Essential surfaces in graph pairs

By Henry Wilton

Abstract

A well-known question of Gromov asks whether every one-ended hyperbolic group has a surface subgroup. We give a positive answer when is the fundamental group of a graph of free groups with cyclic edge groups. As a result, Gromov’s question is reduced (modulo a technical assumption on 2-torsion) to the case when is rigid. We also find surface subgroups in limit groups. It follows that a limit group with the same profinite completion as a free group must in fact be free, which answers a question of Remeslennikov in this case.

This paper addresses a well-known question about hyperbolic groups, usually attributed to Gromov.

Question 0.1.

Does every one-ended hyperbolic group contain a surface subgroup?

Here, a surface subgroup is a subgroup isomorphic to the fundamental group of a closed surface of nonpositive Euler characteristic. Various motivations for Gromov’s question can be given. It generalizes the famous Surface Subgroup conjecture for hyperbolic 3-manifolds, but it is also a natural challenge when one considers that the Ping-Pong lemma makes free subgroups very easy to construct in hyperbolic groups, whereas a theorem of Gromov, Sela, and Delzant Reference 17 asserts that a one-ended group has at most finitely many images (up to conjugacy) in a hyperbolic group. More recently, Markovic proposed finding surface subgroups as a route to proving the Cannon conjecture Reference 31.

Several important cases of Gromov’s question have recently been resolved. Most famously, Kahn and Markovic proved the Surface Subgroup conjecture Reference 23. Extending their work has been the topic of a great deal of recent research (see Reference 21Reference 28, for instance). In another dramatic development, Calegari and Walker answered Gromov’s question affirmatively for random groups Reference 10, following similar results for random ascending HNN extensions of free groups (Calegari and Walker Reference 11) and random graphs of free groups with edge groups of rank at least two (by Calegari and Wilton Reference 12).

In this paper we resolve Gromov’s question for a contrasting class of hyperbolic groups—graphs of free groups with cyclic edge groups. Our main theorem answers Gromov’s question affirmatively in this case.

Theorem A.

Let be the fundamental group of a graph of free groups with cyclic edge groups. If is one ended and hyperbolic, then contains a quasi-convex surface subgroup.

In fact, using a result of Wise, we are able to find surface subgroups in graphs of virtually free groups with virtually cyclic edge groups; see Theorem 6.1 below.

Numerous special cases of this result are already known. Calegari used his work on the rationality of stable commutator length in free groups Reference 9 to show that surface subgroups exist when Reference 7. Infinite classes of examples were found by the author in joint works with Gordon Reference 18 and with Kim Reference 27. Kim and Oum found surface subgroups in doubles of free groups of rank two Reference 26. The author answered a weaker version of Gromov’s question for this class of groups by showing that every such is either a surface group or contains a finitely generated, one-ended subgroup of infinite index Reference 42.

Although the class of hyperbolic groups covered by Theorem A is quite specific, the theorem has wider consequences for Gromov’s question. We call a group rigid if it does not admit a nontrivial splitting with a virtually cyclic edge group. Using strong accessibility Reference 29, we can, modulo a technical hypothesis on 2-torsion, reduce Gromov’s question to the rigid case, using the following corollary.

Corollary B.

Let be a one-ended hyperbolic group without -torsion. Either contains a quasi-convex surface subgroup or contains a quasi-convex rigid subgroup.

See Corollary 6.4 for full details. By a theorem of Bowditch Reference 4 a one-ended hyperbolic group is rigid if and only if its Gromov boundary does not contain local cutpoints (unless is a finite extension of a triangle group). Corollary B should be useful in any attempt at a general answer to Gromov’s question, since local cutpoints in the boundary present extra technical challenges for the ergodic techniques of Reference 23 and the probabilistic techniques of Reference 10, as witnessed by the difficulties resolved in Kahn and Markovic’s proof of the Ehrenpreis conjecture Reference 24.

A limit group is a finitely generated, fully residually free group—that is, a finitely generated group in which every finite subset can be mapped injectively into a free group by a group homomorphism. Limit groups play a central role in the study of algebraic geometry and logic over free groups; see Reference 36, and others in this series, in which they were defined, and also the parallel project Reference 25, and others in this series. Theorem A addresses the key case for the problem of finding surface subgroups of limit groups, and so we can also answer Gromov’s question in that context.

Corollary C.

Let be a limit group. If is one-ended, then contains a surface subgroup.

See Corollary 6.3 for details. Note that limit groups are not all hyperbolic, but they are all toral relatively hyperbolic Reference 1Reference 16. In particular, nonhyperbolic limit groups contain a subgroup, so the hyperbolic case is the one of interest.

These results have interesting applications to a different structural problem in group theory. Recall that the profinite completion, , of a group is the closure of the image of in the direct product of its finite quotients (endowed with the product topology). If two groups and have isomorphic profinite completions, then it is natural to ask whether and must be isomorphic.

In general, the anwer is “no”. There are even examples of nonisomorphic pairs of virtually cyclic groups with isomorphic profinite completions Reference 2. Nevertheless, many important questions of this type remain open, of which the following question of Remeslennikov is one of the most notable Reference 32, Question 15.

Question 0.2 (Remeslennikov).

Suppose that is a finitely generated, nonabelian free group and that is finitely generated and residually finite. If , does it follow that ?

It is particularly natural to consider Question 0.2 when is a limit group. Indeed, limit groups are closely related to free groups (for instance, Remeslennikov showed that they are precisely the existentially free groups Reference 34) and are frequently hard to distinguish from them. Bridson, Conder, and Reid Reference 6 pointed out that Corollary C, combined with the results of Reference 41, would resolve Remeslennikov’s question in this case.

Corollary D.

If is a limit group and not free, then the profinite completion is not isomorphic to the profinite completion of any free group.

The same ideas give a new proof of a theorem of Puder and Parzanchevski Reference 33, Corollary 1.5. Recall that a word in a free group is called primitive if splits as a free product . Similarly, an element of the profinite free group is called primitive if decomposes as a coproduct in the category of profinite groups. Puder and Parzanchevski showed that an element of that is primitive in the profinite completion is already primitive in Reference 33, Corollary 1.5. This can be thought of as answering a relative version of Question 0.2.

In fact, we can generalize their result, from words to multiwords (i.e., finite indexed sets of words). Let us call a multiword in primitive if splits as a free product for some and make the corresponding definition of a primitive multiword in .

Corollary E.

Let be a finitely generated free group. If a multiword is primitive in the profinite completion then it is primitive in .

It is interesting to contrast the techniques of this paper with those of Reference 33. Puder and Parzanchevski deduce their result from their beautiful characterization of primitive words in free groups as precisely the measure-preserving words Reference 33, Theorem 1.1. The proof given here is cohomological and goes via the fact that the virtual second cohomology of the profinite completion of a nonfree limit group is nonzero. Corollaries D and E follow quickly from Theorem 7.1. We refer the reader to that theorem and the subsequent remarks for details.

Let us now turn to discuss the proof of Theorem A. Our main technical result addresses a relative version of Gromov’s question, finding surfaces in free groups relative to families of cyclic subgroups. To state it concisely, we need to introduce some definitions.

Consider a graph of spaces in the sense of Scott and Wall Reference 35, and let be a vertex with incident edges . The vertex space , together with the maps of incident edge spaces , defines a space pair . In the case of interest, the vertex space will always be a graph (usually denoted by ), and the edge spaces will be circles ; such a is called a graph pair.

Global properties of the graph of spaces can be characterized locally, using properties of the pairs associated to vertex spaces. For instance, is called irreducible if does not split over a finite subgroup; we may correspondingly define an irreducible pair (see Definition 2.9), and a lemma of Shenitzer asserts that if the pairs associated to the vertices are irreducible, then so is Reference 42, Theorem 18. Corresponding to the notion of a -injective map of graphs of spaces , we have an essential map of pairs , and indeed if a morphism of graphs of spaces is essential on each space pair associated to a vertex, then the morphism is itself -injective (see Proposition 1.7).

We are now ready to state the main technical result.

Theorem F.

If is an irreducible graph pair, then there is a compact surface with boundary and an essential map of pairs .

In fact, we obtain a bit more control than this—the surface is also admissible, meaning that every point of the domain of has the same number of preimages in ; see Theorem 5.11. It has been well known for a while that a result like Theorem F would imply the existence of surface subgroups in graphs of free groups with cyclic edge groups; see, for instance, Reference 7 or Reference 27.

Let us now briefly sketch the proof of Theorem F. It can be thought of as a combination of the techniques of Reference 9 and Reference 42.

First, we study irreducible pairs that map into the irreducible pair . The irreducibility of the pairs and is characterized using Whitehead graphs. We would like to study essential maps , but it turns out to be difficult to simultaneously characterize both the fact that is irreducible and the fact that the map is essential. In order to recognize both these properties simultaneously, we work with -immersions, which are compositions . We can recognize if the pair is locally irreducible, and this guarantees that is (weakly) irreducible.

The idea behind the proof of Theorem F is that, among all irreducible pairs mapping to , the pairs of surface type should be the ones of most negative Euler characteristic. To make this precise, we define a positive polyhedral cone in a finite-dimensional vector space, such that the integer points in correspond to admissible -immersions of (weakly) irreducible pairs . We also define the projective -rank function on the projectivization as a quotient of two linear functionals: the (negation of the) Euler characteristic of and the degree with which covers . In particular, achieves its maximum value at some vertex of the polyhedron , which is necessarily a rational line in . Since this rational line contains an integer point, an admissible -immersion of an irreducible pair exists that maximizes . We call such a pair maximal.

This approach is similar to the argument of Reference 9, in which a polyhedral cone is defined whose integer points correspond to certain maps of surfaces . The hypothesis in Reference 7 that rational second homology is nonzero is needed to ensure that this cone is nonzero. In contrast, the cone is guaranteed to be nonzero since, whenever is irreducible, the identity map leads to an admissible -immersion of an irreducible pair.

The final step of the proof applies the ideas of Reference 42 to the relative JSJ decomposition of a maximal pair . The conclusion is that any maximal pair has no rigid vertices in its JSJ decomposition. It follows that the JSJ decomposition is built from surface pieces, and one quickly concludes that a pair of surface type exists. Thus, we deduce the existence of an admissible, -essential surface .

