Assume towards a contradiction that 𝒞i=\mathcal{C}_{i}=\emptyset for some iIi\in I such that YiY_{i} has positive measure. In other words (|Yi,ρS)(\mathcal{B}_{|Y_{i}},\rho_{S}) is irreducible. As ρS\rho_{S} has trivial kernel in restriction to \mathcal{H}^{\prime}, and as \mathcal{H}^{\prime} (which is contained in \mathcal{H}) normalizes \mathcal{B}, Lemma 3.12 implies that |Yi\mathcal{H}^{\prime}_{|Y_{i}} is amenable. But FF is a nonabelian free group and ρ\rho is action-type, so we get a contradiction to Lemma 3.6.

It follows that for every iIi\in I, there exists a conull Borel subset YiYiY_{i}^{*}\subseteq Y_{i} such that ρS(|Yi)={1}\rho_{S}(\mathcal{B}_{|Y_{i}^{*}})=\{1\}. As SS was an arbitrary connected component of Σ\Sigma, this precisely means that \mathcal{B} is stably contained in 𝒜\mathcal{A}. ∎

We now check Assertion (d) from Definition 3.16.

Lemma 3.20.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V). Let cc be a nonseparating meridian, and let \mathcal{H} be the (𝒢,ρ)(\mathcal{G},\rho)-stabilizer of the isotopy class of cc.

If \mathcal{H}^{\prime} is a subgroupoid of 𝒢\mathcal{G} which is everywhere nonamenable and contains a stably normal amenable subgroupoid of infinite type, and if \mathcal{H} is stably contained in \mathcal{H}^{\prime}, then \mathcal{H} is stably equal to \mathcal{H}^{\prime}.

Proof.

Let 𝒜\mathcal{A}^{\prime} be an amenable subgroupoid of 𝒢\mathcal{G} of infinite type which is contained in \mathcal{H}^{\prime} and stably normal in \mathcal{H}^{\prime}. By Lemma 3.9, we can find a partition Y=iIYiY=\sqcup_{i\in I}Y_{i} into at most countably many Borel subsets such that for every iIi\in I, the pair (𝒜|Yi,ρ)(\mathcal{A}^{\prime}_{|Y_{i}},\rho) has a (possibly empty) canonical reduction set 𝒞i\mathcal{C}_{i}. For every iIi\in I, we let XiX_{i} be the (possibly empty) boundary multicurve of 𝒞i\mathcal{C}_{i}. As 𝒜\mathcal{A}^{\prime} is stably normal in \mathcal{H}^{\prime}, up to refining the above partition, we can assume that for every iIi\in I, the set 𝒞i\mathcal{C}_{i} is (|Yi,ρ)(\mathcal{H}^{\prime}_{|Y_{i}},\rho)-invariant (Lemma 3.10), and therefore so is the multicurve XiX_{i}. As \mathcal{H} is stably contained in \mathcal{H}^{\prime}, we will also assume up to refining the above partition once more that for every iIi\in I, one has |Yi|Yi\mathcal{H}_{|Y_{i}}\subseteq\mathcal{H}^{\prime}_{|Y_{i}}. In particular XiX_{i} is (|Yi,ρ)(\mathcal{H}_{|Y_{i}},\rho)-invariant, which implies that either Xi=X_{i}=\emptyset or Xi=cX_{i}=c.

Let iIi\in I be such that YiY_{i} has positive measure. If Xi=X_{i}=\emptyset, then as 𝒜\mathcal{A}^{\prime} is of infinite type and ρ\rho has trivial kernel, we deduce that 𝒞i=\mathcal{C}_{i}=\emptyset, i.e. (𝒜|Yi,ρ)(\mathcal{A}^{\prime}_{|Y_{i}},\rho) is irreducible. Lemma 3.12 then implies that |Yi\mathcal{H}^{\prime}_{|Y_{i}} is amenable, a contradiction. Therefore Xi=cX_{i}=c, so the isotopy class of cc is (|Yi,ρ)(\mathcal{H}^{\prime}_{|Y_{i}},\rho)-invariant. As this is true for every iIi\in I such that YiY_{i} has positive measure, we deduce that \mathcal{H}^{\prime} is stably contained in \mathcal{H}. ∎

We finally check Assertion (e) from Definition 3.16.

Lemma 3.21.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V). Let cc be a nonseparating meridian, and let \mathcal{H} be the (𝒢,ρ)(\mathcal{G},\rho)-stabilizer of the isotopy class of cc.

Then for every Borel subset UYU\subseteq Y of positive measure, the groupoid |U\mathcal{H}_{|U} is not contained in any product-like subgroupoid of 𝒢|U\mathcal{G}_{|U}.

Proof.

Let UYU\subseteq Y be a Borel subset of positive measure. Assume towards a contradiction that |U\mathcal{H}_{|U} is contained in a product-like subgroupoid 𝒫\mathcal{P} of 𝒢|U\mathcal{G}_{|U}. Let 𝒫1,𝒫2,𝒜11,𝒜12,𝒜21,𝒜22𝒫\mathcal{P}_{1},\mathcal{P}_{2},\mathcal{A}_{1}^{1},\mathcal{A}_{1}^{2},\mathcal{A}_{2}^{1},\mathcal{A}_{2}^{2}\subseteq\mathcal{P} be as in the definition of a product-like subgroupoid (Definition 3.15).

Up to restricting to a Borel subset of UU of positive measure, we can assume that (𝒫,ρ)(\mathcal{P},\rho) has a canonical reduction multicurve XX. As |U𝒫\mathcal{H}_{|U}\subseteq\mathcal{P}, the isotopy class of XX is (|U,ρ)(\mathcal{H}_{|U},\rho)-invariant, so X=X=\emptyset or X=cX=c, and the kernel of the induced cocycle ρ:𝒫Mod0(VX)\rho^{\prime}:\mathcal{P}\to\mathrm{Mod}^{0}(\partial V\setminus X) is amenable (it is trivial if X=X=\emptyset, and contained in ρ1(Tc)\rho^{-1}(\langle T_{c}\rangle) if X=cX=c). As 𝒫\mathcal{P} is everywhere nonamenable, it follows that ρ(𝒫|U){1}\rho^{\prime}(\mathcal{P}_{|U^{*}})\neq\{1\} for every conull Borel subset UUU^{*}\subseteq U (as otherwise 𝒫|U\mathcal{P}_{|U^{*}} would be equal to the kernel and therefore amenable). In particular, the subsurface VX\partial V\setminus X is active for (𝒫,ρ)(\mathcal{P},\rho), and therefore (𝒫,ρ)(\mathcal{P},\rho^{\prime}) is irreducible. As 𝒫2\mathcal{P}_{2} is everywhere nonamenable, we also have ρ((𝒫2)|U){1}\rho^{\prime}((\mathcal{P}_{2})_{|U^{\prime}})\neq\{1\} for every positive measure Borel subset UUU^{\prime}\subseteq U. As 𝒫2\mathcal{P}_{2} is stably normal in 𝒫\mathcal{P}, Corollary 3.11 therefore ensures that (𝒫2,ρ)(\mathcal{P}_{2},\rho^{\prime}) is also irreducible.

By definition of a strongly Schottky pair (applied to (𝒜11,𝒜12)(\mathcal{A}_{1}^{1},\mathcal{A}_{1}^{2})), there exists a Borel subset UUU^{\prime}\subseteq U of positive measure such that every normal amenable subgroupoid of (𝒜11)|U,(𝒜12)|U\langle(\mathcal{A}_{1}^{1})_{|U^{\prime}},(\mathcal{A}_{1}^{2})_{|U^{\prime}}\rangle is stably trivial. In particular, the kernel of ρ\rho^{\prime} restricted to the subgroupoid (𝒜11)|U,(𝒜12)|U\langle(\mathcal{A}_{1}^{1})_{|U^{\prime}},(\mathcal{A}_{1}^{2})_{|U^{\prime}}\rangle is stably trivial. As 𝒜11\mathcal{A}_{1}^{1} is of infinite type, it follows that for every Borel subset U′′UU^{\prime\prime}\subseteq U^{\prime} of positive measure, we have ρ((𝒜11)|U′′){1}\rho^{\prime}((\mathcal{A}_{1}^{1})_{|U^{\prime\prime}})\neq\{1\}. As (𝒜11)|U(\mathcal{A}_{1}^{1})_{|U^{\prime}} is stably normalized by (𝒫2)|U(\mathcal{P}_{2})_{|U^{\prime}}, Corollary 3.11 ensures that ((𝒜11)|U,ρ)((\mathcal{A}_{1}^{1})_{|U^{\prime}},\rho^{\prime}) is irreducible.

By definition of a strongly Schottky pair (applied to (𝒜21,𝒜22)(\mathcal{A}_{2}^{1},\mathcal{A}_{2}^{2})), there exists a Borel subset WUW\subseteq U^{\prime} of positive measure such that every normal amenable subgroupoid of (𝒜21)|W,(𝒜22)|W\langle(\mathcal{A}_{2}^{1})_{|W},(\mathcal{A}_{2}^{2})_{|W}\rangle is stably trivial. In particular, the kernel of ρ\rho^{\prime} restricted to the subgroupoid (𝒜21)|W,(𝒜22)|W\langle(\mathcal{A}_{2}^{1})_{|W},(\mathcal{A}_{2}^{2})_{|W}\rangle is stably trivial. This implies that we can find a positive measure Borel subset WWW^{\prime}\subseteq W such that ρ\rho^{\prime} has trivial kernel in restriction to (𝒜21)|W,(𝒜22)|W\langle(\mathcal{A}_{2}^{1})_{|W^{\prime}},(\mathcal{A}_{2}^{2})_{|W^{\prime}}\rangle. As (𝒜11)|W(\mathcal{A}_{1}^{1})_{|W^{\prime}} is stably normalized by (𝒜21)|W,(𝒜22)|W\langle(\mathcal{A}_{2}^{1})_{|W^{\prime}},(\mathcal{A}_{2}^{2})_{|W^{\prime}}\rangle, it thus follows from Lemma 3.12 that (𝒜21)|W,(𝒜22)|W\langle(\mathcal{A}_{2}^{1})_{|W^{\prime}},(\mathcal{A}_{2}^{2})_{|W^{\prime}}\rangle is amenable, which yields the desired contradiction. ∎

3.6 Stabilizers of separating meridians do not satisfy Property (Pnsep)(\mathrm{P}_{\mathrm{nsep}})

Lemma 3.22.

Let 𝒢\mathcal{G} be a measured groupoid, equipped with a strict action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V), and let \mathcal{H} be a measured subgroupoid of 𝒢\mathcal{G}. Let cc be a separating meridian, and assume that the isotopy class of cc is (,ρ)(\mathcal{H},\rho)-invariant.

Then \mathcal{H} does not satisfy Property (Pnsep)(\mathrm{P}_{\mathrm{nsep}}).

Proof.

We first assume that one complementary component Σ\Sigma of cc is a once-holed torus. Then Σ\Sigma contains, up to isotopy, a unique nonseparating meridian δ\delta (Lemma 1.2), so \mathcal{H} is contained in the (𝒢,ρ)(\mathcal{G},\rho)-stabilizer \mathcal{H}^{\prime} of the isotopy class of δ\delta. In addition \mathcal{H}^{\prime} is everywhere nonamenable and contains ρ1(Tδ)\rho^{-1}(\langle T_{\delta}\rangle) as a normal amenable subgroupoid of infinite type. Finally \mathcal{H}^{\prime} is not stably contained in \mathcal{H} because Vδ\partial V\setminus\delta supports a pseudo-Anosov handlebody group element gg, and no nontrivial power of gg preserves the isotopy class of cc. So Assumption (d) from Definition 3.16 fails.

