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

Covers of surfaces

Ian Biringer Ian Biringer, Boston College, 140 Commonwealth Ave, Chestnut Hill, MA 02467 [email protected] Yassin Chandran Yassin Chandran, CUNY Graduate Center, 365 Fifth Avenue, New York, NY 10016 [email protected] Tommaso Cremaschi Tommaso Cremaschi, School of Mathematics, Trinity College Dublin, 17 Westland Row, Dublin 2, Ireland [email protected] Jing Tao Jing Tao, Department of Mathematics, University of Oklahoma, 601 Elm Ave, Room 423, Norman, OK 73069 [email protected] Nicholas G. Vlamis Nicholas G. Vlamis, Department of Mathematics, CUNY Graduate Center, 365 Fifth Avenue, New York, NY 10016, and Department of Mathematics, CUNY Queens College, 65-30 Kissena Blvd., Flushing, NY 11367 [email protected] Mujie Wang Mujie Wang, Boston College, 140 Commonwealth Ave, Chestnut Hill, MA 02467 [email protected]  and  Brandis Whitfield Brandis Whitfield, Temple University, 1805 N. Broad Street, Philadelphia PA, 19122 [email protected]
Abstract.

We study the homeomorphism types of certain covers of (always orientable) surfaces, usually of infinite-type. We show that every surface with non-abelian fundamental group is covered by every noncompact surface, we identify the universal abelian covers and the /n\mathbb{Z}/n\mathbb{Z}-homology covers of surfaces, and we show that non-locally finite characteristic covers of surfaces have four possible homeomorphism types.

1. Introduction

We are interested in the question of when a given surface covers another given surface, either in general or with constraints on the type of covering. Our first theorem is as follows.

Theorem 1.1 (Everything covers everything).

Suppose that SS is an orientable, borderless surface with non-abelian fundamental group. Then SS is covered by any noncompact borderless orientable surface.

A borderless surface is a topological 22-manifold with empty boundary. We stick with orientable surfaces for simplicity. The only orientable surfaces SS not covered by the theorem are the disk, the annulus, and the torus, each of which is only covered by itself and the former surface. A version of Theorem 1.1 also appears in [8] in which the author shows that any infinite-type surface covers any compact surface with negative euler characteristic. Moreover, they also allow for non-orientable surfaces.

If SS is any surface, there are certain natural covers coming from homology. The universal abelian cover of SS is the cover SabSS^{ab}\to S corresponding to the commutator subgroup of π1(S)\pi_{1}(S). We prove:

Theorem 1.2 (Universal abelian covers).

Let SS be an orientable, borderless surface, and let SabS^{ab} be its universal abelian cover. Then

  1. (1)

    if SS is 2\mathbb{R}^{2}, the annulus, or the torus then Sab2S^{ab}\cong\mathbb{R}^{2},

  2. (2)

    if SS is the sphere, so is SabS^{ab},

  3. (3)

    if SS is the once-punctured torus, then SabS^{ab} is a flute surface,

  4. (4)

    if SS is a finite-type surface with one puncture and genus at least 22, then SabS^{ab} is the spotted Loch Ness monster surface,

  5. (5)

    otherwise, SabS^{ab} is the Loch Ness monster surface.

Note that the commutator subgroup of π1(S)\pi_{1}(S) is the kernel of the map π1(S)H1(S,),\pi_{1}(S)\to H_{1}(S,\mathbb{Z}), so the universal abelian cover could also be called the “\mathbb{Z}-homology cover” of SS. We also consider the /n\mathbb{Z}/n\mathbb{Z}-homology cover S~S\tilde{S}\to S corresponding to the kernel of the map

π1(S)H1(S,/n).\pi_{1}(S)\to H_{1}(S,\mathbb{Z}/n\mathbb{Z}).

When SS has finite-type, the /n\mathbb{Z}/n\mathbb{Z}-homology cover S~S\tilde{S}\to S is a finite cover, and one can figure out the topological type of S~\tilde{S} via an Euler characteristic argument and by analyzing how boundary components lift. On the other hand, when S~\tilde{S} has infinite-type, its /n\mathbb{Z}/n\mathbb{Z}-homology covers all have infinite degree, and it turns out their topology is quite constrained.

Theorem 1.3 (/n\mathbb{Z}/n\mathbb{Z}-homology covers).

Suppose that SS is an infinite-type orientable borderless surface, n2n\geq 2, and π:S~S\pi:\tilde{S}\to S is the /n\mathbb{Z}/n\mathbb{Z}-homology cover. If SS has no isolated planar ends, then S~\tilde{S} is homeomorphic to the Loch Ness monster surface. Otherwise, S~\tilde{S} is homeomorphic to the spotted Loch Ness monster surface.

Suppose that a cover S~S\tilde{S}\to S corresponds to the subgroup N<π1(S,p)N<\pi_{1}(S,p). We say that the cover is characteristic if NN is characteristic, i.e., if NN is invariant under every automorphism of π1(S,p)\pi_{1}(S,p). The \mathbb{Z} and /n\mathbb{Z}/n\mathbb{Z}-homology covers above are all characteristic. Motivated by Theorems 1.2 and 1.3, we ask:

Question 1.4.

Are all infinite-degree characteristic covers of orientable borderless surfaces homeomorphic to either the disk, the flute surface, the Loch Ness monster surface, or the spotted Loch Ness monster surface?

Note that it is important to say ‘infinite degree’ above. Indeed, if SS has finite type it has many finite-degree characteristic covers S~\tilde{S}, all of which also have finite type. (For these S~\tilde{S}, in practice one can usually figure out the topological type of S~\tilde{S} by analyzing how boundary components lift and using an Euler characteristic argument.) When SS has infinite type, the ‘infinite degree’ hypothesis just prohibits the cover from being trivial (which is obviously necessary), since FF_{\infty}, the free group on a countably infinite alphabet, does not have any proper finite-index characteristic subgroups.

Here is some evidence that the answer to Question 1.4 could be yes.

Theorem 1.5 (Characteristic covers of finite-type surfaces).

Suppose that SS is an orientable borderless surface of finite type and S~S\tilde{S}\to S is an infinite-degree, geometrically characteristic cover. Then S~\tilde{S} is homeomorphic to either a disc, the flute surface, the Loch Ness monster surface, or the spotted Loch Ness monster surface.

Above, an automorphism of π1(S,p)\pi_{1}(S,p) is called geometric if it is induced by a homeomorphism SSS\to S that preserves the basepoint pp. We call a subgroup N<π1(S,p)N<\pi_{1}(S,p) geometrically characteristic if NN is invariant under all geometric automorphisms of π1(S,p)\pi_{1}(S,p). So, ‘geometrically characteristic’ is a weaker condition than ‘characteristic’. A cover S~S\tilde{S}\to S is geometrically characteristic if it corresponds to a geometrically characterstic subgroup of π1(S,p)\pi_{1}(S,p).

It follows quickly from Theorem 1.5 that any non-universal characteristic cover S~\tilde{S} of a closed surface is homeomorphic to the Loch Ness monster surface. This was proved earlier by Atarihuana–García–Hidalgo–Quispe–Maluendas [2, Prop 6.1], using the fact that S~\tilde{S} is also a characteristic cover of a triangle orbifold. However, their argument does not extend to the finite-type setting.

Infinite-type surfaces can be exhausted by finite-type surfaces, and characteristic subgroups of the infinite-type surface group intersect the finite-type surface groups in characteristic subgroups (see §3.2 for details) which suggests that the theorem above can be promoted to apply to infinite-type SS. However, there is a problem: there are infinite-degree characteristic covers of infinite-type surfaces S~\tilde{S} that restrict to finite-degree covers on all finite-type subsurfaces, and Theorem 1.5 does not apply to finite covers. To deal with this, we make the following definition.

Definition 1.6.

If GG is a group, a subgroup N<GN<G is locally finite index if NN intersects every finitely generated subgroup FGF\subset G in a finite-index subgroup of FF. A cover π:S~S\pi:\tilde{S}\to S is locally finite if it corresponds to a locally finite index subgroup of π1(S)F\pi_{1}(S)\cong F_{\infty}. Equivalently, π\pi is locally finite if for every compact subsurface KSK\subset S, each component of π1(K)\pi^{-1}(K) is compact.

When SS has finite-type, a locally finite cover S~S\tilde{S}\to S is just a finite cover. When SS has infinite-type, the /n\mathbb{Z}/n\mathbb{Z}-homology covers discussed above are locally finite, but not finite. Locally finite index subgroups have been studied previously in [3, 9], for instance, but we are not aware of any previous discussion of locally finite covering maps. We prove:

Theorem 1.7 (Characteristic covers that are not locally finite).

Suppose that SS is an orientable, infinite-type borderless surface and S~S\tilde{S}\to S is a characteristic cover that is not locally finite. Then S~\tilde{S} is homeomorphic to either a disc, the flute surface, the Loch Ness monster surface, or the spotted Loch Ness monster surface.

So, to answer Question 1.4, it only remains to understand locally finite characteristic covers of infinite-type surfaces. One can check that the /n\mathbb{Z}/n\mathbb{Z}-homology covers discussed above are the only such covers that are abelian, and we have answered Question 1.4 positively for those covers. An example of a non-abelian, locally finite, characteristic cover of a surface SS is the cover S~\tilde{S} corresponding to the subgroup N<π1(S)N<\pi_{1}(S) generated by all third powers of elements of π1(S)\pi_{1}(S). This cover is obviously characteristic, and it is locally finite because Burnside [6] proved that the group B(d,3)=Fd/w3|wFdB(d,3)=F_{d}/\langle\langle w^{3}\ |\ w\in F_{d}\rangle\rangle is finite for all dd. It is not clear to us what S~\tilde{S} is in this case.

Our interest in Theorem 1.7 is partly inspired by work on big mapping class groups of surfaces. In particular, Aramayona–Leininger–McLeay [1, Proposition 3.3] show that the /2\mathbb{Z}/2\mathbb{Z} homology cover of the Loch Ness monster surface is the Loch Ness monster surface. They use this to show that the mapping class group GG of the once-punctured Loch Ness monster surface admits a continuous injective homomorphism GGG\to G that is not surjective. We note that there is a gap in the published version of their proof of the covering statement, but they were able to fix it quickly when we notified them about it. Alternatively, their Proposition 3.3 follows from Theorem 1.3, or even from Theorem 1.5 or [2, Proposition 6.1], since the Loch Ness monster surface is itself a characteristic cover of a genus two surface.

Surprisingly, to our knowledge there are not many previous results in which the homeomorphism types of infinite covers of surfaces are studied. In particular, we were surprised that a result like Theorem 1.1 does not currently exist in the literature. However, there are a few results out there that are related to ours. First, a classification of the possible homeomorphism types of infinite-degree regular covers of a closed surface is readily deduced from the fact that the cover and the deck group are quasi-isometric (for instance, see [4, Proposition 5.2]). Then there are older papers of Maskit [10] and Papakyriakopoulos [11] on identifying when certain covers of surfaces are planar; their motivation was some interesting applications to 33-manifolds, including a new proof of the Loop Theorem. There are the papers of Aramayona–Leininger–McLeay [1] and Atarihuana et al [2] referenced above, and there is also a forthcoming paper of Ghaswala–McLeay in which (among other things) the authors construct finite-degree covers from certain maps between potential end spaces of surfaces.

1.1. Ideas of the proofs

The proofs of our theorems above have distinct flavors. We will sketch each of them here for the convenience of the reader.

Theorem 1.1, in which we show that every noncompact surface covers every surface with non-abelian fundamental group, is very hands-on. Every surface SS with non-abelian fundamental group has an essential pair of pants and so is covered by an open pair of pants. The universal abelian cover of an open pair of pants is the Loch Ness monster surface, which is covered by the blooming Cantor tree surface, and every noncompact surface S~\tilde{S} can be π1\pi_{1}-injectively embedded in the blooming Cantor tree surface, and hence covers the blooming Cantor tree surface. (Here the embedding is as a component of the complement of a set of closed curves and bi-infinite arcs.) Composing, we get a covering map S~S\tilde{S}\to S.

Theorem 1.2, in which we identify the universal abelian covers SabS^{ab} of surfaces SS, is proved as follows. When SS has finite-type, one can identify SabS^{ab} using the fact that the deck group is free abelian, along with the knowledge of whether the peripheral curves of SS lift or not. We deal with infinite-type SS by exhausting them with finite-type subsurfaces and saying that SabS^{ab} is exhausted by the universal abelian covers of the subsurfaces.

Theorem 1.3, in which we show that the /n\mathbb{Z}/n\mathbb{Z}-homology cover S~S\tilde{S}\to S of any infinite-type surface is either the Loch Ness monster surface or its spotted version, uses the fact that the cover is abelian, together with the fact that it unwraps curves that appear basically everywhere in SS. Assuming for simplicity that SS has no cusps, we show that S~\tilde{S} has one end as follows: take two points p~,q~S~\tilde{p},\tilde{q}\in\tilde{S} that are far away from some chosen basepoint, and a path a~\tilde{a} between them. This a~\tilde{a} may pass close to the basepoint of S~\tilde{S}, but we can construct a new path γ~\tilde{\gamma} between them that stays far from the basepoint, by using abelianness of the cover to form a path-quadrilateral in S~\tilde{S} where a~,γ~\tilde{a},\tilde{\gamma} are a pair of opposite sides. See also the beginning of §3.4.1 for a more detailed sketch of the most important part of the argument.

Theorem 1.5, in which we identify all infinite-degree characteristic covers S~\tilde{S} of finite-type surfaces SS, has the most complicated proof. The argument is inspired by Proposition 3.3 in [1], which we discussed above.

Say for simplicity that SS is closed. We want to prove that S~\tilde{S} has one end. Here, S~\tilde{S} is quasiisometric to the deck group QQ. If QQ has more than one end, it admits an action on a tree TT with finite edge stabilizers and no global fixed point, by Stallings’s Theorem. We then construct a QQ-equivariant map S~T\tilde{S}\to T that is transverse to the union of edge midpoints, and let C~S~\tilde{C}\subset\tilde{S} be the preimage of the set of edge midpoints, which is then a properly embedded 11-submanifold invariant under QQ. Since edge stabilizers are finite, each component of C~\tilde{C} is a simple closed curve. Using similar methods and an inductive argument, we can enlarge C~\tilde{C} until every component of S~C~\tilde{S}\smallsetminus\tilde{C} is one-ended.

If CSC\subset S is the projection of C~\tilde{C}, and γ\gamma is an essential curve in SCS\smallsetminus C, then γ\gamma is ‘at most one ended’, meaning that all lifts of γ\gamma to S~\tilde{S} are either compact or are bi-infinite arcs both of whose ends go out the same end of S~\tilde{S}. Any at most one-ended γ\gamma projects to an element of QQ that acts elliptically on TT. Since the cover S~S\tilde{S}\to S is characteristic, any homeomorphism of SS lifts to a homeomorphism of S~\tilde{S}, and hence the image of γ\gamma under any homeomorphism of SS also is at most one-ended, and therefore acts elliptically on TT. The point is then that this gives a lot of elements of QQ that act elliptically, enough so that Serre’s criterion (see 2.5) implies that the action of QQ on TT has a global fixed point, a contradiction.

