Measure equivalence rigidity of the handlebody groups
Abstract
Let be a connected -dimensional handlebody of finite genus at least . We prove that the handlebody group is superrigid for measure equivalence, i.e. every countable group which is measure equivalent to is in fact virtually isomorphic to . Applications include a rigidity theorem for lattice embeddings of , an orbit equivalence rigidity theorem for free ergodic measure-preserving actions of on standard probability spaces, and a -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 and are measure equivalent if there exists a standard infinite measure space equipped with an action of by measure-preserving Borel automorphisms, such that for every , the action of on 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 are always measure equivalent, by considering the left and right multiplications on 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 of finite-type surfaces are ME-superrigid, i.e. every countable group that is measure equivalent to , is in fact commensurable to up to a finite kernel [Kid10]. This led to further strong rigidity results, for certain amalgamated free products [Kid11], certain subgroups of 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 , the outer automorphism group of a finitely generated free group of rank , is also ME-superrigid [GH21].
In the present paper, we establish a superrigidity theorem for handlebody groups, defined as mapping class groups of connected -dimensional handlebodies , i.e. is a disk-sum of finitely many copies of . These groups are of particular importance in -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 is a closed orientable surface of finite genus , and embeds as a (highly distorted [HH12]) subgroup of ; it also surjects onto 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 be a connected -dimensional handlebody of finite genus at least . Then 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 be a connected -dimensional handlebody of finite genus at least . Let be a locally compact second countable group, equipped with its Haar measure. Let be a finite index subgroup of , and let be an injective homomorphism whose image is a lattice.
Then there exists a homomorphism with compact kernel such that for every , one has .
If is a finite generating set of , then naturally embeds as a lattice in the automorphism group of the Cayley graph , defined as the simplicial graph whose vertices are the elements of , with an edge between two distinct vertices whenever (this convention excludes for instance loop-edges when contains the identity of , or multiple edges if 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 be a connected -dimensional handlebody of finite genus at least , and let be a finite generating set of . Then every graph automorphism of is at bounded distance from the left multiplication by an element of .
If is a torsion-free finite-index subgroup of , and if is a finite generating set of , then the automorphism group of is countable, and in fact embeds as a subgroup of containing .
The torsion-freeness assumption is crucial in the second part of the statement: for every finitely generated group containing a nontrivial torsion element, there exists a finite generating set of such that the automorphism group of 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 be a connected -dimensional handlebody of finite genus at least . Let be a countable group. Let and be two free ergodic measure-preserving group actions by Borel automorphisms on standard probability spaces.
If the actions and 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 be a connected -dimensional handlebody of finite genus at least . Let be a countable group. Let and be two free ergodic measure-preserving group actions by Borel automorphisms on standard probability spaces, and assume that is weakly compact.
If the von Neumann algebras and are isomorphic, then the actions and 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 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 (see Section 1.2 for its precise definition). More precisely, we are given a measured groupoid , which comes from restricting two actions of on standard finite measure spaces to a positive measure Borel subset on which their orbits coincide. The groupoid is thus equipped with two cocycles , given by the two actions: whenever two points are joined by an arrow , there is an element sending to for the first action, and an element sending to for the second action. Our goal is to build a canonical map such that and are cohomologous through : this means that whenever are joined by an arrow , then . In fact, using a theorem of Korkmaz and Schleimer which identifies to the automorphism group of the disk graph of , our goal is to build a (canonical) map . Recall that the disk graph is the graph whose vertices are the isotopy classes of meridians in (i.e. essential simple closed curves that bound a properly embedded disk in ), 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 , the main step is to characterize subgroupoids of that arise as stabilizers of Borel maps in a purely groupoid-theoretic way, i.e. with no reference to the cocycles (so that a vertex stabilizer for is also a vertex stabilizer for ).
In the surface mapping class group setting (where the disk graph is replaced by the curve graph of the surface ), the important observation made by Kida is the following: curve stabilizers inside are characterized as maximal nonamenable subgroups of 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 of which contains an infinite amenable normal subgroup . A first bad situation we encounter is the following: could be generated by a partial pseudo-Anosov, supported on a subsurface , and be its normalizer. In the surface setting considered by Kida [Kid10], such an is not maximal, as it is contained in the stabilizer of the boundary multicurve of . But for us, the group of multitwists about could intersect trivially; in this case may not contain any infinite normal amenable subgroup, so may not violate the maximality of . We resolve this first difficulty by further imposing that 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 is given by the following situation: there is a multicurve , together with a (possibly empty) collection of complementary components of labeled active, is the stabilizer of , and is exactly the active subgroup of , i.e. the subgroup of the stabilizer of acting trivially on all inactive subsurfaces, and it is amenable. This still includes several possibilities: could be a nonseparating meridian and (in which case is the twist subgroup). But (still with ), the multicurve could also be of the form , where and together bound an annulus in (see Figure 1): the cyclic subgroup generated by the product of twists 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 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 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 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 .
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 , we mean a connected orientable -manifold which is a disk-sum of copies of , where is a closed disk and is a circle. The boundary of a handlebody of genus is a closed, connected, orientable surface of the same genus . The handlebody group is the mapping class group of , i.e. the group of all isotopy classes of orientation-preserving homeomorphisms of . There is a restriction homomorphism , which is injective, thus allowing us to view as a subgroup of (see e.g. [Hen, Lemma 3.1]).
Meridians and annuli.
Let be a handlebody. Recall that a simple closed curve on is essential if it is homotopically nontrivial, i.e. it does not bound a disk on . An essential simple closed curve on is a meridian (represented in blue in Figure 1) if it bounds a properly embedded disk in .

If is a meridian, then the Dehn twist associated to belongs to , viewed as a subgroup of – 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 of disjoint nonisotopic essential simple closed curves on is an annulus pair (represented in red in Figure 1) if neither nor is a meridian, and there exists a properly embedded annulus such that . An annulus twist is a mapping class of the form for some annulus pair . Annulus twists belong to ([McC06, Theorem 1] or [Oer02, Theorem 1.11]).
Lemma 1.1.
Let be a meridian. Then every connected component of supports two handlebody group elements which both restrict to a pseudo-Anosov mapping class of and together generate a nonabelian free subgroup.
Proof.
Let be a connected component of which is not a once-holed torus, and denote by a boundary component of (corresponding to one of the sides of ). Then, for any essential simple closed curve which is not boundary parallel in , we can (and shall) choose an essential simple closed curve which is not isotopic to and bounds a pair of pants on together with (here, we are using that is not a once-holed torus). Since is a meridian, are either both meridians, or form an annulus pair. Thus, in either case, the multitwist is a handlebody group element supported in .
By choosing curves which fill we can thus find so that no essential simple closed curve in is fixed by both up to isotopy. This implies that the group generated by contains a pseudo-Anosov ([Iva92], see also the discussion in [Man13, Section 2.4]). Conjugating by yields a second one, and sufficiently high powers of and generate a nonabelian free subgroup. ∎
Lemma 1.1 implies in particular that when has genus at least , the stabilizer in of every meridian in contains a nonabelian free subgroup (because at least one connected component of is not a one-holed torus). The requirement of having genus at least is necessary, as the following shows.
Lemma 1.2.
Suppose that is a separating meridian, and suppose that is a component of which is a once-holed torus. Then contains a unique (nonseparating) meridian which is not peripheral in up to isotopy, and therefore
If the genus of is at least , then is the only other meridian whose stabiliser contains (or even a finite-index subgroup of ).
Proof.
The subsurface is the boundary of a once-spotted genus handlebody . Hence, there is a nonseparating meridian contained in . 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 . This inclusion is strict, because there exists a handlebody group element which fixes and restricts to a pseudo-Anosov homeomorphism on the complementary subsurface, in particular does not fix the isotopy class of .
To show the final claim, recall from Lemma 1.1 that there are elements in restricting to pseudo-Anosov elements on any component of which is not a once-holed torus. If the genus of is at least , the complement of will be such a component. Hence, is the unique other meridian fixed by (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 is rotationless (or pure) if the following holds: if a power of fixes the isotopy class of a simple closed curve , then actually fixes the oriented isotopy class of .
Let be a surface obtained from a closed, connected, orientable surface by removing at most finitely many points. We denote by the (finite index) subgroup of consisting of mapping classes acting trivially on homology mod of the surface. It is well-known [Iva92] that every is rotationless.