We now assume that both complementary components Σ1,Σ2\Sigma_{1},\Sigma_{2} of cc have genus at least 22. Let 𝒫\mathcal{P} be the (𝒢,ρ)(\mathcal{G},\rho)-stabilizer of cc. Then \mathcal{H} is contained in 𝒫\mathcal{P}, and we will prove that 𝒫\mathcal{P} is product-like (which will imply that Assumption (e) from Definition 3.16 fails). For every i{1,2}i\in\{1,2\}, let PiP_{i} be the subgroup of Mod0(V)\mathrm{Mod}^{0}(V) made of elements that have a representative supported in Σi\Sigma_{i}, and let 𝒫i=ρ1(Pi)\mathcal{P}_{i}=\rho^{-1}(P_{i}). Then 𝒫i\mathcal{P}_{i} is normal in 𝒫\mathcal{P}. For every i{1,2}i\in\{1,2\}, let fi1f_{i}^{1} and fi2f_{i}^{2} be two elements of PiP_{i} that generate a nonabelian free subgroup of Mod0(V)\mathrm{Mod}^{0}(V). For every i{1,2}i\in\{1,2\} and every j{1,2}j\in\{1,2\}, let 𝒜ij=ρ1(fij)\mathcal{A}_{i}^{j}=\rho^{-1}(\langle f_{i}^{j}\rangle). Then 𝒜ij\mathcal{A}_{i}^{j} is normalized by 𝒫3i\mathcal{P}_{3-i}, and Lemma 3.6 ensures that (𝒜i1,𝒜i2)(\mathcal{A}_{i}^{1},\mathcal{A}_{i}^{2}) is a strongly Schottky pair of subgroupoids of 𝒢\mathcal{G}. This completes our proof. ∎

3.7 Admissible decorated multicurves and their active subgroups

A decorated multicurve is a pair (X,𝔄)(X,\mathfrak{A}), where XX is a multicurve on V\partial V, and 𝔄\mathfrak{A} is a subset of the set of complementary components of XX in V\partial V. We make the following definition.

Definition 3.23.

Let (X,𝔄)(X,\mathfrak{A}) be a decorated multicurve. The subgroup AA of Mod0(V)\mathrm{Mod}^{0}(V) made of all elements that preserve the isotopy class of XX and act trivially on all complementary subsurfaces not in 𝔄\mathfrak{A} is called the active subgroup of (X,𝔄)(X,\mathfrak{A}).

The decorated multicurve (X,𝔄)(X,\mathfrak{A}) is admissible if its active subgroup AA is amenable, XX is the boundary multicurve of AA, and 𝔄\mathfrak{A} is its set of active complementary components.

Here is an example. If XX consists of a single nonseparating meridian on V\partial V, or a separating meridian none of whose complementary components is a once-holed torus, and 𝔄=\mathfrak{A}=\emptyset, then (X,𝔄)(X,\mathfrak{A}) is admissible, and its active subgroup is the corresponding twist subgroup. In the case where one of the complementary components of a meridian cc is a once-holed torus (and contains a unique nonseparating meridian dd up to isotopy), then (cd,)(c\cup d,\emptyset) is admissible, with active subgroup the twist subgroup about cdc\cup d. Other examples come from annulus pairs instead of meridians.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with an action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V). We say that a pair (,𝒜)(\mathcal{H},\mathcal{A}) of subgroupoids of 𝒢\mathcal{G} is admissible with respect to ρ\rho if there exist a conull Borel subset YYY^{*}\subseteq Y and a partition Y=iIYiY^{*}=\sqcup_{i\in I}Y_{i} into at most countably many Borel subsets, such that for every iIi\in I, there exist a multicurve XiX_{i} on V\partial V, and a subset 𝔄i\mathfrak{A}_{i} of the set of all complementary components of XiX_{i} such that (Xi,𝔄i)(X_{i},\mathfrak{A}_{i}) is admissible, |Yi\mathcal{H}_{|Y_{i}} is equal to the (𝒢|Yi,ρ)(\mathcal{G}_{|Y_{i}},\rho)-stabilizer of the isotopy class of XiX_{i}, and denoting by AiMod0(V)A_{i}\subseteq\mathrm{Mod}^{0}(V) the active subgroup of (Xi,𝔄i)(X_{i},\mathfrak{A}_{i}), one has 𝒜|Yi=ρ1(Ai)|Yi\mathcal{A}_{|Y_{i}}=\rho^{-1}(A_{i})_{|Y_{i}}. Notice that, although the above partition is not unique (one can always pass to a further partition), the map sending yYiy\in Y_{i} to the isotopy class of (Xi,𝔄i)(X_{i},\mathfrak{A}_{i}) is uniquely determined by (,𝒜)(\mathcal{H},\mathcal{A}); we call it the decomposition map of (,𝒜)(\mathcal{H},\mathcal{A}).

Lemma 3.24.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V). Let 𝒜,\mathcal{A},\mathcal{H} be measured subgroupoids of 𝒢\mathcal{G}, with 𝒜\mathcal{A}\subseteq\mathcal{H}.

If (,𝒜)(\mathcal{H},\mathcal{A}) satisfies Property (Qnsep)(\mathrm{Q}_{\mathrm{nsep}}), then (,𝒜)(\mathcal{H},\mathcal{A}) is an admissible pair.

Proof.

By Assumption (b) from Definition 3.16, the groupoid 𝒜\mathcal{A} is amenable, of infinite type, and stably normal in \mathcal{H}. Up to a countable partition of the base space YY, we will assume that 𝒜\mathcal{A} is normal in \mathcal{H}. Up to a further partition, we can also assume that (𝒜,ρ)(\mathcal{A},\rho) has a canonical reduction set 𝒞\mathcal{C} (Lemma 3.9). Let XX be the boundary multicurve of 𝒞\mathcal{C}, let 𝔄\mathfrak{A} be the set of all active complementary components for (𝒜,ρ)(\mathcal{A},\rho), and let 𝔄¯\overline{\mathfrak{A}} be the set of all complementary components of XX not in 𝔄\mathfrak{A}. Up to replacing YY by a conull Borel subset, we will assume using Lemma 3.10 that ρ()StabMod0(V)(X)\rho(\mathcal{H})\subseteq\mathrm{Stab}_{\mathrm{Mod}^{0}(V)}(X).

We will first prove that (X,𝔄)(X,\mathfrak{A}) is admissible, so let us assume towards a contradiction that it is not. Let AMod0(V)A\subseteq\mathrm{Mod}^{0}(V) be the active subgroup of (X,𝔄)(X,\mathfrak{A}). Then there exists a conull Borel subset YYY^{*}\subseteq Y such that ρ(𝒜|Y)A\rho(\mathcal{A}_{|Y^{*}})\subseteq A. Therefore 𝒞\mathcal{C} is exactly the set of all curves whose isotopy class is AA-invariant, so XX is the boundary multicurve of AA and 𝔄\mathfrak{A} is its set of active complementary components. Therefore, our assumption that (X,𝔄)(X,\mathfrak{A}) is not admissible implies that AA is not amenable, so it contains two elements f,gf,g which together generate a nonabelian free group (by the Tits alternative for mapping class groups [McC85a, Iva92]).

Let Σ1\Sigma_{1} be the union of all subsurfaces in 𝔄\mathfrak{A}, viewed as a (possibly disconnected) surface of finite type. Let ρ1:Mod0(Σ1)\rho_{1}:\mathcal{H}\to\mathrm{Mod}^{0}(\Sigma_{1}) be the cocycle obtained by composing ρ\rho with the restriction to Σ1\Sigma_{1}. We now observe that for every UYU\subseteq Y of positive measure, the restriction to UU of kernel of ρ1\rho_{1} is nontrivial: otherwise (𝒜|U,ρ1)(\mathcal{A}_{|U},\rho_{1}) is irreducible and |U\mathcal{H}_{|U} normalizes 𝒜|U\mathcal{A}_{|U}, so Lemma 3.12 ensures that |U\mathcal{H}_{|U} is amenable, a contradiction to Assumption (a) from Definition 3.16.

Let \mathcal{B} be the kernel of ρ1\rho_{1}. The groupoid \mathcal{B} is normal in \mathcal{H}. We first assume that \mathcal{B} is amenable, and reach a contradiction in this case. Assumption (c) from Definition 3.16 ensures that there exists a Borel subset UYU\subseteq Y of positive measure such that |U𝒜|U\mathcal{B}_{|U}\subseteq\mathcal{A}_{|U}. But the ρ\rho-image of every element of |U\mathcal{B}_{|U} acts trivially on all components in 𝔄\mathfrak{A}, while the ρ\rho-image of every element of 𝒜|U\mathcal{A}_{|U} acts trivially on all components in 𝔄¯\overline{\mathfrak{A}}. It follows that for every g|Ug\in\mathcal{B}_{|U}, the element ρ(g)\rho(g) is a multitwist around curves in XX. As ρ\rho has trivial kernel and |U\mathcal{B}_{|U} is nontrivial, it follows that the subgroup Tw\mathrm{Tw} of Mod0(V)\mathrm{Mod}^{0}(V) consisting of all multitwists about the curves in XX is infinite. Let =ρ1(StabMod0(V)(X))\mathcal{H}^{\prime}=\rho^{-1}(\mathrm{Stab}_{\mathrm{Mod}^{0}(V)}(X)). Then |U|U\mathcal{H}_{|U}\subseteq\mathcal{H}^{\prime}_{|U}, and |U\mathcal{H}^{\prime}_{|U} is everywhere nonamenable (it contains |U\mathcal{H}_{|U}) and contains ρ1(Tw)|U\rho^{-1}(\mathrm{Tw})_{|U} as a normal amenable subgroupoid of infinite type. So Assumption (d) from Definition 3.16 ensures that there exists a Borel subset UUU^{\prime}\subseteq U of positive measure such that |U=|U\mathcal{H}^{\prime}_{|U^{\prime}}=\mathcal{H}_{|U^{\prime}}. Now, the groupoid ρ1(f,g)|U\rho^{-1}(\langle f,g\rangle)_{|U^{\prime}} is contained in |U\mathcal{H}_{|U^{\prime}}, so it normalizes 𝒜|U\mathcal{A}_{|U^{\prime}}, and ρ1\rho_{1} has trivial kernel in restriction to ρ1(f,g)|U\rho^{-1}(\langle f,g\rangle)_{|U^{\prime}} (as otherwise f,g\langle f,g\rangle would contain an infinite amenable normal subgroup). As (𝒜|U,ρ1)(\mathcal{A}_{|U^{\prime}},\rho_{1}) is irreducible, Lemma 3.12 implies that ρ1(f,g)|U\rho^{-1}(\langle f,g\rangle)_{|U^{\prime}} is amenable, a contradiction to Lemma 3.6.

We now assume that \mathcal{B} is nonamenable, and also reach a contradiction in this case. As ρ\rho has trivial kernel, the subgroup P2P_{2} of Mod0(V)\mathrm{Mod}^{0}(V) made of all elements that fix the isotopy class of XX and act trivially on all connected components in 𝔄\mathfrak{A} is nonamenable, and therefore contains a nonabelian free subgroup. Let P=StabMod0(V)(X)P=\mathrm{Stab}_{\mathrm{Mod}^{0}(V)}(X), and let 𝒫=ρ1(P)\mathcal{P}=\rho^{-1}(P) (i.e. 𝒫=\mathcal{P}=\mathcal{H}^{\prime} with the notation from above). We will now reach a contradiction to Assumption (e) from Definition 3.16 by proving that 𝒫\mathcal{P} is a product-like subgroupoid of 𝒢\mathcal{G} (in which \mathcal{H} is contained).

Let P1PP_{1}\unlhd P be the normal subgroup made of all elements of PP that act trivially on all components in 𝔄¯\overline{\mathfrak{A}}, and recall that P2PP_{2}\unlhd P is the normal subgroup made of all elements of PP acting trivially on all components in 𝔄\mathfrak{A}. Then 𝒫i=ρ1(Pi)\mathcal{P}_{i}=\rho^{-1}(P_{i}) is normal in 𝒫=ρ1(P)\mathcal{P}=\rho^{-1}(P) for every i{1,2}i\in\{1,2\}. Notice that P1P_{1} contains the nonabelian free subgroup f,g\langle f,g\rangle, and we saw in the previous paragraph that P2P_{2} also contains a nonabelian free subgroup. For every i{1,2}i\in\{1,2\}, let Ai1,Ai2A_{i}^{1},A_{i}^{2} be two cyclic subgroups of PiP_{i} that generate a nonabelian free subgroup, and for j{1,2}j\in\{1,2\}, let 𝒜ij=ρ1(Aij)\mathcal{A}_{i}^{j}=\rho^{-1}(A_{i}^{j}). As P1P_{1} and P2P_{2} centralize each other, it follows that each 𝒜ij\mathcal{A}_{i}^{j} is normalized by 𝒫3i\mathcal{P}_{3-i}. In addition, Lemma 3.6 ensures that (𝒜i1,𝒜i2)(\mathcal{A}_{i}^{1},\mathcal{A}_{i}^{2}) is a strongly Schottky pair of subgroupoids of 𝒢\mathcal{G}. So 𝒫\mathcal{P} is a product-like subgroupoid of 𝒢\mathcal{G}, which is the desired contradiction.