There are some complications in extending this argument to the setting of noncompact finite-type surfaces. For instance, one needs a version of Stallings’s Theorem that works relative to the peripheral groups, in a certain sense. Luckily, there is indeed a relative version of Stallings’s theorem due to Swarup [16], but it is phrased in terms of group cohomology. In order to apply Swarup’s theorem in our context, we wrote an Appendix (see §A) translating between the group cohomological interpretation of ends and the version that is natural for our paper. We could not find a readable treatment of this in the existing literature, and we hope this can be a useful reference for others.

Theorem 1.7, in which we identify non-locally finite characteristic covers of infinite-type surfaces, is proved by exhausting by finite-type surfaces and using Theorem 1.5, in a way analogous to the infinite-type case of Theorem 1.2.

1.2. Outline of the Paper

In §2 we review the necessary background. We define the surfaces mentioned in the introduction, set conventions, and review actions on trees, ends, Stallings’s Theorem, and the classification of infinite-type surfaces. In §3 we study characteristic covers of surfaces, proving Theorems 1.5, 1.7, 1.2 and 1.3 in subsections 3.1, 3.2, 3.3 and 3.4, respectively. In §4 we prove Theorem 1.1, and then §A is our appendix on ends via group cohomology.

1.3. Acknowledgements

IB was partially supported by NSF CAREER award 1654114. TC was partially supported by MSCA grant 101107744–DefHyp. NGV was partially supported by NSF award 2212922 and PSC-CUNY awards 66435-00 54 and 67380-00 55. JT was partially supported by NSF award 2304920.

2. Background

2.1. Conventions

All surfaces within this work are assumed to be connected, second countable, orientable, and possibly with boundary. We say the surface is borderless if it has no boundary, and remind the reader that some of the surfaces which appear in our proofs may have non-empty boundary with possibly non-compact boundary components. All borderless surfaces, whether finite or infinite type, are classified by their genus and the topology of their space of ends, see Theorem 2.2. We let Sg,pbS^{b}_{g,p} denote the finite-type surface with bb boundary components, gg genus, and pp punctures.

There are a handful of special infinite-type surfaces that occur in our main theorem statements:

  1. (1)

    The flute surface is homeomorphic to 22\mathbb{R}^{2}\smallsetminus\mathbb{Z}^{2}.

    [Uncaptioned image]
  2. (2)

    The Loch Ness monster surface is the unique borderless orientable infinite-type surface with one end.

    [Uncaptioned image]
  3. (3)

    The spotted Loch Ness monster surface is obtained from the Loch Ness monster surface by removing an infinite, discrete set of points.

    [Uncaptioned image]
  4. (4)

    The Cantor tree surface is homeomorphic to S2𝒞S^{2}\smallsetminus\mathcal{C}, where 𝒞\mathcal{C} is an embedding of a Cantor set.

  5. (5)

    The blooming Cantor tree surface is the surface with infinitely many genus whose ends space, (S)\mathcal{E}(S), and ends space accumulated by genus, g(S)\mathcal{E}_{g}(S), satisfy g(S)=(S)𝒞\mathcal{E}_{g}(S)=\mathcal{E}(S)\cong\mathcal{C}.

    [Uncaptioned image][Uncaptioned image]

2.2. Ends of a topological space

Given a topological space XX, its space of ends is defined to be

(X)=limπ0(XK),\mathcal{E}(X)=\lim_{\longleftarrow}\pi_{0}(X\smallsetminus K),

where the inverse limit is taken over all compact subsets KXK\subset X.

For a finitely generated group GG, consider its Cayley graph Cay(G,𝒮)Cay(G,\mathcal{S}) with respect to a finite generating set 𝒮\mathcal{S}. We define the space of ends of GG by (G)=(Cay(G,𝒮))\mathcal{E}(G)=\mathcal{E}({Cay}(G,\mathcal{S})). The following proposition shows that (G)\mathcal{E}(G) is independent of the choice of a finite generating set of GG.

Proposition 2.1.

[5, Prop. 8.29] Given two proper geodesic metric spaces X1X_{1} and X2X_{2}, then any quasi-isometry f:X1X2f:X_{1}\to X_{2} induces a homeomorphism f:(X1)(X2)f_{\mathcal{E}}:\mathcal{E}(X_{1})\to\mathcal{E}(X_{2}).

2.3. Classification of surfaces

Let SS be an orientable surface with genus g(S){0,,}g(S)\in\{0,\ldots,\infty\} and end space (S)\mathcal{E}(S). Recall that (S)\mathcal{E}(S) is always compact, totally disconnected, and metrizable. We may take a neighborhood of an end to be an open subset UU of SS containing the end. An end is planar if it has a neighborhood that embeds into the plane, and accumulated by genus otherwise. Let g(S)(S)\mathcal{E}_{g}(S)\subset\mathcal{E}(S) be the closed subset consisting of all ends that are accumulated by genus. The following classification theorem was first proved by Kérekjarto, and the proof was later simplified by Richards, see [12].

Theorem 2.2.

Two borderless, orientable surfaces SS and TT are homeomorphic if and only if g(S)=g(T)g(S)=g(T) and there is a homeomorphism of pairs

((S),g(S))((T),g(T)).(\mathcal{E}(S),\mathcal{E}_{g}(S))\to(\mathcal{E}(T),\mathcal{E}_{g}(T)).

Moreover, for every n{0,,}n\in\{0,\ldots,\infty\} and for every compact, totally disconnected metrizable space XX, and closed subset CXC\subset X, there is an orientable surface SS such that g(S)=ng(S)=n and ((S),g(S))(\mathcal{E}(S),\mathcal{E}_{g}(S)) is homeomorphic to (X,C)(X,C).

Below, we’ll call the pair ((S),g(S))(\mathcal{E}(S),\mathcal{E}_{g}(S)) the genus-marked end space of SS.

2.4. Preliminaries on covers

By a cover of SS, we mean a space S~\tilde{S} equipped with a surjective map π:S~S\pi:\tilde{S}\to S so that every point pSp\in S has a neighborhood UU so that for each component UαU_{\alpha} of π1(U)\pi^{-1}(U), the restriction π|Uα:UαU\pi|_{U_{\alpha}}:U_{\alpha}\to U is a homeomorphism. We will occasionally use the word “cover” just to refer to the covering surface S~\tilde{S}. The Galois correspondence associates subgroups of π1(S)\pi_{1}(S) with (equivalence classes of) connected covers of SS.

Remark 2.3.

Suppose SS is a surface with non-abelian π1(S)\pi_{1}(S), and Hπ1(S)H\leq\pi_{1}(S) is a nontrivial, infinite-index, normal subgroup. Then the cover S~\tilde{S} corresponding to HH is an infinite-type surface.

We record a general proof that the cover of a surface corresponding to the fundamental group of an essential subsurface is homeomorphic to the interior of the subsurface. For us, a subsurface of a surface SS is always a properly embedded, closed submanifold of SS. Given a π1\pi_{1}-injective subsurface YY of a surface SS and a point xYx\in Y, we identify π1(Y,x)\pi_{1}(Y,x) with its image under the monomorphism π1(Y,x)π1(S,x)\pi_{1}(Y,x)\to\pi_{1}(S,x) induced by the inclusion YSY\hookrightarrow S. When necessary, we will keep track of base points, otherwise we omit them and simply write π1(Y)\pi_{1}(Y) and view it as a subgroup of π1(S)\pi_{1}(S).

Lemma 2.4.

Let YY be a π1\pi_{1}–injective subsurface of a borderless surface SS. Then the cover Y~\tilde{Y} of SS corresponding to π1(Y)\pi_{1}(Y) is homeomorphic to the interior of YY.

Proof.

Let π:Y~S\pi:\tilde{Y}\to S be the cover corresponding to π1(Y)\pi_{1}(Y) and observe that the inclusion YSY\to S lifts to an embedding YY~Y\to\tilde{Y}, which induces an isomorphism π1(Y)π1(Y~)\pi_{1}(Y)\to\pi_{1}(\tilde{Y}). Therefore, any loop in Y~\tilde{Y} is a homotopic to one in the interior of YY. This implies that each boundary component of YY is separating—otherwise, there is an essential loop of Y~\tilde{Y} which essentially intersects a component of Y\partial Y. Further, each component of Y~Y\tilde{Y}\smallsetminus Y is topologically a product. This shows Y~Y(Y×[0,))\tilde{Y}\cong Y\cup(\partial Y\times[0,\infty)). ∎

2.5. Group actions on trees

Let TT be a simplicial tree, and let GG be a group acting on TT. By convention, we assume all such group actions are by simplicial isomorphisms and without edge inversions. Here, an edge inversion is a tree automorphism that leaves invariant some edge but reverses its orientation. An element gGg\in G acts elliptically on TT if gg fixes a vertex of TT, and hyperbolically otherwise, in which case it translates along a geodesic axis. We direct the reader to Serre’s book on Trees [13] for more information about group actions on trees.

2.5.1. Stallings’s Theorem and its relative version

Theorem 2.5 (Stallings’s Theorem, [15]).

Suppose QQ is a finitely generated group with more than one end. Then QQ admits an action on a simplicial tree TT with finite edge stablizers, no edge inversions, and no global fixed point.

We also need the following relative version of Stallings’s Theorem.

Theorem 2.6 (Relative Stallings’s Theorem).

Let QQ be a finitely generated group and let H1,,HmH_{1},\ldots,H_{m} be finitely generated subgroups of QQ. Suppose there is a nonconstant, continuous function f:(Q)f:\mathcal{E}(Q)\to\mathbb{Z} that is constant on the boundary Hig¯(Q)\overline{H_{i}g}\cap\mathcal{E}(Q) of each right coset HigH_{i}g, where i=1,,mi=1,\ldots,m and gQg\in Q. Then QQ admits an action on a simplicial tree TT with finite edge stabilizers, no inversions, no global fixed point, and where for each ii, the action of HiH_{i} on TT has a global fixed point.

This is really a result of Swarup [16], but Swarup’s result is stated in terms of group cohomology with group ring coefficients. In Appendix A, we give a gentle introduction to ends via group cohomology and show how the formulation above follows from Swarup’s result.

The intuition behind the relative version is as follows. Let {\sim} be the smallest equivalence relation on (Q)\mathcal{E}(Q) such that

  1. (1)

    for every right coset HjgH_{j}g, all points in Hjg¯(Q)\overline{H_{j}g}\cap\mathcal{E}(Q) are equivalent,

  2. (2)

    {\sim}–equivalence classes are closed in (G)\mathcal{E}(G).

Taking the quotient (Q)/\mathcal{E}(Q)/{{\sim}}, we are continuously collapsing the ends of all these cosets. The hypothesis in Theorem 2.6 is that (G)/\mathcal{E}(G)/{\sim} has more than one element, which intuitively means that QQ has more than one end even up to collapsing cosets of the HiH_{i}. So, there should be an action of QQ on a tree TT, as in Stallings’s Theorem, in which the HiH_{i} are not really used.

In this paper, we use Theorem 2.6 in the following rather explicit context. The group acting will be the deck group QQ of a regular cover S~S\tilde{S}\to S, where SS is a compact surface with boundary satifying ~(Q)~(S~)\tilde{\mathcal{E}}(Q)\cong\tilde{\mathcal{E}}(\tilde{S}). The subgroups Hi<QH_{i}<Q will be stabilizers of noncompact boundary components of S~\tilde{S}. Deleting these boundary components creates a new surface S~\tilde{S}_{\circ}, in which the two ends of each such boundary component become a single end of S~\tilde{S}_{\circ}. So, the end space (S~)\mathcal{E}(\tilde{S}_{\circ}) is the same as the quotient (Q)/\mathcal{E}(Q)/{\sim} discussed in the previous paragraph, and the hypothesis in Theorem 2.6 is that S~\tilde{S}_{\circ} has more than one end. In the paper, we phrase all of this in a way that appeals directly to the statement of Theorem 2.6, though, rather than talking about the relation {\sim} above.

In addition to Stalling’s Theorem and our formulation of its relative version, our main arguments in Section 3.1 uses Serre’s criterion to cook up a proof by contradiction to ensure the surface S~\tilde{S}_{\circ} is one-ended.

Lemma 2.7 (Serre’s Criterion, see Corollary 2 in §6.5 of [13]).

Let GG be a group acting on a tree TT. If GG admits a generating set {a1,,an}\{a_{1},\ldots,a_{n}\} such that aia_{i} and aiaja_{i}a_{j} act elliptically on TT for each distinct ii and jj, then the GG-action has a global fixed point.

3. Characteristic covers

In this section we study characteristic covers of surface, proving Theorems 1.5, 1.7, 1.2 and 1.3 in sections 3.1, 3.2, 3.3 and 3.4, respectively.

3.1. Finite-type surfaces

Here, we prove the following theorem from the introduction.

Theorem 1.5.

Any infinite-degree, geometrically characteristic cover of an orientable, borderless, finite-type surface is either a disc, the flute surface, the Loch Ness monster surface, or the spotted Loch Ness monster surface.

Throughout the section, we use the following notation: Given a surface SS, we let SS_{\circ} denote the surface obtained by removing every non-compact boundary component of SS. We’ll prove the following statement.

Proposition 3.1.

Let SS be a compact, orientable surface. If S~S\tilde{S}\to S is a geometrically characteristic cover of infinite degree, then S~\tilde{S}_{\circ} has one end.

Assuming Proposition 3.1, here is how to derive Theorem 1.5.

Proof of Theorem 1.5.

Note that any finite-type surface can be identified as the interior of some compact surface SS. Further, a geometrically characteristic cover of the interior of SS is the interior of a geometrically characteristic cover S~S\tilde{S}\to S. Thus, the goal is to show that the interior of S~\tilde{S} is one of the four surfaces listed.

Let QQ be the associated deck group of S~\tilde{S}. First assume that S~\tilde{S} has no compact boundary components, so int(S~)=S~\text{int}(\tilde{S})=\tilde{S}_{\circ}. As QQ is infinite and acts cocompactly on S~\tilde{S}, the genus of int(S~)\text{int}(\tilde{S}) is either 0 or infinity, so the classification of surfaces (Theorem 2.2) implies that it is either the disc or the Loch Ness monster surface. Now, suppose that S~\tilde{S} has a compact boundary component. As QQ is infinite, S~\tilde{S} has infinitely many compact boundary components. Gluing discs to each compact boundary component, we obtain a surface without boundary, call it YY. By Proposition 3.1, YY has a single end, and again the classification of surfaces (Theorem 2.2) implies that YY is either the disc or the Loch Ness monster surface. To finish, observe that int(S~)\text{int}(\tilde{S}) is obtained from YY by puncturing all the infinitely many discs, so then int(S~)\text{int}(\tilde{S}) is either the flute surface or the spotted Loch Ness monster surface. ∎

In the proof of Proposition 3.1, we will start by assuming S~\tilde{S}_{\circ} has more than one end. We then use a relative version of Stallings’ Theorem to produce an action on a tree TT, and decompose S~\tilde{S} equivariantly into subsurfaces essentially corresponding to the vertices of TT. We use the induced decomposition of SS to construct elements of π1(S)\pi_{1}(S) that act elliptically on TT, then use the fact that the cover is characteristic to construct many more elements that act elliptically, in fact so many that the action on TT must have a global fixed point, a contradiction.