The paper is structured as follows. In Section 1 we define pairs of groups, spaces and graphs, the natural notions of maps between them, and various properties of those maps. In Section 2 we adapt the classical theory of Whitehead graphs to the setting of a graph pair . The main result here is a converse to Whitehead’s lemma (Lemma 2.11), which asserts that an irreducible pair can always be unfolded to a locally irreducible pair, in which the irreducibility is recognized by the Whitehead graphs at the vertices. In Section 3 we characterize admissible -immersions from locally irreducible graph pairs into as precisely those maps that can be built from a certain finite set of pieces. We define the cone and note that there is a surjective map from admissible -immersions of locally irreducible pairs to the integer points of . In Section 4 we define the projective -rank function and prove that it attains its extremal values at rational points of . We deduce the existence of a maximal, admissible -immersion from a locally irreducible pair. In Section 5 we apply the results of Reference 42 to study admissible -immersions with maximal projective -rank. The main result is that there is such a maximal pair of (weak) surface type (Theorem 5.6). Theorem 5.11, and hence Theorem F, follow quickly. In Section 6 we deduce Theorem A and Corollaries B and C. Finally, in Section 7 we deduce Corollaries D and E.

1. Pairs

We will make heavy use of graphs of groups and Bass–Serre theory, as detailed in Serre’s standard work on the subject Reference 37 to which the reader is referred for details. To fix notation, we recall the definition of a graph.

Definition 1.1.

A graph consists of a vertex set , an edge set , a fixed-point free involution denoted by , and an origin map . The terminus map is defined by .

The edges of are thus equipped with orientations, and the unoriented edges are the pairs .

As well as using graphs of groups, we will also frequently adopt the topological point of view, in which a graph of groups is viewed as the fundamental group of a graph of spaces Reference 35. Graphs of spaces are not required to be connected, which will present some technical advantages, although the attaching maps are required to be injective on fundamental groups. Analogously, we may also work with disconnected graphs of groups, as long as we are careful to choose a basepoint before talking about the fundamental group.

1.1. Group pairs

It is particularly important for us to work with relative versions of graphs of groups and spaces, which characterize the relationship between a vertex group (or space) and its incident edge groups (or spaces). To this end we define various notions of pairs. We start with pairs of groups.

Definition 1.2.

A group pair is a pair , where is a group and is a -set. It is often convenient to choose a finite set of orbit representatives , to let , and to specify the pair via the data . We will use both the notations and to specify group pairs, without fear of confusion.

The key example of a group pair arises when considering a vertex of a graph of groups . Having fixed a lift of to the Bass–Serre tree, one takes to be the vertex stabilizer and to be the set of edges incident at .

Definition 1.3.

A morphism of graphs of groups is a morphism of the underlying graphs, accompanied by associated maps of vertex groups and edge groups that intertwine with the attaching maps. This is most easily thought of by passing to the Bass–Serre tree. A morphism of graphs of groups induces a homomorphism of fundamental groups and lifts to an equivariant map on Bass–Serre trees.

This motivates the following definition for pairs.

Definition 1.4.

A morphism of group pairs consists of a set map and a homomorphism that intertwines . That is, we require that

for all and .

In particular, a morphism of graphs of groups defines morphisms of the group pairs at each vertex, and, conversely, morphisms of pairs that satisfy an obvious compatibility condition can be pieced together to give a morphism of a graph of groups.

It is convenient if we can detect global properties of morphisms of graphs of groups by looking at local properties of the induced maps on group pairs. We are particularly concerned with -injectivity, and so we need to develop corresponding notions for group pairs. Requiring that the map of groups be injective is clearly significant. The following condition is also important.

Definition 1.5.

A morphism of group pairs is -essential if the map

induced by is injective.

When applied to pairs associated to graphs of groups, this condition guarantees that the induced map on Bass–Serre trees does not factor through a fold. Putting this together with injectivity, we have the notion of an essential morphism.

Definition 1.6.

A morphism of group pairs is essential if the homomorphism is injective and the morphism is also -essential.

From this, one easily deduces a local criterion for morphisms of graphs of groups to be -injective.

Proposition 1.7.

Suppose that is a morphism of graphs of groups. If induces essential morphisms on the group pairs corresponding to vertices, then induces an injective map on fundamental groups.

Proof.

Suppose that a group element is in the kernel of . Since the map on Bass–Serre trees does not factor through a fold, if acts hyperbolically on the Bass–Serre tree, then so does its image, contradicting the fact that is in the kernel. Therefore is elliptic, but since is injective on vertex stabilizers, it follows that .

1.2. Space pairs

We next make analogous definitions for spaces.

Definition 1.8.

A space pair consists of cell complexes together with a continuous map . We will frequently take to be an index set and we let

denote the restriction of to , the path component of corresponding to . We will often use the notation to denote such a space pair.

In most of what follows, we will take to be a graph and to be a disjoint union of circles. However, it is useful to allow the extra flexibility of the general definition.

If is path connected, then a space pair naturally defines a group pair . Let be the universal cover, and consider the fiber product

Taking and , we see that acts naturally on , and so is a group pair.

This definition is more transparent if one thinks of as a vertex space of a graph of spaces . The universal cover of inherits a decomposition as a graph of spaces; appears as a vertex space of , and the fiber product is the disjoint union of the edge spaces of incident at .

We next define morphisms of space pairs analogously to morphisms of group pairs.

Definition 1.9.

Let and define space pairs and , respectively. A morphism of space pairs consists of continuous maps and so that .

As in the case of groups, compatible collections of maps of pairs can be glued together to construct a map of graphs of spaces. Again, we will need a definition of a -essential morphism.

Definition 1.10.

Consider a morphism of space pairs . Let be the universal cover of and let be the corresponding covering space of , obtained by pulling back the covering map along . Consider the fiber products

and note that the map lifts to a map . The morphism is called -essential if induces an injective map .

Again, we combine this with injectivity on vertex groups to obtain a notion of an essential morphism.

Definition 1.11.

A morphism of space pairs is essential if it is -essential and is -injective.

Finally, we note that our two definitions of -essential pairs coincide.

Lemma 1.12.

Let be a morphism of space pairs, inducing the corresponding morphism of pairs on fundamental groups. The morphism is -essential if and only if the morphism is -essential. Hence, is essential if and only if is essential.

Proof.

The quotient of by the action of is , and the corresponding covering map induces a map .

We need to show that two path components and of have the same image under this map if and only if they are in the same orbit of . The “if” direction is clear. For the converse we choose compatible basepoints and suppose that and have the same image. Then the images of their basepoints in are joined by a concatenation of paths , where maps to a loop in and is the image of a lift of a loop from to . These define group elements and so that translates to , and so and are indeed in the same orbit of .

The lemma follows immediately.

1.3. Graph pairs

In the setting of Theorem A, the groups are finitely generated free groups, so the spaces can be taken to be graphs. We may therefore apply the techniques of Stallings Reference 38.

Definition 1.13.

Let be a graph. The star of a vertex is the set , the set of edges with initial vertex . (We will also write for when there is no fear of confusion.) A morphism of graphs is an immersion if the induced maps on stars are injective. In this case, we write .

Stallings famously observed that immersions are -injective and that any morphism of finite graphs factors through a canonical immersion

where the map is a composition of finitely many folds Reference 38, §§3.3.

Definition 1.14.

A multicycle in a graph is an immersion of graphs , where is a disjoint union of graphs homeomorphic to circles. The components of are denoted by and the restriction of to is denoted by .

A graph pair is a space pair , where is a finite graph without vertices of valence one and is a multicycle. Note that we do not require the graph to be connected.

Again, we will need a notion of morphism for graph pairs. As for Stallings, for us a morphism of graphs takes vertices to vertices and edges to edges. Since we insist that the maps are immersions, we make a corresponding requirement for morphisms of graph pairs.

Definition 1.15.

Let and be graph pairs. A morphism of space pairs is a morphism of graph pairs if the map is a morphism of graphs and the map is an immersion.

The first advantage of this setting is that we can certify -injective maps using immersions. Note that a morphism of graphs is an immersion if and only if the lift to universal covers is injective. Similarly, we may define an immersion of graph pairs.

Definition 1.16.

A map of graphs pairs is an immersion if the lifts and are injective. In this case, we write .

A map of graph pairs factors through a canonical immersion, just as maps of graphs do.

Lemma 1.17.

A map of graphs pairs factors through a canonical immersion . The immersion has the universal property that, whenever factors through an immersion , also factors through .

Proof.

Let be the image of in . We take to be the image of in ; likewise, we take to be the image of in . The group acts naturally on each of these, and we take and to be the respective quotients by the action of .

The next lemma provides a means of locally certifying that a map is -essential.

Lemma 1.18.

If a map of graph pairs is -essential and is the corresponding canonical immersion, then the induced map is injective. Conversely, if a map of graph pairs factors through an immersion as

and is injective, then is -essential.

Proof.

By definition, if is -essential, then the corresponding map is injective on . The components of these spaces are lines, so the map is itself injective, and so is injective too. Finally, since is obtained by quotienting the domain and the range by , it is also injective.

For the converse, if factors as hypothesized, then the lift of factors as

The first map is a lift of an injection, hence an injection, and the second map is a lift of an immersion, hence injective. The result follows since a composition of injective maps is injective.

Thus, we can use a map to an immersed pair as a certificate that a morphism of pairs is -essential. We call the data of this certificate a -immersion.

Definition 1.19.

A -immersion is a concatenation

where is bijective.

2. Whitehead graphs and folds

Given a group pair , where is some nontrivial element of a free group , it is natural to ask ask whether or not is a free factor of . The standard way of answering this question uses the Whitehead graph, which was defined by J. H. C. Whitehead in his original paper on automorphisms of free groups Reference 40. (See also Reference 14 and the references therein for a modern account of Whitehead graphs.) The definition of the Whitehead graph given in Reference 40 implicitly involves representing as the fundamental group of a rose—a graph with a single vertex.

Here we develop the theory of Whitehead graphs for general graph pairs. We are not aware that this approach has been taken in the literature before, but it is similar to the approaches to Whitehead graphs given by Cashen and Macura Reference 14 and Manning Reference 30.

Definition 2.1.

Consider a graph pair and a vertex of . The Whitehead graph at is denoted by . Its set of vertices is the star . The unoriented edges of are the vertices of that map to ; the edge corresponding to joins the vertices and of , where and are the two edges of with .