This contradiction shows that (X,𝔄)(X,\mathfrak{A}) is admissible. Now, let 𝒜=ρ1(A)\mathcal{A}^{\prime}=\rho^{-1}(A), and let \mathcal{H}^{\prime} be the (𝒢,ρ)(\mathcal{G},\rho)-stabilizer of the isotopy class of XX. Then \mathcal{H} is contained in \mathcal{H}^{\prime}, and \mathcal{H}^{\prime} contains 𝒜\mathcal{A}^{\prime} as a normal amenable subgroupoid of infinite type. So Assertion (d) from Definition 3.16 ensures that \mathcal{H} is stably equal to \mathcal{H}^{\prime}. And Assertion (c) then implies that 𝒜\mathcal{A} is stably equal to 𝒜\mathcal{A}^{\prime}. This proves that (,𝒜)(\mathcal{H},\mathcal{A}) is an admissible pair. ∎

3.8 Compatibility

Two decorated multicurves (X,𝔄)(X,\mathfrak{A}) and (X,𝔄)(X^{\prime},\mathfrak{A}^{\prime}) are compatible if XX and XX^{\prime} are disjoint up to isotopy, and given any two components S𝔄S\in\mathfrak{A} and S𝔄S^{\prime}\in\mathfrak{A}^{\prime}, either SS and SS^{\prime} are isotopic, or they are disjoint up to isotopy. We start with the following observation.

Lemma 3.25.

Let (X,𝔄)(X,\mathfrak{A}) and (X,𝔄)(X^{\prime},\mathfrak{A}^{\prime}) be two decorated multicurves, with respective active subgroups A,AA,A^{\prime}.

If (X,𝔄)(X,\mathfrak{A}) and (X,𝔄)(X^{\prime},\mathfrak{A}^{\prime}) are compatible, then A,A\langle A,A^{\prime}\rangle is amenable.

Proof.

Otherwise, using the Tits alternative for mapping class groups [McC85a, Iva92], there exist φ,ψA,A\varphi,\psi\in\langle A,A^{\prime}\rangle that together generate a nonabelian free group. Let SS be the union of all subsurfaces in 𝔄𝔄\mathfrak{A}\cap\mathfrak{A}^{\prime}. The conclusion is clear if SS is empty (as AA and AA’ commute in this case), so we will assume otherwise.

We observe that φ|S,ψ|S\langle\varphi_{|S},\psi_{|S}\rangle is not virtually abelian: otherwise, as AA and AA^{\prime} are amenable subgroups of Mod(V)\mathrm{Mod}(V), there would exist two other virtually abelian subgroups B,BMod(V)B,B^{\prime}\subseteq\mathrm{Mod}(\partial V) (made of mapping classes supported on ΣS\Sigma\setminus S and ΣS\Sigma^{\prime}\setminus S, respectively), which commute and both commute with φ|S,ψ|S\langle\varphi_{|S},\psi_{|S}\rangle, such that every element in φ,ψ\langle\varphi,\psi\rangle is a product of an element in BB, an element in BB^{\prime}, and an element in φ|S,ψ|S\langle\varphi_{|S},\psi_{|S}\rangle. This would contradict the fact that φ,ψ\langle\varphi,\psi\rangle is a nonabelian free group.

Therefore φ|S,ψ|S\langle\varphi_{|S},\psi_{|S}\rangle contains a nonabelian free subgroup f,g\langle f,g\rangle, and the commutator subgroup of f,g\langle f,g\rangle is a nonabelian free group contained in AAA\cap A^{\prime}. This contradiction completes our proof. ∎

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY which admits an action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V). Two admissible pairs (,𝒜)(\mathcal{H},\mathcal{A}) and (,𝒜)(\mathcal{H}^{\prime},\mathcal{A}^{\prime}) (with respect to ρ\rho) are compatible with respect to ρ\rho if, denoting by (X,𝔄)(X,\mathfrak{A}) and (X,𝔄)(X^{\prime},\mathfrak{A}^{\prime}) their respective decomposition maps, for a.e. yYy\in Y, the pairs (X(y),𝔄(y))(X(y),\mathfrak{A}(y)) and (X(y),𝔄(y))(X^{\prime}(y),\mathfrak{A}^{\prime}(y)) are compatible. The following proposition gives a purely groupoid-theoretic characterization of compatibility (i.e. with no reference to the cocycle ρ\rho).

Proposition 3.26.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V). Let (,𝒜)(\mathcal{H},\mathcal{A}) and (,𝒜)(\mathcal{H}^{\prime},\mathcal{A}^{\prime}) be two admissible pairs with respect to ρ\rho. Then the following are equivalent.

  1. 1.

    (,𝒜)(\mathcal{H},\mathcal{A}) and (,𝒜)(\mathcal{H}^{\prime},\mathcal{A}^{\prime}) are compatible with respect to ρ\rho;

  2. 2.

    for every Borel subset UYU\subseteq Y of positive measure, there exists a Borel subset VUV\subseteq U of positive measure such that 𝒜|V,𝒜|V\langle\mathcal{A}_{|V},\mathcal{A}^{\prime}_{|V}\rangle is amenable.

Proof of Proposition 3.26.

Let Y=iIYiY^{*}=\sqcup_{i\in I}Y_{i} be a countable Borel partition of a conull Borel subset YYY^{*}\subseteq Y such that for every iIi\in I, there exist admissible pairs (Xi,𝔄i)(X_{i},\mathfrak{A}_{i}) and (Xi,𝔄i)(X^{\prime}_{i},\mathfrak{A}^{\prime}_{i}) such that |Yi=ρ1(StabMod0(V)(Xi))\mathcal{H}_{|Y_{i}}=\rho^{-1}(\mathrm{Stab}_{\mathrm{Mod}^{0}(V)}(X_{i})) and |Yi=ρ1(StabMod0(V)(Xi))\mathcal{H}^{\prime}_{|Y_{i}}=\rho^{-1}(\mathrm{Stab}_{\mathrm{Mod}^{0}(V)}(X^{\prime}_{i})), and letting Ai,AiMod0(V)A_{i},A^{\prime}_{i}\subseteq\mathrm{Mod}^{0}(V) be the active subgroups of (Xi,𝔄i)(X_{i},\mathfrak{A}_{i}) and (Xi,𝔄i)(X^{\prime}_{i},\mathfrak{A}^{\prime}_{i}) respectively, we have 𝒜|Yi=ρ1(Ai)\mathcal{A}_{|Y_{i}}=\rho^{-1}(A_{i}) and 𝒜|Yi=ρ1(Ai)\mathcal{A}^{\prime}_{|Y_{i}}=\rho^{-1}(A^{\prime}_{i}).

We first prove that ¬1¬2\neg 1\Rightarrow\neg 2. If 1.1. fails, then there exists i0Ii_{0}\in I such that Yi0Y_{i_{0}} has positive measure and (Xi0,𝔄i0)(X_{i_{0}},\mathfrak{A}_{i_{0}}) and (Xi0,𝔄i0)(X^{\prime}_{i_{0}},\mathfrak{A}^{\prime}_{i_{0}}) are not compatible. Then there exist gi0Ai0g_{i_{0}}\in A_{i_{0}} and gi0Ai0g^{\prime}_{i_{0}}\in A^{\prime}_{i_{0}} that generate a nonabelian free subgroup of Mod0(V)\mathrm{Mod}^{0}(V), as follows from [Kob12, Theorem 1.8, and the sentence following it]. Lemma 3.6 ensures that for every Borel subset VYi0V\subseteq Y_{i_{0}} of positive measure, the groupoid ρ1(gi0)|V,ρ1(gi0)|V\langle\rho^{-1}(\langle g_{i_{0}}\rangle)_{|V},\rho^{-1}(\langle g^{\prime}_{i_{0}}\rangle)_{|V}\rangle is nonamenable. Therefore 𝒜|V,𝒜|V\langle\mathcal{A}_{|V},\mathcal{A}^{\prime}_{|V}\rangle is also nonamenable for every Borel subset VYi0V\subseteq Y_{i_{0}} of positive measure, so 2.2. fails.

We now prove that 121\Rightarrow 2. If 1.1. holds, then for every iIi\in I such that YiY_{i} has positive measure, the pairs (Xi,𝔄i)(X_{i},\mathfrak{A}_{i}) and (Xi,𝔄i)(X^{\prime}_{i},\mathfrak{A}^{\prime}_{i}) are compatible, so Ai,Ai\langle A_{i},A^{\prime}_{i}\rangle is amenable (Lemma 3.25). Let now UYU\subseteq Y be a Borel subset of positive measure, and let VUV\subseteq U of positive measure be contained in Yi0Y_{i_{0}} for some i0Ii_{0}\in I. Then 𝒜|V,𝒜|V\langle\mathcal{A}_{|V},\mathcal{A}^{\prime}_{|V}\rangle is contained in ρ1(Ai0,Ai0)|V\rho^{-1}(\langle A_{i_{0}},A^{\prime}_{i_{0}}\rangle)_{|V}, which is amenable because Ai0,Ai0\langle A_{i_{0}},A^{\prime}_{i_{0}}\rangle is and ρ\rho has trivial kernel. ∎

Let ,\mathcal{H},\mathcal{H}^{\prime} be two measured subgroupoids of 𝒢\mathcal{G} of meridian type with respect to an action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V). We say that \mathcal{H} and \mathcal{H}^{\prime} are compatible with respect to ρ\rho if, denoting by φ,φ\varphi,\varphi^{\prime} their respective meridian maps with respect to ρ\rho, for a.e. yYy\in Y, the meridians φ(y)\varphi(y) and φ(y)\varphi^{\prime}(y) are disjoint up to isotopy.

Corollary 3.27.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with two strict action-type cocycles ρ1,ρ2:𝒢Mod0(V)\rho_{1},\rho_{2}:\mathcal{G}\to\mathrm{Mod}^{0}(V), and let ,\mathcal{H},\mathcal{H}^{\prime} be two measured subgroupoids of 𝒢\mathcal{G} of meridian type with respect to both ρ1\rho_{1} and ρ2\rho_{2}.

Then \mathcal{H} and \mathcal{H}^{\prime} are compatible with respect to ρ1\rho_{1} if and only if they are compatible with respect to ρ2\rho_{2}.

Proof.

Let Y=jJYjY^{*}=\sqcup_{j\in J}Y_{j} be a partition of a conull Borel subset YYY^{*}\subseteq Y into at most countably many Borel subsets such that for every i{1,2}i\in\{1,2\} and every jJj\in J, there exist meridians ci,j,ci,jc_{i,j},c^{\prime}_{i,j} such that |Yj,|Yj\mathcal{H}_{|Y_{j}},\mathcal{H}^{\prime}_{|Y_{j}} are equal to the (𝒢|Yj,ρi)(\mathcal{G}_{|Y_{j}},\rho_{i})-stabilizers of the isotopy classes of ci,j,ci,jc_{i,j},c^{\prime}_{i,j}, respectively.