For clarity of exposition, though, before starting the proof properly in §3.1.1, we isolate some of its components into a series of preliminary results. First, the following will allow us to construct the action on the tree referenced above.

Lemma 3.2 (A peripherally elliptic action).

Let π:S~S\pi:\tilde{S}\to S be a regular cover of a compact surface SS. If S~\tilde{S}_{\circ} has more than one end, then there is an action of the deck group QQ on a tree TT with no edge inversions and no global fixed point, such that the projection to QQ of every peripheral element of π1(S)\pi_{1}(S) acts elliptically on TT.

Proof.

Let ι¯:(S~)(S~)\bar{\iota}:\mathcal{E}(\tilde{S})\to\mathcal{E}(\tilde{S}_{\circ}) be defined as follows: given ξ(S~)\xi\in\mathcal{E}(\tilde{S}), take a sequence of points xiS~0S~x_{i}\in\tilde{S}_{0}\subset\tilde{S} with xiξx_{i}\to\xi, and define ι¯(ξ)\bar{\iota}(\xi) to be the limit in (S~)\mathcal{E}(\tilde{S}_{\circ}). It is readily verified that ι¯\bar{\iota} is a well-defined continuous surjection.

As the action of QQ on S~\tilde{S} is co-compact and proper, the end space of QQ, denoted (Q)\mathcal{E}(Q), is homeomorphic to (S~)\mathcal{E}(\tilde{S}). If S~\tilde{S}_{\circ} has more than one end, then there is a non-constant continuous function ϕ:(S~)\phi:\mathcal{E}(\tilde{S}_{\circ})\to\mathbb{Z}. Consider the composition

(Q)(S~)ι¯(S~)ϕ.\mathcal{E}(Q)\overset{\cong}{\to}\mathcal{E}(\tilde{S})\overset{\bar{\iota}}{\to}\mathcal{E}(\tilde{S}_{\circ})\overset{\phi}{\to}\mathbb{Z}.

Pick cyclic subgroups of π1(S)\pi_{1}(S) representing the boundary components, and let HjH_{j} denote their projections to QQ. For any right coset HjgH_{j}g, the intersection Hjg¯(Q)\overline{H_{j}g}\cap\mathcal{E}(Q) is either empty (if HjH_{j} is finite, which happens when that boundary component lifts to a closed curve in S~\tilde{S}) or projects to the endpoints in (S~)\mathcal{E}(\tilde{S}) of some non-compact boundary component of S~\tilde{S}, both of which (if they are distinct) map to the same element of (S~)\mathcal{E}(\tilde{S}_{\circ}).

We have established that ϕ\phi is non-constant but that it is constant on Hjg¯(Q)\overline{H_{j}g}\cap\mathcal{E}(Q) for each jj and each right coset HjgH_{j}g. Therefore, by Theorem 2.6 there exists an action of QQ on a tree TT, with no edge inversions and finite edge stabilizers, and no global fixed point, such that all the subgroups HjH_{j} act elliptically. Every peripheral element of π1(S)\pi_{1}(S) projects to an element of QQ that is conjugate into one of the HjH_{j}’s and hence acts elliptically on TT as desired. ∎

Given a peripherally elliptic action of QQ on a tree TT, the following proposition constructs a QQ-equivariant collection of simple closed curves in S~\tilde{S} corresponding (perhaps multiple-to-one) to the edges of TT. Cutting along these curves decomposes S~\tilde{S} into subsurfaces corresponding to the vertices of TT.

Proposition 3.3.

Let π:S~S\pi:\tilde{S}\to S be a regular cover of a compact surface SS with deck group QQ. Suppose QQ acts on a tree TT with finite edge stabilizers, no edge inversions, no global fixed point, and such that the projection to QQ of every peripheral element of π1(S)\pi_{1}(S) acts elliptically on TT. Then there exists a QQ-equivariant map f~:S~T\tilde{f}:\tilde{S}\to T such that each component of S~\partial\tilde{S} maps to a vertex under f~\tilde{f}. Moreover, if ZZ is the set of edge midpoints in TT, then the preimage C~:=f~1(Z)\tilde{C}:=\tilde{f}^{-1}(Z) is a properly embedded 11-submanifold of S~\tilde{S}, each component of which is a simple closed curve, and its projection C:=π(C~)C:=\pi(\tilde{C}) is a compact 11-submanifold of SS each component of which is non-peripheral and essential.

Above, there may be components of CC that are parallel, which is why we call CC a 11-submanifold and not a multicurve.

Proof.

We begin by constructing a prototype for the map f~\tilde{f}, which we call f~1\tilde{f}_{1}. Following Shalen [14], we construct the QQ-equivariant map f~1:S~T\tilde{f}_{1}\mskip 0.5mu\colon\thinspace\tilde{S}\to T as follows. Fix a triangulation τ\tau of SS, and consider its associated QQ-invariant triangulation τ~\tilde{\tau} of S~\tilde{S}. Let X~0T(0)\tilde{X}_{0}\subset T^{(0)} contain a single representative for each orbit equivalence class of τ~(0)\tilde{\tau}^{(0)}, and pick a map

h~(0):X~0T(0)\tilde{h}^{(0)}\mskip 0.5mu\colon\thinspace\tilde{X}_{0}\to T^{(0)}

that is arbitrary, except that for every component βS~\beta\subset\partial\tilde{S}, we require that h~(0)\tilde{h}^{(0)} takes X~0β\tilde{X}_{0}\cap\beta to a single vertex of TT that is fixed by the stabilizer QβQQ_{\beta}\subset Q; this can be done as QβQ_{\beta} is the projection to QQ of a peripheral \mathbb{Z}-subgroup of π1(S)\pi_{1}(S), so acts with a global fixed point on TT by assumption. (Note that QβQ_{\beta} must fix a vertex, since it acts simplicially without edge inversions.)

Refer to caption
Figure 1. The map f~1\tilde{f}_{1} from Proposition 3.3 restricted to a triangle in the triangulation τ~\tilde{\tau}; the green, blue, and red regions map onto the green, blue, and red segments, respectively, in the tripod; the restriction to any one of the six subtriangles is an orthogonal projection map. The preimage of an edge midpoint zTz\in T is shown in orange.

There is a unique QQ-equivariant extension of h~(0)\tilde{h}^{(0)} to a map

f~1(0):τ(0)T(0).\tilde{f}_{1}^{(0)}\mskip 0.5mu\colon\thinspace\tau^{(0)}\to T^{(0)}.

The map f~1(0)\tilde{f}_{1}^{(0)} is readily extended to a piecewise linear map f~1:S~T\tilde{f}_{1}\mskip 0.5mu\colon\thinspace\tilde{S}\to T (see Figure 1). Observe that, under this piecewise linear extension, each boundary component β\beta of S~\partial\tilde{S} is mapped to a point by f~1\tilde{f}_{1}, as its vertices are all mapped to the same point by construction.

Let ZZ be the set of all edge midpoints of TT, let C~1=f~11(Z)\tilde{C}_{1}=\tilde{f}_{1}^{-1}(Z), and let C1=π(C1)C_{1}=\pi(C_{1}). We claim that each component C~1\tilde{C}_{1} (resp., C1C_{1}) is a simple closed curve. By the construction of f~1\tilde{f}_{1}, we know that the components of C~1\tilde{C}_{1} form a locally finite family of pairwise-disjoint 1-submanifolds of S~\tilde{S}. As π\pi is a local homeomorphism and C~1\tilde{C}_{1} is invariant under the action of QQ, we have that C1C_{1} is a compact 1-submanifold of SS; in particular, each component of C1C_{1} is a simple closed curve. Now, let c~\tilde{c} be a component of C~1\tilde{C}_{1}. Then π(c~)\pi(\tilde{c}) is compact, and as the action of QQ on TT has finite edge stabilizers, the stabilizer of c~\tilde{c} is finite, implying that c~\tilde{c} is compact. Therefore, C~1\tilde{C}_{1} is a union of simple closed curves. We now need to edit f~1\tilde{f}_{1} so that no component of C~1\tilde{C}_{1} nor C1C_{1} is peripheral or bounds a disk.

First, suppose that there is a component of C1C_{1} that bounds a disk. Let Δ1,,Δk\Delta_{1},\ldots,\Delta_{k} be the outermost disks bounded by components of C1C_{1}, so that if cc is a component of C1C_{1} bounding a disk, then there exists a unique j{1,,k}j\in\{1,\ldots,k\} such that cΔjc\subset\Delta_{j}. Choose a lift Δ~j\tilde{\Delta}_{j} in S~\tilde{S} for each of the Δj\Delta_{j}, and for qQq\in Q, let Δ~jq=qΔ~j\tilde{\Delta}_{j}^{q}=q\cdot\tilde{\Delta}_{j}. Let ejqe_{j}^{q} be the edge with vertices vjq,wjqTv_{j}^{q},w_{j}^{q}\in T whose midpoint is f1~(~Δjq)\tilde{f_{1}}(\tilde{\partial}\Delta_{j}^{q}). By the construction of C1C_{1}, all the 11-cells in τ~\tilde{\tau} that intersect ~Δjq\tilde{\partial}\Delta_{j}^{q} are mapped to segments in TT that all contain the edge ejqe_{j}^{q}. If the 0-cells in Δ~jq\tilde{\Delta}_{j}^{q} are mapped into the vjqv_{j}^{q} side of T{ejq}T\smallsetminus\{e_{j}^{q}\}, we can define a new map

f~2(0):τ~(0)T(0)\tilde{f}_{2}^{(0)}:\tilde{\tau}^{(0)}\to T^{(0)}

that maps the 0-cells in Δjq\Delta_{j}^{q} to wjqw_{j}^{q}, but agrees with f~1(0)\tilde{f}_{1}^{(0)} otherwise. We can now extend f~2(0)\tilde{f}_{2}^{(0)} to a QQ-invariant piecewise linear map f~2:S~T\tilde{f}_{2}\mskip 0.5mu\colon\thinspace\tilde{S}\to T in such a way that f~21(z)=f~11(z)\tilde{f}_{2}^{-1}(z)=\tilde{f}_{1}^{-1}(z) for any zZ(Δ~jq)z\in Z\smallsetminus(\bigcup\tilde{\Delta}_{j}^{q}). Setting C~2=f~21(Z)\tilde{C}_{2}=\tilde{f}_{2}^{-1}(Z) and C2=π(C~2)C_{2}=\pi(\tilde{C}_{2}), we have that C2C1C_{2}\subset C_{1} and no component of C2C_{2} bounds a disk.

Now, suppose that C2C_{2} contains a peripheral component. Then each peripheral component of C2C_{2} bounds an annulus with a component of S\partial S. Let A1,,AmA_{1},\ldots,A_{m} be the outermost annuli, so that if cc is a component of C2C_{2} bounding an annulus with a component of S\partial S, then there exists a unique j{1,,m}j\in\{1,\ldots,m\} such that cAjc\subset A_{j}. Each lift of an AjA_{j} is again an annulus in S~\tilde{S} as the component of C~2\tilde{C}_{2} are compact. We now proceed identically as we did above with the disks Δj\Delta_{j} replaced with the annuli AjA_{j}. The result is a QQ-equivariant map f~:S~T\tilde{f}\mskip 0.5mu\colon\thinspace\tilde{S}\to T so that C~:=f~1(Z)\tilde{C}:=\tilde{f}^{-1}(Z) and C:=π(C~)C:=\pi(\tilde{C}) have neither peripheral nor inessential components. ∎

Combining Lemma 3.2 and Proposition 3.3, we obtain:

Corollary 3.4.

Let π:S~S\pi:\tilde{S}\to S be a regular cover of a compact surface. If S~\tilde{S}_{\circ} has more than one end, then there is a non-peripheral, essential, simple closed curve γ\gamma on SS such that every component of π1(γ)\pi^{-1}(\gamma) is compact.

Proof.

Since the action of the deck group QQ produced by Lemma 3.2 has no global fixed point, the image of the map in Proposition 3.3 has to contain an edge midpoint, the 11-submanifold CSC\subset S it produces is non-empty. Take γ\gamma to be any component of CC. Any component of π1(γ)\pi^{-1}(\gamma) is a component of C~\tilde{C}, and hence is compact. ∎

Given a properly embedded 11-submanifold CC on a surface SS, we write SCS\mathbin{{\smallsetminus}\mspace{-5.0mu}{\smallsetminus}}C to refer to the surface obtained by cutting SS along CC, that is, the surface obtained by removing an open regular neighborhood of CC from SS.

Lemma 3.5 (Cutting covers into one-ended pieces).

If π:S~S\pi:\tilde{S}\to S is a regular cover of a compact surface and CSC\subset S is a compact 11-submanifold such that every component of π1(C)S~\pi^{-1}(C)\subset\tilde{S} is compact. Then there exists a compact 11-submanifold DSD\subset S containing CC such that

  • every component of π1(D)\pi^{-1}(D) is compact, and

  • every component of S~π1(D)\tilde{S}_{\circ}\mathbin{{\smallsetminus}\mspace{-5.0mu}{\smallsetminus}}\pi^{-1}(D) has at most one end.

Proof.

It suffices to show that if some component of S~π1(C)\tilde{S}_{\circ}\mathbin{{\smallsetminus}\mspace{-5.0mu}{\smallsetminus}}\pi^{-1}(C) has more than one end, then there is an essential, nonperipheral simple closed curve γSC\gamma\subset S\mathbin{{\smallsetminus}\mspace{-5.0mu}{\smallsetminus}}C such that every component of π1(γ)S~\pi^{-1}(\gamma)\subset\tilde{S} is compact. For then we can add such curves recursively to CC to produce DD as required. Since the number of isotopy classes in a collection of pairwise disjoint curves on SS is bounded, this process terminates.