Note that the requirement that the multicycle is an immersion implies that the endpoints of any edge of are distinct. However, each pair of vertices may be joined by many edges.

We can collect together all the Whitehead graphs at the vertices of into a global Whitehead graph for the pair . Figure 1 shows an example of a Whitehead graph.

Definition 2.2.

The Whitehead graph of the pair is the disjoint union

Note that comes equipped with two additional structures:

(i)

the components of are naturally partitioned: two components are equivalent if they are both components of some ;

(ii)

the fixed-point free involution on the edges of defines a fixed-point free involution on the vertices of that extends to a bijection .

We will always think of as equipped with these extra structures.

Remark 2.3.

The partition on the components of and the involutions are enough information to reconstruct the pair .

Stallings studied morphisms of graphs by observing that they always factor as a composition of a sequence of folds followed by an immersion. Recall that a fold identifies a pair of edges with . It is therefore natural to study the effect that a fold has on Whitehead graphs.

Definition 2.4.

Let be a graph, and let be a pair of vertices. (When we apply this, will be a disjoint union of Whitehead graphs.) A wedge is a quotient map that identifies and and leaves the rest of unchanged. We write for the quotient graph . If with then we write for . The reverse move, which replaces by , is called an unwedge.

The following lemma shows that, at the level of Whitehead graphs, folds correspond to wedges.

Lemma 2.5.

Let be the morphism given by a fold which identifies two edges of with a common initial vertex to an edge of . Let and let and . Suppose further that is a homotopy equivalence, i.e., . Then

(Note that the unions in this expression may not be disjoint, since may equal for at most one , in which case also equals In particular, is a cut vertex of

Proof.

This follows immediately from the definitions. (The case in which is illustrated in Figure 2.)

Remark 2.6.

In the above lemma the hypothesis that the map is a morphism of pairs is essential: if the induced map were not an immersion, then after folding one would need to tighten to an immersion, which might destroy the cutpoint structure of the Whitehead graphs.

Remark 2.7.

In the setting of Lemma 2.5, for any vertex of , the map induced by is injective on edges.

In fact, the implication of Lemma 2.5 can be reversed: if one of the Whitehead graphs has a cut vertex, then we can unfold.

Lemma 2.8.

Let be a graph pair, and suppose that some edge defines a cut vertex in (where . Then there is a graph pair and a morphism of pairs defining a homotopy-equivalent fold that identifies a pair of edges to .

Proof.

The hypothesis tells us that , where is a vertex with image in the wedge. Let (and note that and are not necessarily distinct). We will define the pair via its Whitehead graphs (appealing to Remark 2.3). The proof divides into two similar cases, depending on whether or not .

Suppose first that . For any vertex of not equal to or , we take a vertex for with Whitehead graph isomorphic to . The remaining vertices of are denoted by . We define to be for . Finally, is defined so that

That is, is obtained from by dividing the vertex into two vertices, . The edges of incident at and are defined so that they respect the natural bijections between the stars of the and the stars of the . There is then a natural lift of the bijections on stars in to bijections on stars in , and this completes the construction of .

The case in which is similar. Again, for any vertex of not equal to or , we take a vertex for with Whitehead graph isomorphic to . The remaining vertices of are denoted by . Since , the vertex is contained in and, without loss of generality, we may take . We now define and so that

As before, this means that is obtained by dividing the vertex into two vertices , and the edges of incident at the are defined to respect the natural bijections between the stars of the and the stars of the . Again, this completes the construction of .

In either case identifying and defines a fold . Since , the fold is a homotopy equivalence. Note also that, by construction, the fold is a morphism of pairs.

Whitehead introduced Whitehead graphs to recognize basis elements of free groups and, more generally, free splittings. (More generally still, Whitehead gave an algorithm to find the shortest element in an orbit of the automorphism group.) We will use Whitehead graphs to recognize (weakly) irreducible pairs.

Definition 2.9.

Consider a graph pair with a finite connected graph. By Grushko’s theorem, splits canonically as

where for each , there is an index set so that is conjugate into for all , each does not split freely relative to the set , and no is conjugate into . A factor is called cyclic if is a singleton and, up to conjugacy, generates .

The pair is called weakly irreducible if there are no cyclic factors; otherwise it is called strongly reducible. The pair is called reducible if it is weakly irreducible and ; otherwise it is called reducible. When is disconnected, the pair is called (weakly) irreducible or (strongly) reducible if and only if each component has that property.

The point of the above definition is that the pair is reducible if and only if the fundamental group of the double , obtained as a graph of spaces with two vertex spaces homeomorphic to and edge maps given by , admits a nontrivial free splitting. Cyclic factors are relevant because they give rise to factors of the double. Even a single cyclic factor is reducible, since splits as an HNN extension of the trivial group.

The following lemma is the key result for recognizing reducible pairs. It is quite standard, but we give a proof using folds and wedges as a sample application of the above ideas.

Lemma 2.10 (Whitehead).

If is reducible, then there is a vertex of so that one of the following holds:

is disconnected;

has a leaf, i.e., a vertex of valence ; or

has a cut vertex, i.e., a vertex so that is disconnected.

Proof.

The case in which is cyclic and is a -isomorphism is easy and is left as an exercise. Suppose therefore that admits a free splitting relative to . It follows that there is a morphism of graphs which is a homotopy equivalence, so that has a vertex with disconnected. The morphism now factors as a sequence of homotopy-equivalent folds; in particular, whenever and with are identified, we have . Consider the final such fold, which identifies a pair of distinct vertices to a vertex . Lemma 2.5 implies that has a cut vertex.

Motivated by Whitehead’s lemma, we call a Whitehead graph reducible if it satisfies one of the three conclusions of the lemma; otherwise, we call irreducible. We call the pair locally irreducible if, for every vertex of , the Whitehead graph is irreducible. Whitehead’s lemma therefore says that a locally irreducible pair is irreducible.

The converse to this statement is not quite true—there are irreducible pairs that are not locally irreducible. To construct an example, take an irreducible pair and apply one fold. We therefore must allow ourselves to unfold in order to prove a converse to Whitehead’s lemma.

Lemma 2.11 (Converse to Whitehead’s lemma).

If a pair is irreducible, then there is a locally irreducible pair and a map of pairs which is a homotopy equivalence.

Proof.

If some vertex of has either a disconnected Whitehead graph or a leaf, then is reducible. Suppose therefore that there is a vertex so that has a cut vertex . By Lemma 2.8 there is a fold so that the edge  is unfolded to a pair of edges .

Note: the number of edges of is equal to the number of edges of ; the number of vertices of is greater than the number of vertices of ; for each vertex of , the Whitehead graph is connected without leaves. In particular, the number of vertices of is at most the number of edges of . It follows that only finitely many unfoldings of this form can be performed.

When no further unfoldings can be performed, the final pair is locally irreducible, as claimed.

The final lemma of this section shows that we can use locally irreducible -immersions to recognize weakly irreducible immersions.

Lemma 2.12.

If is a -immersion and is locally irreducible, then the pair is weakly irreducible.

Proof.

By repeatedly applying Lemma 2.8 as in the proof of Lemma 2.11, there is a homotopy-equivalent morphism of pairs so that every Whitehead graph of has no cut vertices. Since is locally irreducible, the morphism lifts to a morphism ; note that this map is bijective on edges of Whitehead graphs and surjective on vertices. If some component of a Whitehead graph of had at most one edge, a component of a Whitehead graph of that mapped to it would also, contradicting the hypothesis that is locally irreducible.

3. Admissible -immersions

Consider an irreducible graph pair as above. In this section we will study -immersions

where is a locally irreducible pair. We will impose one additional condition on our -immersions.

Definition 3.1.

A map of graph pairs is called admissible if there is a positive integer so that every point in has exactly preimages in . A -immersion is called admissible if the composition is admissible.

Calegari uses the term “admissible” similarly for maps of surfaces with boundary Reference 8, p. 37. Note that, in his context, the integer counts preimages with a sign determined by orientation, whereas in our context, the count is unsigned.

We next write down a finite set of pieces, from which -immersions of locally irreducible pairs can be constructed.

Definition 3.2.

Let be the set of components of . The set of pieces (over consists of all pairs of maps of graphs (up to graph isomorphism)

such that is a disjoint union of irreducible graphs, the map is bijective on edges, the map is injective, and . When is an element of , we will often abuse notation and write , since the map determines .

Remark 3.3.

Note that is finite. This trivial observation is of crucial importance.

We will study admissible -immersions of locally irreducible pairs by looking at how they are constructed from the pieces . The involutions on the stars of vertices of define relations on the elements of , as follows.

Definition 3.4.

Consider

elements of . Suppose that is a vertex of and that is a vertex of . Let be the set of vertices of that map to , and let be the set of vertices of that map to . We write

if the following hold:

(i)

; and

(ii)

up to reordering of indices, restricts to bijections for all .

The relation can be interpreted in terms of Manning’s splicing operation Reference 30. It says that and can be spliced at the sets of vertices and and that and can be spliced at and .

To record how the elements of are glued together, we introduce -stars.

Definition 3.5.

A -star consists of the following data:

(i)

a piece in ;

(ii)

for each vertex of , a choice of piece in and a vertex of so that .

We write and, for each vertex of , we write .

Let be the (finite) set of all -stars. Let and let be the nonnegative orthant. An admissible -immersion

of a locally irreducible pair defines an integer vector in a natural way, as follows. For each vertex of , let be the preimages of in and let be the image vertex in . The piece is then defined to be

If is an edge of with and , then . Therefore, for each vertex of , we can define to be the corresponding -star associated to the labels of the neighboring vertices:

(i)

; and

(ii)

for each , .

Now define to be the vector so that

for each .

The image of is not arbitrary—in fact, it is precisely the set of integer points of a certain cone . We will describe this cone using systems of equations: the gluing equations and the admissibility equations. We start with the gluing equations. A nonnegative integer vector that satisfies the gluing equations necessarily comes from a -immersion of a locally irreducible pair.

Definition 3.6.

We write for an element of . For each pair of pieces and the edge satisfying we have the gluing equation

where each sum is taken over all -stars satisfying the conditions.

We next describe the admissibility equations, which force any -pair that defines a vector to be admissible.

Definition 3.7.

Let be an edge of . For a piece

in , set to be the number of preimages of in . (Note that this is either 0 or 1, by definition.) We now define a linear map by setting

The admissibility equations assert that for all edges and in .