We observe that if is rotationless, and is a subsurface which is preserved by , then the restriction is rotationless – however, if then need not be contained in .
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 the mod--homology cover of the surface . Let be the subgroup of those mapping classes which admit a lift to which acts trivially on .
Then is a finite index subgroup of , and if is any element preserving a connected subsurface , then the restriction is an element of .
The proof uses the following covering argument.
Lemma 1.4.
Let be an essential connected subsurface. Denote by the mod--homology cover, and let be a connected component of . Then the map
induced by the inclusion is injective, and the same is true with replaced with for any .
Proof.
Choose a subsurface with one boundary component, so that is a bordered sphere. Denote by the boundary components of (which are all contained in ). We then have that
where the latter summand is . The first summand injects into , since is a subsurface of with one boundary component.
We now aim to show that for all there is a curve which is disjoint from , intersects each in a single point, and is disjoint from all other (i.e. that the complement of is connected). This will show that are linearly independent from each other and from in thus showing the lemma.
For simplicity of notation, we will perform the construction only for . Choose a basepoint in , and let be its image in . Since the mod- homology cover is normal (Galois), the preimage is exactly the orbit of under the deck group .
To describe the intersection , first observe that since is connected, a point is contained in exactly if there is a path connecting to contained in . Such paths are exactly the lifts of loops based at which are contained in . So is contained in if and only if the deck group element mapping to is the image of some . The image of in the deck group is exactly the subgroup . Together this shows that .
Similarly, the components of can be identified with the cosets of the subgroup .
To describe the cover more precisely, we choose curves based at in the following way:
-
(1)
The homology classes form a basis of ,
-
(2)
is a basis of , and the curves are contained in .
-
(3)
The curves for intersect in exactly two points.
To see that these curves exist, we argue as follows. Denote by the components of . Choose a curve . The connectivity of implies that for every , the curve is homologically nontrivial (in ) exactly if has more than one component. For each boundary curve we can find a loop based at which intersects in two points, one on and one on . We can thus choose independent homology classes defined by curves intersecting in at most two points, so that for any there is a linear combination of the , so that has algebraic intersection number with all curves in . Any such class can be realised by a multicurve disjoint from . Since every homology class defined by a curve (without specified basepoint) in can be realised by a loop based at which intersects in two points, and every curve in can be realised by a loop disjoint from the desired existence follows.
Lifting a curve of the type in (2) at a point stays in the same connected component , while lifting a curve of the type in (3) joins to and intersects in a single point. To see that last claim observe that a lift of a curve as in (3) cannot join two points of , as the image of that curve in would then be contained in , contradicting (1) and (2).
For every , denote by the component of adjacent to . There are , so that are surfaces adjacent to . Namely, either has genus (which is automatically the case if is separating), and contains a curve defining one of the (of the second type), or is a punctured sphere so that for the boundary component there is some (of the third type) which intersects it once (namely, if all would interesect in an even number of points, the could not be a basis of , since is nonseparating). In both cases the desired component is . Choose paths joining to which intersect only among the .
Since is a subgroup of generated by a subset of the generators, there is a path in the Cayley graph of from to which is disjoint from the Cayley graph of the subgroup . Each edge in such a path corresponds to a right multiplication , and we can choose a corresponding path joining to which is disjoint from . By concatenating these paths with (in the right order) we then find the desired path . ∎
Lemma 1.3 is now an immediate consequence of the following corollary.
Corollary 1.5.
Suppose that is a mapping class so that
-
1.
admits a lift to the mod--homology-cover , which acts trivially on
-
2.
preserves a subsurface
Then the restriction acts trivially on .
Proof.
Let be a simple closed curve which is part of a basis for . Then there is a power so that lifts to a curve (with notation as in the previous lemma).
Denote by the restriction of the covering map (which is then also a covering). We have . Since is invertible mod , there is a multiple so that mod .
By Lemma 1.4, is a subspace of . Since acts trivially on , this implies that the restriction acts trivially on . Hence, we have . Since is a lift of this implies . ∎
In the sequel of the paper, we will always let , where is as in Lemma 1.3.
1.3 Infinite conjugacy classes
A countable group is said to be ICC (standing for infinite conjugacy classes) if the conjugacy class of every nontrivial element of is infinite.
Lemma 1.6.
Let be a handlebody of genus at least , and let be a handlebody group element. Then either the conjugacy class of is infinite, or fixes the isotopy class of every meridian.
In particular, when the genus of is at least , the group is ICC.
We remark that in genus , the hyperelliptic involution fixes the isotopy class of every essential simple closed curve on , and its conjugacy class is finite in .
Proof.
Suppose that is an element with finite conjugacy class. For any meridian , consider the elements for . By finiteness of the conjugacy class, two of these have to be equal, and thus there is some so that
or equivalently,
This implies is isotopic to , see e.g. [FM12, Section 3.3]. The first part of the lemma follows since was arbitrary. The fact that is ICC when the genus is at least 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 (whose elements are thought of as being arrows) equipped with two Borel maps towards a standard Borel space (giving the source and range of an arrow), and coming with a measurable composition law and inverse map and with a unit element per . The Borel space is called the base space of the groupoid . All Borel groupoids considered in the present paper are assumed to be discrete, i.e. there are countably many arrows in 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 can always be written as a countable disjoint union of bisections, i.e. Borel subsets of on which and are injective (in which case and are Borel subsets of , see [Kec95, Corollary 15.2]). A Borel groupoid with base space is trivial if .
A finite Borel measure on is quasi-invariant for the groupoid if for every bisection , one has if and only if . A measured groupoid is a Borel groupoid together with a quasi-invariant finite Borel measure on its base space .
An important example of a measured groupoid to keep in mind is the following: when a countable group acts on a standard probability space by Borel automorphisms in a quasi-measure-preserving way, then has a natural structure of a measured groupoid over , denoted by : the source and range maps are given by and , the composition law is , the inverse of is and the units are .
A Borel subset which is stable under composition and inverse and contains all unit elements of has the structure of a measured subgroupoid of over the same base space . Given a Borel subset , the restriction is the measured groupoid over defined by only keeping the arrows whose source and range both belong to . Given two subgroupoids , we denote by the subgroupoid of generated by and , i.e. the smallest subgroupoid of containing and (it is made of all arrows obtained as finite compositions of arrows in and arrows in ).
A measured groupoid with base space is of infinite type if for every positive measure Borel subset and almost every , there are infinitely many elements of with source . Observe that if is of infinite type, then for every Borel subset of positive measure, the restricted groupoid is again of infinite type.
Let be a measured groupoid over a standard probability space , and let be a countable group. A strict cocycle is a Borel map such that for all , if the source of is equal to the range of (so that the product is well-defined), then . The kernel of a cocycle is the subgroupoid of made of all such that . We say that has trivial kernel if its kernel is equal to the trivial subgroupoid of , i.e. it only consists of the unit elements of . We say that a strict cocycle is action-type if has trivial kernel, and whenever is an infinite subgroup, and is a Borel subset of positive measure, then is a subgroupoid of of infinite type. An important example is that given a measure-preserving -action on a standard probability space , the natural cocycle is action-type [Kid09, Proposition 2.26]. We warn the reader that in the latter example, it is important that the -action on preserves the measure, as opposed to only quasi-preserving it.
Given a Polish space equipped with a -action by Borel automorphisms, we say that a measurable map is -equivariant if there exists a conull Borel subset such that for every , one has . We say that an element is -invariant if the constant map with value is -equivariant (equivalently, there exists a conull Borel subset such that ). The -stabilizer of is the subgroupoid of made of all elements such that . A measurable map is stably -equivariant if one can partition into at most countably many Borel subsets such that for every , the map is -equivariant.
Given two measured subgroupoids , we say that is stably contained in if there exist a conull Borel subset and a partition into at most countably many Borel subsets such that for every , one has . We say that and are stably equal if there exist a conull Borel subset and a partition as above such that for every , one has . We say that is stably trivial if it is stably equal to the trivial subgroupoid of .