So, let Σ\Sigma be a component of SCS\mathbin{{\smallsetminus}\mspace{-5.0mu}{\smallsetminus}}C, and assume that we have a component Σ~S~C~\tilde{\Sigma}\subset\tilde{S}\mathbin{{\smallsetminus}\mspace{-5.0mu}{\smallsetminus}}\tilde{C} such that Σ~\tilde{\Sigma}_{\circ} has more than one end. We claim that the restriction of π\pi to Σ~\tilde{\Sigma} is a regular covering map Σ~Σ.\tilde{\Sigma}\to\Sigma. Indeed, if two points p,qΣ~p,q\in\tilde{\Sigma} have the same projection to Σ\Sigma, then there is a deck transformation gg of S~S\tilde{S}\to S with g(p)=qg(p)=q, and since C~\tilde{C} is invariant under the deck group, we have g(Σ~)=Σ~g(\tilde{\Sigma})=\tilde{\Sigma}, implying gg restricts to a deck transformation of Σ~Σ\tilde{\Sigma}\to\Sigma taking pp to qq.

Corollary 3.4 then says that there is some essential, non-peripheral curve in γΣ\gamma\subset\Sigma such that every component of π|Σ~1(γ)\pi|_{\tilde{\Sigma}}^{-1}(\gamma) is compact. But since the cover π:S~S\pi:\tilde{S}\to S is regular, this implies that every component of π1(γ)S~\pi^{-1}(\gamma)\subset\tilde{S} is compact, so we are done. ∎

3.1.1. Proof of Proposition 3.1

We argue by way of contradiction: assume that S~\tilde{S}_{\circ} has more than one end. Let QQ be the deck group associated to the covering π\pi. Apply Proposition 3.3 to obtain a tree TT, a peripherally elliptic action of QQ on TT with no global fixed point, and the 1-submanifolds C~\tilde{C} and CC of S~\tilde{S} and SS, respectively, and apply Lemma 3.5 to get the 11-submanifold DD on SS. Recall that the action of QQ on TT has finite edge stabilizers, no edge inversions, no global fixed point, and every element of QQ that is the image of a peripheral element in π1(S)\pi_{1}(S) acts elliptically on TT. The contradiction will arise by showing that the action of QQ has a global fixed point.

We say that a closed curve γ\gamma is at most one ended in S~\tilde{S}_{\circ} if for some (equivalently, any) component γ~\tilde{\gamma} of π1(γ)\pi^{-1}(\gamma), we have that γ~\tilde{\gamma} is either compact or it is non-compact and its two ends go out the same end of S~\tilde{S}_{\circ}. In particular, any closed curve disjoint from DD is at most one ended in S~\tilde{S}_{\circ}.

We say two closed curves γ1\gamma_{1} and γ2\gamma_{2} in SS have the same homeomorphism type if there exists a homeomorphism h:SSh\mskip 0.5mu\colon\thinspace S\to S such h(γ1)=γ2h(\gamma_{1})=\gamma_{2}. As π\pi is geometrically characteristic, every homeomorphism h:SSh\mskip 0.5mu\colon\thinspace S\to S lifts to a homeomorphism h:S~S~h\mskip 0.5mu\colon\thinspace\tilde{S}\to\tilde{S}. Also note that the restriction of any homeomorphism S~S~\tilde{S}\to\tilde{S} to S~\tilde{S}_{\circ} induces a homeomorphism S~S~\tilde{S}_{\circ}\to\tilde{S}_{\circ}. Therefore, if γ\gamma is a closed curve that has the same homeomorphism type as a closed curve disjoint from DD, then γ\gamma is at most one ended in S~\tilde{S}_{\circ}.

Now, we claim that if γ\gamma is an essential closed curve on SS that is at most one ended, then γ\gamma acts elliptically on TT, i.e., any element of π1(S)\pi_{1}(S) in the free homotopy class of γ\gamma acts elliptically on TT, where we let π1(S)\pi_{1}(S) act on TT via the action factoring through QQ.

To see this, pick some gQg\in Q that is the projection of an element of π1(S)\pi_{1}(S) in the free homotopy class of γ\gamma, and hoping for a contradiction, suppose that gg acts hyperbolically on TT. Now gg stabilizes some component γ~π1(γ)\tilde{\gamma}\subset\pi^{-1}(\gamma). Since gg acts hyperbolically on TT, the image f~(γ~)\tilde{f}(\tilde{\gamma}) is a proper bi-infinite path in TT, whose ends go out distinct ends of TT. Pick some edge midpoint mTm\in T that separates those two ends, and let M=f~1(m).M=\tilde{f}^{-1}(m). Then MM is a union of components of C~\tilde{C}.

We claim that MM has finitely many components. If not, then since the projection CC has only finitely many components, there are infinitely many elements of QQ that send a component of MM to a component of MM, and hence mTm\in T has infinite stabilizer, contradicting our assumption that QQ acts with finite edge stabilizers. Hence, MM has finitely many components.

The two ends of the path γ~S~\tilde{\gamma}\subset\tilde{S} lie in different components of S~M\tilde{S}\smallsetminus M, since they map to different components of TmT\smallsetminus m. Since MM is compact, this means that the two ends of γ~\tilde{\gamma} converge to distinct ends of S~\tilde{S}_{\circ}, a contradiction. This establishes the claim that γ\gamma acts elliptically on TT.

Thus far, we have established that any essential closed curve on SS that has the same homeomorphism type as a curve disjoint from DD acts elliptically on TT. In the rest of the proof, using this plethora of elliptic elements, we apply Serre’s Criterion (Lemma 2.7) to show that the action QTQ\curvearrowright T has a global fixed point.

First, let us assume that there is no non-separating simple closed curve on SS that is disjoint from DD. Then the components of SDS\smallsetminus D are all planar and are connected together in the pattern of a tree, since if there is a cycle in the component graph, you can realize it as a simple closed curve on SS intersecting some component of DD once, and then that component of DD is non-separating, contradicting that there are no non-separating curves disjoint from DD. It follows that SS is planar, and hence its fundamental group can be generated by a set of elements representing all but one of the boundary components of SS. By the construction of the action of QQ on TT, all such curves act elliptically on TT.

A boundary product is a simple closed curve on SS that bounds a pair of pants with two boundary components of SS. Note that all boundary products in SS differ by homeomorphisms of SS. If we choose orientations for the generators above correctly, all products of pairs of distinct generators are homotopic to boundary products. Looking within a component of SDS\smallsetminus D that is a leaf of the associated component tree, we can find a boundary product disjoint from DD, so it follows from the above that all boundary products act elliptically on TT, and hence the action has a global fixed point by Serre’s Criterion (Lemma 2.7), a contradiction.

Now assume there is a non-separating simple closed curve disjoint from DD. Then all non-separating curves on SS act elliptically on TT. Consider a ‘standard’ generating set for π1(S)\pi_{1}(S) consisting of two non-separating curves intersecting once for each genus, plus generators for all but one peripheral curve (see Figure 2). All these generators act elliptically on TT. A product of a pair of generators is freely homotopic to either a boundary product or a non-separating simple closed curve. To use Serre’s Criterion again to get a contradiction, it suffices to prove that all boundary products on SS act elliptically on TT.

Refer to caption
Figure 2. Standard generators for π1(S)\pi_{1}(S) for SS on a genus two surface with three punctures are given by the elements a1,,a6a_{1},\ldots,a_{6}.
Refer to caption
Figure 3. A generating set for a thrice-punctured torus consisting of non-separating curves, namely a1,a2,a3,a4a_{1},a_{2},a_{3},a_{4}, where products of pairs of distinct generators are either non-separating, or are freely homotopic to δ\delta. For a similar example on a twice-punctured torus, just delete one of the two boundary components that is not δ\delta.

Fix a boundary product γ\gamma. Recall that this means that γ\gamma bounds a pair of pants in SS. Call δ\delta and δ\delta^{\prime} the two peripheral curves of this pair of pants. Since SS has a non-separating curve, it has positive genus. It follows that there is an essential genus one subsurface XSX\subset S containing γ\gamma that has either two or three boundary components. Indeed, if SS has genus one and only two boundary components, then set X=SX=S. Otherwise, we can find a separating curve in SS that cuts off a subsurface XX of genus one containing γ\gamma, and XX only has three boundary components. In either case, both δ\delta and δ\delta^{\prime} are peripheral curves of XX.

Given such an XX, we can find a generating set for π1(X)\pi_{1}(X) consisting of non-separating curves, such that every product of distinct pairs of generators is either non-separating or is freely homotopic to δ\delta. See Figure 3. Since non-separating curves and peripheral curves in SS act elliptically on TT, by Serre’s criterion we have that the action π1XT\pi_{1}X\curvearrowright T has a global fixed point. In particular, γ\gamma acts elliptically on TT. Since γ\gamma was an arbitrary boundary product, all boundary products act elliptically on TT.

We have now established that the action of QQ on TT has a global fixed point, a contradiction. This finishes the proof of Proposition 3.1.

3.2. Non-locally finite covers of infinite-type surfaces

Recall that a subgroup NN of a group GG is locally finite index if NHN\cap H is finite index in HH whenever H<GH<G is finitely generated; a cover π:S~S\pi\mskip 0.5mu\colon\thinspace\tilde{S}\to S is locally finite if π(π1S~)\pi_{*}(\pi_{1}\tilde{S}) is locally finite index in π1(S)\pi_{1}(S). In this section we prove:

Theorem 1.7.

Let SS be an orientable, infinite-type borderless surface. If S~S\tilde{S}\to S is a characteristic cover that is not locally finite, then S~\tilde{S} is homeomorphic to either a disc, the flute surface, the Loch Ness monster surface, or the spotted Loch Ness monster surface.

We first start with some terminology and three lemmas. Recall that a subgroup AA of a group GG is a free factor if there exists another subgroup BB of GG such that GG is the internal free product of AA and BB. The three lemmas below tell us the following: given a non-locally finite characteristic cover π:S~S\pi\mskip 0.5mu\colon\thinspace\tilde{S}\to S and a “large” compact subsurface YY of SS, the restriction of π\pi to a component Y~\tilde{Y} of π1(Y)\pi^{-1}(Y) is an infinite-sheeted characteristic cover Y~Y\tilde{Y}\to Y.

Lemma 3.6.

Let SS be a noncompact surface, and let YY be a compact subsurface of SS. If each component of Y\partial Y is a separating curve and each component of Sint(Y)S\smallsetminus int(Y) is noncompact, then π1(Y)\pi_{1}(Y) is a free factor of π1(S)\pi_{1}(S).

Proof.

We observe that if α\alpha is a separating curve in SS, such that both sides of AA and BB of SαS\smallsetminus\alpha are not precompact, then α\alpha can be extended to a free basis for π1(A)\pi_{1}(A) and π1(B)\pi_{1}(B). By Van Kampen’s theorem, π1(S)=π1(A)π1(B)\pi_{1}(S)=\pi_{1}(A)*\pi_{1}(B). It is not hard to see from this that any basis for π1(Y)\pi_{1}(Y) can be extended to a basis for π1(S)\pi_{1}(S). ∎

Lemma 3.7.

Let SS be a noncompact surface, and let π:S~S\pi\mskip 0.5mu\colon\thinspace\tilde{S}\to S be a non-locally finite characteristic cover. Then there exists R>0R>0 such that if YSY\subset S is an π1\pi_{1}-injective subsurface whose fundamental group contains a free factor of π1(S)\pi_{1}(S) of rank RR, then each component of π1(Y)\pi^{-1}(Y) is not compact.

Proof.

Let N=π(π1(S~))N=\pi_{*}(\pi_{1}(\tilde{S})). As NN fails to be locally finite index in π1(S)\pi_{1}(S), there exists a finitely generated group H<π1(S)H<\pi_{1}(S) such that [H:HN][H:H\cap N] is infinite. Using the fact that HH is finitely generated, we can find a free factor AA of π1(S)\pi_{1}(S) containing HH; let RR denote the rank of AA. Let YSY\subset S be a π1\pi_{1}-injective subsurface so that π1(Y)\pi_{1}(Y) contains a free factor BB of π1(S)\pi_{1}(S) of rank RR. Then there exists an automorphism φ\varphi of π1(S)\pi_{1}(S) such that φ(A)=B\varphi(A)=B. As NN is characteristic, φ(N)=N\varphi(N)=N, implying that

[φ(H):φ(H)N]=[φ(H):φ(HN)]=[H:HN].[\varphi(H):\varphi(H)\cap N]=[\varphi(H):\varphi(H\cap N)]=[H:H\cap N].

In particular, as φ(H)<π1(Y)\varphi(H)<\pi_{1}(Y), we have that the index of π1(Y)N\pi_{1}(Y)\cap N in π1(Y)\pi_{1}(Y) is infinite. Given a component Y~\tilde{Y} of π1(Y)\pi^{-1}(Y), the restriction of π\pi to Y~\tilde{Y} is a covering Y~Y\tilde{Y}\to Y with corresponding subgroup π1(Y)N\pi_{1}(Y)\cap N, and hence Y~\tilde{Y} is an infinite-sheeted cover of YY, implying it is noncompact. ∎

Lemma 3.8.

Let SS be a noncompact surface, and let YSY\subset S be a compact subsurface such that each component of Y\partial Y is a separating curve and each component of Sint(Y)S\smallsetminus int(Y) is noncompact. If N<π1(S)N<\pi_{1}(S) is characteristic, then Nπ1(Y)N\cap\pi_{1}(Y) is a characteristic subgroup of π1(Y)\pi_{1}(Y).

Proof.

Let φAut(π1(Y))\varphi\in\mathrm{Aut}(\pi_{1}(Y)). By Lemma 3.6, π1(Y)\pi_{1}(Y) is a free factor of π1(S)\pi_{1}(S), and so there exists B<π1(S)B<\pi_{1}(S) such that π1(S)\pi_{1}(S) is isomorphic to π1(Y)B\pi_{1}(Y)*B. We can therefore extend φ\varphi to an automorphism φ^\hat{\varphi} of π1(S)\pi_{1}(S) by declaring φ^(b)=b\hat{\varphi}(b)=b for all bBb\in B and φ^|π1(Y)=φ\hat{\varphi}|_{\pi_{1}(Y)}=\varphi. So,

φ(Nπ1(Y))\displaystyle\varphi(N\cap\pi_{1}(Y)) =φ^(Nπ1(Y))\displaystyle=\hat{\varphi}(N\cap\pi_{1}(Y))
=φ^(N)φ^(π1(Y))\displaystyle=\hat{\varphi}(N)\cap\hat{\varphi}(\pi_{1}(Y))
=Nπ1(Y).\displaystyle=N\cap\pi_{1}(Y).

Hence, Nπ1(Y)N\cap\pi_{1}(Y) is characteristic in π1(Y)\pi_{1}(Y). ∎

We are now ready to prove Theorem 1.7.

Proof of Theorem 1.7.

Let SS be an infinite-type surface, and let S¯\bar{S} denote the surface obtained by removing an open annular neighborhood of each isolated planar end of SS. That is, S¯\bar{S} is obtained by replacing the isolated planar ends of SS with compact boundary components.