For every i{1,2}i\in\{1,2\} and every jJj\in J, let Ai,jA_{i,j} (resp. Ai,jA^{\prime}_{i,j}) be the subgroup of Mod0(V)\mathrm{Mod}^{0}(V) made of all elements that act trivially in restriction to every connected component of Vci,j\partial V\setminus c_{i,j} (resp. Vci,j\partial V\setminus c^{\prime}_{i,j}) which is not a once-holed torus. Notice that Ai,j,Ai,jA_{i,j},A^{\prime}_{i,j} are the active subgroups of some admissible decorated multicurves (Xi,j,𝔄i,j),(Xi,j,𝔄i,j)(X_{i,j},\mathfrak{A}_{i,j}),(X^{\prime}_{i,j},\mathfrak{A}^{\prime}_{i,j}), by letting Xi,jX_{i,j} and Xi,jX^{\prime}_{i,j} be obtained from ci,jc_{i,j} and ci,jc^{\prime}_{i,j} by adding the unique nonseparating meridian in every complementary component which is a once-holed torus, and letting 𝔄i,j=𝔄i,j=\mathfrak{A}_{i,j}=\mathfrak{A}^{\prime}_{i,j}=\emptyset. See the examples right after Definition 3.23. Notice that ci,jc_{i,j} and ci,jc^{\prime}_{i,j} are disjoint up to isotopy if and only if (Xi,j,)(X_{i,j},\emptyset) and (Xi,j,)(X^{\prime}_{i,j},\emptyset) are compatible.

For every i{1,2}i\in\{1,2\}, let 𝒜i\mathcal{A}_{i}\subseteq\mathcal{H} be a subgroupoid such that (𝒜i)|Yj=ρi1(Ai,j)|Yj(\mathcal{A}_{i})_{|Y_{j}}=\rho_{i}^{-1}(A_{i,j})_{|Y_{j}} for every jJj\in J, and let 𝒜i\mathcal{A}^{\prime}_{i}\subseteq\mathcal{H}^{\prime} be defined in the same way, using Ai,jA^{\prime}_{i,j} in place of Ai,jA_{i,j}. Then (,𝒜i)(\mathcal{H},\mathcal{A}_{i}) and (,𝒜i)(\mathcal{H}^{\prime},\mathcal{A}^{\prime}_{i}) are admissible pairs with respect to ρi\rho_{i}. Lemma 3.19 thus ensures that 𝒜1\mathcal{A}_{1} and 𝒜2\mathcal{A}_{2} are stably equal (as they are both stably maximal for the property of being a stably normal amenable subgroupoid of \mathcal{H}), and likewise 𝒜1\mathcal{A}^{\prime}_{1} and 𝒜2\mathcal{A}^{\prime}_{2} are stably equal. The conclusion therefore follows from Proposition 3.26. ∎

3.9 Characterizing subgroupoids of nonseparating-meridian type

The goal of this section is to prove the following proposition.

Proposition 3.28.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with two strict action-type cocycles ρ1,ρ2:𝒢Mod0(V)\rho_{1},\rho_{2}:\mathcal{G}\to\mathrm{Mod}^{0}(V), and let 𝒢\mathcal{H}\subseteq\mathcal{G} be a measured subgroupoid.

Then \mathcal{H} is of nonseparating-meridian type with respect to ρ1\rho_{1} if and only if it is of nonseparating-meridian type with respect to ρ2\rho_{2}.

A decorated multicurve (X,𝔄)(X,\mathfrak{A}) is clean if it is not of the form (c,)(c,\emptyset) for some separating meridian cc. The graph of clean admissible decorated multicurves 𝕄\mathbb{M} is the graph whose vertices correspond to isotopy classes of clean admissible decorated multicurves, where two distinct vertices are joined by an edge if the corresponding decorated multicurves are compatible. The graph of nonseparating meridians 𝔻nsep\mathbb{D}^{\mathrm{nsep}} is the graph whose vertices correspond to isotopy classes of nonseparating meridians, where two distinct vertices are joined by an edge if the corresponding meridians are disjoint up to isotopy. Notice that 𝔻nsep\mathbb{D}^{\mathrm{nsep}} is naturally a subgraph of 𝕄\mathbb{M}, by sending a nonseparating meridian cc to the pair (c,)(c,\emptyset).

Lemma 3.29.

Every injective graph map222i.e. preserving adjacency and non-adjacency from 𝔻nsep\mathbb{D}^{\mathrm{nsep}} to 𝕄\mathbb{M} takes its values in 𝔻nsep\mathbb{D}^{\mathrm{nsep}} (viewed as a subgraph of 𝕄\mathbb{M} via the natural inclusion).

Proof.

Let vV(𝔻nsep)v\in V(\mathbb{D}^{\mathrm{nsep}}) be a vertex. By completing vv to a pair of pants decomposition made of nonseparating meridians, we can find 3g33g-3 pairwise distinct, pairwise adjacent vertices v=v1,,v3g3v=v_{1},\dots,v_{3g-3} such that for every i{1,,3g3}i\in\{1,\dots,3g-3\}, one has

Lk𝔻nsep({v1,,v3g3})Lk𝔻nsep({v1,,vi1,vi+1,,v3g3}){vi}.\mathrm{Lk}_{\mathbb{D}^{\mathrm{nsep}}}(\{v_{1},\dots,v_{3g-3}\})\subsetneq\mathrm{Lk}_{\mathbb{D}^{\mathrm{nsep}}}(\{v_{1},\dots,v_{i-1},v_{i+1},\dots,v_{3g-3}\})\setminus\{v_{i}\}.

So the same property should hold for their images in 𝕄\mathbb{M}, which correspond to decorated multicurves (X1,𝔄1),,(X3g3,𝔄3g3)(X_{1},\mathfrak{A}_{1}),\dots,(X_{3g-3},\mathfrak{A}_{3g-3}). For every i{1,,3g3}i\in\{1,\dots,3g-3\}, we let Σi\Sigma_{i} be the subsurface of V\partial V equal to the union of all subsurfaces in 𝔄i\mathfrak{A}_{i}, together with all annuli around curves in XiX_{i} that are not boundary curves of any subsurface in 𝔄i\mathfrak{A}_{i}. Notice that the set {Σ1,,Σ3g3}\{\Sigma_{1},\dots,\Sigma_{3g-3}\} cannot contain both a subsurface SS and the collar neighborhood AA of one of its boundary components, as otherwise removing AA from the collection does not change the link. More generally, for every i{1,,3g3}i\in\{1,\dots,3g-3\}, one of the connected components ΣiΣi\Sigma^{\prime}_{i}\subseteq\Sigma_{i} is not a connected component of some Σj\Sigma_{j} with jij\neq i, and is also not the collar neighborhood of a boundary curve of Σj\Sigma_{j} – otherwise removing (Xi,𝔄i)(X_{i},\mathfrak{A}_{i}) from the collection does not change its link. So the subsurfaces Σ1,,Σ3g3\Sigma^{\prime}_{1},\dots,\Sigma^{\prime}_{3g-3} are pairwise nonisotopic and {Σ1,,Σ3g3}\{\Sigma^{\prime}_{1},\dots,\Sigma^{\prime}_{3g-3}\} does not contain a subsurface together with the collar neighborhood of one of its boundary components. For every i{1,,3g3}i\in\{1,\dots,3g-3\}, let {bi,1,,bi,ki}\{b_{i,1},\dots,b_{i,k_{i}}\} be the set of all boundary curves of Σi\Sigma^{\prime}_{i}, and let {di,1,,di,i}\{d_{i,1},\dots,d_{i,\ell_{i}}\} be a set of isotopy classes of essential simple closed curves on Σi\Sigma^{\prime}_{i} that form a pair of pants decomposition of Σi\Sigma^{\prime}_{i} (with the convention that in the case of an annulus, the former set contains two isotopic curves, and the latter set is empty). The tuple consisting of all bi,jb_{i,j} and di,jd_{i,j} contains at least 6g66g-6 curves, each being repeated at most twice up to isotopy (and the di,jd_{i,j} are not isotopic to any other curve in the collection). So every subsurface Σi\Sigma^{\prime}_{i} contributes exactly two curves that are both of the form ci,jc_{i,j}, and is therefore an annular subsurface. Furthermore, since there are 3g33g-3 such, and no Σi\Sigma^{\prime}_{i} appears as a subsurface of Σj,ji\Sigma_{j},j\neq i, we actually have Σi=Σi\Sigma^{\prime}_{i}=\Sigma_{i} for all ii. Therefore (Xi,𝔄i)=(ci,)(X_{i},\mathfrak{A}_{i})=(c_{i},\emptyset), where cic_{i} is the core curve of the annulus Σi\Sigma_{i}. As (Xi,𝔄i)(X_{i},\mathfrak{A}_{i}) is admissible, some power of the twist around cic_{i} must belong to the handlebody group, so cic_{i} is a meridian by [Oer02, Theorem 1.11] or [McC06, Theorem 1]. As (Xi,𝔄i)(X_{i},\mathfrak{A}_{i}) is clean, the meridian cic_{i} is nonseparating, and the conclusion follows. ∎

Proof of Proposition 3.28.

Let vV(𝔻nsep)v\in V(\mathbb{D}^{\mathrm{nsep}}), in other words vv is the isotopy class of a nonseparating meridian. Let v\mathcal{H}_{v} be the (𝒢,ρ1)(\mathcal{G},\rho_{1})-stabilizer of vv, and let 𝒜v=ρ11(Tv)\mathcal{A}_{v}=\rho_{1}^{-1}(\langle T_{v}\rangle). Then (v,𝒜v)(\mathcal{H}_{v},\mathcal{A}_{v}) satisfies Property (Qnsep)(\mathrm{Q}_{\mathrm{nsep}}) (by Proposition 3.18, applied to the cocycle ρ1\rho_{1}). Lemma 3.24, applied to the cocycle ρ2\rho_{2}, implies that (v,𝒜v)(\mathcal{H}_{v},\mathcal{A}_{v}) is an admissible pair with respect to ρ2\rho_{2}. So there exist a conull Borel subset YYY^{*}\subseteq Y and a partition Y=iIYv,iY^{*}=\sqcup_{i\in I}Y_{v,i} into at most countably many Borel subsets such that for every iIi\in I, there exists a (unique) admissible pair (Xv,i,𝔄v,i)(X_{v,i},\mathfrak{A}_{v,i}) such that (v)|Yv,i(\mathcal{H}_{v})_{|Y_{v,i}} is the (𝒢|Yv,i,ρ2)(\mathcal{G}_{|Y_{v,i}},\rho_{2})-stabilizer of Xv,iX_{v,i} and, denoting by Av,iA_{v,i} the active subgroup of (Xv,i,𝔄v,i)(X_{v,i},\mathfrak{A}_{v,i}), one has (𝒜v)|Yv,i=ρ1(Av,i)|Yv,i(\mathcal{A}_{v})_{|Y_{v,i}}=\rho^{-1}(A_{v,i})_{|Y_{v,i}}. In addition, Lemma 3.22 ensures that (Xv,i,𝔄v,i)(X_{v,i},\mathfrak{A}_{v,i}) is clean. For every yYy\in Y and every vV(𝔻nsep)v\in V(\mathbb{D}^{\mathrm{nsep}}), we then let θ(y,v)=(Xv,i,𝔄v,i)\theta(y,v)=(X_{v,i},\mathfrak{A}_{v,i}) whenever yYv,iy\in Y_{v,i}. This defines a Borel map θ:Y×V(𝔻nsep)V(𝕄)\theta:Y\times V(\mathbb{D}^{\mathrm{nsep}})\to V(\mathbb{M}).

We claim that for almost every yYy\in Y, the map θ(y,)\theta(y,\cdot) determines a graph embedding 𝔻nsep𝕄\mathbb{D}^{\mathrm{nsep}}\to\mathbb{M}. Let us first explain how to complete the proof of the proposition from this claim. By Lemma 3.29, every graph embedding 𝔻nsep𝕄\mathbb{D}^{\mathrm{nsep}}\to\mathbb{M} sends nonseparating meridians to nonseparating meridians. Therefore, if vv is a nonseparating meridian, then Xv,iX_{v,i} is a nonseparating meridian (and 𝔄v,i=\mathfrak{A}_{v,i}=\emptyset) whenever Yv,iY_{v,i} has positive measure, and the proposition follows.