Let be a measured subgroupoid of , and be a bisection. We say that is -invariant if there exists a conull Borel subset such that for every and every such that the composition is well-defined, we have if and only if . Let now be another measured subgroupoid of . The groupoid is normalized by if can be covered by countably many bisections in such a way that is -invariant for every . The subgroupoid is stably normalized by if one can partition into at most countably many Borel subsets in such a way that for every , the groupoid is normalized by . When , we will simply say that is stably normal in .
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 is amenable and comes with a cocycle towards a countable group , and if acts by homeomorphisms on a compact metrizable space , then there exists a -equivariant Borel map , see [Kid09, Proposition 4.14]. Here denotes the set of Borel probability measures on , equipped with the weak- topology coming from the duality with the space of real-valued continuous functions on given by the Riesz–Markov–Kakutani theorem. Second, whenever is a cocycle with trivial kernel, and is an amenable subgroup of , then is an amenable subgroupoid of (see e.g. [GH21, Corollary 3.39]). Amenability is stable under subgroupoids and restrictions. Furthermore, if there exists a conull Borel subset and a partition into at most countably many Borel subsets such that for every , the groupoid is amenable, then 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 over a standard probability space is everywhere nonamenable if for every Borel subset of positive measure, the groupoid 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 be a handlebody of genus at least . Then is ME-superrigid.
Using the fact that 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 be a handlebody of genus at least . Let be a measured groupoid over a standard probability space (with source map and range map ), and let be two strict action-type cocycles.
Then there exist a Borel map and a conull Borel subset such that for all , one has .
Remark 3.3.
The case where is reduced to a point is actually already relevant: if is an automorphism of , then the group , viewed as a groupoid over a point, comes equipped with two (action-type) cocycles towards , given by the identity and . The conclusion in this case is that every automorphism of is a conjugation. More generally, a consequence of Theorem 3.2 is that the natural map from to its abstract commensurator is an isomorphism (using that 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 with two action-type cocycles towards , we ultimately aim to show that subgroupoids of corresponding to meridian stabilizers for - in the precise sense that they are of meridian type as in Definition 3.4 below - are also of meridian type with respect to . 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 from the base space of the groupoid 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 to conclude. We make the following definition.
Definition 3.4 (Subgroupoids of meridian type).
Let be a measured groupoid over a standard probability space , and let be a strict cocycle. A measured subgroupoid of is of meridian type with respect to if there exists a conull Borel subset and a partition into at most countably many Borel subsets such that for every , the groupoid is equal to the -stabilizer of the isotopy class of a meridian .
When can be written as in Definition 3.4, we say that the map sending every to the isotopy class of the meridian is a meridian map for . The essential uniqueness of this map (i.e. the fact that, up to measure , it does not depend on the choice of a partition and meridians 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 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 be a measured groupoid over a standard probability space . A strongly Schottky pair of subgroupoids of is a pair of amenable subgroupoids of of infinite type such that for every Borel subset of positive measure, there exists a Borel subset of positive measure such that every normal amenable subgroupoid of is stably trivial.
We observe that this notion is stable under restrictions: if is a strongly Schottky pair of subgroupoids of , then for every Borel subset of positive measure, the pair is a strongly Schottky pair of subgroupoids of . In addition, this notion is stable under stabilization: given a pair of subgroupoids of , and a partition into at most countably many Borel subsets, if is a strongly Schottky pair of subgroupoids of for every , then is a strongly Schottky pair of subgroupoids of .
Notice that the last conclusion implies in particular that is nonamenable. So the existence of a strongly Schottky pair of subgroupoids of forces 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 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 be a countable group, and let be two elements that generate a nonabelian free subgroup of . Let be a measured groupoid over a standard probability space , equipped with a strict action-type cocycle .
Then is a strongly Schottky pair of subgroupoids of (in particular is everywhere nonamenable).
In the following proof, whenever is a Polish space, the set of all Borel probability measures on is equipped with the topology generated by the maps , where varies over the set of all real-valued bounded continuous functions. When is compact, this is nothing but the weak- topology coming from the duality given by the Riesz–Markov–Kakutani theorem. When 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 which justify the measurability of all maps in the following proof.
Proof of Lemma 3.6.
As and are amenable subgroups of and has trivial kernel, the subgroupoids and are amenable. As and are infinite and is action-type, the subgroupoids and are of infinite type.
Now it is enough to prove that if is a Borel subset of positive measure, and is a normal amenable subgroupoid of , then is stably trivial.
Let be the Cayley tree of the free group , with respect to the generating set . The -action on by isometries extends to an -action on by homeomorphisms. As is amenable and is compact and metrizable, there exists an -equivariant Borel map , see [Kid09, Proposition 4.14].
We claim that we can find a (possibly null) Borel subset of maximal measure such that there exists a Borel map which is stably -equivariant and such that for every , the support of the measure has cardinality at least . Indeed, let be the supremum of the measures of all Borel subsets of having the above property, and let be a measure-maximizing sequence of such sets – in particular, for every , we have a Borel map as above. Then the countable union of all subsets has measure at least , and we claim that it also satisfies the above property. To see this, inductively define and . Then the map , defined to coincide with when restricted to , is stably -equivariant, and for every , the support of the measure has cardinality at least . This concludes the proof of our claim.
From now on, we choose a Borel subset as in the above claim. We first prove that is stably trivial. Up to partitioning into at most countably many Borel subsets, we can assume that the map is -equivariant (and not just stably equivariant). For every , the probability measure on gives positive measure to the -invariant subset made of pairwise distinct triples. Thus, after restricting this measure to and renormalizing to turn this restricted measure into a probability measure, we get an -equivariant Borel map . Now, denoting by the vertex set of , there is a natural -equivariant barycenter map . By pushing the probability measures through this map, we get an -equivariant Borel map . Let be the set of all nonempty finite subsets of . As is countable, there is also a natural -equivariant Borel map , sending a probability measure to the finite subset of made of all vertices that have maximal -measure. We thus derive an -equivariant Borel map . As is countable, we can then find a Borel partition into at most countably many Borel subsets such that for every , the map is constant, with value a nonempty finite set of vertices of . In other words, there exists a conull Borel subset such that is contained in the -stabilizer of . As this stabilizer is trivial and has trivial kernel, it follows that is stably trivial.
We will now prove that is a null set, which will conclude the proof of the lemma. So assume towards a contradiction that has positive measure. We know that there exists an -equivariant Borel map , and that for every such map and almost every , the support of has cardinality at most . Let be the set of all nonempty subsets of of cardinality at most . As in [Ada94, Lemma 3.2], we can thus find an -equivariant Borel map which is maximal in the sense that for every other -equivariant Borel map and a.e. , one has . Being canonical, the map is then equivariant under the groupoid which normalizes . Recall that the groupoid is amenable and of infinite type. Therefore, repeating the argument from the present proof shows that there exists a maximal -equivariant Borel map , and this must then be the constant map with value . Likewise, the constant map with value is the maximal -equivariant Borel map . As , 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 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 be a (possibly disconnected) orientable surface of finite type, i.e. is obtained from the disjoint union of finitely many closed connected orientable surfaces by removing at most finitely many points. We define as the group of all isotopy classes of orientation-preserving diffeomorphisms of that do not permute the connected components of , and act trivially on the homology mod of each connected component; in other words , where are the connected components of .
Definition 3.7 (Irreducibility).
Let be a measured groupoid over a standard probability space , equipped with a strict cocycle .
We say that is reducible if there exist a Borel subset of positive measure and an essential simple closed curve on such that the isotopy class of is -invariant.
Otherwise, we say that is irreducible.
Definition 3.8 (Canonical reduction set).
Let be a measured groupoid over a standard probability space , equipped with a strict cocycle . A (possibly infinite) set of isotopy classes of essential simple closed curves on is a canonical reduction set for if
-
1.
every is -invariant, and
-
2.
for every Borel subset of positive measure, every isotopy class of essential simple closed curves which is -invariant belongs to .
Note that is irreducible if and only if is a canonical reduction set for . Notice also that if a canonical reduction set for 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 be a measured groupoid over a standard probability space , equipped with a strict cocycle .
Then there exist a partition into at most countably many Borel subsets such that for every , has a canonical reduction set.
Proof.
Let be a Borel subset of maximal measure such that there exists a partition into at most countably many Borel subsets such that for every , the set of all isotopy classes of essential simple closed curves that are -invariant is nonempty. We emphasise that is possibly infinite, and may contain non-disjoint curves. Notice that such a Borel subset exists, because if is a measure-maximizing sequence of such sets, then their countable union also satisfies the same property. Let now . The maximality of the measure of ensures that is irreducible.