The cone is now defined to be the subset of that satisfies the gluing equations and the admissibility equations.

Lemma 3.8.

An integer vector is the image of an admissible -immersion from a locally irreducible pair under if and only if it is in .

Proof.

Let be locally irreducible, and let

be a -immersion. First we show that satisfies the gluing equations. Indeed, the expression in the gluing equations is just two different ways of evaluating the number of edges of with and . The admissibility equations are satisfied since each evaluates to .

Conversely, given an integer vector , we need to construct an admissible -immersion

with locally irreducible. By Remark 2.3, it is enough to describe

together with their pairings on stars of vertices. For each star , contains copies of the piece . This determines the graphs and maps ; it remains to determine the pairings. Consider the pieces

and suppose that is an edge of and is an edge of . The gluing equations imply that there is a bijection between the number of -stars so that and and the number of -stars so that and which satisfy the condition that the bijection restricts to a bijection , and thence to bijections of the stars of the preimages in and . Choosing a bijection between these -stars then determines the required bijection between vertices of the copies of and in these -stars, and the bijections between stars are then determined by the relation .

By construction, is locally irreducible and

is a -immersion. Finally, the admissibility equations immediately imply that this -immersion is admissible.

Thus, we have seen that admissible -immersions of locally irreducible pairs correspond naturally to nonzero integer vectors in or, equivalently, to rational points in the projectivization . Motivated by Calegari’s work on stable commutator length (see Reference 9 and Reference 8, and also Reference 5), we will study these rational points via rational functions on .

4. The rationality theorem

We start by writing down two natural linear maps on . For an admissible -immersion

the corresponding linear maps are (minus) the Euler characteristic of , and the degree with which covers . The key observation is that both of these can be computed from the vector .

First, the admissibility equations imply that the linear map is independent of . We therefore write , evidently a linear map which is nonzero on .

Second, for a piece

in , we let denote the number of connected components of , and we let denote the number of vertices of . We then define by

It is well known that the Euler characteristic of a graph can be computed as the sum over the vertices of one minus half the valence, and from this we see that, for an admissible -immersion

of a locally irreducible pair , we have .

Definition 4.1.

Since is a quotient of two linear maps on and the denominator is nonzero, it yields a well-defined function on the projectivization . We call this function

the projective -rank function on .

In analogy with stable commutator length (see Reference 8), we may use the rational function to define an invariant of a multicycle in a graph .

Definition 4.2.

The maximal -rank of a pair is denoted by and is defined to be

Note that this maximum is indeed realized, since is compact. Similarly, the minimal -rank, , is defined to the minimum of over the same domain.

Since is compact, the maximal and minimal -ranks are certainly attained as long as is nonempty (i.e., as long as is nonzero). In fact, since is a quotient of linear maps, the maximal and minimal -ranks are attained on rational points of and, hence, are realized by admissible -immersions.

Theorem 4.3.

If , then the maximal and minimal -ranks are realized by admissible -immersions of locally irreducible pairs; that is, there exist locally irreducible pairs and admissible -immersions

so that . In particular, are positive rational numbers.

Proof.

We prove the result for the maximal -rank; the proof for the minimal -rank is identical. If then the projectivization is nonempty. From the definition of , we may normalize and restrict our attention to the rational polytope , so

But is linear and so attains its maximum on a vertex of . Since is rational and is a projective function, there is some integer vector , a multiple of , on which attains its maximum. Since is an integer vector, it is equal to for some admissible -pair , which therefore realizes , as required.

An admissible -immersion

of a locally irreducible pair for which is called maximal. (Similarly, if then the -immersion is called minimal.)

5. Maximal -rank and surfaces

Our results so far imply that every irreducible pair admits a maximal -immersion.

Lemma 5.1.

If is irreducible, then there exists a maximal, admissible -immersion

for a locally irreducible pair .

Proof.

Since is irreducible, Lemma 2.11 guarantees a locally irreducible pair . The map consists of a -isomorphism and a homeomorphism , so it is certainly admissible and essential. In particular,

is an admissible -immersion of a locally irreducible pair, so and . Theorem 4.3 now implies that a maximal -immersion exists.

In this section, we shall use the relative JSJ decomposition together with the results of Reference 42 to show that maximal -immersions are closely related to surfaces.

Definition 5.2.

A group pair is said to be of surface type if it arises as the fundamental group of a space pair , where is a compact surface with boundary. It is said to be of weak surface type if it is a free product of pairs of surface type. A graph pair is of (weak) surface type if the corresponding group pair is of (weak) surface type.

A theorem of Culler Reference 15 shows that any pair of surface type can be unfolded to a fatgraph —a graph pair in which every Whitehead graph is a cycle. One may therefore equivalently think of pairs of surface type as given by fatgraphs. Likewise, a pair of weak surface type can be unfolded to a graph pair in which every Whitehead graph is a disjoint union of cycles.

Fundamental groups of pairs of surface type can typically be decomposed as graphs of groups in many ways. In order to discuss this, we introduce some terminology for graph-of-groups decompositions of pairs.

Definition 5.3.

Let be a group pair. A decomposition of is a graph of groups with fundamental group such that, for every , the stabilizer is conjugate into a vertex group of .

Let be a vertex of , and fix a preimage of in the Bass–Serre tree . Let be the stabilizer of . Set

and let be the set of edges of incident at . The induced pair at is defined to be , which is defined up to conjugacy in . The vertex is called peripheral if is nonempty.

If every edge group of is cyclic, then is said to be a cyclic decomposition of . As usual, the graph of groups is called trivial if is the stabilizer of some vertex of the Bass–Serre tree.

We will only be concerned with cyclic decompositions of graph pairs , with . We will abuse notation and write the corresponding group pair as .

Pairs of surface type can be contrasted with rigid pairs, which only have trivial decompositions.

Definition 5.4.

An irreducible graph pair is rigid if every cyclic decomposition of is trivial and if is not of surface type. (This last requirement is to rule out the pair of pants, which is of surface type but admits no cyclic decompositions.) A group pair is rigid if some (any) corresponding graph pair is rigid.

The main theorem of this section is phrased in terms of the relative JSJ decomposition of the group pair . This is a canonical decomposition of the pair , which in a sense encodes all cyclic decompositions. The absolute version of this decomposition was described in the hyperbolic case by Bowditch Reference 4; the relative version in the free case was described by Cashen Reference 13. See also the work of Guirardel and Levitt Reference 20, who explain how to construct this JSJ decomposition as a tree of cylinders.

Theorem 5.5 (Relative JSJ decomposition).

Let be an irreducible group pair. There is a canonical cyclic decomposition for with the following properties.

The underlying graph of has three kinds of vertices—rigid, surface and cyclic—such that the following hold:

if a vertex is of rigid type, then the induced pair at is a rigid group pair;

if a vertex is of surface type, then the induced pair at is of surface type;

if a vertex is cyclic, then the vertex group is (infinite) cyclic.

The underlying graph of is bipartite, with red vertices cyclic and green vertices either rigid or surface. In particular, every edge adjoins exactly one cyclic vertex.

Every peripheral subgroup is conjugate into a unique cyclic vertex group. (These cyclic vertices are called peripheral.)

The decomposition guaranteed by the theorem is called the relative JSJ decomposition of the pair . For an irreducible graph pair , we will refer to the disjoint union of the relative JSJ decompositions of the fundamental groups of the components as the relative JSJ decomposition of the pair .

We are now ready to state the main theorem of this section, which describes the relative JSJ decompositions of maximal -immersions.

Theorem 5.6.

If is locally irreducible and

is a maximal, admissible -immersion, then for each irreducible free factor of the corresponding group pair , the relative JSJ decomposition of has no rigid vertices. Furthermore, if there is such a maximal -immersion, then there is a maximal, admissible -immersion so that is of weak surface type.

The proof is based on ideas from Reference 42; the main technical result of that paper is as follows Reference 42, Theorem 8.

Theorem 5.7.

If is a rigid graph pair, then there is a finite-sheeted cover such that, whenever a finite-sheeted cover factors through , the pair is irreducible for any component of .

We will also need a relative analogue of Shenitzer’s lemma (see, for instance, Reference 39, Corollary 1.1), which we state here in the terminology of this paper.

Lemma 5.8 (Relative Shenitzer’s lemma).

Consider a decomposition of a group pair . If the induced pair at every vertex of is irreducible, then the pair is irreducible.

We now assemble the lemmas that we will need to prove Theorem 5.6. The first shows how to use a rigid vertex to increase irreducible rank. Its proof is illustrated in Figure 3.

Lemma 5.9.

Let be an irreducible graph pair. If the relative JSJ decomposition of has a rigid vertex, then there is a locally irreducible pair and an admissible, essential map with

Proof.

Consider the relative JSJ decomposition of the pair . By Theorem 5.7 and Marshall Hall’s theorem Reference 22, after replacing with a finite-sheeted cover, we may assume that every rigid vertex has the property guaranteed by Theorem 5.7. Note that, if denotes the valence of the vertex , then every rigid vertex has .

To construct , we realize the relative JSJ decomposition of as a graph of spaces . We may take the vertex spaces of to be graphs and the edge spaces to be circles, although this is not important for the subsequent argument. We now construct a new graph of spaces and an essential map .

Let , where the product is taken over all rigid vertices of . For each nonrigid vertex , we take copies of the induced pair at . Consider a rigid vertex , with incident edges . For each edge incident at , we take copies of the pair . Note that every edge space of appears exactly times in this collection of pairs. We may therefore glue up the resulting collection of pairs to form a graph of spaces .

By construction, is naturally equipped with a map , which is -injective by Proposition 1.7. Furthermore, is naturally equipped with exactly copies of component of ; we call this map of circles . Let be the disjoint unions of the cores (in the sense of Reference 38) of the covers of corresponding to the components of . We may realize as a collection of cycles in . Then is an admissible, essential map. The pair is irreducible by Lemma 5.8, and hence by Lemma 2.11, can be unfolded to a locally irreducible pair .

Finally, we compute Euler characteristics. We have , while

where ranges over all the rigid vertices of and ranges over all the nonrigid vertices. This completes the proof.

A similar argument shows that there are always maximal pairs of surface type.

Lemma 5.10.