We are now left with proving the above claim. First, for almost every yYy\in Y, the map θ(y,)\theta(y,\cdot) is injective. Indeed otherwise, as V(𝔻nsep)V(\mathbb{D}^{\mathrm{nsep}}) is countable, there exist a Borel subset UYU\subseteq Y of positive measure and two non-isotopic nonseparating meridians c,cc,c^{\prime} such that for every yUy\in U, one has θ(y,c)=θ(y,c)\theta(y,c)=\theta(y,c^{\prime}) (we denote by (X,𝔄)(X,\mathfrak{A}) the common image). In particular, the (𝒢|U,ρ1)(\mathcal{G}_{|U},\rho_{1})-stabilizer of cc is stably equal to the (𝒢|U,ρ1)(\mathcal{G}_{|U},\rho_{1})-stabilizer of cc^{\prime}, since they are both stably equal to the (𝒢|U,ρ2)(\mathcal{G}_{|U},\rho_{2})-stabilizer of XX. This contradicts Lemma 3.13.

Second, Proposition 3.26 ensures that for almost every yYy\in Y, the map θ(y,)\theta(y,\cdot) is a graph map, i.e. it preserves both adjacency and non-adjacency. ∎

3.10 Characterizing subgroupoids of separating-meridian type

In this section, we establish a purely groupoid-theoretic characterization of subgroupoids of separating-meridian type with respect to a strict action-type cocycle towards Mod0(V)\mathrm{Mod}^{0}(V), and derive that being of separating-meridian type is a notion that does not depend on the choice of such a cocycle.

3.10.1 Property (Psep)(\mathrm{P}_{\mathrm{sep}})

We proved in Proposition 3.28 that for a subgroupoid 𝒢\mathcal{H}\subseteq\mathcal{G}, being of nonseparating meridian type does not depend of the choice of an action-type cocycle 𝒢Mod0(V)\mathcal{G}\to\mathrm{Mod}^{0}(V). Also, it follows from Corollary 3.27 that compatibility of two subgroupoids of nonseparating meridian type is also independent of such a choice. Thus, the following notion is a purely groupoid-theoretic property.

Definition 3.30 (Property (Psep)(\mathrm{P}_{\mathrm{sep}})).

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY which admits a strict action-type cocycle towards Mod0(V)\mathrm{Mod}^{0}(V). A measured subgroupoid 𝒢\mathcal{H}\subseteq\mathcal{G} satisfies Property (Psep)(\mathrm{P}_{\mathrm{sep}}) if

  1. 1.

    \mathcal{H} contains a strongly Schottky pair of subgroupoids;

  2. 2.

    there exists a stably normal amenable subgroupoid \mathcal{B}\subseteq\mathcal{H} of infinite type, such that for every measured subgroupoid 𝒢\mathcal{H}^{\prime}\subseteq\mathcal{G} of nonseparating-meridian type, and every stably normal amenable subgroupoid 𝒜\mathcal{A}\subseteq\mathcal{H}^{\prime} of infinite type, the intersection 𝒜\mathcal{A}\cap\mathcal{B} is stably trivial;

  3. 3.

    given any two subgroupoids 1,2𝒢\mathcal{H}_{1},\mathcal{H}_{2}\subseteq\mathcal{G} of nonseparating-meridian type, and any Borel subset UYU\subseteq Y of positive measure, assuming that |U(12)|U\mathcal{H}_{|U}\subseteq(\mathcal{H}_{1}\cap\mathcal{H}_{2})_{|U}, then (1)|U(\mathcal{H}_{1})_{|U} and (2)|U(\mathcal{H}_{2})_{|U} are stably equal;

  4. 4.

    there exist 3g43g-4 measured subgroupoids 1,,3g4\mathcal{H}_{1},\dots,\mathcal{H}_{3g-4} of 𝒢\mathcal{G} of nonseparating-meridian type, which are pairwise compatible, such that

    1. (a)

      for every Borel subset UYU\subseteq Y of positive measure, and any two distinct i,j{1,,3g4}i,j\in\{1,\dots,3g-4\}, one has (i)|U(j)|U(\mathcal{H}_{i})_{|U}\neq(\mathcal{H}_{j})_{|U}, and

    2. (b)

      for every j{1,,3g4}j\in\{1,\dots,3g-4\}, every stably normal amenable subgroupoid 𝒜j\mathcal{A}_{j} of j\mathcal{H}_{j} is stably contained in \mathcal{H}.

Notice that this notion is stable under restriction to a positive measure subset. Also, if \mathcal{H} is a subgroupoid of 𝒢\mathcal{G}, and if there exists a partition Y=iIYiY=\sqcup_{i\in I}Y_{i} into at most countably many Borel subsets such that for every iIi\in I, the subgroupoid |Yi\mathcal{H}_{|Y_{i}} of 𝒢|Yi\mathcal{G}_{|Y_{i}} satisfies Property (Psep)(\mathrm{P}_{\mathrm{sep}}), then \mathcal{H} (as a subgroupoid of 𝒢\mathcal{G}) satisfies Property (Psep)(\mathrm{P}_{\mathrm{sep}}).

To motivate the definition, we begin by proving that subgroupoids of separating meridian type have this property.

Proposition 3.31.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V). Let cc be a separating meridian, and let \mathcal{H} be the (𝒢,ρ)(\mathcal{G},\rho)-stabilizer of the isotopy class of cc.

Then \mathcal{H} satisfies Property (Psep)(\mathrm{P}_{\mathrm{sep}}).

Proof.

For Property (Psep)(1)(\mathrm{P}_{\mathrm{sep}})(1), Lemma 1.1 ensures that the stabilizer in Mod0(V)\mathrm{Mod}^{0}(V) of any separating meridian contains a nonabelian free subgroup (recall that we are assuming that the genus of VV is at least 33). Lemma 3.6 thus implies that \mathcal{H} contains a strongly Schottky pair of subgroupoids.

For Property (Psep)(2)(\mathrm{P}_{\mathrm{sep}})(2), notice that the cyclic subgroup Tc\langle T_{c}\rangle generated by the Dehn twist about cc is normal in the stabilizer of cc. Therefore =ρ1(Tc)\mathcal{B}=\rho^{-1}(\langle T_{c}\rangle) is a normal subgroupoid of \mathcal{H}, which is amenable as ρ\rho has trivial kernel, and of infinite type because ρ\rho is action-type. Let now \mathcal{H}^{\prime} be a measured subgroupoid of 𝒢\mathcal{G} of nonseparating meridian type, and let 𝒜\mathcal{A}\subseteq\mathcal{H}^{\prime} be a stably normal amenable subgroupoid of infinite type. By Lemma 3.19, we can find a partition Y=iIYiY^{*}=\sqcup_{i\in I}Y_{i} of a conull Borel subset YYY^{*}\subseteq Y into at most countably many Borel subsets such that for every iIi\in I, there exists a nonseparating meridian did_{i} such that 𝒜|Yiρ1(Tdi)|Yi\mathcal{A}_{|Y_{i}}\subseteq\rho^{-1}(\langle T_{d_{i}}\rangle)_{|Y_{i}}. It follows that (𝒜)|Yi(\mathcal{A}\cap\mathcal{B})_{|Y_{i}} is trivial, so 𝒜\mathcal{A}\cap\mathcal{B} is stably trivial.

Property (Psep)(3)(\mathrm{P}_{\mathrm{sep}})(3) follows from the fact that for every separating meridian cc, there is at most one nonseparating meridian dd such that every infinite-order element of StabMod0(V)(c)\mathrm{Stab}_{\mathrm{Mod}^{0}(V)}(c) has a power that fixes dd (in fact, the existence of such a dd occurs precisely when one of the two connected components of Vc\partial V\setminus c is a once-holed torus, in which case it contains a unique nonseparating meridian up to isotopy, and we take dd as such – notice that we are using the fact that the genus of VV is at least 33 here; see Lemma 1.2).

We now prove that \mathcal{H} satisfies Property (Psep)(4)(\mathrm{P}_{\mathrm{sep}})(4). Let {c1,,c3g4}\{c_{1},\dots,c_{3g-4}\} be a set of 3g43g-4 pairwise non-isotopic nonseparating meridians which together with cc form a pair of pants decomposition of V\partial V. For every j{1,,3g4}j\in\{1,\dots,3g-4\}, let j\mathcal{H}_{j} be the (𝒢,ρ)(\mathcal{G},\rho)-stabilizer of the isotopy class of cjc_{j}. Then 1,,3g4\mathcal{H}_{1},\dots,\mathcal{H}_{3g-4} are of nonseparating meridian type. Lemma 3.13 ensures that they satisfy Assertion (4.a)(4.a). Finally, Lemma 3.19 ensures that every stably normal amenable subgroupoid 𝒜j\mathcal{A}_{j} of j\mathcal{H}_{j} is stably contained in ρ1(Tcj)\rho^{-1}(\langle T_{c_{j}}\rangle). In particular each 𝒜j\mathcal{A}_{j} is stably contained in \mathcal{H}. ∎

3.10.2 Characterization

Our goal is now to characterise subgroups of separating meridian-type by proving the following proposition.

Proposition 3.32.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V). Let \mathcal{H} be a measured subgroupoid of 𝒢\mathcal{G}. The following assertions are equivalent.

  1. 1.

    The subgroupoid \mathcal{H} is of separating-meridian type with respect to ρ\rho.

  2. 2.

    The subgroupoid \mathcal{H} satisfies Property (Psep)(\mathrm{P}_{\mathrm{sep}}), and is stably maximal among all measured subgroupoids of 𝒢\mathcal{G} with respect to this property, i.e. if \mathcal{H}^{\prime} is another subgroupoid satisfying Property (Psep)(\mathrm{P}_{\mathrm{sep}}), and if \mathcal{H} is stably contained in \mathcal{H}^{\prime}, then \mathcal{H} is stably equal to \mathcal{H}^{\prime}.

Before turning to the proof of Proposition 3.32, we record the following consequence.

Corollary 3.33.

Let 𝒢\mathcal{G} be a measured groupoid over a base space YY, equipped with two strict action-type cocycles ρ1,ρ2:𝒢Mod0(V)\rho_{1},\rho_{2}:\mathcal{G}\to\mathrm{Mod}^{0}(V), and let 𝒢\mathcal{H}\subseteq\mathcal{G} be a measured subgroupoid.

Then \mathcal{H} is of separating-meridian type with respect to ρ1\rho_{1} if and only if it is of separating-meridian type with respect to ρ2\rho_{2}. ∎

Our goal is now to prove Proposition 3.32. Our first lemma exploits the first two assumptions of Property (Psep)(\mathrm{P}_{\mathrm{sep}}) in order to derive information about the possible canonical reduction multicurves of \mathcal{B}.

Lemma 3.34.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V). Let ,\mathcal{B},\mathcal{H} be measured subgroupoids of 𝒢\mathcal{G}, with \mathcal{B}\subseteq\mathcal{H}. Assume that

  1. 1.

    \mathcal{H} contains a strongly Schottky pair of subgroupoids;

  2. 2.

    \mathcal{B} is amenable and of infinite type, and stably normal in \mathcal{H};

  3. 3.

    for every measured subgroupoid 𝒢\mathcal{H}^{\prime}\subseteq\mathcal{G} of nonseparating-meridian type, and every stably normal amenable subgroupoid 𝒜\mathcal{A}\subseteq\mathcal{H}^{\prime} of infinite type, the intersection 𝒜\mathcal{A}\cap\mathcal{B} is stably trivial.

Then for every Borel subset UYU\subseteq Y of positive measure, the pair (|U,ρ)(\mathcal{B}_{|U},\rho) cannot have a canonical reduction multicurve consisting of a single nonseparating meridian.

Proof.

Assume towards a contradiction that (|U,ρ)(\mathcal{B}_{|U},\rho) has a canonical reduction multicurve which is reduced to a single nonseparating meridian cc. As \mathcal{B} is stably normal in \mathcal{H}, up to restricting to a positive measure Borel subset of UU, we can assume that cc is (|U,ρ)(\mathcal{H}_{|U},\rho)-invariant. In particular, letting Σ=Vc\Sigma=\partial V\setminus c, we have a natural cocycle ρ:|UMod0(Σ)\rho^{\prime}:\mathcal{H}_{|U}\to\mathrm{Mod}^{0}(\Sigma). The kernel of ρ\rho^{\prime} is contained in ρ1(Tc)|U\rho^{-1}(\langle T_{c}\rangle)_{|U}. As ρ\rho has trivial kernel, it follows that the kernel of ρ\rho^{\prime} is amenable. In particular, letting (𝒜1,𝒜2)(\mathcal{A}^{1},\mathcal{A}^{2}) be a strongly Schottky pair of subgroupoids of \mathcal{H}, there exists a positive measure Borel subset VUV\subseteq U such that ρ\rho^{\prime} has trivial kernel when restricted to 𝒜|V1,𝒜|V2\langle\mathcal{A}^{1}_{|V},\mathcal{A}^{2}_{|V}\rangle.