For every , we then let be the elementwise stabilizer of in : this is a proper subgroup of (because ). Repeating the above argument, for every , there exists a Borel partition such that
-
1.
for every Borel subset of positive measure, every -invariant isotopy class of essential simple closed curve belongs to ,
-
2.
there exists a partition into at most countably many Borel subsets such that for every , the set of all isotopy classes of essential simple closed curves that are -invariant contains properly.
For every , we then let be the elementwise stabilizer of in . We observe that is a proper subgroup of . Indeed, there exists a conull Borel subset such that . If , then every curve in is -invariant, and therefore -invariant, contradicting the definition of . This contradiction shows that .
We now repeat the above procedure inductively. Since there is a bound, only depending on the topology of , on a chain (for inclusion) of collections of curves on with pairwise distinct elementwise stabilizers in , we attain a partition of 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 be a measured groupoid over a standard probability space , equipped with a strict cocycle , and let be a measured subgroupoid of . Assume that has a canonical reduction set .
Then for every measured subgroupoid of that normalizes , the set is -invariant. In other words, denoting by the global stabilizer of in , there exists a conull Borel subset such that .
Proof.
Since normalizes , there exists a covering of by countably many bisections that all leave invariant. Up to subdividing the bisections , we will assume that for every , the -image of is a single element . For every , we let and be the source and range of .
Every isotopy class is -invariant, so is -invariant. If has positive measure, the maximality condition in the definition of a canonical reduction set ensures that . By reversing the arrows in the bisection , we also derive that if . Let be a conull Borel subset which avoids each of the countably many subsets and of zero measure. Then . This concludes our proof. ∎
Given a (possibly infinite) set of isotopy classes of essential simple closed curves on , there is up to isotopy a unique essential subsurface such that every curve in is isotopic into , and if is another subsurface with this property, then up to isotopy . We call the subsurface of filled by . The multicurve , obtained from by only keeping one curve in each isotopy class, is called the boundary multicurve of .
Corollary 3.11.
Let be a measured groupoid over a standard probability space , equipped with a strict cocycle . Let be measured subgroupoids. Assume that is stably normalized by , and that for every Borel subset of positive measure, one has .
If is reducible, then so is .
Proof.
Since is reducible, we can find a Borel subset of positive measure such that has a nonempty canonical reduction set . As , the set does not fill , so the boundary multicurve of is nonempty. Up to restricting to a Borel subset of of positive measure, we can assume that is normalized by . Then Lemma 3.10 ensures that is -invariant. In particular is -invariant, showing that is reducible. ∎
When is the canonical reduction set for , the boundary multicurve of will be called the canonical reduction multicurve of . A connected component of is then called active for if it contains an essential simple closed curve whose isotopy class does not belong to , and inactive for otherwise (because in the latter case, every element in the elementwise stabilizer of acts trivially on ).
We give a few examples of active and inactive subsurfaces in the case that the essential image of is a cyclic subgroup generated by .
-
i)
If is a partial pseudo-Anosov supported on a connected subsurface , possibly composed with Dehn twists about curves contained in , then the canonical reduction multicurve is , and is the only active complementary component.
-
ii)
If is a Dehn twist about a curve , then the canonical reduction multicurve is , 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 or to subsurfaces of . We include a proof to explain how to deal with disconnected subsurfaces.
Lemma 3.12 (Kida [Kid08a]).
Let be a (possibly disconnected) surface of finite type, so that every connected component has negative Euler characteristic. Let be a measured groupoid, equipped with a strict cocycle . Let be a measured subgroupoid of such that has trivial kernel.
If stably normalizes an amenable subgroupoid of , with irreducible, then is amenable.
Proof.
Up to a countable Borel partition of the base space of (which does not affect the conclusion), we will assume that normalizes .
Let be the connected components of . Then decomposes as . For , let be the cocycle obtained by post-composing with the projection.
Let . Then acts on the compact metrizable space of projective measured laminations on . As is amenable, there exists an -equivariant Borel map . The space has a -invariant Borel partition into the subspace made of arational laminations, and the subspace made of non-arational laminations.
Let us first assume towards a contradiction that there exists a Borel subset of positive measure such that for all , the measure gives positive measure to . After restricting to and renormalizing it to get a probability measure, we obtain an -equivariant Borel map . Let be the countable set of all nonempty finite sets of isotopy classes of essential simple closed curves on . There is a -equivariant map , 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 -equivariant Borel map . As is countable, there is also a -equivariant map , sending a probability measure to the union of all finite sets with maximal -measure. In summary, we have found an -equivariant Borel map . Let be a Borel subset of positive measure where this map is constant, with value a finite set . As we are working in the finite-index subgroup , every curve in is -invariant, contradicting the irreducibility of .
Therefore determines an -equivariant Borel map . Klarreich’s description [Kla18] of the boundary of the curve graph of yields a continuous -equivariant map , so we get an -equivariant Borel map . Denoting by the space of pairwise distinct triples, Kida proved in [Kid08a, Section 4.1] the existence of a -equivariant Borel map . Using again the irreducibility of , together with an Adams-type argument as in the proof of Lemma 3.6, we deduce that there exists a Borel map which is both -equivariant and -equivariant.
Combining all these maps as varies in yields an -equivariant Borel map
For every , the action of on is Borel amenable [Kid08a, Ham09], and therefore so is the action of on (see e.g. [HH20b, Section 3.4.1] for the relevant background). As has trivial kernel, it then follows from [GH21, Proposition 3.38] (originially due to Kida [Kid08a, Proposition 4.33]) that is amenable. ∎
3.4 Uniqueness statements
Lemma 3.13.
Let be a measured groupoid over a standard probability space , equipped with a strict action-type cocycle . Let be a measured subgroupoid of .
Let be two meridians, with nonseparating. Assume that there exists a Borel subset of positive measure such that is equal to the -stabilizer of the isotopy class of , and the isotopy class of is -invariant.
Then (up to isotopy).
Proof.
The stabilizer of in contains an element which restricts to a pseudo-Anosov element on (Lemma 1.1). The groupoid is contained in , and it is of infinite type since is action-type. Therefore is fixed by some positive power of , which implies that 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 of a separating meridian may fix other nonseparating meridians, is the unique separating meridian it fixes by Lemma 1.2).
Lemma 3.14.
Let be a measured groupoid over a standard probability space , equipped with a strict action-type cocycle . Let be a measured subgroupoid of .
Let be two separating meridians. Assume that there exists a Borel subset of positive measure such that is equal to the -stabilizer of the isotopy class of , and the isotopy class of is -invariant.
Then (up to isotopy). ∎
3.5 Property and subgroupoids of non-separating meridian type
We make the following definition.
Definition 3.15 (Product-like subgroupoid).
A measured groupoid is product-like if there exist two subgroupoids which are both stably normal in , such that for every , the groupoid contains a strongly Schottky pair of subgroupoids , with and both stably normalized by .
Notice that this notion is stable under restrictions and stabilization. In the terminology from [GH21, Definition 13.5], the subgroupoids and 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 and are stably normal in an ambient groupoid .
We now introduce the following properties, which will be useful in order to detect subgroupoids of nonseparating-meridian type.
Definition 3.16.
Let be a measured groupoid, and let be measured subgroupoids of , with .
-
1.
We say that the pair satisfies Property if the following conditions hold:
-
(a)
is everywhere nonamenable;
-
(b)
is amenable, of infinite type, and stably normal in ;
-
(c)
if is a stably normal amenable subgroupoid of , then is stably contained in ;
-
(d)
if is another subgroupoid of which is everywhere nonamenable and contains a stably normal amenable subgroupoid of infinite type, and if is stably contained in , then is stably equal to ;
-
(e)
for every Borel subset of positive measure, the groupoid is not contained in any product-like subgroupoid of .
-
(a)
-
2.
We say that satisfies Property if there exists a measured subgroupoid such that satisfies Property .
Remark 3.17.
These properties are stable under restrictions and stabilization. Also, if satisfies Property , then a subgroupoid such that satisfies Property is “stably unique” in the following sense: if and are two such subgroupoids, there exist a conull Borel subset and a partition into at most countably many Borel subsets such that for every , one has . 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 satisfy Property .