Let be an irreducible graph pair. If the relative JSJ decomposition of has no rigid vertices, then there is a locally irreducible group pair of surface type and an admissible, essential with

Proof.

Consider the JSJ decomposition of the pair . By Marshall Hall’s theorem, we may assume that the attaching maps at cyclic vertices are all isomorphisms.

We realize as a graph of spaces in the natural way, taking each surface vertex to be a compact surface and each cyclic vertex to be a circle, and we define a new graph of spaces as follows. We take two copies of each surface vertex . We take copies of each nonperipheral cyclic vertex space . We take two copies of each peripheral cyclic vertex space of ; these will each be a peripheral vertex of . Finally, we take further copies of each peripheral vertex space ; these will be nonperipheral vertices of . It is now easy to see that we can assemble these to form so that every nonperipheral cyclic vertex group has exactly two incident edges and all the attaching maps are isomorphisms. As before, the natural map is -injective by Proposition 1.7.

Every nonperipheral cyclic vertex is adjacent to exactly two surface vertices and is identified with two boundary components of these. We may therefore contract the two edges adjacent to to obtain a larger surface vertex.

Thus, the resulting graph of spaces is homeomorphic to a surface . The peripheral cyclic vertices equip with exactly two copies of each component of ; we call this collection of cyclic subgroups . We note that and that, since only surface vertices contribute to Euler characteristic, . Therefore, if we replace by a locally irreducible graph pair as in the previous lemma, .

At this stage we have a locally irreducible, admissible pair of surface type, satisfying the required constraints on the Euler characteristic, so that each component of is conjugate into some component of (and every component of contains a component of ). To make this pair of surface type, we need to be identified bijectively with . To ensure this, we first invoke Marshall Hall’s theorem again, replacing with a finite-sheeted cover and with its pullback, so that each component of maps isomorphically to the component of that contains it. For each component , let be the number of components of contained in . Replacing with two copies of itself, we may assume that each is even. Without loss of generality, we may also assume that is minimal among the . We now take copies of , and equip each boundary component with exactly one component of . We may then add annuli to to pair up the remaining components of . This completes the proof.

We can now apply these two lemmas to prove that we can always find a maximal -immersion of surface type.

Proof of Theorem 5.6.

By Lemma 2.12 is weakly irreducible, so it can be unfolded to a pair (without loss of generality, ) which is a wedge of locally irreducible graph pairs . That is, there is a finite set equipped with maps , so that

where for all , and . For each , let us fix choices of lifts of the maps to maps .

Suppose that, for some (without loss of generality, ), the relative JSJ decomposition of the group pair has a rigid vertex. Lemma 5.9 applied to yields an essential map from an irreducible pair that satisfies , where is the degree of the covering map . By Lemma 2.11, after unfolding, we may take the pair to be locally irreducible. Let consist of copies of , for each . Let . Let , and choose a map so that, for each , indexes the preimages of in . Let .

The maps now define a wedge , where for all , which folds (preserving Euler characteristic) to an immersion . Let be the composition of with the natural quotient map , and let be the composition of with the quotient map . We therefore have an admissible -immersion

such that is locally irreducible with and .

We now compare Euler characteristics:

whereas

so . The -immersion

therefore has greater projective -rank, contradicting the maximality hypothesis.

The second part of the theorem follows in the same way, using Lemma 5.10 instead of Lemma 5.9.

Our main technical theorem follows immediately.

Theorem 5.11.

Let be an irreducible graph pair. There exists a compact surface with boundary and an admissible, essential map .

Proof.

By Lemma 5.1 a maximal, admissible -immersion

exists. By Theorem 5.6 there is such a maximal, admissible -immersion so that is of weak surface type. After passing to the disjoint union of the free factors, we obtain a maximal, admissible -immersion so that is of surface type.

6. Surface subgroups and hierarchies

In this section we deduce the claimed consequences of Theorem 5.11. We start with graphs of virtually free groups with virtually cyclic edge groups. The deduction of the existence of surface subgroups from a result like Theorem 5.11 is well known (cf. Reference 7 or Reference 42, for instance); we include an argument here for completeness.

Theorem 6.1.

Let be the fundamental group of a graph of virtually free groups with virtually cyclic edge groups. If is hyperbolic and one-ended, then contains a quasi-convex surface subgroup.

Proof.

By Reference 43 is residually finite and so virtually torsion-free. We may therefore assume that is the fundamental group of a graph of free groups with cyclic edge groups. We call the vertices of the underlying graph of noncyclic, and we subdivide each edge, putting a cyclic vertex in the middle with vertex group .

Consider the induced pair for a noncyclic vertex . Since is one ended, is irreducible. By Theorem 5.11 we can replace each by an admissible, essential map of a surface pair . By gluing these to the adjacent cyclic vertices, we define a new graph of free groups with cyclic edge groups , with every noncyclic vertex of surface type. Note that for all .

The fundamental group of is equipped with a natural map , and by Proposition 1.7 is injective. In particular, contains no Baumslag–Solitar subgroups, since is hyperbolic.

The graph of groups is a graph of surfaces glued along their boundary components to circles, which we realize in the natural way as a graph of spaces . Note that . By Reference 43, after replacing by a subgroup of finite index, we may assume that the attaching maps are all homeomorphisms. It is then easy to see that can be thickened to a 3-manifold with boundary. Since closed 3-manifolds have zero Euler characterisitic,

and so we may choose a component of with . Inclusion induces a natural map . But is one ended by Shenitzer’s lemma, and Dehn’s lemma then implies that the map (and hence the composition ) is injective.

Finally, is locally quasi-convex Reference 3, Theorem D and, hence, the surface subgroup is quasi-convex. This completes the proof.

A group is called rigid if it does not split over a (possibly finite) virtually cyclic subgroup. Given a group , a virtually cyclic hierarchy for is a set of subgroups of obtained by passing to the vertex groups of a splitting of over virtually cyclic edge groups and then repeating this operation on those subgroups recursively. If a finite virtually cyclic hierarchy exists, terminating in (possibly finite) rigid subgroups, then we shall say that has a finite hierarchy. Graphs of virtually free groups with virtually cyclic edge groups play a special role in the subgroup theory of groups that have finite hierarchies.

Remark 6.2.

Let be a one-ended group with a finite hierarchy, and let be a one-ended subgroup in the hierarchy of with no one-ended subgroups below it. Then either is rigid or is a graph of virtually free groups over virtually cyclic edge groups.

In Reference 36 Sela showed that limit groups have a finite hierarchy. He also showed that a limit group without a subgroup is hyperbolic and that nonabelian limit groups are never rigid. We thus obtain

Corollary 6.3.

Every one-ended limit group contains a surface subgroup.

Louder and Touikan showed that a hyperbolic group without 2-torsion has a finite hierarchy Reference 29. (The restriction on 2-torsion is technical and, conjecturally, can be removed.) We thus obtain the following contribution toward the complete resolution of Gromov’s question.

Corollary 6.4.

Every one-ended hyperbolic group without -torsion either contains a surface subgroup or contains a quasi-convex, infinite, rigid subgroup.

In particular, Gromov’s question is reduced to the rigid case (modulo the technical issue of 2-torsion).

7. Applications to profinite rigidity

In this section we discuss applications to Question 0.2 and related problems. As explained in Reference 6, Theorem 4.17, Corollary 6.3 resolves the question for limit groups.

We include the proof for completeness. The key point is that it follows that the profinite completions of nonfree limit groups have nonzero virtual second cohomology. We work with continuous cohomology, with coefficients in .

Theorem 7.1.

If is a limit group and not free, then there is a subgroup of finite index in such that .

Proof.

By Corollary 6.3 contains a subgroup isomorphic to the fundamental group of a closed surface of nonpositive Euler characteristic; in particular, (with coefficients in ). Since surface groups are good in the sense of Serre Reference 19, it follows that the continuous cohomology is also nonzero. By Reference 41 is a virtual retract of , so there is a finite-index subgroup containing and a retraction . Let be the inclusion map, so . Both and extend by continuity to maps and , and . Therefore the induced maps on cohomology satisfy . In particular, since is nonzero, is also nonzero, as claimed.

Since every open subgroup of a profinite free group is profinite free and, hence, has zero second cohomology, Corollary D follows immediately. Corollary E also follows quickly from Theorem 7.1.

Proof of Corollary E.

To prove the contrapositive, we assume that is not primitive in . Recall that the double is the fundamental group of the graph of groups with two vertices labeled by , and it has one edge between them for each component of . The double is a limit group and hence, by Theorem 7.1, has a subgroup of finite index with . If were primitive in then would be free profinite, hence so would , and therefore would be zero, a contradiction.

Acknowledgments

Alan Reid asked me whether one-ended limit groups have surface subgroups in 2006. I have worked on finding surface subgroups with a variety of collaborators and am grateful to them all: Lars Louder, Sang-hyun Kim, Cameron Gordon, Danny Calegari, Ben Barrett. Not all of these projects led to publications, but I learned a lot from each of them. A conversation with Frédéric Haglund and Pierre Pansu led to the discovery of a serious mistake in an earlier attempted proof. Thanks are also due to Daniel Groves for comments on an early version of this paper. I am especially grateful to Lars Louder for spotting a subtle error in the first version of this paper.

Figures

Figure 1.

The Whitehead graph of the Baumslag–Solitar word

Graphic without alt textGraphic without alt text
Figure 2.

The effect of a fold on Whitehead graphs. Note that is obtained by wedging , and is obtained by wedging .

Graphic without alt textGraphic without alt text
Figure 3.

The graph of spaces has one rigid vertex, one surface vertex, one nonperipheral cyclic vertex, and three peripheral cyclic vertices. By taking multiple copies of and deleting complementary components of the rigid vertices, we construct a new graph of spaces with greater projective -rank.

Graphic without alt text

Mathematical Fragments

Theorem A.

Let be the fundamental group of a graph of free groups with cyclic edge groups. If is one ended and hyperbolic, then contains a quasi-convex surface subgroup.

Corollary B.

Let be a one-ended hyperbolic group without -torsion. Either contains a quasi-convex surface subgroup or contains a quasi-convex rigid subgroup.

Corollary C.

Let be a limit group. If is one-ended, then contains a surface subgroup.

Question 0.2 (Remeslennikov).

Suppose that is a finitely generated, nonabelian free group and that is finitely generated and residually finite. If , does it follow that ?