Our third assumption, applied by taking for \mathcal{H}^{\prime} the (𝒢,ρ)(\mathcal{G},\rho)-stabilizer of cc, and with 𝒜=ρ1(Tc)\mathcal{A}=\rho^{-1}(\langle T_{c}\rangle), ensures that ρ1(Tc)\mathcal{B}\cap\rho^{-1}(\langle T_{c}\rangle) is stably trivial. Let WVW\subseteq V be a Borel subset of positive measure such that (ρ1(Tc))|W(\mathcal{B}\cap\rho^{-1}(\langle T_{c}\rangle))_{|W} is trivial. Then ρ\rho^{\prime} also has trivial kernel when restricted to |W\mathcal{B}_{|W}. In particular (|W,ρ)(\mathcal{B}_{|W},\rho^{\prime}) is irreducible, and Lemma 3.12 implies that 𝒜|W1,𝒜|W2\langle\mathcal{A}^{1}_{|W},\mathcal{A}^{2}_{|W}\rangle is amenable, a contradiction. ∎

Lemma 3.35.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, equipped with a strict action-type cocycle ρ:𝒢Mod0(V)\rho:\mathcal{G}\to\mathrm{Mod}^{0}(V). Let \mathcal{H} be a measured subgroupoid of 𝒢\mathcal{G} which satisfies Property (Psep)(\mathrm{P}_{\mathrm{sep}}).

Then there exists a partition Y=iIYiY=\sqcup_{i\in I}Y_{i} into at most countably many Borel subsets such that for every iIi\in I, there exists an (|Yi,ρ)(\mathcal{H}_{|Y_{i}},\rho)-invariant isotopy class of separating meridian.

Proof.

Let 1,,3g4\mathcal{H}_{1},\dots,\mathcal{H}_{3g-4} be subgroupoids of 𝒢\mathcal{G} provided by Property (Psep)(4)(\mathrm{P}_{\mathrm{sep}})(4). Up to partitioning YY into at most countably many Borel subsets, we can assume that for every j{1,,3g4}j\in\{1,\dots,3g-4\}, the groupoid j\mathcal{H}_{j} is equal to the (𝒢,ρ)(\mathcal{G},\rho)-stabilizer of the isotopy class of a nonseparating meridian djd_{j}.

Let \mathcal{B}\subseteq\mathcal{H} be as in Property (Psep)(2)(\mathrm{P}_{\mathrm{sep}})(2). By Lemma 3.9, we can find a partition Y=iIYiY=\sqcup_{i\in I}Y_{i} into at most countably many Borel subsets of positive measure such that for every iIi\in I, the pair (|Yi,ρ)(\mathcal{B}_{|Y_{i}},\rho) has a canonical reduction set 𝒞i\mathcal{C}_{i}, with boundary multicurve XiX_{i}. As \mathcal{B} is stably normal in \mathcal{H}, up to refining this partition, we can assume that for every iIi\in I, the isotopy class of the multicurve XiX_{i} is (|Yi,ρ)(\mathcal{H}_{|Y_{i}},\rho)-invariant.

We first observe that for every iIi\in I, one has 𝒞i\mathcal{C}_{i}\neq\emptyset. Indeed, otherwise (|Yi,ρ)(\mathcal{B}_{|Y_{i}},\rho) is irreducible. As \mathcal{B} is amenable and stably normal in \mathcal{H}, and ρ\rho has trivial kernel, Lemma 3.12 implies that |Yi\mathcal{H}_{|Y_{i}} is amenable, contradicting Property (Psep)(1)(\mathrm{P}_{\mathrm{sep}})(1).

For every j{1,,3g4}j\in\{1,\dots,3g-4\}, let TjT_{j} be the Dehn twist about the meridian djd_{j}. Then 𝒜j=ρ1(Tj)\mathcal{A}_{j}=\rho^{-1}(\langle T_{j}\rangle) is a normal amenable subgroupoid of j\mathcal{H}_{j}. Property (Psep)(4.b)(\mathrm{P}_{\mathrm{sep}})(4.b) thus ensures that 𝒜j\mathcal{A}_{j} is stably contained in \mathcal{H}. Therefore, for every curve cc in XiX_{i}, there exists a positive integer kk such that the isotopy class of cc is fixed by TjkT_{j}^{k}. This implies that XiX_{i} is disjoint (up to isotopy) from all meridians djd_{j}.

We now claim that for every iIi\in I, the multicurve XiX_{i} contains at most one of the curves djd_{j}. Indeed, assume by contradiction that it contains two curves dj1d_{j_{1}} and dj2d_{j_{2}}. Then |Yi\mathcal{H}_{|Y_{i}} is contained in (j1j2)|Yi(\mathcal{H}_{j_{1}}\cap\mathcal{H}_{j_{2}})_{|Y_{i}}. As j1\mathcal{H}_{j_{1}} and j2\mathcal{H}_{j_{2}} are of nonseparating meridian type with respect to ρ\rho, Property (Psep)(3)(\mathrm{P}_{\mathrm{sep}})(3) implies that there exists a positive measure Borel subset UYiU\subseteq Y_{i} such that (j1)|U=(j2)|U(\mathcal{H}_{j_{1}})_{|U}=(\mathcal{H}_{j_{2}})_{|U}, contradicting Property (Psep)(4.a)(\mathrm{P}_{\mathrm{sep}})(4.a).

As {d1,,d3g4}\{d_{1},\dots,d_{3g-4}\} is a set of 3g43g-4 pairwise disjoint and pairwise non-isotopic nonseparating simple closed curves on V\partial V, one of the complementary components of the union of all curves djd_{j} is a 44-holed sphere SS. Notice that every essential simple closed curve contained in SS is a meridian, and XiX_{i} may contain such a curve. This leaves three possibilities for the canonical reduction multicurve of (|Yi,ρ)(\mathcal{B}_{|Y_{i}},\rho), namely:

  1. 1.

    a single nonseparating meridian (either one of the meridians djd_{j}, or else a nonseparating meridian contained in SS);

  2. 2.

    the union of a nonseparating meridian djd_{j} and a nonseparating essential simple closed curve (in fact a meridian) contained in SS;

  3. 3.

    a separating (on V\partial V) essential simple closed curve (in fact a meridian) contained in SS, possibly together with a meridian djd_{j}.

The first case is excluded by Lemma 3.34, the second case is excluded using Property (Psep)(3)(\mathrm{P}_{\mathrm{sep}})(3) and Lemma 3.13, and the last case leads to the desired conclusion of our lemma. ∎

Proof of Proposition 3.32.

We first prove that (1)(2)(1)\Rightarrow(2). Let \mathcal{H} be a measured subgroupoid of 𝒢\mathcal{G} of separating meridian type with respect to ρ\rho, and let Y=iIYiY^{*}=\sqcup_{i\in I}Y_{i} be a partition of a conull Borel subset YYY^{*}\subseteq Y into at most countably many Borel subsets, such that for every iIi\in I, the groupoid |Yi\mathcal{H}_{|Y_{i}} is equal to the (𝒢|Yi,ρ)(\mathcal{G}_{|Y_{i}},\rho)-stabilizer of the isotopy class of a separating meridian cic_{i}.

Proposition 3.31 implies that \mathcal{H} satisfies Property (Psep)(\mathrm{P}_{\mathrm{sep}}). We need to check that \mathcal{H} is stably maximal among all measured subgroupoids of 𝒢\mathcal{G} that satisfy Property (Psep)(\mathrm{P}_{\mathrm{sep}}). So let \mathcal{H}^{\prime} be a measured subgroupoid of 𝒢\mathcal{G} which satisfies Property (Psep)(\mathrm{P}_{\mathrm{sep}}), and such that \mathcal{H} is stably contained in \mathcal{H}^{\prime}. By Lemma 3.35, up to refining the above partition of YY, we can assume that for every iIi\in I, there exists a separating meridian cic^{\prime}_{i} whose isotopy class is (|Yi,ρ)(\mathcal{H}^{\prime}_{|Y_{i}},\rho)-invariant. Lemma 3.14 implies that ci=cic_{i}=c^{\prime}_{i} for every iIi\in I. It follows that \mathcal{H}^{\prime} is stably contained in \mathcal{H}, so they are stably equal. This completes our proof of the implication (1)(2)(1)\Rightarrow(2).

We now prove that (2)(1)(2)\Rightarrow(1), so let \mathcal{H} be a measured groupoid of 𝒢\mathcal{G} that satisfies Assertion (2)(2). By Lemma 3.35, we can find a partition Y=iIYiY=\sqcup_{i\in I}Y_{i} into at most countably many Borel subsets such that for every iIi\in I, there exists a separating meridian cic_{i} whose isotopy class is (|Yi,ρ)(\mathcal{H}_{|Y_{i}},\rho)-invariant. Let \mathcal{H}^{\prime} be a measured subgroupoid of 𝒢\mathcal{G} such that for every iIi\in I, the groupoid |Yi\mathcal{H}^{\prime}_{|Y_{i}} is equal to the (𝒢|Yi,ρ)(\mathcal{G}_{|Y_{i}},\rho)-stabilizer of the isotopy class of cic_{i}. Then \mathcal{H} is stably contained in \mathcal{H}^{\prime}. In addition, \mathcal{H}^{\prime} is of separating meridian type, so Proposition 3.31 shows that \mathcal{H}^{\prime} satisfies Property (Psep)(\mathrm{P}_{\mathrm{sep}}). The maximality assumption on \mathcal{H} therefore implies that \mathcal{H} is stably equal to \mathcal{H}^{\prime}. Hence \mathcal{H} itself is of separating meridian type, which concludes our proof. ∎

3.11 Conclusion

Before concluding the proof of our main theorem, we first record the following easy consequence of Propositions 3.28 and 3.32.

Proposition 3.36.

Let 𝒢\mathcal{G} be a measured groupoid, equipped with two strict action-type cocycles ρ1,ρ2:𝒢Mod0(V)\rho_{1},\rho_{2}:\mathcal{G}\to\mathrm{Mod}^{0}(V), and let 𝒢\mathcal{H}\subseteq\mathcal{G} be a measured subgroupoid.

Then \mathcal{H} is of meridian type with respect to ρ1\rho_{1} if and only if it is of meridian type with respect to ρ2\rho_{2}. ∎

We will now simply say that \mathcal{H} is of meridian type to mean that it is of meridian type with respect to any action-type cocycle 𝒢Mod0(V)\mathcal{G}\to\mathrm{Mod}^{0}(V). We are now in position to complete the proof of Theorem 3.2, which as we already explained at the beginning of this section yields the measure equivalence superrigidity of handlebody groups in genus at least 33.

Proof of Theorem 3.2.

Let 𝒢\mathcal{G} be a measured groupoid over a standard probability space YY, and let ρ1,ρ2:𝒢Mod0(V)\rho_{1},\rho_{2}:\mathcal{G}\to\mathrm{Mod}^{0}(V) be two strict action-type cocycles. Let 𝔻\mathbb{D} be the disk graph of VV: we recall that its vertices are the isotopy classes of meridians in V\partial V, and two such isotopy classes are joined by an edge if they have disjoint representatives.

Proposition 3.36 ensures that for every vertex vV(𝔻)v\in V(\mathbb{D}), there exists a Borel map ϕv:YV(𝔻)\phi_{v}:Y\to V(\mathbb{D}) such that for every wV(𝔻)w\in V(\mathbb{D}), letting Yv,w=ϕv1(w)Y_{v,w}=\phi_{v}^{-1}(w), the (𝒢|Yv,w,ρ1)(\mathcal{G}_{|Y_{v,w}},\rho_{1})-stabilizer of vv is stably equal to the (𝒢|Yv,w,ρ2)(\mathcal{G}_{|Y_{v,w}},\rho_{2})-stabilizer of ww. Lemmas 3.13 and 3.14 ensure that the map ϕv\phi_{v} is essentially unique.

