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

One-relator hierarchies

Marco Linton University of Oxford, Oxford, OX2 6GG, UK [email protected]
Abstract.

We prove that one-relator groups with negative immersions are hyperbolic and virtually special; this resolves a recent conjecture of Louder and Wilton. As a consequence, one-relator groups with negative immersions are residually finite, linear and have isomorphism problem decidable among one-relator groups. Using the fact that parafree one-relator groups have negative immersions, we answer a question of Baumslag’s from 1986. The main new tool we develop is a refinement of the classic Magnus–Moldavanskii hierarchy for one-relator groups. We introduce the notions of \mathbb{Z}-stable HNN-extensions and \mathbb{Z}-stable hierarchies. We then show that a one-relator group is hyperbolic and has a quasi-convex one-relator hierarchy if and only if it does not contain a Baumslag–Solitar subgroup and has a \mathbb{Z}-stable one-relator hierarchy.

1. Introduction

One-relator groups, despite their simple definition, have been stubbornly resistant to geometric characterisations. A well known conjecture attributed to Gersten states that a one-relator group is hyperbolic if and only if it does not contain Baumslag–Solitar subgroups [Ger92, AG99]. Modifying the hypothesis or conclusion of this conjecture appears to be problematic. For example, conjectures attempting to classify one-relator groups that are automatic [Ger92, MUW11] or act freely on a CAT(0) cube complex [Wis14] have been posed, but both have been disproven [GW19]. Conjectures attempting to classify when (subgroups of) groups of finite type are hyperbolic have also been posed [Bra99, Bes04], but have also been disproven [IMM23]. More variations can be found, for example, in [Gro93, Wis05, GKL21].

Recently, a different type of geometric characterisation for one-relator groups has emerged from work of Helfer and Wise [HW16] and, independently, Louder and Wilton [LW17]. Namely, if XX is the presentation complex of a one-relator group, then:

  1. (1)

    XX has not-too-positive immersions [LW17].

  2. (2)

    XX has non-positive immersions if and only if π1(X)\pi_{1}(X) is torsion-free [HW16].

  3. (3)

    XX has negative immersions if and only if every two-generator subgroup of π1(X)\pi_{1}(X) is free [LW22].

Moreover, each of these properties is decidable from XX [KMS60, LW22]. Since one-relator groups with negative immersions cannot contain Baumslag–Solitar subgroups, this led Louder and Wilton to make the following conjecture [LW22, Conjecture 1.9].

Conjecture.

Every one-relator group with negative immersions is hyperbolic.

The original notions of non-positive and negative immersions are due to Wise [Wis03, Wis04]; note that they are not the same as Louder and Wilton’s [LW24]. The aforementioned conjecture can be considered as a special case of an older conjecture of Wise [Wis04, Conjecture 14.2]. Another related property is that of uniform negative immersions, introduced in [LW24]. Results analogous to those proved by Louder and Wilton [LW22, LW24] for one-relator groups with negative immersions have also been shown for fundamental groups of two-complexes with a stronger form of uniform negative immersions by Wise [Wis20]. However, having uniform negative immersions turns out to be equivalent to having negative immersions in the case of one-relator complexes [LW24, Theorem C].

Louder and Wilton’s conjecture has been experimentally verified for all one-relator groups with negative immersions that admit a one-relator presentation with relator of length less than 17 [CH21]. In this article we verify their conjecture and in fact prove more.

{hyperbolicity_theorem}

One-relator groups with negative immersions are hyperbolic and virtually special.

Despite the maturity of the theory of one-relator groups, the isomorphism problem has remained almost untouched. A subclass of one-relator groups that is often mentioned when demonstrating the difficulty of this problem is that of parafree one-relator groups [CM82, BFR19]. Indeed, Baumslag asked [Bau86, Problem 4] whether the isomorphism problem for parafree one-relator groups is solvable. See also [FRMW21, Question 3]. A large body of work has been carried out on distinguishing parafree one-relator groups; see, for example, [FRS97, HK17, HK20, Che21] and [LL94, BCH04] for computational experiments. For several families of examples of parafree one-relator groups, see also [BC06]. A consequence of Theorem 7.2 is the following.

{isomorphism}

The isomorphism problem for one-relator groups with negative immersions is decidable within the class of one-relator groups.

By showing that parafree one-relator groups have negative immersions, we may answer Baumslag’s question in the affirmative.