Corollary D.

If is a limit group and not free, then the profinite completion is not isomorphic to the profinite completion of any free group.

Corollary E.

Let be a finitely generated free group. If a multiword is primitive in the profinite completion then it is primitive in .

Theorem F.

If is an irreducible graph pair, then there is a compact surface with boundary and an essential map of pairs .

Proposition 1.7.

Suppose that is a morphism of graphs of groups. If induces essential morphisms on the group pairs corresponding to vertices, then induces an injective map on fundamental groups.

Remark 2.3.

The partition on the components of and the involutions are enough information to reconstruct the pair .

Lemma 2.5.

Let be the morphism given by a fold which identifies two edges of with a common initial vertex to an edge of . Let and let and . Suppose further that is a homotopy equivalence, i.e., . Then

(Note that the unions in this expression may not be disjoint, since may equal for at most one , in which case also equals In particular, is a cut vertex of

Lemma 2.8.

Let be a graph pair, and suppose that some edge defines a cut vertex in (where . Then there is a graph pair and a morphism of pairs defining a homotopy-equivalent fold that identifies a pair of edges to .

Definition 2.9.

Consider a graph pair with a finite connected graph. By Grushko’s theorem, splits canonically as

where for each , there is an index set so that is conjugate into for all , each does not split freely relative to the set , and no is conjugate into . A factor is called cyclic if is a singleton and, up to conjugacy, generates .

The pair is called weakly irreducible if there are no cyclic factors; otherwise it is called strongly reducible. The pair is called reducible if it is weakly irreducible and ; otherwise it is called reducible. When is disconnected, the pair is called (weakly) irreducible or (strongly) reducible if and only if each component has that property.

Lemma 2.11 (Converse to Whitehead’s lemma).

If a pair is irreducible, then there is a locally irreducible pair and a map of pairs which is a homotopy equivalence.

Lemma 2.12.

If is a -immersion and is locally irreducible, then the pair is weakly irreducible.

Theorem 4.3.

If , then the maximal and minimal -ranks are realized by admissible -immersions of locally irreducible pairs; that is, there exist locally irreducible pairs and admissible -immersions

so that . In particular, are positive rational numbers.

Lemma 5.1.

If is irreducible, then there exists a maximal, admissible -immersion

for a locally irreducible pair .

Theorem 5.6.

If is locally irreducible and

is a maximal, admissible -immersion, then for each irreducible free factor of the corresponding group pair , the relative JSJ decomposition of has no rigid vertices. Furthermore, if there is such a maximal -immersion, then there is a maximal, admissible -immersion so that is of weak surface type.

Theorem 5.7.

If is a rigid graph pair, then there is a finite-sheeted cover such that, whenever a finite-sheeted cover factors through , the pair is irreducible for any component of .

Lemma 5.8 (Relative Shenitzer’s lemma).

Consider a decomposition of a group pair . If the induced pair at every vertex of is irreducible, then the pair is irreducible.

Lemma 5.9.

Let be an irreducible graph pair. If the relative JSJ decomposition of has a rigid vertex, then there is a locally irreducible pair and an admissible, essential map with

Lemma 5.10.

Let be an irreducible graph pair. If the relative JSJ decomposition of has no rigid vertices, then there is a locally irreducible group pair of surface type and an admissible, essential with

Theorem 5.11.

Let be an irreducible graph pair. There exists a compact surface with boundary and an admissible, essential map .

Theorem 6.1.

Let be the fundamental group of a graph of virtually free groups with virtually cyclic edge groups. If is hyperbolic and one-ended, then contains a quasi-convex surface subgroup.

Corollary 6.3.

Every one-ended limit group contains a surface subgroup.

Corollary 6.4.

Every one-ended hyperbolic group without -torsion either contains a surface subgroup or contains a quasi-convex, infinite, rigid subgroup.

Theorem 7.1.

If is a limit group and not free, then there is a subgroup of finite index in such that .