Let π:S~S¯\pi\mskip 0.5mu\colon\thinspace\tilde{S}\to\bar{S} be an infinite-sheeted characteristic cover. We will prove that S~\tilde{S}_{\circ} is one ended (recall that S~\tilde{S}_{\circ} is obtained by deleting the non-compact boundary components of S~\tilde{S}). Note that we are actually interested in the topology of int(S~)int(\tilde{S}); however, if S~\tilde{S}_{\circ} is one ended, then the result is a consequence of the classification of surfaces.

Let KS~K\subset\tilde{S}_{\circ} be a compact subset. By possibly enlarging KK, we can assume that no component of S~K\tilde{S}_{\circ}\smallsetminus K is precompact. Our goal is to show S~K\tilde{S}_{\circ}\smallsetminus K is connected.

Let RR be the constant of Lemma 3.7. We can exhaust S¯\bar{S} by compact subsurfaces whose boundary components are separating and whose complementary components are not precompact. Thus, we may find two nested compact subsurfaces YXS¯Y\subset X\subset\bar{S} with the following properties.

  • The interior of YY contains π(K)\pi(K).

  • Each boundary component of XX is separating, and each component of S¯int(X)\bar{S}\smallsetminus int(X) is noncompact,

  • Each component of Xint(Y)X\smallsetminus int(Y) has a fundamental group containing a free factor of rank RR.

Now let Y~\tilde{Y} be the component of π1(Y)\pi^{-1}(Y) containing KK, and let X~Y~\tilde{X}\supset\tilde{Y} be the associated component of π1(X)\pi^{-1}(X). By Lemma 3.8, the restriction π\pi to X~\tilde{X} gives a characteristic cover X~X\tilde{X}\to X, so it follows from Proposition 3.1 that X~\tilde{X}_{\circ} is one ended. In particular, X~K\tilde{X}_{\circ}\smallsetminus K has only one component which is not precompact. On the other hand, by Lemma 3.7, no component of X~Y~\tilde{X}\smallsetminus\tilde{Y} is precompact, so the same is true of X~K\tilde{X}_{\circ}\smallsetminus K. This shows the complement of KK in X~\tilde{X}_{\circ}, and hence in S~\tilde{S}_{\circ}, is connected. ∎

3.3. The universal abelian cover

If SS is an orientable borderless surface with fundamental group π1(S)\pi_{1}(S), one example of a characteristic cover of SS is the universal abelian cover SabS^{ab}, which is defined to be the cover corresponding to the commutator subgroup [π1(S),π1(S)][\pi_{1}(S),\pi_{1}(S)]. Unless SS is a disk or sphere, the universal abelian cover SabS^{ab} has infinite degree and is not locally finite, so Theorem 1.7 implies that SabS^{ab} is homeomorphic to either the disk, the flute surface, the Loch Ness monster surface, or the spotted Loch Ness monster surface. Given a surface SS, the next theorem tells you which of these surfaces is its universal abelian cover.

Theorem 1.2 (Universal abelian covers).

For a surface SS, let SabS^{ab} be its universal abelian cover. Then

  1. (1)

    if SS is 2\mathbb{R}^{2}, the annulus, or the torus, then Sab2S^{ab}\cong\mathbb{R}^{2},

  2. (2)

    if SS is the sphere, then so is SabS^{ab},

  3. (3)

    if SS is the once-punctured torus, then SabS^{ab} is the flute surface,

  4. (4)

    if SS is a finite-type surface with one puncture and genus at least two, then SabS^{ab} is the spotted Loch Ness monster surface,

  5. (5)

    otherwise, SabS^{ab} is the Loch Ness monster surface.

Proof.

Cases (1) and (2) are trivial, so let us assume from now on that SS has non-abelian fundamental group.

First, assume that SS is a finite-type genus gg surface with nn punctures. We can then realize SS as the interior of a compact surface S¯\bar{S} with nn boundary components. The abelianization of π1S¯\pi_{1}\bar{S} is 2g+n1\mathbb{Z}^{2g+n-1}, so 2g+n1\mathbb{Z}^{2g+n-1} acts properly and cocompactly on S¯ab\bar{S}^{ab}; hence, the end space of 2g+n1\mathbb{Z}^{2g+n-1} and the end space of S¯ab\bar{S}^{ab} are homeomorphic. As we are assuming that π1(S)\pi_{1}(S) is not abelian, we have that 2g+n122g+n-1\geq 2, so that 2g+n1\mathbb{Z}^{2g+n-1}, and hence S¯ab\bar{S}^{ab}, is one ended.

If n=0n=0, then S¯=S\bar{S}=S, and the classification of surfaces implies that SabS^{ab} is the Loch Ness monster surface. If n>1n>1, then the preimage of S¯\partial\bar{S} in S¯ab\bar{S}^{ab} is a collection of properly embedded lines. Deleting these lines cannot increase the number of ends, so SabS^{ab} is also one-ended. Hence, SabS^{ab} is the Loch Ness monster surface.

Now, suppose n=1n=1, which implies that g1g\geq 1. The single boundary component of S¯\bar{S} lifts homeomorphically, and so S¯ab\partial\bar{S}^{ab} is an infinite collection of circles. Gluing on a disk to each of these circles results in a borderless one-ended surface S~cap\tilde{S}_{cap} and a covering S~capScap\tilde{S}_{cap}\to S_{cap} with deck group 2g\mathbb{Z}^{2g}, where ScapS_{cap} is the closed surface obtained by gluing on a disk to S¯\bar{S}. By the classification of surfaces, S~cap\tilde{S}_{cap} is either the plane or the Loch Ness monster surface. If it is the plane, then this covering map is the universal covering map and the fundamental group of the surface is free abelian; hence, ScapS_{cap} is a torus, so SS is the once-punctured torus and its universal abelian cover is obtained by puncturing the plane in an infinite discrete set of points, implying SabS^{ab} is the flute surface. Otherwise, S~cap\tilde{S}_{cap} is the Loch Ness monster surface, and SabS^{ab} is obtained by puncturing the Loch Ness monster surface in an infinite discrete set of points, implying that SabS^{ab} is the spotted Loch Ness monster surface.

Finally, assume that SS is an infinite-type surface. The universal abelian cover SabSS^{ab}\to S is not locally finite, as the deck group is free abelian and hence not a torsion group. By Theorem 1.7 and the fact that π1(Sab)\pi_{1}(S^{ab}) is nontrivial, the only options are the flute surface, the Loch Ness monster surface, or the spotted Loch Ness monster surface. We claim there are no isolated planar ends in SabS^{ab}. Indeed, regularity of the cover implies that any such end covers an isolated planar end of SS. But any peripheral curve in SS is part of a free basis for π1(S)\pi_{1}(S), and hence no power of it lies in the commutator subgroup, so the preimage of any annular neighborhood of an isolated planar end of SS has simply connected components in SabS^{ab}. Hence, SabS^{ab} is the Loch Ness monster surface. ∎

3.4. /n\mathbb{Z}/n\mathbb{Z}-homology covers

Recall from the introduction that the /n\mathbb{Z}/n\mathbb{Z}-homology cover of a surface SS is the cover corresponding to the kernel of the map π1(S)H1(S,/n)\pi_{1}(S)\to H_{1}(S,\mathbb{Z}/n\mathbb{Z}).

Theorem 1.3 (/n\mathbb{Z}/n\mathbb{Z}-homology covers).

Say SS is an infinite-type orientable surface and π:S~S\pi:\tilde{S}\to S is its /n\mathbb{Z}/n\mathbb{Z}-homology cover, where n2n\geq 2. If SS has no isolated planar ends, then S~\tilde{S} is homeomorphic to the Loch Ness monster surface. Otherwise, S~\tilde{S} is homeomorphic to the spotted Loch Ness monster surface.

The main point is to show the following lemma:

Lemma 3.9.

With π:S~S\pi:\tilde{S}\to S as above, the cover S~\tilde{S} has at most one end that is not an isolated planar end.

Deferring the proof of Lemma 3.9, we now prove prove Theorem 1.3.

Proof Theorem 1.3.

The first step is to prove:

Claim 3.10.

S~\tilde{S} has infinite genus.

Proof.

Since π\pi is a regular cover of infinite degree, it suffices to prove that S~\tilde{S} has positive genus.

If SS has positive genus, we can pick an embedded once punctured torus TST\subset S. Since SS has infinite-type, STS\smallsetminus T is noncompact, so Mayer-Vietoris implies that H1(T,)H1(S,)H_{1}(T,\mathbb{Z})\hookrightarrow H_{1}(S,\mathbb{Z}). Hence, picking a component

T~π1(T)S~,\tilde{T}\subset\pi^{-1}(T)\subset\tilde{S}~{},

the cover T~T\tilde{T}\to T is the /n\mathbb{Z}/n\mathbb{Z}-homology cover of TT. As π1(T)F2\pi_{1}(T)\cong F_{2}, we have

H1(T,/n)(/n)2,H_{1}(T,\mathbb{Z}/n\mathbb{Z})\cong(\mathbb{Z}/n\mathbb{Z})^{2}~{},

so the cover T~T\tilde{T}\to T has degree n2n^{2}, so χ(T~)=n2\chi(\tilde{T})=-n^{2}. However, since T\partial T is trivial in homology, it lifts homeomorphically to T~\tilde{T}, and hence T~\tilde{T} has n2n^{2} boundary components, implying T~\tilde{T} is a n2n^{2}-punctured torus111One can also just visualize the cover directly by unwrapping the meridian and longitude direction nn-times., so S~\tilde{S} has positive genus.

If SS has genus zero, take some essential 44-punctured sphere TST\subset S. Then STS\smallsetminus T has four components. Moreover, all four components of STS\smallsetminus T are noncompact: any compact component is a planar surface with one boundary component, and hence is a disk, contradicting that TT is essential. It follows that H1(T,)H1(S,)H_{1}(T,\mathbb{Z})\hookrightarrow H_{1}(S,\mathbb{Z}). Hence, if T~π1(T)\tilde{T}\subset\pi^{-1}(T) is a component, then T~T\tilde{T}\to T is the /n\mathbb{Z}/n\mathbb{Z}-homology covering of TT, which has degree n3n^{3}, so χ(T~)=2n3\chi(\tilde{T})=-2n^{3}. Each component of T\partial T is primitive in H1(S,)H_{1}(S,\mathbb{Z}), so unwraps nn times in T~\tilde{T}, and hence there are n3/n=n2n^{3}/n=n^{2} lifts of each, for a total of 4n24n^{2} boundary components. A quick computation shows that T~\tilde{T} has genus g=112(2n3+4n2)g=1-\frac{1}{2}(-2n^{3}+4n^{2}), which is never zero for n2n\geq 2. ∎

By Lemma 3.9 and Claim 3.10, S~\tilde{S} is either the Loch Ness monster surface or its spotted version. If SS has isolated planar ends, these ends lift to isolated planar ends in S~\tilde{S}, so S~\tilde{S} is the spotted Loch Ness monster surface. Conversely, since the cover is regular, any isolated planar end in S~\tilde{S} covers an isolated planar end in SS. So if SS has no isolated planar ends, S~\tilde{S} is the Loch Ness monster surface. ∎

3.4.1. The proof of Lemma 3.9

Before formally proving the Lemma 3.9, we give a brief heuristic of the proof.

The rough idea here is to use the fact that the cover S~S\tilde{S}\to S is abelian, and the fact that homologically essential curves are ubiquitous in SS, and such curves do not lift homeomorphically to S~\tilde{S}. Slightly more precisely, and assuming for simplicity that neither SS nor S~\tilde{S} has isolated planar ends, we take a compact set K~S~\tilde{K}\subset\tilde{S} that projects to KSK\subset S, a path a~\tilde{a} in S~\tilde{S} with endpoints outside of K~\tilde{K}, and we claim that there exists another path γ~\tilde{\gamma} with the same endpoints that does not enter K~\tilde{K} (this will show that S~\tilde{S} is one-ended). To construct such a γ~\tilde{\gamma} we work as follows. Writing a=π(a~)a=\pi(\tilde{a}) and γ=π(γ~)\gamma=\pi(\tilde{\gamma}), it may be that there is not really another way to travel within SS between the endpoints of aa without roughly following aa, which might take us through KK. So, we try to construct γ~\tilde{\gamma} so that at some point its projection γ\gamma follows aa, but when it does so, γ~\tilde{\gamma} still avoids going through K~\tilde{K}. To do this, we first make γ\gamma loop around some homologically essential curve bb, so that when γ\gamma tracks aa the lift γ~\tilde{\gamma} is actually tracking a translate of a~\tilde{a} rather than a~\tilde{a} itself. We then loop backwards around bb at the end, and use that the cover is abelian to say that this cancels out the first loop around bb, so that γ~\tilde{\gamma} terminates at an endpoint of a~\tilde{a} rather than at some translate of that endpoint.

We now prove Lemma 3.9 by contradiction. Suppose that S~\tilde{S} has more than one end that is not isolated and planar. We first show:

Claim 3.11.

There exists a connected, finite-type subsurface XSX\subset S such that

  1. (1)

    every component of SXS\smallsetminus X has infinite type, and

  2. (2)

    there is a finite-type connected component X~\tilde{X} of π1(X)S~\pi^{-1}(X)\subset\tilde{S} such that the complement S~X~\tilde{S}\smallsetminus\tilde{X} contains two distinct infinite-type connected components U~+\tilde{U}_{+} and U~\tilde{U}_{-}.

Proof.

Since S~\tilde{S} has more than one end that is not isolated and planar, there is a compact subsurface K~S~\tilde{K}\subset\tilde{S} that has two distinct infinite-type connected components. The projection π(K~)S\pi(\tilde{K})\subset S is compact, so there is a finite-type subsurface XSX\subset S containing the projection. We can assume that all components of SXS\smallsetminus X have infinite type, by adding any finite type components into XX. Since π\pi is locally finite (see §1.6), all components of π1(X)\pi^{-1}(X) have finite type, and the component X~\tilde{X} containing K~\tilde{K} has at least two infinite-type complementary components. ∎

Pick an oriented path a~\tilde{a} in X~\tilde{X} that starts at a point x~U~\tilde{x}_{-}\in\partial\tilde{U}_{-} and ends at a point x~+U~+\tilde{x}_{+}\in\partial\tilde{U}_{+} and let a,x,x+a,x_{-},x_{+} be the projections to SS. Then x±x_{\pm} lies in X\partial X and is adjacent to a component V±SXV_{\pm}\subset S\smallsetminus X. As V±V_{\pm} has infinite type, there is a nonseparating bi-infinite arc α±\alpha_{\pm} in V±V_{\pm}. If V+=VV_{+}=V_{-}, we can choose the arcs α+,α\alpha_{+},\alpha_{-} so that the union α+α\alpha_{+}\cup\alpha_{-} is nonseparating in V+=VV_{+}=V_{-}. Let b±b_{\pm} be an oriented loop in V±V_{\pm} that starts and ends at x±x_{\pm}, and that intersects α±\alpha_{\pm} exactly once. If V+=VV_{+}=V_{-}, we can assume that b±b_{\pm} is disjoint from α\alpha_{\mp}.