Proposition 3.18.
Let be a measured groupoid over a standard probability space , equipped with a strict action-type cocycle . Let be a nonseparating meridian, let be the -stabilizer of the isotopy class of , and let .
Then satisfies Property .
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 be a measured groupoid, equipped with a strict action-type cocycle . Let be a meridian, and let be the -stabilizer of the isotopy class of . Let be the union of all components of which are not once-holed tori. Let be the kernel of the restriction homomorphism to 111notice that if is nonseparating, or if is separating and both complementary components have genus at least , then , and let .
Then is everywhere nonamenable, is a normal amenable subgroupoid of of infinite type, and every stably normal amenable subgroupoid of is stably contained in .
Proof.
Notice that the subsurface is nonempty because the genus of is at least . Lemma 1.1 ensures that contains a nonabelian free subgroup, so Lemma 3.6 shows that is everywhere nonamenable.
Normality of in follows from the normality of in . Notice that is amenable (using Lemma 1.2 in the case where one of the complementary components of is a once-holed torus). As has trivial kernel, it follows that is amenable. And is of infinite type because is infinite (it always contains a power of ) and is action-type.
Let now be a stably normal amenable subgroupoid of . Let be a connected component of . Let be the cocycle obtained by post-composing with the restriction homomorphism. Let also be a nonabelian free subgroup which embeds into under the restriction homomorphism, and whose image in contains a pseudo-Anosov mapping class (this exists because is not a once-holed torus, see Lemma 1.1). Let .
By Lemma 3.9, we can find a partition into at most countably many Borel subsets such that for every , the pair has a canonical reduction set . As is stably normal in , up to refining the above partition, we can assume that for every , the groupoid is normal in . Lemma 3.10 thus ensures that is -invariant, so either or fills .
Assume towards a contradiction that for some such that has positive measure. In other words is irreducible. As has trivial kernel in restriction to , and as (which is contained in ) normalizes , Lemma 3.12 implies that is amenable. But is a nonabelian free group and is action-type, so we get a contradiction to Lemma 3.6.
It follows that for every , there exists a conull Borel subset such that . As was an arbitrary connected component of , this precisely means that is stably contained in . ∎
We now check Assertion (d) from Definition 3.16.
Lemma 3.20.
Let be a measured groupoid over a standard probability space , equipped with a strict action-type cocycle . Let be a nonseparating meridian, and let be the -stabilizer of the isotopy class of .
If is a subgroupoid of which is everywhere nonamenable and contains a stably normal amenable subgroupoid of infinite type, and if is stably contained in , then is stably equal to .
Proof.
Let be an amenable subgroupoid of of infinite type which is contained in and stably normal in . By Lemma 3.9, we can find a partition into at most countably many Borel subsets such that for every , the pair has a (possibly empty) canonical reduction set . For every , we let be the (possibly empty) boundary multicurve of . As is stably normal in , up to refining the above partition, we can assume that for every , the set is -invariant (Lemma 3.10), and therefore so is the multicurve . As is stably contained in , we will also assume up to refining the above partition once more that for every , one has . In particular is -invariant, which implies that either or .
Let be such that has positive measure. If , then as is of infinite type and has trivial kernel, we deduce that , i.e. is irreducible. Lemma 3.12 then implies that is amenable, a contradiction. Therefore , so the isotopy class of is -invariant. As this is true for every such that has positive measure, we deduce that is stably contained in . ∎
We finally check Assertion (e) from Definition 3.16.
Lemma 3.21.
Let be a measured groupoid over a standard probability space , equipped with a strict action-type cocycle . Let be a nonseparating meridian, and let be the -stabilizer of the isotopy class of .
Then for every Borel subset of positive measure, the groupoid is not contained in any product-like subgroupoid of .
Proof.
Let be a Borel subset of positive measure. Assume towards a contradiction that is contained in a product-like subgroupoid of . Let be as in the definition of a product-like subgroupoid (Definition 3.15).
Up to restricting to a Borel subset of of positive measure, we can assume that has a canonical reduction multicurve . As , the isotopy class of is -invariant, so or , and the kernel of the induced cocycle is amenable (it is trivial if , and contained in if ). As is everywhere nonamenable, it follows that for every conull Borel subset (as otherwise would be equal to the kernel and therefore amenable). In particular, the subsurface is active for , and therefore is irreducible. As is everywhere nonamenable, we also have for every positive measure Borel subset . As is stably normal in , Corollary 3.11 therefore ensures that is also irreducible.
By definition of a strongly Schottky pair (applied to ), there exists a Borel subset of positive measure such that every normal amenable subgroupoid of is stably trivial. In particular, the kernel of restricted to the subgroupoid is stably trivial. As is of infinite type, it follows that for every Borel subset of positive measure, we have . As is stably normalized by , Corollary 3.11 ensures that is irreducible.
By definition of a strongly Schottky pair (applied to ), there exists a Borel subset of positive measure such that every normal amenable subgroupoid of is stably trivial. In particular, the kernel of restricted to the subgroupoid is stably trivial. This implies that we can find a positive measure Borel subset such that has trivial kernel in restriction to . As is stably normalized by , it thus follows from Lemma 3.12 that is amenable, which yields the desired contradiction. ∎
3.6 Stabilizers of separating meridians do not satisfy Property
Lemma 3.22.
Let be a measured groupoid, equipped with a strict action-type cocycle , and let be a measured subgroupoid of . Let be a separating meridian, and assume that the isotopy class of is -invariant.
Then does not satisfy Property .
Proof.
We first assume that one complementary component of is a once-holed torus. Then contains, up to isotopy, a unique nonseparating meridian (Lemma 1.2), so is contained in the -stabilizer of the isotopy class of . In addition is everywhere nonamenable and contains as a normal amenable subgroupoid of infinite type. Finally is not stably contained in because supports a pseudo-Anosov handlebody group element , and no nontrivial power of preserves the isotopy class of . So Assumption (d) from Definition 3.16 fails.
We now assume that both complementary components of have genus at least . Let be the -stabilizer of . Then is contained in , and we will prove that is product-like (which will imply that Assumption (e) from Definition 3.16 fails). For every , let be the subgroup of made of elements that have a representative supported in , and let . Then is normal in . For every , let and be two elements of that generate a nonabelian free subgroup of . For every and every , let . Then is normalized by , and Lemma 3.6 ensures that is a strongly Schottky pair of subgroupoids of . This completes our proof. ∎
3.7 Admissible decorated multicurves and their active subgroups
A decorated multicurve is a pair , where is a multicurve on , and is a subset of the set of complementary components of in . We make the following definition.
Definition 3.23.
Let be a decorated multicurve. The subgroup of made of all elements that preserve the isotopy class of and act trivially on all complementary subsurfaces not in is called the active subgroup of .
The decorated multicurve is admissible if its active subgroup is amenable, is the boundary multicurve of , and is its set of active complementary components.
Here is an example. If consists of a single nonseparating meridian on , or a separating meridian none of whose complementary components is a once-holed torus, and , then is admissible, and its active subgroup is the corresponding twist subgroup. In the case where one of the complementary components of a meridian is a once-holed torus (and contains a unique nonseparating meridian up to isotopy), then is admissible, with active subgroup the twist subgroup about . Other examples come from annulus pairs instead of meridians.
Let be a measured groupoid over a standard probability space , equipped with an action-type cocycle . We say that a pair of subgroupoids of is admissible with respect to if there exist a conull Borel subset and a partition into at most countably many Borel subsets, such that for every , there exist a multicurve on , and a subset of the set of all complementary components of such that is admissible, is equal to the -stabilizer of the isotopy class of , and denoting by the active subgroup of , one has . Notice that, although the above partition is not unique (one can always pass to a further partition), the map sending to the isotopy class of is uniquely determined by ; we call it the decomposition map of .
Lemma 3.24.
Let be a measured groupoid over a standard probability space , equipped with a strict action-type cocycle . Let be measured subgroupoids of , with .
If satisfies Property , then is an admissible pair.
Proof.