References

Reference [1]
Emina Alibegović, A combination theorem for relatively hyperbolic groups, Bull. London Math. Soc. 37 (2005), no. 3, 459–466, DOI 10.1112/S0024609304004059. MR2131400,
Show rawAMSref \bib{alibegovic_combination_2005}{article}{ author={Alibegovi\'c, Emina}, title={A combination theorem for relatively hyperbolic groups}, journal={Bull. London Math. Soc.}, volume={37}, date={2005}, number={3}, pages={459--466}, issn={0024-6093}, review={\MR {2131400}}, doi={10.1112/S0024609304004059}, }
Reference [2]
Gilbert Baumslag, Residually finite groups with the same finite images, Compositio Math. 29 (1974), 249–252. MR0357615,
Show rawAMSref \bib{baumslag_residually_1974}{article}{ author={Baumslag, Gilbert}, title={Residually finite groups with the same finite images}, journal={Compositio Math.}, volume={29}, date={1974}, pages={249--252}, issn={0010-437X}, review={\MR {0357615}}, }
Reference [3]
Hadi Bigdely and Daniel T. Wise, Quasiconvexity and relatively hyperbolic groups that split, Michigan Math. J. 62 (2013), no. 2, 387–406, DOI 10.1307/mmj/1370870378. MR3079269,
Show rawAMSref \bib{bigdely_quasiconvexity_2013}{article}{ author={Bigdely, Hadi}, author={Wise, Daniel T.}, title={Quasiconvexity and relatively hyperbolic groups that split}, journal={Michigan Math. J.}, volume={62}, date={2013}, number={2}, pages={387--406}, issn={0026-2285}, review={\MR {3079269}}, doi={10.1307/mmj/1370870378}, }
Reference [4]
Brian H. Bowditch, Cut points and canonical splittings of hyperbolic groups, Acta Math. 180 (1998), no. 2, 145–186, DOI 10.1007/BF02392898. MR1638764,
Show rawAMSref \bib{bowditch_cut_1998}{article}{ author={Bowditch, Brian H.}, title={Cut points and canonical splittings of hyperbolic groups}, journal={Acta Math.}, volume={180}, date={1998}, number={2}, pages={145--186}, issn={0001-5962}, review={\MR {1638764}}, doi={10.1007/BF02392898}, }
Reference [5]
Noel Brady, Matt Clay, and Max Forester, Turn graphs and extremal surfaces in free groups, Topology and geometry in dimension three, Contemp. Math., vol. 560, Amer. Math. Soc., Providence, RI, 2011, pp. 171–178, DOI 10.1090/conm/560/11098. MR2866930,
Show rawAMSref \bib{brady_turn_2011}{article}{ author={Brady, Noel}, author={Clay, Matt}, author={Forester, Max}, title={Turn graphs and extremal surfaces in free groups}, conference={ title={Topology and geometry in dimension three}, }, book={ series={Contemp. Math.}, volume={560}, publisher={Amer. Math. Soc., Providence, RI}, }, date={2011}, pages={171--178}, review={\MR {2866930}}, doi={10.1090/conm/560/11098}, }
Reference [6]
M. R. Bridson, M. D. E. Conder, and A. W. Reid, Determining Fuchsian groups by their finite quotients, Israel J. Math. 214 (2016), no. 1, 1–41, DOI 10.1007/s11856-016-1341-6. MR3540604,
Show rawAMSref \bib{bridson_determining_2016}{article}{ author={Bridson, M. R.}, author={Conder, M. D. E.}, author={Reid, A. W.}, title={Determining Fuchsian groups by their finite quotients}, journal={Israel J. Math.}, volume={214}, date={2016}, number={1}, pages={1--41}, issn={0021-2172}, review={\MR {3540604}}, doi={10.1007/s11856-016-1341-6}, }
Reference [7]
Danny Calegari, Surface subgroups from homology, Geom. Topol. 12 (2008), no. 4, 1995–2007, DOI 10.2140/gt.2008.12.1995. MR2431013,
Show rawAMSref \bib{calegari_surface_2008}{article}{ author={Calegari, Danny}, title={Surface subgroups from homology}, journal={Geom. Topol.}, volume={12}, date={2008}, number={4}, pages={1995--2007}, issn={1465-3060}, review={\MR {2431013}}, doi={10.2140/gt.2008.12.1995}, }
Reference [8]
Danny Calegari, scl, MSJ Memoirs, vol. 20, Mathematical Society of Japan, Tokyo, 2009, DOI 10.1142/e018. MR2527432,
Show rawAMSref \bib{calegari_scl_2009}{book}{ author={Calegari, Danny}, title={scl}, series={MSJ Memoirs}, volume={20}, publisher={Mathematical Society of Japan, Tokyo}, date={2009}, pages={xii+209}, isbn={978-4-931469-53-2}, review={\MR {2527432}}, doi={10.1142/e018}, }
Reference [9]
Danny Calegari, Stable commutator length is rational in free groups, J. Amer. Math. Soc. 22 (2009), no. 4, 941–961, DOI 10.1090/S0894-0347-09-00634-1. MR2525776,
Show rawAMSref \bib{calegari_stable_2009}{article}{ author={Calegari, Danny}, title={Stable commutator length is rational in free groups}, journal={J. Amer. Math. Soc.}, volume={22}, date={2009}, number={4}, pages={941--961}, issn={0894-0347}, review={\MR {2525776}}, doi={10.1090/S0894-0347-09-00634-1}, }
Reference [10]
Danny Calegari and Alden Walker, Random groups contain surface subgroups, J. Amer. Math. Soc. 28 (2015), no. 2, 383–419, DOI 10.1090/S0894-0347-2014-00802-X. MR3300698,
Show rawAMSref \bib{calegari_random_2015}{article}{ author={Calegari, Danny}, author={Walker, Alden}, title={Random groups contain surface subgroups}, journal={J. Amer. Math. Soc.}, volume={28}, date={2015}, number={2}, pages={383--419}, issn={0894-0347}, review={\MR {3300698}}, doi={10.1090/S0894-0347-2014-00802-X}, }
Reference [11]
Danny Calegari and Alden Walker, Surface subgroups from linear programming, Duke Math. J. 164 (2015), no. 5, 933–972, DOI 10.1215/00127094-2877511. MR3332895,
Show rawAMSref \bib{calegari_surface_2015}{article}{ author={Calegari, Danny}, author={Walker, Alden}, title={Surface subgroups from linear programming}, journal={Duke Math. J.}, volume={164}, date={2015}, number={5}, pages={933--972}, issn={0012-7094}, review={\MR {3332895}}, doi={10.1215/00127094-2877511}, }
Reference [12]
Danny Calegari and Henry Wilton, Random graphs of free groups contain surface subgroups, Math. Res. Lett. (to appear).
Reference [13]
Christopher H. Cashen, Splitting line patterns in free groups, Algebr. Geom. Topol. 16 (2016), no. 2, 621–673, DOI 10.2140/agt.2016.16.621. MR3493403,
Show rawAMSref \bib{cashen_splitting_2016}{article}{ author={Cashen, Christopher H.}, title={Splitting line patterns in free groups}, journal={Algebr. Geom. Topol.}, volume={16}, date={2016}, number={2}, pages={621--673}, issn={1472-2747}, review={\MR {3493403}}, doi={10.2140/agt.2016.16.621}, }
Reference [14]
Christopher H. Cashen and Nataša Macura, Line patterns in free groups, Geom. Topol. 15 (2011), no. 3, 1419–1475, DOI 10.2140/gt.2011.15.1419. MR2825316,
Show rawAMSref \bib{cashen_line_2011}{article}{ author={Cashen, Christopher H.}, author={Macura, Nata\v sa}, title={Line patterns in free groups}, journal={Geom. Topol.}, volume={15}, date={2011}, number={3}, pages={1419--1475}, issn={1465-3060}, review={\MR {2825316}}, doi={10.2140/gt.2011.15.1419}, }
Reference [15]
Marc Culler, Using surfaces to solve equations in free groups, Topology 20 (1981), no. 2, 133–145, DOI 10.1016/0040-9383(81)90033-1. MR605653,
Show rawAMSref \bib{culler_using_1981}{article}{ author={Culler, Marc}, title={Using surfaces to solve equations in free groups}, journal={Topology}, volume={20}, date={1981}, number={2}, pages={133--145}, issn={0040-9383}, review={\MR {605653}}, doi={10.1016/0040-9383(81)90033-1}, }
Reference [16]
François Dahmani, Combination of convergence groups, Geom. Topol. 7 (2003), 933–963, DOI 10.2140/gt.2003.7.933. MR2026551,
Show rawAMSref \bib{dahmani_combination_2003}{article}{ author={Dahmani, Fran\c {c}ois}, title={Combination of convergence groups}, journal={Geom. Topol.}, volume={7}, date={2003}, pages={933--963}, issn={1465-3060}, review={\MR {2026551}}, doi={10.2140/gt.2003.7.933}, }
Reference [17]
T. Delzant, L’image d’un groupe dans un groupe hyperbolique (French), Comment. Math. Helv. 70 (1995), no. 2, 267–284, DOI 10.1007/BF02566008. MR1324630,
Show rawAMSref \bib{delzant_limage_1995}{article}{ author={Delzant, T.}, title={L'image d'un groupe dans un groupe hyperbolique}, language={French}, journal={Comment. Math. Helv.}, volume={70}, date={1995}, number={2}, pages={267--284}, issn={0010-2571}, review={\MR {1324630}}, doi={10.1007/BF02566008}, }
Reference [18]
Cameron Gordon and Henry Wilton, On surface subgroups of doubles of free groups, J. Lond. Math. Soc. (2) 82 (2010), no. 1, 17–31, DOI 10.1112/jlms/jdq007. MR2669638,
Show rawAMSref \bib{gordon_surface_2010}{article}{ author={Gordon, Cameron}, author={Wilton, Henry}, title={On surface subgroups of doubles of free groups}, journal={J. Lond. Math. Soc. (2)}, volume={82}, date={2010}, number={1}, pages={17--31}, issn={0024-6107}, review={\MR {2669638}}, doi={10.1112/jlms/jdq007}, }
Reference [19]
F. Grunewald, A. Jaikin-Zapirain, and P. A. Zalesskii, Cohomological goodness and the profinite completion of Bianchi groups, Duke Math. J. 144 (2008), no. 1, 53–72, DOI 10.1215/00127094-2008-031. MR2429321,
Show rawAMSref \bib{grunewald_cohomological_2008}{article}{ author={Grunewald, F.}, author={Jaikin-Zapirain, A.}, author={Zalesskii, P. A.}, title={Cohomological goodness and the profinite completion of Bianchi groups}, journal={Duke Math. J.}, volume={144}, date={2008}, number={1}, pages={53--72}, issn={0012-7094}, review={\MR {2429321}}, doi={10.1215/00127094-2008-031}, }
Reference [20]
Vincent Guirardel and Gilbert Levitt, JSJ decompositions of groups (English, with English and French summaries), Astérisque 395 (2017), vii+165. MR3758992,
Show rawAMSref \bib{guirardel_JSJ_2017}{article}{ author={Guirardel, Vincent}, author={Levitt, Gilbert}, title={JSJ decompositions of groups}, language={English, with English and French summaries}, journal={Ast\'erisque}, number={395}, date={2017}, pages={vii+165}, issn={0303-1179}, isbn={978-2-85629-870-1}, review={\MR {3758992}}, }
Reference [21]
Ursula Hamenstädt, Incompressible surfaces in rank one locally symmetric spaces, Geom. Funct. Anal. 25 (2015), no. 3, 815–859, DOI 10.1007/s00039-015-0330-y. MR3361773,
Show rawAMSref \bib{hamenstaedt_incompressible_2015}{article}{ author={Hamenst\"adt, Ursula}, title={Incompressible surfaces in rank one locally symmetric spaces}, journal={Geom. Funct. Anal.}, volume={25}, date={2015}, number={3}, pages={815--859}, issn={1016-443X}, review={\MR {3361773}}, doi={10.1007/s00039-015-0330-y}, }
Reference [22]
Marshall Hall Jr., Subgroups of finite index in free groups, Canadian J. Math. 1 (1949), 187–190. MR0028836,
Show rawAMSref \bib{hall_jr_subgroups_1949}{article}{ author={Hall, Marshall, Jr.}, title={Subgroups of finite index in free groups}, journal={Canadian J. Math.}, volume={1}, date={1949}, pages={187--190}, issn={0008-414X}, review={\MR {0028836}}, }
Reference [23]
Jeremy Kahn and Vladimir Markovic, Immersing almost geodesic surfaces in a closed hyperbolic three manifold, Ann. of Math. (2) 175 (2012), no. 3, 1127–1190, DOI 10.4007/annals.2012.175.3.4. MR2912704,
Show rawAMSref \bib{kahn_immersing_2012}{article}{ author={Kahn, Jeremy}, author={Markovic, Vladimir}, title={Immersing almost geodesic surfaces in a closed hyperbolic three manifold}, journal={Ann. of Math. (2)}, volume={175}, date={2012}, number={3}, pages={1127--1190}, issn={0003-486X}, review={\MR {2912704}}, doi={10.4007/annals.2012.175.3.4}, }
Reference [24]
Jeremy Kahn and Vladimir Markovic, The good pants homology and the Ehrenpreis conjecture, Ann. of Math. (2) 182 (2015), no. 1, 1–72, DOI 10.4007/annals.2015.182.1.1. MR3374956,