Consider now the path γ=bab+1a1b1ab+\gamma=b_{-}ab_{+}^{-1}a^{-1}b_{-}^{-1}ab_{+} in SS, which starts at xx_{-} and ends at x+x_{+}, see Figure 4 for a configuration.

Refer to caption
KK
VV_{-}
V+V_{+}
α+\alpha_{+}
α\alpha_{-}
xx_{-}
x+x_{+}
Refer to caption
AA
aa
bb_{-}
b+b_{+}
Refer to caption
Figure 4. The compact subsurface KK, the reference bi-infinite arcs α\alpha_{-} and α+\alpha_{+}, and the arcs bb_{-}, b+b_{+}, and aa used to form γ\gamma.

Then γa1\gamma a^{-1} is a loop that starts and ends at xx_{-}, and since

γa1=b(ab+1a1)b1(ab+a1)=[b,ab+1a1],\gamma a^{-1}=b_{-}(ab_{+}^{-1}a^{-1})b_{-}^{-1}(ab_{+}a^{-1})=[b_{-},ab_{+}^{-1}a^{-1}],

this loop is trivial in homology and therefore lifts to S~\tilde{S}. Choose the lift such that the terminal a1a^{-1} lifts to the already defined path a~1\tilde{a}^{-1}, which starts at x~+\tilde{x}_{+} and ends at x~\tilde{x}_{-}. If we do this, then the resulting lift γ~\tilde{\gamma} of γ\gamma starts at x~\tilde{x}_{-} and ends at x~+\tilde{x}_{+}.

Claim 3.12.

γ~\tilde{\gamma} is disjoint from int(X~)\text{int}(\tilde{X}).

This claim will finish the proof of Lemma 3.9, as it contradicts the fact that x~\tilde{x}_{-} and x~+\tilde{x}_{+} lie on the boundaries of distinct connected components of S~X~\tilde{S}\smallsetminus\tilde{X}. So, all that is left is to prove the claim.

Proof of Claim 3.12.

Below, when we talk about ‘the lift of’ a path like b,b_{-}, or a,a, or b+1b_{+}^{-1}, etc., we mean the lift that appears as the associated subpath of γ~\tilde{\gamma}. Since there are two traversals of aa in γ\gamma, we refer to the two lifts as the ‘first lift’ of aa and the ‘second lift’ of aa, using the order they appear in the word. In this terminology, orientations matter: the ‘lift of b1b_{-}^{-1}’ may not be obtained by traversing the ‘lift of bb_{-}’ backwards.

First, observe that all the lifts of b,b1,b+,b+1b_{-},b_{-}^{-1},b_{+},b_{+}^{-1} are disjoint from int(X~)\text{int}(\tilde{X}), as their projections are disjoint from XX by construction. Now, suppose the first lift of aa intersects int(X~)\text{int}(\tilde{X}). Then we can create a loop in S~\tilde{S} by concatenating the lift of bb_{-}, part of the lift of aa, and a path in X~\tilde{X}. The projection of this loop to SS intersects α\alpha_{-} exactly once, as the latter two parts project into XX. This is impossible, since the projection must be trivial in /n\mathbb{Z}/n\mathbb{Z}-homology. The same argument shows that the lift of a1a^{-1} is disjoint from int(X~)\text{int}(\tilde{X}): if not, one can construct a loop that is trivial in /n\mathbb{Z}/n\mathbb{Z}-homology and intersects α\alpha_{-} once.

A similar argument shows that the second lift of aa is disjoint from int(X~)\text{int}(\tilde{X}): if not, one can construct a loop that is trivial in /n\mathbb{Z}/n\mathbb{Z}-homology and intersects α+\alpha_{+} once. Note that above, we could have used α\alpha_{-} or α+\alpha_{+}, but here we need to use α+\alpha_{+}, as both bb_{-} and b1b_{-}^{-1} appear before the second occurrence of aa in the word defining γ\gamma. ∎

4. Arbitrary covers

In this section we prove the following theorem.

Theorem 1.1 (Everything covers everything).

Suppose that SS is an orientable, borderless surface with non-abelian fundamental group. Then SS is covered by any noncompact borderless orientable surface.

The idea for the proof of Theorem 1.1 is to show that any surface SS with non-abelian fundamental group is covered by the blooming Cantor tree surface \mathcal{B} and that any infinite-type surface admits a π1\pi_{1}-injective embedding into \mathcal{B}, see Proposition 4.2. Then, Lemma 2.4 gives us Theorem 1.1.

Before starting the proof of Theorem 1.1 we show that the blooming Cantor tree surface regularly covers the Loch Ness monster surface.

Lemma 4.1.

The blooming Cantor tree surface regularly covers the Loch Ness monster surface.

Proof.

Let GG be the Cayley graph of 2\mathbb{Z}^{2}. Note that if we replace every vertex of GG by a four-punctured torus and we glue them according to the edges relation we obtain a surface LL that, by classification, is homeomorphic to the Loch Ness monster surface. Let G~\tilde{G} be the universal cover of GG, i.e., G~\tilde{G} is a complete 4-valent tree. Then, if we replace each vertex of G~\tilde{G} by a four punctured torus we obtain a surface S~\tilde{S} that, again by classification, is homeomorphic to the blooming Cantor tree surface. Then the induced covering map from the graphs induces the required covering map from S~\tilde{S} to SS.∎

We now show that any infinite-type surface can be embedded in the blooming Cantor tree surface \mathcal{B} as a connected component of the complement of a collection of simple closed curves and bi-infinite arcs.

Proposition 4.2.

For any borderless non-compact orientable surface SS, there is a π1\pi_{1}-injective embedding of SS into the blooming Cantor tree \mathcal{B}.

Proof.

We assume throughout the proof that SS has either zero or infinite genus, since an arbitrary surface can be written as the connected sum S#TS\#T of such an SS and a closed orientable surface TT, and a π1\pi_{1}-injective embedding SS\hookrightarrow\mathcal{B} induces a π1\pi_{1}-injective embedding of S#T#TS\#T\hookrightarrow\mathcal{B}\#T\cong\mathcal{B}.

Let (,g)(\mathcal{E},\mathcal{E}_{g}) be the genus-marked end space of SS, which is nonempty since SS is non-compact. Since the end space ()\mathcal{E}(\mathcal{B}) is a Cantor set, we can fix an embedding ι:()\iota:\mathcal{E}\rightarrow\mathcal{E}(\mathcal{B}), and we let E=ι()E=\iota(\mathcal{E}) and Eg=ι(g)E_{g}=\iota(\mathcal{E}_{g}). We will construct an open, connected π1\pi_{1}-injective subsurface S′′S^{\prime\prime}\subset\mathcal{B} such that

  1. (a)

    the inclusion S′′S^{\prime\prime}\hookrightarrow\mathcal{B} induces a homeomorphism ((S′′),g(S′′))(E,Eg)(\mathcal{E}(S^{\prime\prime}),\mathcal{E}_{g}(S^{\prime\prime}))\cong(E,E_{g}),

  2. (b)

    the genus of S′′S^{\prime\prime} is the same as the genus of SS.

By classification of surfaces, see Theorem 2.2, such a surface S′′S^{\prime\prime} will be homeomorphic to SS, proving the proposition.

As a first step, we construct a π1\pi_{1}-injective closed subsurface SS^{\prime}\subset\mathcal{B} (with boundary) that has properties (a) and (b). Let Λ\Lambda\subset\mathcal{B} be a union of simple closed curves that cuts \mathcal{B} into a collection of thrice-punctured tori, connected together in the pattern of a trivalent tree. For each of the thrice-punctured tori CΛC\subset\mathcal{B}\setminus\Lambda, cut CC along some simple closed curve into a punctured torus TCT_{C}, and a four-punctured sphere XCCX_{C}\subset C. See Figure 5 for such a configuration.

Refer to caption
Figure 5. The blooming Cantor tree surface with multicurve Λ\Lambda depicted in red.

Fixing some component λ0Λ\lambda_{0}\subset\Lambda as the ‘root’, we then define SS^{\prime}\subset\mathcal{B} to be the union of:

  1. (1)

    all XC¯\overline{X_{C}} such that CC separates λ0\lambda_{0} from some point of EE,

  2. (2)

    all TC¯\overline{T_{C}} such that CC separates λ0\lambda_{0} from some point of EgE_{g},

here XC¯\overline{X_{C}} and TC¯\overline{T_{C}} are just the closures of the open subsurfaces mentioned above. Observe that SS^{\prime} is connected, it π1\pi_{1}-injects into \mathcal{B}, there is a homeomorphism ((S),g(S))(E,Eg)(\mathcal{E}(S^{\prime}),\mathcal{E}_{g}(S^{\prime}))\simeq(E,E_{g}) as above and the genus of SS^{\prime} is the same as that of SS. The boundary S\partial S^{\prime} consists of countably many circles. Our goal now is to remove the compact boundary components of SS^{\prime} so that they are replaced by half-planes.

We claim that for any orientable surface SS^{\prime} such that S\partial S^{\prime} is a union of simple closed curves, there is a π1\pi_{1}-injective borderless subsurface S′′SS^{\prime\prime}\subset S^{\prime}, with the same genus, such that the inclusion induces a homeomorphism of genus-marked end spaces. Let {γi}i\{\gamma_{i}\}_{i\in\mathbb{N}} be the collection of boundary components of SS^{\prime} and let {ri}i\{r_{i}\}_{i\in\mathbb{N}} be a locally finite collection of pairwise disjoint properly embedded rays on SS^{\prime} such that rir_{i} starts on γi\gamma_{i}, and let PiP_{i} be a closed regular neighborhood of γiri\gamma_{i}\cup r_{i} in SS^{\prime}. If we take these regular neighborhoods small enough, they are all disjoint and we obtain a surface S′′:=SiPiS^{\prime\prime}:=S^{\prime}\setminus\cup_{i}P_{i}. Observe that the end space of S′′S^{\prime\prime} agrees with that of SS^{\prime}, so it also satisfies the required properties.

Composing the inclusions S′′SS^{\prime\prime}\hookrightarrow S^{\prime}\hookrightarrow\mathcal{B} gives an open, π1\pi_{1}-injective subsurface of \mathcal{B} satisfying (a) and (b) above, which proves the proposition. ∎

Refer to caption
Figure 6. The once-punctured Loch Ness monster surface as a subsurface of \mathcal{B} according to our construction. The left side of the figure represents the sole planar end, while the right side represents the end accumulated by genus.
Proof of Theorem 1.1.

By Proposition 4.2 and Lemma 2.4 it suffices to show that given SS there exists a covering space BB homeomorphic to a blooming Cantor tree surface. Since SS has non-abelian π1\pi_{1} we can find a π1\pi_{1}-injective pair of pants PSP\subset S. Then, by Lemma 2.4 we get a cover S′′SS^{\prime\prime}\to S where S′′S^{\prime\prime} is an open pair of pants. By Theorem 1.2 S′′S^{\prime\prime} is covered by a Loch Ness monster surface LL and we conclude by Lemma 4.1. ∎

Appendix A Appendix

The goal of this appendix is to show how Theorem 2.6 is a restatement of the main result in Swarup’s paper [16]. To do so, we relate three ways of looking at ends of a group GG. In the background, we have the space of ends (G)\mathcal{E}(G) as defined in §2.2. We want to relate this to H1(G,G)H^{1}(G,\mathbb{Z}G), which is the definition of the space of ends in Swarup’s paper [16]. To do so, we relate both to a combinatorial description of ends that is similar to a description in Stallings’s paper [15]. All of this material is likely well known to experts, but we did not readily find a reference for the portions relevant to the goals of this article, and so we present it here in hopes of providing a useful reference to others.

A.1. Group cohomology

Let GG be a group. A GG-module is an abelian group that comes equipped with an action of GG by automorphisms. For instance, any abelian group AA with the trivial GG action is a GG-module. On the other end of the spectrum, let

G:={ϕ:G},\mathbb{Z}^{G}:=\{\phi\mskip 0.5mu\colon\thinspace G\to\mathbb{Z}\},

which we consider as an abelian group under addition of functions. We realize G\mathbb{Z}^{G} as a GG-module by defining the (left) GG-action (gϕ)(h)=ϕ(gh)(g\phi)(h)=\phi(gh). So, the action of gGg\in G precomposes ϕ\phi with left translation by gg. From this definition we observe the following equality, which will be used several times: if xGx\in\mathbb{Z}^{G} and g,hGg,h\in G, then

(1) (gx)(h)=x(g1h)(gx)(h)=x(g^{-1}h)

The group ring of GG, denoted by G\mathbb{Z}G, is the submodule of G\mathbb{Z}^{G} given by

G:={f:G|f has finite support},\mathbb{Z}G:=\{f\mskip 0.5mu\colon\thinspace G\to\mathbb{Z}\ |\ f\text{ has finite support}\},

where the support of ff is the set of all gGg\in G with f(g)0f(g)\neq 0. We can express the elements of G\mathbb{Z}^{G} as formal \mathbb{Z}-linear combinations

gGcgg,\sum_{g\in G}c_{g}\cdot g,

where cgc_{g}\in\mathbb{Z} and—abusing notation—we view gg as the indicator function that assigns 11 to gg and 0 to all other group elements. The elements of G\mathbb{Z}G can then be realized as the formal sums in which cg=0c_{g}=0 for all but finitely many g. We will switch back and forth between formal \mathbb{Z}-linear combinations and functions as convenient.

Suppose that MM is a GG-module. We will now define the cohomology groups Hn(G,M)H^{n}(G,M) with coefficients in MM. An nn-cochain with coefficients in MM is a function ϕ:GnM\phi:G^{n}\to M, where Gn=G××GG^{n}=G\times\cdots\times G, and by convention G0={1}G^{0}=\{1\}. Let CnC^{n} denote the set of nn-cochains. We then define the coboundary maps n:CnCn+1\partial^{n}:C^{n}\to C^{n+1} as follows: for n>0n>0, set

(nϕ)(g1,,gn+1)=\displaystyle(\partial^{n}\phi)(g_{1},\ldots,g_{n+1})\ \ \ =\ \ \ \ g1ϕ(g2,,gn+1)\displaystyle g_{1}\phi(g_{2},\ldots,g_{n+1})
+\displaystyle+\ i=1n(1)iϕ(g1,,gigi+1,,gn+1)\displaystyle\sum_{i=1}^{n}(-1)^{i}\phi(g_{1},\ldots,g_{i}g_{i+1},\ldots,g_{n+1})
+\displaystyle+\ (1)n+1ϕ(g1,,gn).\displaystyle(-1)^{n+1}\phi(g_{1},\ldots,g_{n}).

and for n=0n=0, set

(2) (0ϕ)(g)=gϕ(1)ϕ(1)=(g1)ϕ(1).(\partial^{0}\phi)(g)=g\cdot\phi(1)-\phi(1)=(g-1)\phi(1).