By Assumption (b) from Definition 3.16, the groupoid is amenable, of infinite type, and stably normal in . Up to a countable partition of the base space , we will assume that is normal in . Up to a further partition, we can also assume that has a canonical reduction set (Lemma 3.9). Let be the boundary multicurve of , let be the set of all active complementary components for , and let be the set of all complementary components of not in . Up to replacing by a conull Borel subset, we will assume using Lemma 3.10 that .
We will first prove that is admissible, so let us assume towards a contradiction that it is not. Let be the active subgroup of . Then there exists a conull Borel subset such that . Therefore is exactly the set of all curves whose isotopy class is -invariant, so is the boundary multicurve of and is its set of active complementary components. Therefore, our assumption that is not admissible implies that is not amenable, so it contains two elements which together generate a nonabelian free group (by the Tits alternative for mapping class groups [McC85a, Iva92]).
Let be the union of all subsurfaces in , viewed as a (possibly disconnected) surface of finite type. Let be the cocycle obtained by composing with the restriction to . We now observe that for every of positive measure, the restriction to of kernel of is nontrivial: otherwise is irreducible and normalizes , so Lemma 3.12 ensures that is amenable, a contradiction to Assumption (a) from Definition 3.16.
Let be the kernel of . The groupoid is normal in . We first assume that is amenable, and reach a contradiction in this case. Assumption (c) from Definition 3.16 ensures that there exists a Borel subset of positive measure such that . But the -image of every element of acts trivially on all components in , while the -image of every element of acts trivially on all components in . It follows that for every , the element is a multitwist around curves in . As has trivial kernel and is nontrivial, it follows that the subgroup of consisting of all multitwists about the curves in is infinite. Let . Then , and is everywhere nonamenable (it contains ) and contains as a normal amenable subgroupoid of infinite type. So Assumption (d) from Definition 3.16 ensures that there exists a Borel subset of positive measure such that . Now, the groupoid is contained in , so it normalizes , and has trivial kernel in restriction to (as otherwise would contain an infinite amenable normal subgroup). As is irreducible, Lemma 3.12 implies that is amenable, a contradiction to Lemma 3.6.
We now assume that is nonamenable, and also reach a contradiction in this case. As has trivial kernel, the subgroup of made of all elements that fix the isotopy class of and act trivially on all connected components in is nonamenable, and therefore contains a nonabelian free subgroup. Let , and let (i.e. with the notation from above). We will now reach a contradiction to Assumption (e) from Definition 3.16 by proving that is a product-like subgroupoid of (in which is contained).
Let be the normal subgroup made of all elements of that act trivially on all components in , and recall that is the normal subgroup made of all elements of acting trivially on all components in . Then is normal in for every . Notice that contains the nonabelian free subgroup , and we saw in the previous paragraph that also contains a nonabelian free subgroup. For every , let be two cyclic subgroups of that generate a nonabelian free subgroup, and for , let . As and centralize each other, it follows that each is normalized by . In addition, Lemma 3.6 ensures that is a strongly Schottky pair of subgroupoids of . So is a product-like subgroupoid of , which is the desired contradiction.
This contradiction shows that is admissible. Now, let , and let be the -stabilizer of the isotopy class of . Then is contained in , and contains as a normal amenable subgroupoid of infinite type. So Assertion (d) from Definition 3.16 ensures that is stably equal to . And Assertion (c) then implies that is stably equal to . This proves that is an admissible pair. ∎
3.8 Compatibility
Two decorated multicurves and are compatible if and are disjoint up to isotopy, and given any two components and , either and are isotopic, or they are disjoint up to isotopy. We start with the following observation.
Lemma 3.25.
Let and be two decorated multicurves, with respective active subgroups .
If and are compatible, then is amenable.
Proof.
Otherwise, using the Tits alternative for mapping class groups [McC85a, Iva92], there exist that together generate a nonabelian free group. Let be the union of all subsurfaces in . The conclusion is clear if is empty (as and ’ commute in this case), so we will assume otherwise.
We observe that is not virtually abelian: otherwise, as and are amenable subgroups of , there would exist two other virtually abelian subgroups (made of mapping classes supported on and , respectively), which commute and both commute with , such that every element in is a product of an element in , an element in , and an element in . This would contradict the fact that is a nonabelian free group.
Therefore contains a nonabelian free subgroup , and the commutator subgroup of is a nonabelian free group contained in . This contradiction completes our proof. ∎
Let be a measured groupoid over a standard probability space which admits an action-type cocycle . Two admissible pairs and (with respect to ) are compatible with respect to if, denoting by and their respective decomposition maps, for a.e. , the pairs and are compatible. The following proposition gives a purely groupoid-theoretic characterization of compatibility (i.e. with no reference to the cocycle ).
Proposition 3.26.
Let be a measured groupoid over a standard probability space , equipped with a strict action-type cocycle . Let and be two admissible pairs with respect to . Then the following are equivalent.
-
1.
and are compatible with respect to ;
-
2.
for every Borel subset of positive measure, there exists a Borel subset of positive measure such that is amenable.
Proof of Proposition 3.26.
Let be a countable Borel partition of a conull Borel subset such that for every , there exist admissible pairs and such that and , and letting be the active subgroups of and respectively, we have and .
We first prove that . If fails, then there exists such that has positive measure and and are not compatible. Then there exist and that generate a nonabelian free subgroup of , as follows from [Kob12, Theorem 1.8, and the sentence following it]. Lemma 3.6 ensures that for every Borel subset of positive measure, the groupoid is nonamenable. Therefore is also nonamenable for every Borel subset of positive measure, so fails.
We now prove that . If holds, then for every such that has positive measure, the pairs and are compatible, so is amenable (Lemma 3.25). Let now be a Borel subset of positive measure, and let of positive measure be contained in for some . Then is contained in , which is amenable because is and has trivial kernel. ∎
Let be two measured subgroupoids of of meridian type with respect to an action-type cocycle . We say that and are compatible with respect to if, denoting by their respective meridian maps with respect to , for a.e. , the meridians and are disjoint up to isotopy.
Corollary 3.27.
Let be a measured groupoid over a standard probability space , equipped with two strict action-type cocycles , and let be two measured subgroupoids of of meridian type with respect to both and .
Then and are compatible with respect to if and only if they are compatible with respect to .
Proof.
Let be a partition of a conull Borel subset into at most countably many Borel subsets such that for every and every , there exist meridians such that are equal to the -stabilizers of the isotopy classes of , respectively.
For every and every , let (resp. ) be the subgroup of made of all elements that act trivially in restriction to every connected component of (resp. ) which is not a once-holed torus. Notice that are the active subgroups of some admissible decorated multicurves , by letting and be obtained from and by adding the unique nonseparating meridian in every complementary component which is a once-holed torus, and letting . See the examples right after Definition 3.23. Notice that and are disjoint up to isotopy if and only if and are compatible.
For every , let be a subgroupoid such that for every , and let be defined in the same way, using in place of . Then and are admissible pairs with respect to . Lemma 3.19 thus ensures that and are stably equal (as they are both stably maximal for the property of being a stably normal amenable subgroupoid of ), and likewise and 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 be a measured groupoid over a standard probability space , equipped with two strict action-type cocycles , and let be a measured subgroupoid.
Then is of nonseparating-meridian type with respect to if and only if it is of nonseparating-meridian type with respect to .
A decorated multicurve is clean if it is not of the form for some separating meridian . The graph of clean admissible decorated multicurves 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 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 is naturally a subgraph of , by sending a nonseparating meridian to the pair .
Lemma 3.29.
Every injective graph map222i.e. preserving adjacency and non-adjacency from to takes its values in (viewed as a subgraph of via the natural inclusion).
Proof.
Let be a vertex. By completing to a pair of pants decomposition made of nonseparating meridians, we can find pairwise distinct, pairwise adjacent vertices such that for every , one has
So the same property should hold for their images in , which correspond to decorated multicurves . For every , we let be the subsurface of equal to the union of all subsurfaces in , together with all annuli around curves in that are not boundary curves of any subsurface in . Notice that the set cannot contain both a subsurface and the collar neighborhood of one of its boundary components, as otherwise removing from the collection does not change the link. More generally, for every , one of the connected components is not a connected component of some with , and is also not the collar neighborhood of a boundary curve of – otherwise removing from the collection does not change its link. So the subsurfaces are pairwise nonisotopic and does not contain a subsurface together with the collar neighborhood of one of its boundary components. For every , let be the set of all boundary curves of , and let be a set of isotopy classes of essential simple closed curves on that form a pair of pants decomposition of (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 and contains at least curves, each being repeated at most twice up to isotopy (and the are not isotopic to any other curve in the collection). So every subsurface contributes exactly two curves that are both of the form , and is therefore an annular subsurface. Furthermore, since there are such, and no appears as a subsurface of , we actually have for all . Therefore , where is the core curve of the annulus . As is admissible, some power of the twist around must belong to the handlebody group, so is a meridian by [Oer02, Theorem 1.11] or [McC06, Theorem 1]. As is clean, the meridian is nonseparating, and the conclusion follows. ∎
Proof of Proposition 3.28.