Show rawAMSref \bib{kahn_good_2015}{article}{ author={Kahn, Jeremy}, author={Markovic, Vladimir}, title={The good pants homology and the Ehrenpreis conjecture}, journal={Ann. of Math. (2)}, volume={182}, date={2015}, number={1}, pages={1--72}, issn={0003-486X}, review={\MR {3374956}}, doi={10.4007/annals.2015.182.1.1}, }
Reference [25]
Olga Kharlampovich and Alexei Myasnikov, Tarski’s problem about the elementary theory of free groups has a positive solution, Electron. Res. Announc. Amer. Math. Soc. 4 (1998), 101–108, DOI 10.1090/S1079-6762-98-00047-X. MR1662319,
Show rawAMSref \bib{kharlampovich_tarskis_1998}{article}{ author={Kharlampovich, Olga}, author={Myasnikov, Alexei}, title={Tarski's problem about the elementary theory of free groups has a positive solution}, journal={Electron. Res. Announc. Amer. Math. Soc.}, volume={4}, date={1998}, pages={101--108}, issn={1079-6762}, review={\MR {1662319}}, doi={10.1090/S1079-6762-98-00047-X}, }
Reference [26]
Sang-hyun Kim and Sang-il Oum, Hyperbolic surface subgroups of one-ended doubles of free groups, J. Topol. 7 (2014), no. 4, 927–947, DOI 10.1112/jtopol/jtu004. MR3286893,
Show rawAMSref \bib{kim_hyperbolic_2014}{article}{ author={Kim, Sang-hyun}, author={Oum, Sang-il}, title={Hyperbolic surface subgroups of one-ended doubles of free groups}, journal={J. Topol.}, volume={7}, date={2014}, number={4}, pages={927--947}, issn={1753-8416}, review={\MR {3286893}}, doi={10.1112/jtopol/jtu004}, }
Reference [27]
Sang-Hyun Kim and Henry Wilton, Polygonal words in free groups, Q. J. Math. 63 (2012), no. 2, 399–421, DOI 10.1093/qmath/haq045. MR2925298,
Show rawAMSref \bib{kim_polygonal_2012}{article}{ author={Kim, Sang-Hyun}, author={Wilton, Henry}, title={Polygonal words in free groups}, journal={Q. J. Math.}, volume={63}, date={2012}, number={2}, pages={399--421}, issn={0033-5606}, review={\MR {2925298}}, doi={10.1093/qmath/haq045}, }
Reference [28]
Yi Liu and Vladimir Markovic, Homology of curves and surfaces in closed hyperbolic 3-manifolds, Duke Math. J. 164 (2015), no. 14, 2723–2808, DOI 10.1215/00127094-3167744. MR3417184,
Show rawAMSref \bib{liu_homology_2015}{article}{ author={Liu, Yi}, author={Markovic, Vladimir}, title={Homology of curves and surfaces in closed hyperbolic 3-manifolds}, journal={Duke Math. J.}, volume={164}, date={2015}, number={14}, pages={2723--2808}, issn={0012-7094}, review={\MR {3417184}}, doi={10.1215/00127094-3167744}, }
Reference [29]
Larsen Louder and Nicholas Touikan, Strong accessibility for finitely presented groups, Geom. Topol. 21 (2017), no. 3, 1805–1835, DOI 10.2140/gt.2017.21.1805. MR3650082,
Show rawAMSref \bib{louder_strong_2017}{article}{ author={Louder, Larsen}, author={Touikan, Nicholas}, title={Strong accessibility for finitely presented groups}, journal={Geom. Topol.}, volume={21}, date={2017}, number={3}, pages={1805--1835}, issn={1465-3060}, review={\MR {3650082}}, doi={10.2140/gt.2017.21.1805}, }
Reference [30]
Jason Fox Manning, Virtually geometric words and Whitehead’s algorithm, Math. Res. Lett. 17 (2010), no. 5, 917–925, DOI 10.4310/MRL.2010.v17.n5.a9. MR2727618,
Show rawAMSref \bib{manning_virtually_2010}{article}{ author={Manning, Jason Fox}, title={Virtually geometric words and Whitehead's algorithm}, journal={Math. Res. Lett.}, volume={17}, date={2010}, number={5}, pages={917--925}, issn={1073-2780}, review={\MR {2727618}}, doi={10.4310/MRL.2010.v17.n5.a9}, }
Reference [31]
Vladimir Markovic, Criterion for Cannon’s conjecture, Geom. Funct. Anal. 23 (2013), no. 3, 1035–1061, DOI 10.1007/s00039-013-0228-5. MR3061779,
Show rawAMSref \bib{markovic_criterion_2013}{article}{ author={Markovic, Vladimir}, title={Criterion for Cannon's conjecture}, journal={Geom. Funct. Anal.}, volume={23}, date={2013}, number={3}, pages={1035--1061}, issn={1016-443X}, review={\MR {3061779}}, doi={10.1007/s00039-013-0228-5}, }
Reference [32]
G. A. Noskov, V. N. Remeslennikov, and V. A. Roman′kov, Infinite groups (Russian), Algebra. Topology. Geometry, Vol. 17 (Russian), Akad. Nauk SSSR, Vsesoyuz. Inst. Nauchn. i Tekhn. Informatsii, Moscow, 1979, pp. 65–157, 308. MR584569,
Show rawAMSref \bib{noskov_infinite_1979}{article}{ author={Noskov, G. A.}, author={Remeslennikov, V. N.}, author={Roman\cprime kov, V. A.}, title={Infinite groups}, language={Russian}, conference={ title={Algebra. Topology. Geometry, Vol. 17 (Russian)}, }, book={ publisher={Akad. Nauk SSSR, Vsesoyuz. Inst. Nauchn. i Tekhn. Informatsii, Moscow}, }, date={1979}, pages={65--157, 308}, review={\MR {584569}}, }
Reference [33]
Doron Puder and Ori Parzanchevski, Measure preserving words are primitive, J. Amer. Math. Soc. 28 (2015), no. 1, 63–97, DOI 10.1090/S0894-0347-2014-00796-7. MR3264763,
Show rawAMSref \bib{puder_measure_2015}{article}{ author={Puder, Doron}, author={Parzanchevski, Ori}, title={Measure preserving words are primitive}, journal={J. Amer. Math. Soc.}, volume={28}, date={2015}, number={1}, pages={63--97}, issn={0894-0347}, review={\MR {3264763}}, doi={10.1090/S0894-0347-2014-00796-7}, }
Reference [34]
V. N. Remeslennikov, -free groups (Russian), Sibirsk. Mat. Zh. 30 (1989), no. 6, 193–197, DOI 10.1007/BF00970922; English transl., Siberian Math. J. 30 (1989), no. 6, 998–1001 (1990). MR1043446,
Show rawAMSref \bib{remeslennikov_exists-free_1989}{article}{ author={Remeslennikov, V. N.}, title={$\exists $-free groups}, language={Russian}, journal={Sibirsk. Mat. Zh.}, volume={30}, date={1989}, number={6}, pages={193--197}, issn={0037-4474}, translation={ journal={Siberian Math. J.}, volume={30}, date={1989}, number={6}, pages={998--1001 (1990)}, issn={0037-4466}, }, review={\MR {1043446}}, doi={10.1007/BF00970922}, }
Reference [35]
Peter Scott and Terry Wall, Topological methods in group theory, Homological group theory (Proc. Sympos., Durham, 1977), London Math. Soc. Lecture Note Ser., vol. 36, Cambridge Univ. Press, Cambridge-New York, 1979, pp. 137–203. MR564422,
Show rawAMSref \bib{scott_topological_1979}{article}{ author={Scott, Peter}, author={Wall, Terry}, title={Topological methods in group theory}, conference={ title={Homological group theory}, address={Proc. Sympos., Durham}, date={1977}, }, book={ series={London Math. Soc. Lecture Note Ser.}, volume={36}, publisher={Cambridge Univ. Press, Cambridge-New York}, }, date={1979}, pages={137--203}, review={\MR {564422}}, }
Reference [36]
Zlil Sela, Diophantine geometry over groups. I. Makanin-Razborov diagrams, Publ. Math. Inst. Hautes Études Sci. 93 (2001), 31–105, DOI 10.1007/s10240-001-8188-y. MR1863735,
Show rawAMSref \bib{sela_diophantine_2001}{article}{ author={Sela, Zlil}, title={Diophantine geometry over groups. I. Makanin-Razborov diagrams}, journal={Publ. Math. Inst. Hautes \'Etudes Sci.}, number={93}, date={2001}, pages={31--105}, issn={0073-8301}, review={\MR {1863735}}, doi={10.1007/s10240-001-8188-y}, }
Reference [37]
Jean-Pierre Serre, Arbres, amalgames, (French), Société Mathématique de France, Paris, 1977. Avec un sommaire anglais; Rédigé avec la collaboration de Hyman Bass; Astérisque, No. 46. MR0476875,
Show rawAMSref \bib{serre_arbres_1977}{book}{ author={Serre, Jean-Pierre}, title={Arbres, amalgames, ${\rm SL}_{2}$}, language={French}, note={Avec un sommaire anglais; R\'edig\'e avec la collaboration de Hyman Bass; Ast\'erisque, No. 46}, publisher={Soci\'et\'e Math\'ematique de France, Paris}, date={1977}, pages={189 pp. (1 plate)}, review={\MR {0476875}}, }
Reference [38]
John R. Stallings, Topology of finite graphs, Invent. Math. 71 (1983), no. 3, 551–565, DOI 10.1007/BF02095993. MR695906,
Show rawAMSref \bib{stallings_topology_1983}{article}{ author={Stallings, John R.}, title={Topology of finite graphs}, journal={Invent. Math.}, volume={71}, date={1983}, number={3}, pages={551--565}, issn={0020-9910}, review={\MR {695906}}, doi={10.1007/BF02095993}, }
Reference [39]
Nicholas Touikan, On the one-endedness of graphs of groups, Pacific J. Math. 278 (2015), no. 2, 463–478. MR3407182,
Show rawAMSref \bib{touikan_one-endedness_2015}{article}{ author={Touikan, Nicholas}, title={On the one-endedness of graphs of groups}, journal={Pacific J. Math.}, volume={278}, date={2015}, number={2}, pages={463--478}, issn={0030-8730}, review={\MR {3407182}}, }
Reference [40]
J. H. C. Whitehead, On equivalent sets of elements in a free group, Ann. of Math. (2) 37 (1936), no. 4, 782–800, DOI 10.2307/1968618. MR1503309,
Show rawAMSref \bib{whitehead_equivalent_1936}{article}{ author={Whitehead, J. H. C.}, title={On equivalent sets of elements in a free group}, journal={Ann. of Math. (2)}, volume={37}, date={1936}, number={4}, pages={782--800}, issn={0003-486X}, review={\MR {1503309}}, doi={10.2307/1968618}, }
Reference [41]
Henry Wilton, Hall’s theorem for limit groups, Geom. Funct. Anal. 18 (2008), no. 1, 271–303, DOI 10.1007/s00039-008-0657-8. MR2399104,
Show rawAMSref \bib{wilton_halls_2008}{article}{ author={Wilton, Henry}, title={Hall's theorem for limit groups}, journal={Geom. Funct. Anal.}, volume={18}, date={2008}, number={1}, pages={271--303}, issn={1016-443X}, review={\MR {2399104}}, doi={10.1007/s00039-008-0657-8}, }
Reference [42]
Henry Wilton, One-ended subgroups of graphs of free groups with cyclic edge groups, Geom. Topol. 16 (2012), no. 2, 665–683, DOI 10.2140/gt.2012.16.665. MR2928980,
Show rawAMSref \bib{wilton_one-ended_2011}{article}{ author={Wilton, Henry}, title={One-ended subgroups of graphs of free groups with cyclic edge groups}, journal={Geom. Topol.}, volume={16}, date={2012}, number={2}, pages={665--683}, issn={1465-3060}, review={\MR {2928980}}, doi={10.2140/gt.2012.16.665}, }
Reference [43]
Daniel T. Wise, Subgroup separability of graphs of free groups with cyclic edge groups, Q. J. Math. 51 (2000), no. 1, 107–129, DOI 10.1093/qmathj/50.1.107. MR1760573,
Show rawAMSref \bib{wise_subgroup_2000}{article}{ author={Wise, Daniel T.}, title={Subgroup separability of graphs of free groups with cyclic edge groups}, journal={Q. J. Math.}, volume={51}, date={2000}, number={1}, pages={107--129}, issn={0033-5606}, review={\MR {1760573}}, doi={10.1093/qmathj/50.1.107}, }

Article Information

MSC 2010
Primary: 20F65 (Geometric group theory), 20F67 (Hyperbolic groups and nonpositively curved groups), 57M07 (Topological methods in group theory)
Author Information
Henry Wilton
DPMMS, Centre for Mathematical Sciences, Wilberforce Road, Cambridge CB3 0WB, United Kingdom
h.wilton@maths.cam.ac.uk
MathSciNet
Additional Notes

The author was supported by EPSRC Standard Grant EP/L026481/1.

Journal Information
Journal of the American Mathematical Society, Volume 31, Issue 4, ISSN 1088-6834, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2018 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/jams/901
  • MathSciNet Review: 3836561
  • Show rawAMSref \bib{3836561}{article}{ author={Wilton, Henry}, title={Essential surfaces in graph pairs}, journal={J. Amer. Math. Soc.}, volume={31}, number={4}, date={2018-10}, pages={893-919}, issn={0894-0347}, review={3836561}, doi={10.1090/jams/901}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.