The cohomology groups are now defined to be

Hn(G,M):=kern/Imn1.H^{n}(G,M):=\mathrm{ker}\ \partial^{n}/\mathrm{Im}\ \partial^{n-1}.

This definition relies on the fact that Im(n1)kern\mathrm{Im}(\partial^{n-1})\subset\ker\partial^{n}, a fact we have not checked; we will verify this for n=1n=1 below, which is the only case we will use in this appendix.

Now, 0-chains are functions ϕ:{1}M\phi:\{1\}\to M, or in other words, a choice of an element of MM. By (2), 0ϕ=0\partial^{0}\phi=0 exactly when the element ϕ(1)M\phi(1)\in M is invariant under the action of GG, and hence we get

H0(G,M){xM|gx=x,gG}.H^{0}(G,M)\cong\{x\in M\ |\ gx=x,\forall g\in G\}.

Here are some examples for different GG-modules MM.

  1. (1)

    If MM is an abelian group with the trivial GG-action, then H0(G,M)MH^{0}(G,M)\cong M.

  2. (2)

    If M=GM=\mathbb{Z}^{G}, then H0(G,M)H^{0}(G,M)\cong\mathbb{Z} since the GG-invariant elements of G\mathbb{Z}^{G} are exactly those where all the coefficients are the same.

  3. (3)

    If M=GM=\mathbb{Z}G, then

    H0(G,G){ if G is finite 0 if G is infinite. H^{0}(G,\mathbb{Z}G)\cong\begin{cases}\mathbb{Z}&\text{ if }G\text{ is finite }\\ 0&\text{ if }G\text{ is infinite. }\end{cases}

    Indeed, if [f]H0(G,G)=ker0[f]\in H^{0}(G,\mathbb{Z}G)=\ker\partial^{0}, then the support of ff is finite and f(g)=f(h)f(g)=f(h) for all g,hGg,h\in G.

To compute H1(G,M)H^{1}(G,M), we need to understand the image of 0\partial^{0} (i.e., the set of coboundaries) and the kernel of 1\partial^{1} (i.e., the set of cocycles). Given xMx\in M, define the 1-cochain ϕx:GM\phi_{x}\mskip 0.5mu\colon\thinspace G\to M by

ϕx(g)=(g1)x.\phi_{x}(g)=(g-1)x.

In (2), we computed the image of 0\partial^{0}, establishing that the set of 1-coboundaries is {ϕx:xM}\{\phi_{x}:x\in M\}.

To understand 11-cocycles, first note that if ϕ\phi is a 11-cochain, then

(1ϕ)(g,h)=gϕ(h)ϕ(gh)+ϕ(g).(\partial^{1}\phi)(g,h)=g\phi(h)-\phi(gh)+\phi(g).

Setting the right-hand side to zero, we see that ϕ\phi is a 11-cocycle if and only if

ϕ(gh)=gϕ(h)+ϕ(g),\phi(gh)=g\phi(h)+\phi(g),

i.e., ϕ\phi satisfies the cocycle condition. Let us verify that every 1-coboundary is a 1-cocycle, i.e., Im(0)ker1\mathrm{Im}(\partial^{0})\subset\ker\partial^{1}. Fix a 1-cobounday ϕx\phi_{x}, where xMx\in M. Then

(3) ϕx(gh)=(gh1)x=g(h1)x+(g1)x=gϕx(h)+ϕx(g)\phi_{x}(gh)=(gh-1)x=g(h-1)x+(g-1)x=g\phi_{x}(h)+\phi_{x}(g)

implying that ϕx\phi_{x} satisfies the cocycle condition and is therefore a 1-cocycle.

Let us work out two example computations of H1(G,M)H^{1}(G,M).

  1. (A)

    Suppose that MM is an abelian group with a trivial GG-action. In this case, the cocycle condition becomes ϕ(gh)=ϕ(g)+ϕ(h)\phi(gh)=\phi(g)+\phi(h), i.e. ϕ\phi is a homomorphism. Given a coboundary ϕ\phi, there exists xMx\in M such that ϕ=ϕx\phi=\phi_{x}. It follows that ϕ(g)=(g1)x=gxx=xx=0\phi(g)=(g-1)x=gx-x=x-x=0, and hence ϕ=0\phi=0. As all coboundaries vanish,

    H1(G,M)=Hom(G,M).H^{1}(G,M)=\mathrm{Hom}(G,M).

    In the case M=M=\mathbb{Z}, we have that H1(G,)H^{1}(G,\mathbb{Z}) is isomorphic to the free part of the abelianization of GG.

  2. (B)

    Suppose that M=GM=\mathbb{Z}^{G}. We claim that H1(G,G)=0H^{1}(G,\mathbb{Z}^{G})=0. To see this, suppose ϕ:GG\phi\mskip 0.5mu\colon\thinspace G\to\mathbb{Z}^{G} is a 1-cocycle. Define xGx\in\mathbb{Z}^{G} by

    x(g)=ϕ(g1)(1).x(g)=\phi(g^{-1})(1).

    We claim that ϕ=ϕx\phi=\phi_{x}, and hence ϕ\phi is a coboundary. Indeed, if hGh\in G, then:

    ϕx(g)(h)\displaystyle\phi_{x}(g)(h) =(g1)x(h)\displaystyle=(g-1)x(h)
    =gx(h)x(h)\displaystyle=gx(h)-x(h)
    =x(g1h)x(h)\displaystyle=x(g^{-1}h)-x(h)
    =ϕ(h1g)(1)ϕ(h1)(1)\displaystyle=\phi(h^{-1}g)(1)-\phi(h^{-1})(1)
    =(ϕ(h1g)ϕ(h1))(1)\displaystyle=(\phi(h^{-1}g)-\phi(h^{-1}))(1)
    =h1ϕ(g)(1)\displaystyle=h^{-1}\phi(g)(1)
    =ϕ(g)(h)\displaystyle=\phi(g)(h)

    where the third equality uses (1) and the second-to-last equality uses the cocycle condition. Therefore, ϕ=ϕx\phi=\phi_{x}, implying that ϕ\phi is a coboundary. It follows that H1(G,G)=0H^{1}(G,\mathbb{Z}^{G})=0.

A.2. Ends via group cohomology

In this section we describe how the set of ends of GG is encoded in the cohomology group H1(G,G)H^{1}(G,\mathbb{Z}G). It is well known that the number of ends of an infinite finitely generated group GG is equal to 1+rank(H1(G,G))1+\mathrm{rank}(H^{1}(G,\mathbb{Z}G)), where rank denotes the cardinality of a maximal \mathbb{Z}-linearly independent subset (see for example [7, Theorem 13.5.5]). However, we require a more transparent proof of this fact that will allow us to more easily generalize to the relative setting discussed in Theorem 2.6.

Suppose GG is generated by a finite set SS that is symmetric under inversion. Let Cay(G){Cay}(G) be the associated left Cayley graph for GG, that is, the graph whose vertex set is GG and where gg and hh are adjacent if h=sgh=sg for some sSs\in S. Note that this is opposite the usual construction of Cayley graphs; however, we need to use left Cayley graphs here since we chose to consider G\mathbb{Z}G as a left GG-module.

Given xGx\in\mathbb{Z}^{G}, the boundary of xx is the subset xG\partial x\subset G consisting of all gGg\in G such that x(sg)x(g)x(sg)\neq x(g) for some sSs\in S. For example, if G=G=\mathbb{Z}, S={±1}S=\{\pm 1\}, and

x(n)={1if n00if n<0x(n)=\left\{\begin{array}[]{ll}1&\text{if }n\geq 0\\ 0&\text{if }n<0\end{array}\right.

then x={1,0}.\partial x=\{-1,0\}.

Recall that, given xGx\in\mathbb{Z}^{G}, we defined ϕx:GG\phi_{x}\mskip 0.5mu\colon\thinspace G\to\mathbb{Z}^{G} by

ϕx(g)=(g1)x.\phi_{x}(g)=(g-1)x.

The next two lemmas tell us that if xx has finite boundary, then ϕx\phi_{x} is a 1-cocycle with G\mathbb{Z}G coefficients and that these account for all 1-cocycles with G\mathbb{Z}G coefficients. The third lemma characterizes the 1-coboundaries with G\mathbb{Z}G coefficients.

Lemma A.1.

If xGx\in\mathbb{Z}^{G}, then ϕx\phi_{x} is a 1-cochain with G\mathbb{Z}G coefficients if and only if x\partial x is finite.

Proof.

If gxg\in\partial x, then there exists sSs\in S such that x(sg)x(g)x(sg)\neq x(g). By (1), it follows that s1x(g)x(g)s^{-1}x(g)\neq x(g) and hence s1x(g)x(g)0s^{-1}x(g)-x(g)\neq 0, which says that gg is an element in the support of (s11)x(s^{-1}-1)x. It follows that x\partial x is contained in the union of the supports of the elements (s1)xG(s-1)x\in\mathbb{Z}^{G} for sSs\in S. If we assume that ϕx\phi_{x} is a 1-cochain with G\mathbb{Z}G coefficients, then (s1)x=ϕx(s)G(s-1)x=\phi_{x}(s)\in\mathbb{Z}G for all sSs\in S; it follows that x\partial x is finite, establishing the forwards direction.

Now, assume x\partial x is finite. Fix gGg\in G. If g1=s1smg^{-1}=s_{1}\cdots s_{m}, where siSs_{i}\in S, we claim that (g1)x(g-1)x is supported within the mm-neighborhood of xG\partial x\subset G, with respect to the word metric. Indeed, if hh lies outside this neighborhood, then for each ii, the elements sismhs_{i}\cdots s_{m}h does not lie in x\partial x, so x(sismh)=x(si1smh)x(s_{i}\cdots s_{m}h)=x(s_{i-1}\cdots s_{m}h), and inductively we get x(g1h)=x(h)x(g^{-1}h)=x(h). Using (1), we have that gx(h)x(h)=0gx(h)-x(h)=0, so (g1)x(h)=0(g-1)x(h)=0 implying that ϕx(g)=(g1)xG\phi_{x}(g)=(g-1)x\in\mathbb{Z}G. As gGg\in G was arbitrary, we conclude that ϕx\phi_{x} is a 1-cochain with G\mathbb{Z}G coefficients. ∎

Lemma A.2 (11-cocycles with G\mathbb{Z}G-coefficients).

A 1-cochain ϕ\phi with G\mathbb{Z}G coefficients is a 1-cocycle if and only if there exists xGx\in\mathbb{Z}^{G} with finite boundary such that ϕ=ϕx\phi=\phi_{x}.

Proof.

Let xGx\in\mathbb{Z}^{G} with finite boundary. Lemma A.1 tells us that ϕx\phi_{x} is a 1-cochain with G\mathbb{Z}G coefficients, and then (2) tells us that ϕx\phi_{x} is a 1-cocycle with G\mathbb{Z}G coefficients. Conversely, if ϕ\phi is a 1-cocycle with G\mathbb{Z}G coefficients, then the computation in (B) above tells us that ϕ=ϕx\phi=\phi_{x} with xGx\in\mathbb{Z}^{G} defined by x(g)=ϕ(g1)(1)x(g)=\phi(g^{-1})(1). It is left to check that xx has finite boundary: as ϕx=ϕ\phi_{x}=\phi, we have that (g1)x=ϕ(g)G(g-1)x=\phi(g)\in\mathbb{Z}G for all gGg\in G; hence, by Lemma A.1, xx has finite boundary. ∎

Lemma A.3.

Let xGx\in\mathbb{Z}^{G}. If ϕx\phi_{x} is a 1-cocycle with G\mathbb{Z}G coefficients, then ϕx\phi_{x} is a 1-coboundary with G\mathbb{Z}G coefficients if and only if xx is constant outside a finite set of GG.

Proof.

By definition, ϕx\phi_{x} is a 1-coboundary if and only if there exists yGy\in\mathbb{Z}G such that ϕx=ϕy\phi_{x}=\phi_{y}. For the forwards direction, suppose that ϕx\phi_{x} is a 1-coboundary. Then there exists yGy\in\mathbb{Z}G such that (g1)x=(g1)y(g-1)x=(g-1)y for every gGg\in G. In particular, g(xy)=xyg(x-y)=x-y for each gGg\in G, implying that xyx-y is constant. Hence, xx is constant off the finite support of yy. Conversely, suppose that xx is constant outside a finite subset YY of GG. Let dd\in\mathbb{Z} such that x(g)=dx(g)=d for all gYg\notin Y. Define yGy\in\mathbb{Z}G by y(g)=0y(g)=0 if gYg\notin Y and y(g)=x(g)dy(g)=x(g)-d if gYg\in Y. Then xyx-y is constant, and by repeating the computation above in reverse, ϕx=ϕy\phi_{x}=\phi_{y}, implying that ϕx\phi_{x} is a 1-coboundary ∎

Now, suppose that GG is an infinite finitely generated group. Let (G)\mathcal{E}(G) be the space of ends of Cay(G)Cay(G), and equip Cay(G)(G)Cay(G)\cup\mathcal{E}(G) with the usual topology. Identifying GG with the vertex set of Cay(G)Cay(G), we can view G(G)G\cup\mathcal{E}(G) as a topological space. It is an exercise to show that (G)\mathcal{E}(G) is either a singleton, doubleton, or a perfect set; we define e(G)e(G) to be one, two, or \infty, respectively.

If xGx\in\mathbb{Z}^{G} has finite boundary, then any end of GG has a neighborhood in GG on which xx is constant, so there is a unique function x¯:(G)\bar{x}\mskip 0.5mu\colon\thinspace\mathcal{E}(G)\to\mathbb{Z} such that xx¯:G(G)x\cup\bar{x}\mskip 0.5mu\colon\thinspace G\cup\mathcal{E}(G)\to\mathbb{Z} is continuous.

Let C((G),)C(\mathcal{E}(G),\mathbb{Z}) be the set of all continuous functions (G)\mathcal{E}(G)\to\mathbb{Z}, which is an abelian group under addition. Let ZC((G),)Z\subset C(\mathcal{E}(G),\mathbb{Z}) be the constant functions. As mentioned above, the rank of an abelian group AA, denoted rank(A)\mathrm{rank}(A), is the cardinality of a maximal \mathbb{Z}-linearly independent subset of AA.

Theorem A.4.

Let GG be an infinite finitely generated group. The map

H1(G,G)C((G),)/ZH^{1}(G,\mathbb{Z}G)\to C(\mathcal{E}(G),\mathbb{Z})/Z

defined by [ϕ][x¯][\phi]\longmapsto[\bar{x}] is an isomorphism, where xGx\in\mathbb{Z}^{G} satisfies ϕ=ϕx\phi=\phi_{x}. Consequently, we have

e(G)=1+rank(H1(G,G)).e(G)=1+\mathrm{rank}(H^{1}(G,\mathbb{Z}G)).