For every yYy\in Y and every vV(𝔻)v\in V(\mathbb{D}), we then let ψ(y,v)=ϕv(y)\psi(y,v)=\phi_{v}(y). This defines a Borel map ψ:Y×V(𝔻)V(𝔻)\psi:Y\times V(\mathbb{D})\to V(\mathbb{D}).

We claim that for a.e. yYy\in Y, the map ψ(y,)\psi(y,\cdot) is a graph automorphism of 𝔻\mathbb{D}. Indeed, injectivity follows from the same argument as in the proof of Proposition 3.28, and the fact that ψ(y,)\psi(y,\cdot) is almost everywhere a graph map follows from Corollary 3.27. We now show that for almost every yYy\in Y, the map ψ(y,)\psi(y,\cdot) is surjective. So let cc be a meridian. By Proposition 3.36, there exists a Borel partition of a conull Borel subset YYY^{*}\subseteq Y into at most countably many Borel subsets YiY_{i} such that for every ii, the (𝒢|Yi,ρ2)(\mathcal{G}_{|Y_{i}},\rho_{2})-stabilizer of cc coincides with the (𝒢|Yi,ρ1)(\mathcal{G}_{|Y_{i}},\rho_{1})-stabilizer of some ciV(𝔻)c_{i}\in V(\mathbb{D}). It follows that ψ(y,ci)=c\psi(y,c_{i})=c for almost every yYiy\in Y_{i}. Surjectivity follows.

By the main theorem of [KS09], the natural map Mod(V)Aut(𝔻)\mathrm{Mod}(V)\to\mathrm{Aut}(\mathbb{D}) is an isomorphism (noting again that the genus of VV is at least 33). We can thus find a Borel map θ:YMod(V)\theta:Y\to\mathrm{Mod}(V) so that for a.e. yYy\in Y we have that ψ(y,δ)=θ(y)(δ)\psi(y,\delta)=\theta(y)(\delta) for all meridians.

We are left with showing that θ\theta satisfies the equivariance condition required in Theorem 3.2. This amounts to proving that there exists a conull Borel subset YYY^{*}\subseteq Y such that for every g𝒢|Yg\in\mathcal{G}_{|Y^{*}} and every vertex vV(𝔻)v\in V(\mathbb{D}), one has ψ(r(g),ρ1(g)v)=ρ2(g)ψ(s(g),v).\psi(r(g),\rho_{1}(g)v)=\rho_{2}(g)\psi(s(g),v). As 𝒢\mathcal{G} is a countable union of bisections, it is enough to prove it for almost every gg in a bisection BB (inducing a Borel isomorphism between U=s(B)U=s(B) and V=r(B)V=r(B)). Up to further partitioning BB, we can assume that (ρ1)|B(\rho_{1})_{|B} and (ρ2)|B(\rho_{2})_{|B} are constant, with values γ1,γ2\gamma_{1},\gamma_{2}, and that ψ(,v)|U\psi(\cdot,v)_{|U} is constant, with value ww. We now aim to show that for almost every yVy\in V, one has ψ(y,γ1v)=γ2w\psi(y,\gamma_{1}v)=\gamma_{2}w. By definition of ψ\psi, the (𝒢|U,ρ1)(\mathcal{G}_{|U},\rho_{1})-stabilizer of vv is stably equal to the (𝒢|U,ρ2)(\mathcal{G}_{|U},\rho_{2})-stabilizer of ww. Conjugating by the bisection, it follows that the (𝒢|V,ρ1)(\mathcal{G}_{|V},\rho_{1})-stabilizer of γ1v\gamma_{1}v is stably equal to the (𝒢|V,ρ2)(\mathcal{G}_{|V},\rho_{2})-stabilizer of γ2w\gamma_{2}w, which is exactly what we wanted to show. ∎

4 Applications

4.1 Lattice embeddings and automorphisms of the Cayley graph

A first consequence of our work is that handlebody groups do not admit any interesting lattice embeddings in locally compact second countable groups.

Theorem 4.1.

Let VV be a handlebody of genus at least 33. Let GG be a locally compact second countable group, equipped with its (left or right) Haar measure. Let Γ\Gamma be a finite index subgroup of Mod(V)\mathrm{Mod}(V), and let σ:ΓG\sigma:\Gamma\to G be an injective homomorphism whose image is a lattice.

Then there exists a homomorphism θ:GMod(V)\theta:G\to\mathrm{Mod}(V) with compact kernel such that for every fΓf\in\Gamma, one has θσ(f)=f\theta\circ\sigma(f)=f.

Proof.

Theorem 3.2 precisely says that Mod(V)\mathrm{Mod}(V) is rigid with respect to action-type cocycles in the sense of [GH21, Definition 4.1]. As Mod(V)\mathrm{Mod}(V) is ICC (Lemma 1.6), the theorem thus follows from [GH21, Theorem 4.7].333Theorem 4.7 from [GH21] records works of Furman [Fur11a] and Kida [Kid10, Theorem 8.1]. The idea behind its proof is that the lattice embedding σ\sigma determines a self measure equivalence coupling of Γ\Gamma (acting on GG equipped with its Haar measure), and the rigidity statement provided by Theorem 3.2 from the present paper ensures that the self coupling of Γ\Gamma on GG factors through the obvious coupling on Mod(V)\mathrm{Mod}(V) where Γ\Gamma acts by left/right multiplication. This yields a Borel map GMod(V)G\to\mathrm{Mod}(V), and some extra work is needed to upgrade it to a continuous homomorphism.

A theorem of Suzuki ensures that Mod(V)\mathrm{Mod}(V) is finitely generated [Suz77] (it is in fact finitely presented by work of Wajnryb [Waj98]). Given a finitely generated group GG and a finite generating set SS of GG, the Cayley graph Cay(G,S)\mathrm{Cay}(G,S) is defined as the simple graph whose vertices are the elements of GG, with an edge between distinct elements g,hg,h if g1hSS1g^{-1}h\in S\cup S^{-1}.

Theorem 4.2.

Let VV be a handlebody of genus at least 33.

  1. 1.

    For every finite generating set SS of Mod(V)\mathrm{Mod}(V), every automorphism of Cay(Mod(V),S)\mathrm{Cay}(\mathrm{Mod}(V),S) is at bounded distance from the left multiplication by an element of Mod(V)\mathrm{Mod}(V).

  2. 2.

    For every torsion-free finite-index subgroup ΓMod(V)\Gamma\subseteq\mathrm{Mod}(V) and every finite generating set SS^{\prime} of Γ\Gamma, the automorphism group of Cay(Γ,S)\mathrm{Cay}(\Gamma,S^{\prime}) is countable (in fact it embeds as a subgroup of Mod(V)\mathrm{Mod}(V) containing Γ\Gamma).

Proof.

Using the fact that Mod(V)\mathrm{Mod}(V) is ICC, this follows from Theorem 3.2 and [GH21, Corollary 4.8] (the idea behind the proof is to view Mod(V)\mathrm{Mod}(V) as a cocompact lattice in the automorphism group of its Cayley graph and apply the previous theorem). ∎

As mentioned in the introduction, torsion-freeness of Γ\Gamma is crucial in the second conclusion in view of [dlST19, Lemma 6.1].

4.2 Orbit equivalence rigidity and von Neumann algebras

Seminal work of Furman [Fur99b] has shown that measure equivalence ridigity is intimately related to orbit equivalence rigidity of ergodic group actions. In fact two countable groups are measure equivalent if and only if they admit stably orbit equivalent free measure-preserving ergodic actions by Borel automorphisms on standard probability spaces, see [Gab02, Proposition 6.2].

Orbit equivalence rigidity.

Let Γ1\Gamma_{1} and Γ2\Gamma_{2} be two countable groups, and for every i{1,2}i\in\{1,2\}, let (Xi,μi)(X_{i},\mu_{i}) be a standard probability space equipped with a free ergodic measure-preserving action of Γi\Gamma_{i}.

The actions Γ1X1\Gamma_{1}\curvearrowright X_{1} and Γ2X2\Gamma_{2}\curvearrowright X_{2} are virtually conjugate (as in [Kid08b, Definition 1.3]) if there exist finite normal subgroups FiΓiF_{i}\unlhd\Gamma_{i}, finite-index subgroups QiΓi/FiQ_{i}\subseteq\Gamma_{i}/F_{i}, and free ergodic measure-preserving actions QiYiQ_{i}\curvearrowright Y_{i} on standard probability spaces, so that Q1Y1Q_{1}\curvearrowright Y_{1} and Q2Y2Q_{2}\curvearrowright Y_{2} are conjugate, and for every i{1,2}i\in\{1,2\}, the action of Γi/Fi\Gamma_{i}/F_{i} on Xi/FiX_{i}/F_{i} is induced from the QiQ_{i}-action on YiY_{i}. This implies in particular that the groups Γ1\Gamma_{1} and Γ2\Gamma_{2} are virtually isomorphic (i.e. commensurable up to finite kernels).

The following is a weaker notion. The actions Γ1X1\Gamma_{1}\curvearrowright X_{1} and Γ2X2\Gamma_{2}\curvearrowright X_{2} are stably orbit equivalent if there exist positive measure Borel subsets A1X1A_{1}\subseteq X_{1} and A2X2A_{2}\subseteq X_{2} and a measure-scaling isomorphism θ:A1A2\theta:A_{1}\to A_{2}444in other words θ\theta induces a measure space isomorphism between the probability spaces 1μ1(A1)A1\frac{1}{\mu_{1}(A_{1})}A_{1} and 1μ2(A2)A2\frac{1}{\mu_{2}(A_{2})}A_{2} such that for almost every xA1x\in A_{1}, one has

θ((Γ1x)A1)=(Γ2θ(x))A2.\theta((\Gamma_{1}\cdot x)\cap A_{1})=(\Gamma_{2}\cdot\theta(x))\cap A_{2}.

A free ergodic measure-preserving action of Γ\Gamma on a standard probability space XX is OE-superrigid if for every countable group Γ\Gamma^{\prime}, and every free ergodic measure-preserving action of Γ\Gamma^{\prime} on a standard probability space XX^{\prime}, if the Γ\Gamma-action on XX is stably orbit equivalent to the Γ\Gamma^{\prime}-action on XX^{\prime}, then the two actions are virtually conjugate (in particular Γ\Gamma and Γ\Gamma^{\prime} are virtually isomorphic).

The following theorem follows from our work in the exact same way as for mapping class groups of surfaces [Kid08b] (see also [Fur11b, Lemma 4.18]).

Theorem 4.3.

Let VV be a handlebody of genus at least 33. Then every free ergodic measure-preserving action of Mod(V)\mathrm{Mod}(V) on a standard probability space is OE-superrigid.

Rigidity of von Neumann algebras.

Let Γ\Gamma be a countable group, and let XX be a standard probability space equipped with a standard ergodic action of Γ\Gamma. Associated to the Γ\Gamma-action on XX is a von Neumann algebra L(X)ΓL^{\infty}(X)\rtimes\Gamma, obtained from the Murray–von Neumann construction [MvN36].

We refer the reader to the work of Ozawa and Popa [OP10, Definition 3.1] for the notion of a weakly compact group action. Let us only mention here that these include profinite actions, i.e. those obtained as inverse limits of actions on finite probability spaces (see [OP10, Proposition 3.2]). For example, this applies to the action of any residually finite countable group on its profinite completion, equipped with the Haar measure. As a subgroup of Mod(V)\mathrm{Mod}(\partial V), the handlebody group Mod(V)\mathrm{Mod}(V) is residually finite by a theorem of Grossman [Gro75].

A free ergodic measure-preserving action of a countable group Γ\Gamma on a standard probability space XX is WwcW^{\ast}_{wc}-superrigid if for every countable group Γ\Gamma^{\prime}, and every weakly compact free ergodic measure-preserving action of Γ\Gamma^{\prime} on a standard probability space XX^{\prime}, if the von Neumann algebras L(X)ΓL^{\infty}(X)\rtimes\Gamma and L(Y)ΓL^{\infty}(Y)\rtimes\Gamma^{\prime} are isomorphic, then the Γ\Gamma-action on XX is virtually conjugate to the Γ\Gamma^{\prime}-action on XX^{\prime}.