Let , in other words is the isotopy class of a nonseparating meridian. Let be the -stabilizer of , and let . Then satisfies Property (by Proposition 3.18, applied to the cocycle ). Lemma 3.24, applied to the cocycle , implies that is an admissible pair with respect to . So there exist a conull Borel subset and a partition into at most countably many Borel subsets such that for every , there exists a (unique) admissible pair such that is the -stabilizer of and, denoting by the active subgroup of , one has . In addition, Lemma 3.22 ensures that is clean. For every and every , we then let whenever . This defines a Borel map .
We claim that for almost every , the map determines a graph embedding . Let us first explain how to complete the proof of the proposition from this claim. By Lemma 3.29, every graph embedding sends nonseparating meridians to nonseparating meridians. Therefore, if is a nonseparating meridian, then is a nonseparating meridian (and ) whenever has positive measure, and the proposition follows.
We are now left with proving the above claim. First, for almost every , the map is injective. Indeed otherwise, as is countable, there exist a Borel subset of positive measure and two non-isotopic nonseparating meridians such that for every , one has (we denote by the common image). In particular, the -stabilizer of is stably equal to the -stabilizer of , since they are both stably equal to the -stabilizer of . This contradicts Lemma 3.13.
Second, Proposition 3.26 ensures that for almost every , the map 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 , 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
We proved in Proposition 3.28 that for a subgroupoid , being of nonseparating meridian type does not depend of the choice of an action-type cocycle . 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 ).
Let be a measured groupoid over a standard probability space which admits a strict action-type cocycle towards . A measured subgroupoid satisfies Property if
-
1.
contains a strongly Schottky pair of subgroupoids;
-
2.
there exists a stably normal amenable subgroupoid of infinite type, such that for every measured subgroupoid of nonseparating-meridian type, and every stably normal amenable subgroupoid of infinite type, the intersection is stably trivial;
-
3.
given any two subgroupoids of nonseparating-meridian type, and any Borel subset of positive measure, assuming that , then and are stably equal;
-
4.
there exist measured subgroupoids of of nonseparating-meridian type, which are pairwise compatible, such that
-
(a)
for every Borel subset of positive measure, and any two distinct , one has , and
-
(b)
for every , every stably normal amenable subgroupoid of is stably contained in .
-
(a)
Notice that this notion is stable under restriction to a positive measure subset. Also, if is a subgroupoid of , and if there exists a partition into at most countably many Borel subsets such that for every , the subgroupoid of satisfies Property , then (as a subgroupoid of ) satisfies Property .
To motivate the definition, we begin by proving that subgroupoids of separating meridian type have this property.
Proposition 3.31.
Let be a measured groupoid over a standard probability space , equipped with a strict action-type cocycle . Let be a separating meridian, and let be the -stabilizer of the isotopy class of .
Then satisfies Property .
Proof.
For Property , Lemma 1.1 ensures that the stabilizer in of any separating meridian contains a nonabelian free subgroup (recall that we are assuming that the genus of is at least ). Lemma 3.6 thus implies that contains a strongly Schottky pair of subgroupoids.
For Property , notice that the cyclic subgroup generated by the Dehn twist about is normal in the stabilizer of . Therefore is a normal subgroupoid of , which is amenable as has trivial kernel, and of infinite type because is action-type. Let now be a measured subgroupoid of of nonseparating meridian type, and let be a stably normal amenable subgroupoid of infinite type. By Lemma 3.19, we can find a partition of a conull Borel subset into at most countably many Borel subsets such that for every , there exists a nonseparating meridian such that . It follows that is trivial, so is stably trivial.
Property follows from the fact that for every separating meridian , there is at most one nonseparating meridian such that every infinite-order element of has a power that fixes (in fact, the existence of such a occurs precisely when one of the two connected components of is a once-holed torus, in which case it contains a unique nonseparating meridian up to isotopy, and we take as such – notice that we are using the fact that the genus of is at least here; see Lemma 1.2).
We now prove that satisfies Property . Let be a set of pairwise non-isotopic nonseparating meridians which together with form a pair of pants decomposition of . For every , let be the -stabilizer of the isotopy class of . Then are of nonseparating meridian type. Lemma 3.13 ensures that they satisfy Assertion . Finally, Lemma 3.19 ensures that every stably normal amenable subgroupoid of is stably contained in . In particular each is stably contained in . ∎
3.10.2 Characterization
Our goal is now to characterise subgroups of separating meridian-type by proving the following proposition.
Proposition 3.32.
Let be a measured groupoid over a standard probability space , equipped with a strict action-type cocycle . Let be a measured subgroupoid of . The following assertions are equivalent.
-
1.
The subgroupoid is of separating-meridian type with respect to .
-
2.
The subgroupoid satisfies Property , and is stably maximal among all measured subgroupoids of with respect to this property, i.e. if is another subgroupoid satisfying Property , and if is stably contained in , then is stably equal to .
Before turning to the proof of Proposition 3.32, we record the following consequence.
Corollary 3.33.
Let be a measured groupoid over a base space , equipped with two strict action-type cocycles , and let be a measured subgroupoid.
Then is of separating-meridian type with respect to if and only if it is of separating-meridian type with respect to . ∎
Our goal is now to prove Proposition 3.32. Our first lemma exploits the first two assumptions of Property in order to derive information about the possible canonical reduction multicurves of .
Lemma 3.34.
Let be a measured groupoid over a standard probability space , equipped with a strict action-type cocycle . Let be measured subgroupoids of , with . Assume that
-
1.
contains a strongly Schottky pair of subgroupoids;
-
2.
is amenable and of infinite type, and stably normal in ;
-
3.
for every measured subgroupoid of nonseparating-meridian type, and every stably normal amenable subgroupoid of infinite type, the intersection is stably trivial.
Then for every Borel subset of positive measure, the pair cannot have a canonical reduction multicurve consisting of a single nonseparating meridian.
Proof.
Assume towards a contradiction that has a canonical reduction multicurve which is reduced to a single nonseparating meridian . As is stably normal in , up to restricting to a positive measure Borel subset of , we can assume that is -invariant. In particular, letting , we have a natural cocycle . The kernel of is contained in . As has trivial kernel, it follows that the kernel of is amenable. In particular, letting be a strongly Schottky pair of subgroupoids of , there exists a positive measure Borel subset such that has trivial kernel when restricted to .
Our third assumption, applied by taking for the -stabilizer of , and with , ensures that is stably trivial. Let be a Borel subset of positive measure such that is trivial. Then also has trivial kernel when restricted to . In particular is irreducible, and Lemma 3.12 implies that is amenable, a contradiction. ∎
Lemma 3.35.
Let be a measured groupoid over a standard probability space , equipped with a strict action-type cocycle . Let be a measured subgroupoid of which satisfies Property .
Then there exists a partition into at most countably many Borel subsets such that for every , there exists an -invariant isotopy class of separating meridian.
Proof.
Let be subgroupoids of provided by Property . Up to partitioning into at most countably many Borel subsets, we can assume that for every , the groupoid is equal to the -stabilizer of the isotopy class of a nonseparating meridian .
Let be as in Property . By Lemma 3.9, we can find a partition into at most countably many Borel subsets of positive measure such that for every , the pair has a canonical reduction set , with boundary multicurve . As is stably normal in , up to refining this partition, we can assume that for every , the isotopy class of the multicurve is -invariant.
We first observe that for every , one has . Indeed, otherwise is irreducible. As is amenable and stably normal in , and has trivial kernel, Lemma 3.12 implies that is amenable, contradicting Property .