Note that we are supposing GG is infinite above, so that (G)\mathcal{E}(G) is nonempty. If GG is finite, then e(G)=0e(G)=0 and, by Lemma A.2, H1(G,G)=0H^{1}(G,\mathbb{Z}G)=0.

Proof of Theorem A.4.

First, we claim the map is well defined: suppose [ϕ]=[ψ][\phi]=[\psi]. Then there exists x,yGx,y\in\mathbb{Z}^{G} such that ϕ=ϕx\phi=\phi_{x}, ψ=ϕy\psi=\phi_{y}, and, by Lemma A.3, xyx-y is constant, implying [x¯]=[y¯][\bar{x}]=[\bar{y}], as desired. The map is a homomorphism as

[ϕx]+[ϕy]=[ϕx+ϕy]=[ϕx+y][\phi_{x}]+[\phi_{y}]=[\phi_{x}+\phi_{y}]=[\phi_{x+y}]

and the map xx¯x\to\bar{x} is linear.

Next, we show the map is injective: suppose xGx\in\mathbb{Z}^{G} with finite boundary such that x¯Z\bar{x}\in Z. As x¯Z\bar{x}\in Z, there exists dd\in\mathbb{Z} such that x¯(ξ)=d\bar{x}(\xi)=d for all ξ(G)\xi\in\mathcal{E}(G). It follows that x(g)=dx(g)=d for every gg that lies in an infinite component of Cay(G)xCay(G)\smallsetminus\partial x. As x\partial x is finite, Lemma A.3 implies that [ϕx]=0[\phi_{x}]=0. Now, we turn to surjectivity. If fC((G),)f\in C(\mathcal{E}(G),\mathbb{Z}), then the continuity of ff and the compactness of (G)\mathcal{E}(G) imply that the image of ff is finite; let {d1,,dm}\{d_{1},\ldots,d_{m}\}\subset\mathbb{Z} be the image of ff. The preimages Ei:=f1(di)E_{i}:=f^{-1}(d_{i}) are clopen sets that partition (G)\mathcal{E}(G); therefore, there exists a finite subset FF of GG such that each of the EiE_{i} are the accumulation points in (G)\mathcal{E}(G) of a component CiC_{i} of Cay(G)FCay(G)\smallsetminus F. Define xGx\in\mathbb{Z}^{G} such that x(g)=dix(g)=d_{i} for all gCig\in C_{i} and x(g)=0x(g)=0 otherwise. The boundary x\partial x is contained in the 11-neighborhood of FF, so x\partial x is finite, and f=x¯f=\bar{x}. By Lemma A.2, ϕx\phi_{x} is a 1-cocycle, and by construction, the image of [ϕx][\phi_{x}] is [f][f], establishing the surjectivity of the map.

We have established that the given map is an isomorphism, so it is left to check the that e(G)=1+rank(H1(G,G))e(G)=1+\mathrm{rank}(H^{1}(G,\mathbb{Z}G)). First suppose that |(G)|=n<|\mathcal{E}(G)|=n<\infty. Let (G)={ξ1,,ξn}\mathcal{E}(G)=\{\xi_{1},\ldots,\xi_{n}\} and define fiC((G),)f_{i}\in C(\mathcal{E}(G),\mathbb{Z}) by fi(ξi)=1f_{i}(\xi_{i})=1 and fi(ξj)=0f_{i}(\xi_{j})=0 if jij\neq i. Then C((G),)C(\mathcal{E}(G),\mathbb{Z}) is generated by {f1,,fn}\{f_{1},\ldots,f_{n}\}, and it is clear that this set of generators is \mathbb{Z}-linearly independent; in particular, C((G),)nC(\mathcal{E}(G),\mathbb{Z})\cong\mathbb{Z}^{n}. Therefore, the quotient C((G),)/Zn1C(\mathcal{E}(G),\mathbb{Z})/Z\cong\mathbb{Z}^{n-1}, yielding the desired result. Now, suppose that (G)\mathcal{E}(G) is infinite. For any nn\in\mathbb{N}, there exists pairwise-disjoint clopen sets P1,,Pn+1P_{1},\ldots,P_{n+1} of (G)\mathcal{E}(G) such that (G)=P1P2Pn+1\mathcal{E}(G)=P_{1}\cup P_{2}\cup\cdots\cup P_{n+1}. Arguing as in the finite end case, the functions fiC((G),)f_{i}\in C(\mathcal{E}(G),\mathbb{Z}) defined to be 1 on PiP_{i} and 0 on the complement of PiP_{i} generate a copy of n+1\mathbb{Z}^{n+1} in C((G),)C(\mathcal{E}(G),\mathbb{Z}), and therefore their equivalence classes generate a copy of n\mathbb{Z}^{n} in C((G),)/ZC(\mathcal{E}(G),\mathbb{Z})/Z. It follows that the rank of C((G),)C(\mathcal{E}(G),\mathbb{Z}) is infinite, yielding the desired result. ∎

It not relevant to the work that follows, but we note that as H1(G,G)H^{1}(G,\mathbb{Z}G) is a free abelian group (see [7, Theorem 13.5.3]), if (G)\mathcal{E}(G) is infinite, then H1(G,G)H^{1}(G,\mathbb{Z}G) is isomorphic to a countably infinite direct sum of copies of \mathbb{Z}.

We can now state Stallings’s theorem in terms of group cohomology.

Theorem A.5.

If GG is a finitely generated group such that H1(G,G)H^{1}(G,\mathbb{Z}G) is nontrivial, then GG admits an action on a simplicial tree TT with finite edge stabilizers, no edge inversions and no global fixed point. ∎

A.3. Swarup’s Theorem

Swarup [16] has proved a relative version of Stallings’s Theorem, which we now describe. Let GG be a finitely generated group, and let HGH\subset G be a subgroup. There is a natural restriction map r:H1(G,G)H1(H,G)r:H^{1}(G,\mathbb{Z}G)\to H^{1}(H,\mathbb{Z}G) given by

(4) r([ϕ])=[ϕ|H]r([\phi])=[\phi|_{H}]

Here, H1(H,G)H^{1}(H,\mathbb{Z}G) is the quotient of the set of all functions ϕ:HG\phi\mskip 0.5mu\colon\thinspace H\to\mathbb{Z}G satisfying ϕ(hh)=ϕ(h)+hϕ(h)\phi(hh^{\prime})=\phi(h)+h\phi(h^{\prime}) by the subset of such ϕ\phi that are of the form ϕ(h)=(h1)y\phi(h)=(h-1)y for some yGy\in\mathbb{Z}G.

Theorem A.6 (Swarup [16]).

Let GG be a finitely generated group, and let H1,,HmH_{1},\ldots,H_{m} be subgroups of GG. For each jj, let

rj:H1(G,G)H1(Hj,G)r_{j}\mskip 0.5mu\colon\thinspace H^{1}(G,\mathbb{Z}G)\to H^{1}(H_{j},\mathbb{Z}G)

denote the restriction map. If the intersection jker(rj)1\bigcap_{j}\mathrm{ker}(r_{j})\neq 1, then there is an action of GG on a simplicial tree TT, with finite edge stabilizers, no edge inversions, no global fixed point, and such that each HjH_{j} is contained in the stabilizer of some vertex pjTp_{j}\in T.

Equivalently, as stated in [16], the conclusion is that GACBG\cong A\star_{C}B or GACG\cong A\star_{C} in such a way that each HjH_{j} is conjugate into AA or BB. The following proposition interprets Swarup’s theorem in terms of end spaces, at least when the subgroups are finitely generated.

Proposition A.7.

Let HGH\subset G be finitely generated, and let rr be the restriction map of (4). Under the isomorphism H1(G,G)C((G),)/ZH^{1}(G,\mathbb{Z}G)\to C(\mathcal{E}(G),\mathbb{Z})/Z given by [ϕx][x¯][\phi_{x}]\longmapsto[\bar{x}] defined in Theorem A.4, the kernel of rr maps to the subset of C((G),)/ZC(\mathcal{E}(G),\mathbb{Z})/Z consisting of all functions f:(G)f:\mathcal{E}(G)\to\mathbb{Z} such that for each right coset HgHg, the function ff is constant on Hg¯(G)\overline{Hg}\cap\mathcal{E}(G).

Here, Hg¯\overline{Hg} is the closure of HgHg in G(G)G\cup\mathcal{E}(G). Given multiple subgroups HjH_{j} as in Swarup’s theorem, the intersection of the kernels is the set of all ff that are constant on Hjg¯(G)\overline{H_{j}g}\cap\mathcal{E}(G) for each jj and each right coset HjgH_{j}g.

Proof of Proposition A.7.

By Lemma A.2, any 11-cocycle with G\mathbb{Z}G-coefficients is of the form ϕx\phi_{x}, where xGx\in\mathbb{Z}^{G} has finite boundary. The map rr takes [ϕx][\phi_{x}] to the class in H1(H,G)H^{1}(H,\mathbb{Z}G) of the restriction ϕx|H:HG\phi_{x}|_{H}:H\to\mathbb{Z}G given by

ϕx|H(h)=(h1),\phi_{x}|_{H}(h)=(h-1),

which is trivial in H1(Hj,G)H^{1}(H_{j},\mathbb{Z}G) exactly when there is an element yGy\in\mathbb{Z}G such that ϕx(h)=ϕy(h)\phi_{x}(h)=\phi_{y}(h) for all hHh\in H. This last equality is equivalent to saying that h(xy)=xyh(x-y)=x-y for all hHh\in H, or saying that xyx-y is invariant under the left action of HH. So, we have that r([ϕx])=0r([\phi_{x}])=0 if and only if the function xx is left HH-invariant outside of a finite subset of GG.

If xx is left HH-invariant outside a finite subset of GG, then on each right coset HgHg, the function xx is constant outside of a finite subset, and hence the boundary map x¯\bar{x} is constant on Hg¯(G)\overline{Hg}\cap\mathcal{E}(G).

Conversely, suppose x¯\bar{x} is constant on Hg¯(G)\overline{Hg}\cap\mathcal{E}(G) for each gGg\in G. Fix a finite generating set SS for GG. As HH is finitely generated, we can pick mm\in\mathbb{N} large enough so that HH is generated by elements of HGH\subset G with word length at most mm in the generating set SS of GG. Note that any two points in a coset HgHg can then be joined in Cay(G)Cay(G) by a path that stays in the mm-neighborhood of HgHg. Call a coset HgHg relevant if it intersects the mm-neighborhood 𝒩m(x)\mathcal{N}_{m}(\partial x) of x\partial x, and call it irrevelant otherwise. As x\partial x is finite, there are only finitely many relevant cosets; moreover, each irrelevant coset lies in a single component of Cay(G)𝒩m(x)Cay(G)\smallsetminus\mathcal{N}_{m}(\partial x). Let FGF\subset G be the union of 𝒩m(x)\mathcal{N}_{m}(\partial x) and all intersections HgUHg\cap U that are finite sets, where HgHg is any relevant coset and UU is any component of Cay(G)𝒩m(x)Cay(G)\smallsetminus\mathcal{N}_{m}(\partial x). Since there are only finitely many UU and finitely many relevant HgHg, the set FF is finite.

We claim that xx is left HH-invariant on the complement of FF, meaning that for every right coset HgHg, the function xx is constant on HgFHg\smallsetminus F. If HgHg is irrelevant, then it lies in a single component of Cay(G)𝒩m(x)Cay(G)\smallsetminus\mathcal{N}_{m}(\partial x), and hence any two points in HgHg can be joined by a path avoiding x\partial x, implying that xx is constant on HgHg as desired. If HgHg is relevant, then since x¯\bar{x} is constant on Hg¯(G)\overline{Hg}\cap\mathcal{E}(G), it follows that on all components of Cay(G)𝒩m(x)Cay(G)\smallsetminus\mathcal{N}_{m}(\partial x) that HgHg intersects in an infinite set, the value of xx is the same constant. Hence xx is constant on HgFHg\smallsetminus F as desired. ∎

Combining Swarup’s theorem with Proposition A.7, we get Theorem 2.6.

References

  • [1] Javier Aramayona, Christopher J Leininger, and Alan McLeay, Big mapping class groups and the co-hopfian property, Michigan Mathematical Journal 1 (2023), no. 1, 1–29.
  • [2] Yasmina Atarihuana, Juan García, Rubén A Hidalgo, Saúl Quispe, and Camilo Ramírez Maluendas, Dessins d’enfants and some holomorphic structures on the Loch Ness monster, The Quarterly Journal of Mathematics 73 (2022), no. 1, 349–369.
  • [3] Taras Banakh, J Higes, and Ihor Zarichnyi, The coarse classification of countable abelian groups, Transactions of the American Mathematical Society 362 (2010), no. 9, 4755–4780.
  • [4] Ara Basmajian and Nicholas G. Vlamis, There are no exotic ladder surfaces, Ann. Fenn. Math. 47 (2022), no. 2, 1007–1023. MR 4453410
  • [5] Martin R. Bridson and André Haefliger, Metric spaces of non-positive curvature, Springer-Verlag Berlin Heidelberg, 1999.
  • [6] William Burnside, On an unsettled question in the theory of discontinuous groups, Quart. J. Pure and Appl. Math. 33 (1902), 230–238.
  • [7] Ross Geoghegan, Topological methods in group theory, vol. 243, Springer Science & Business Media, 2007.
  • [8] Martin E. Goldman, An algebraic classification of noncompact 22-manifolds, Transactions of the American Mathematical Society 156 (1971), 241–258.
  • [9] CR Hampton and Dl S Passman, On the semisimplicity of group rings of solvable groups, Transactions of the American Mathematical Society 173 (1972), 289–301.
  • [10] Bernard Maskit, A theorem on planar covering surfaces with applications to 3-manifolds, Annals of Mathematics 81 (1965), no. 2, 341–355.
  • [11] CD Papakyriakopoulos, Planar regular coverings of orientable closed surfaces, Ann. of Math. Stud 84 (1975), 261–292.
  • [12] Ian Richards, On the classification of noncompact surfaces, Transactions of the American Mathematical Society 106 (1963), no. 2, 259–269.
  • [13] Jean-Pierre Serre, Trees, Springer Monographs in Mathematics, Springer-Verlag, Berlin, 2003, Translated from the French original by John Stillwell, Corrected 2nd printing of the 1980 English translation. MR 1954121
  • [14] Peter B Shalen et al., Representations of 3-manifold groups, Handbook of geometric topology (2002), 955–1044.
  • [15] John Stallings, Group theory and three-dimensional manifolds, Yale Mathematical Monographs, vol. 4, Yale University Press, New Haven, Conn.-London, 1971, A James K. Whittemore Lecture in Mathematics given at Yale University, 1969. MR 415622
  • [16] G Ananda Swarup, Relative version of a theorem of stallings, Journal of Pure and Applied Algebra 11 (1977), no. 1-3, 75–82.