Theorem 4.4.

Let VV be a handlebody of genus at least 33. Then every free ergodic measure-preserving action of Mod(V)\mathrm{Mod}(V) on a standard probability space is WwcW^{*}_{wc}-superrigid.

Proof.

Let XX be a standard probability space equipped with a free ergodic measure-preserving action of Mod(V)\mathrm{Mod}(V), and let XX^{\prime} be a standard probability space equipped with a weakly compact free ergodic measure-preserving action of a countable group Γ\Gamma^{\prime}. Assume that there exists an isomorphism θ:L(X)Mod(V)L(X)Γ\theta:L^{\infty}(X)\rtimes\mathrm{Mod}(V)\to L^{\infty}(X^{\prime})\rtimes\Gamma^{\prime}. By [HHL20, Theorem 7], the group Mod(V)\mathrm{Mod}(V) is properly proximal in the sense of Boutonnet, Ioana and Peterson [BIP21]. It thus follows from [BIP21, Theorem 1.4] that up to unitary conjugacy, the isomorphism θ\theta sends L(X)L^{\infty}(X) to L(X)L^{\infty}(X^{\prime}). This implies that the actions ΓX\Gamma\curvearrowright X and ΓX\Gamma^{\prime}\curvearrowright X^{\prime} are orbit equivalent (see [Sin55]), so the conclusion follows from the orbit equivalence rigidity statement provided by Theorem 4.3. ∎

Remark 4.5.

Beyond the weakly compact case, the only kwown WW^{*}-superrigidity result for handlebody groups concerns their Bernoulli actions, that is, actions of the form Mod(V)X0Mod(V)\mathrm{Mod}(V)\curvearrowright X_{0}^{\mathrm{Mod}(V)}, where X0X_{0} is a standard probability space not reduced to a point, and the action is by shift. More precisely, when VV has genus at least 3, if a Bernoulli action Mod(V)X\mathrm{Mod}(V)\curvearrowright X and a free, ergodic, probability measure-preserving action of a countable group have isomorphic von Neumann algebras, then the actions are conjugate. This follows from [HH21b, Theorem A.2], based on work of Ioana, Popa and Vaes [IPV13, Theorem 10.1], applied by letting Γ0\Gamma_{0} be the cyclic subgroup generated by a Dehn twist about a nonseparating meridian α\alpha, letting Γ1\Gamma_{1} be the stabilizer of the isotopy class of α\alpha, and Γ=Mod(V)\Gamma=\mathrm{Mod}(V). Indeed, to check that [HH21b, Theorem A.2] applies, we only need to find an element gMod(V)g\in\mathrm{Mod}(V) such that gΓ1g1Γ1g\Gamma_{1}g^{-1}\cap\Gamma_{1} is infinite, and Γ1,g\langle\Gamma_{1},g\rangle generates Mod(V)\mathrm{Mod}(V). For this, let β,γ\beta,\gamma be nonseparating meridians such that α,β,γ\alpha,\beta,\gamma are pairwise disjoint, pairwise non-isotopic, and have connected complement. Let gMod(V)g\in\mathrm{Mod}(V) be an element sending α\alpha to β\beta and commuting with the twist TγT_{\gamma}. Then gΓ1g1Γ1g\Gamma_{1}g^{-1}\cap\Gamma_{1} is infinite because it contains TγT_{\gamma}. And Mod(V)\mathrm{Mod}(V) is generated by Γ1\Gamma_{1} and gg because the simplicial graph with vertices the isotopy classes of nonseparating meridians, and edges the nonseparating pairs, is connected (as easily follows from the connectivity of the disk graph) with quotient a single edge.

References

  • [AD13] C. Anantharaman-Delaroche. The Haagerup property for discrete measured groupoids. In Operator algebra and dynamics, volume 58 of Springer Proc. Math. Stat., pages 1–30, Heidelberg, 2013. Springer.
  • [Ada94] S. Adams. Indecomposability of equivalence relations generated by word hyperbolic groups. Topology, 33(4):785–798, 1994.
  • [BIP21] R. Boutonnet, A. Ioana, and J. Peterson. Properly proximal groups and their von Neumann algebras. Ann. Sci. Éc. Norm. Supér., 54(2):445–482, 2021.
  • [CK15] I. Chifan and Y. Kida. OEOE and WW^{*} rigidity results for actions by surface braid groups. Proc. Lond. Math. Soc. (3), 111(6):1431–1470, 2015.
  • [dlST19] M. de la Salle and R. Tessera. Characterizing a vertex-transitive graph by a large ball. J. Topol., 12(3):705–743, 2019.
  • [Dye59] H.A. Dye. On groups of measure preserving transformation. I. Amer. J. Math., 81:119–159, 1959.
  • [Dye63] H.A. Dye. On groups of measure preserving transformations. II. Amer. J. Math., 85:551–576, 1963.
  • [FM12] B. Farb and D. Margalit. A primer on mapping class groups, volume 49 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 2012.
  • [Fur99a] A. Furman. Gromov’s measure equivalence and rigidity of higher-rank lattices. Ann. Math., 150:1059–1081, 1999.
  • [Fur99b] A. Furman. Orbit equivalence rigidity. Ann. of Math. (2), 150(3):1083–1108, 1999.
  • [Fur11a] A. Furman. Mostow–Margulis rigidity with locally compact targets. Geom. Funct. Anal., 11(1):30–59, 2011.
  • [Fur11b] A. Furman. A survey of measured group theory. In Geometry, rigidity, and group actions, Chicago Lectures in Math., pages 296–374, Chicago, IL, 2011. Univ. Chicago Press.
  • [Gab02] D. Gaboriau. Invariants l2l^{2} de relations d’équivalence et de groupes. Publ. Math. Inst. Hautes Études Sci., 95:93–150, 2002.
  • [GH21] V. Guirardel and C. Horbez. Measure equivalence rigidity of Out(FN)\mathrm{{O}ut}({F}_{N}). arXiv:2103.03696, 2021.
  • [Gro93] M. Gromov. Asymptotic invariants of infinite groups. In Geometric group theory, volume 2 of London Math. Soc. Lecture Note Ser., pages 1–295, Cambridge Univ. Press, Cambridge, 1993.
  • [Gro75] E.K. Grossman. On the residual finiteness of certain mapping class groups. J. London Math. Soc. (2), 9:160–164, 1974/75.
  • [Ham09] U. Hamenstädt. Geometry of the mapping class groups I: Boundary amenability. Invent. Math., 175(3):545–609, 2009.
  • [Hen] S. Hensel. A primer on handlebody groups. preprint, available at http://www.mathematik.uni-muenchen.de/ hensel/papers/hno4.pdf.
  • [Hen18] S. Hensel. Rigidity and flexibility for handlebody groups. Comment. Math. Helv., 93(2):335–358, 2018.
  • [Hen21] S. Hensel. (Un)distorted stabilisers in the handlebody group. J. Topol., 14(2):460–487, 2021.
  • [HH12] U. Hamenstädt and S. Hensel. The geometry of the handlebody groups I: Distortion. J. Topol. Anal., 4(1):71–97, 2012.
  • [HH20a] C. Horbez and J. Huang. Boundary amenability and measure equivalence rigidity among two-dimensional Artin groups of hyperbolic type. arXiv:2004.09325, 2020.
  • [HH20b] C. Horbez and J. Huang. Measure equivalence classification of transvection-free right-angled Artin groups. arXiv:2010.03613, 2020.
  • [HH21a] U. Hamenstädt and S. Hensel. The Geometry of the Handlebody Groups II: Dehn functions. Michigan Math. J., 70(1):23–53, 2021.
  • [HH21b] C. Horbez and J. Huang. Orbit equivalence rigidity of irreducible actions of right-angled Artin groups. arXiv:2110.04141, 2021.
  • [HHL20] C. Horbez, J. Huang, and J. Lécureux. Proper proximality in non-positive curvature. arXiv:2005.08756, 2020.
  • [IPV13] A. Ioana, S. Popa, and S. Vaes. A class of superrigid group von Neumann algebras. Ann. of Math. (2), 178(1):231–286, 2013.
  • [Iva92] N. V. Ivanov. Subgroups of Teichmüller modular groups, volume 115 of Translations of Mathematical Monographs. American Mathematical Society, Providence, RI, 1992. Translated from the Russian by E. J. F. Primrose and revised by the author.
  • [Joh95] K. Johannson. Topology and combinatorics of 3-manifolds, volume 1599 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1995.
  • [Kec95] A.S. Kechris. Classical descriptive set theory, volume 156 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1995.
  • [Kid08a] Y. Kida. The mapping class group from the viewpoint of measure equivalence theory. Mem. Amer. Math. Soc., 196(916), 2008.
  • [Kid08b] Y. Kida. Orbit equivalence rigidity for ergodic actions of the mapping class group. Geom. Dedicata, 131:99–109, 2008.
  • [Kid09] Y. Kida. Introduction to measurable rigidity of mapping class groups. In Handbook of Teichmüller theory, volume II, volume 13 of IRMA Lect. Math. Theor. Phys., pages 297–367. Eur. Math. Soc., 2009.
  • [Kid10] Y. Kida. Measure equivalence rigidity of the mapping class group. Ann. Math., 171(3):1851–1901, 2010.
  • [Kid11] Y. Kida. Rigidity of amalgamated free products in measure equivalence. J. Topol., 4(3):687–735, 2011.
  • [Kla18] E. Klarreich. The boundary at infinity of the curve complex and the relative teichmüller space. arXiv:1803.10339, 2018.
  • [Kob12] T. Koberda. Right-angled Artin groups and a generalized isomorphism problem for finitely generated subgroups of mapping class groups. Geom. Funct. Anal., 22(6):1541–1590, 2012.
  • [KS09] M. Korkmaz and S. Schleimer. Automorphisms of the disk complex. arXiv:0910.2038, 2009.
  • [Man13] J. Mangahas. A recipe for short-word pseudo-Anosovs. Amer. J. Math., 135(4):1087–1116, 2013.
  • [McC85a] J. McCarthy. A “Tits-alternative” for subgroups of surface mapping class groups. Trans. Amer. Math. Soc., 291(2):583–612, 1985.
  • [McC85b] D. McCullough. Twist groups of compact 3-manifolds. Topology, 24:461–474, 1985.
  • [McC06] D. McCullough. Homeomorphisms which are Dehn twists on the boundary. Algebr. Geom. Topol., 6:1331–1340, 2006.
  • [MS06] N. Monod and Y. Shalom. Orbit equivalence rigidity and bounded cohomology. Ann. of Math. (2), 164(3):825–878, 2006.
  • [MvN36] F.J. Murray and J. von Neumann. On rings of operators. Ann. of Math. (2), 37(1):116–229, 1936.
  • [Oer02] U. Oertel. Automorphisms of three-dimensional handlebodies. Topology, 41(2):363–410, 2002.
  • [OP10] N. Ozawa and S. Popa. On a class of II1{\rm II}_{1} factors with at most one Cartan subalgebra. Amer. J. Math., 132(3):841–866, 2010.
  • [OW80] D.S. Ornstein and B. Weiss. Ergodic theory of amenable group actions. I. The Rohlin lemma. Bull. Amer. Math. Soc. (N.S.), 2(1):161–164, 1980.
  • [Sin55] I.M. Singer. Automorphisms of finite factors. Amer. J. Math., 77:117–133, 1955.
  • [Suz77] S. Suzuki. On homeomorphisms of a 33-dimensional handlebody. Canadian J. Math., 29(1):111–124, 1977.
  • [Waj98] B. Wajnryb. Mapping class group of a handlebody. Fund. Math., 158(3):195–228, 1998.
  • [Zim80] R.J. Zimmer. Strong rigidity for ergodic actions of semisimple groups. Ann. Math., 112(3):511–529, 1980.
  • [Zim91] R.J. Zimmer. Groups generating transversals to semisimple Lie group actions. Israel J. Math., 73(2):151–159, 1991.

Sebastian Hensel

Mathematisches Institut der Universität München

D-80333 München

e-mail: [email protected]

Camille Horbez

Université Paris-Saclay, CNRS, Laboratoire de mathématiques d’Orsay, 91405, Orsay, France

e-mail:[email protected]