For every , let be the Dehn twist about the meridian . Then is a normal amenable subgroupoid of . Property thus ensures that is stably contained in . Therefore, for every curve in , there exists a positive integer such that the isotopy class of is fixed by . This implies that is disjoint (up to isotopy) from all meridians .
We now claim that for every , the multicurve contains at most one of the curves . Indeed, assume by contradiction that it contains two curves and . Then is contained in . As and are of nonseparating meridian type with respect to , Property implies that there exists a positive measure Borel subset such that , contradicting Property .
As is a set of pairwise disjoint and pairwise non-isotopic nonseparating simple closed curves on , one of the complementary components of the union of all curves is a -holed sphere . Notice that every essential simple closed curve contained in is a meridian, and may contain such a curve. This leaves three possibilities for the canonical reduction multicurve of , namely:
-
1.
a single nonseparating meridian (either one of the meridians , or else a nonseparating meridian contained in );
-
2.
the union of a nonseparating meridian and a nonseparating essential simple closed curve (in fact a meridian) contained in ;
-
3.
a separating (on ) essential simple closed curve (in fact a meridian) contained in , possibly together with a meridian .
The first case is excluded by Lemma 3.34, the second case is excluded using Property 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 . Let be a measured subgroupoid of of separating meridian type with respect to , and let be a partition of a conull Borel subset into at most countably many Borel subsets, such that for every , the groupoid is equal to the -stabilizer of the isotopy class of a separating meridian .
Proposition 3.31 implies that satisfies Property . We need to check that is stably maximal among all measured subgroupoids of that satisfy Property . So let be a measured subgroupoid of which satisfies Property , and such that is stably contained in . By Lemma 3.35, up to refining the above partition of , we can assume that for every , there exists a separating meridian whose isotopy class is -invariant. Lemma 3.14 implies that for every . It follows that is stably contained in , so they are stably equal. This completes our proof of the implication .
We now prove that , so let be a measured groupoid of that satisfies Assertion . By Lemma 3.35, we can find a partition into at most countably many Borel subsets such that for every , there exists a separating meridian whose isotopy class is -invariant. Let be a measured subgroupoid of such that for every , the groupoid is equal to the -stabilizer of the isotopy class of . Then is stably contained in . In addition, is of separating meridian type, so Proposition 3.31 shows that satisfies Property . The maximality assumption on therefore implies that is stably equal to . Hence 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 be a measured groupoid, equipped with two strict action-type cocycles , and let be a measured subgroupoid.
Then is of meridian type with respect to if and only if it is of meridian type with respect to . ∎
We will now simply say that is of meridian type to mean that it is of meridian type with respect to any action-type cocycle . 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 .
Proof of Theorem 3.2.
Let be a measured groupoid over a standard probability space , and let be two strict action-type cocycles. Let be the disk graph of : we recall that its vertices are the isotopy classes of meridians in , and two such isotopy classes are joined by an edge if they have disjoint representatives.
Proposition 3.36 ensures that for every vertex , there exists a Borel map such that for every , letting , the -stabilizer of is stably equal to the -stabilizer of . Lemmas 3.13 and 3.14 ensure that the map is essentially unique.
For every and every , we then let . This defines a Borel map .
We claim that for a.e. , the map is a graph automorphism of . Indeed, injectivity follows from the same argument as in the proof of Proposition 3.28, and the fact that is almost everywhere a graph map follows from Corollary 3.27. We now show that for almost every , the map is surjective. So let be a meridian. By Proposition 3.36, there exists a Borel partition of a conull Borel subset into at most countably many Borel subsets such that for every , the -stabilizer of coincides with the -stabilizer of some . It follows that for almost every . Surjectivity follows.
By the main theorem of [KS09], the natural map is an isomorphism (noting again that the genus of is at least ). We can thus find a Borel map so that for a.e. we have that for all meridians.
We are left with showing that satisfies the equivariance condition required in Theorem 3.2. This amounts to proving that there exists a conull Borel subset such that for every and every vertex , one has As is a countable union of bisections, it is enough to prove it for almost every in a bisection (inducing a Borel isomorphism between and ). Up to further partitioning , we can assume that and are constant, with values , and that is constant, with value . We now aim to show that for almost every , one has . By definition of , the -stabilizer of is stably equal to the -stabilizer of . Conjugating by the bisection, it follows that the -stabilizer of is stably equal to the -stabilizer of , 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 be a handlebody of genus at least . Let be a locally compact second countable group, equipped with its (left or right) Haar measure. Let be a finite index subgroup of , and let be an injective homomorphism whose image is a lattice.
Then there exists a homomorphism with compact kernel such that for every , one has .
Proof.
Theorem 3.2 precisely says that is rigid with respect to action-type cocycles in the sense of [GH21, Definition 4.1]. As 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 determines a self measure equivalence coupling of (acting on equipped with its Haar measure), and the rigidity statement provided by Theorem 3.2 from the present paper ensures that the self coupling of on factors through the obvious coupling on where acts by left/right multiplication. This yields a Borel map , and some extra work is needed to upgrade it to a continuous homomorphism. ∎
A theorem of Suzuki ensures that is finitely generated [Suz77] (it is in fact finitely presented by work of Wajnryb [Waj98]). Given a finitely generated group and a finite generating set of , the Cayley graph is defined as the simple graph whose vertices are the elements of , with an edge between distinct elements if .
Theorem 4.2.
Let be a handlebody of genus at least .
-
1.
For every finite generating set of , every automorphism of is at bounded distance from the left multiplication by an element of .
-
2.
For every torsion-free finite-index subgroup and every finite generating set of , the automorphism group of is countable (in fact it embeds as a subgroup of containing ).
Proof.
As mentioned in the introduction, torsion-freeness of 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 and be two countable groups, and for every , let be a standard probability space equipped with a free ergodic measure-preserving action of .
The actions and are virtually conjugate (as in [Kid08b, Definition 1.3]) if there exist finite normal subgroups , finite-index subgroups , and free ergodic measure-preserving actions on standard probability spaces, so that and are conjugate, and for every , the action of on is induced from the -action on . This implies in particular that the groups and are virtually isomorphic (i.e. commensurable up to finite kernels).
The following is a weaker notion. The actions and are stably orbit equivalent if there exist positive measure Borel subsets and and a measure-scaling isomorphism 444in other words induces a measure space isomorphism between the probability spaces and such that for almost every , one has
A free ergodic measure-preserving action of on a standard probability space is OE-superrigid if for every countable group , and every free ergodic measure-preserving action of on a standard probability space , if the -action on is stably orbit equivalent to the -action on , then the two actions are virtually conjugate (in particular and 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 be a handlebody of genus at least . Then every free ergodic measure-preserving action of on a standard probability space is OE-superrigid.
Rigidity of von Neumann algebras.
Let be a countable group, and let be a standard probability space equipped with a standard ergodic action of . Associated to the -action on is a von Neumann algebra , 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 , the handlebody group is residually finite by a theorem of Grossman [Gro75].
A free ergodic measure-preserving action of a countable group on a standard probability space is -superrigid if for every countable group , and every weakly compact free ergodic measure-preserving action of on a standard probability space , if the von Neumann algebras and are isomorphic, then the -action on is virtually conjugate to the -action on .
Theorem 4.4.
Let be a handlebody of genus at least . Then every free ergodic measure-preserving action of on a standard probability space is -superrigid.
Proof.
Let be a standard probability space equipped with a free ergodic measure-preserving action of , and let be a standard probability space equipped with a weakly compact free ergodic measure-preserving action of a countable group . Assume that there exists an isomorphism . By [HHL20, Theorem 7], the group 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 sends to . This implies that the actions and 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 -superrigidity result for handlebody groups concerns their Bernoulli actions, that is, actions of the form , where is a standard probability space not reduced to a point, and the action is by shift. More precisely, when has genus at least 3, if a Bernoulli action 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 be the cyclic subgroup generated by a Dehn twist about a nonseparating meridian , letting be the stabilizer of the isotopy class of , and . Indeed, to check that [HH21b, Theorem A.2] applies, we only need to find an element such that is infinite, and generates . For this, let be nonseparating meridians such that are pairwise disjoint, pairwise non-isotopic, and have connected complement. Let be an element sending to and commuting with the twist . Then is infinite because it contains . And is generated by and 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. and 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 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 . 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 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 -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]