{parafree}

The isomorphism problem for parafree one-relator groups is decidable within the class of one-relator groups.

We conclude this section by mentioning a couple of other corollaries.

{residual_finiteness}

One-relator groups with negative immersions are residually finite and linear.

{hyperbolicity_corollary}

Every finitely generated subgroup of a one-relator group with negative immersions is hyperbolic.

We now present the main new theorems that go into proving Theorem 7.2.

1.1. Magnus–Moldavanskii–Masters hierarchies

First conceived by Magnus in his thesis [Mag30], the Magnus hierarchy is possibly the oldest general tool in the theory of one-relator groups. After the introduction of the theory of HNN-extensions of groups, the hierarchy was later refined to be called the Magnus–Moldavanskii hierarchy [Mol67]: if GG is a one-relator group, then there is a diagram of monomorphisms of one-relator groups

G=G0{G=G_{0}}G0{G^{\prime}_{0}}G1{G_{1}}G1{G^{\prime}_{1}}{\cdots}{\cdots}GN{G_{N}}

such that GiGi+1ψiG_{i}^{\prime}\cong G_{i+1}*_{\psi_{i}} where ψi\psi_{i} identifies two Magnus subgroups of Gi+1G_{i+1}, and GNG_{N} splits as a free product of cyclic groups. The proofs of many results for one-relator groups then proceed by induction on the length of such a hierarchy. See [MKS66] for a classically flavoured introduction to one-relator groups with many such examples.

In [Mas06], Masters showed that we can dispense with the horizontal homomorphisms. In other words, if GG is a one-relator group, there is a sequence of monomorphisms of one-relator groups:

GNG1G0=GG_{N}\hookrightarrow...\hookrightarrow G_{1}\hookrightarrow G_{0}=G

such that GiGi+1ψiG_{i}\cong G_{i+1}*_{\psi_{i}} where ψi\psi_{i} identifies two Magnus subgroups of Gi+1G_{i+1}, and GNG_{N} splits as a free product of cyclic groups.

1.2. One-relator hierarchies

The versatility of the Magnus–Moldavanskii hierarchy comes from the fact that it may be described very explicitly in terms of one-relator presentations. Masters’ hierarchy is conceptually simpler, but is not so explicit. By working with two-complexes, we may reconcile both of these advantages. Our version of the hierarchy can be stated as follows.

{new_hierarchy}

Let XX be a finite one-relator complex. There exists a finite sequence of immersions of one-relator complexes:

XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X

such that π1(Xi)π1(Xi+1)ψi\pi_{1}(X_{i})\cong\pi_{1}(X_{i+1})*_{\psi_{i}} where ψi\psi_{i} is induced by an identification of Magnus subgraphs, and such that π1(XN)\pi_{1}(X_{N}) is finite cyclic.

Each such HNN-splitting is called a one-relator splitting and we call such a sequence of immersions a one-relator tower for XX. The sequence of immersions are tower maps as defined in [How81]: each immersion factors as an inclusion composed with a cyclic cover. The homomorphisms ψi\psi_{i} are thus induced by an identification of two subcomplexes by the deck group action. A maximal one-relator hierarchy is a maximal tower lifting of the induced map of the closed two-cell to the one-relator complex. Note that not every such tower lifting will induce a hierarchy of HNN-extensions of one-relator groups and thus the novelty of Theorem 4.13 lies in showing that such a tower always exists. The proof of Theorem 4.12 relies on a technical and combinatorial analysis of cyclic covers of two-complexes. Some of its applications not mentioned in this article are explored in the author’s thesis [Lin22].

1.3. \mathbb{Z}-stable one-relator hierarchies

We now introduce the notion of \mathbb{Z}-stable one-relator hierarchies. Let HH be a group and ψ:AB\psi:A\to B an isomorphism between subgroups of HH. Inductively define

𝒜0ψ={[H]},,𝒜i+1ψ={[ψ(AAi)]}Ai[Ai]𝒜iψ,\mathcal{A}^{\psi}_{0}=\{[H]\},\quad\ldots,\quad\mathcal{A}^{\psi}_{i+1}=\{[\psi(A\cap A_{i})]\}_{A_{i}\in[A_{i}]\in\mathcal{A}^{\psi}_{i}},\quad\ldots

where [Ai][A_{i}] denotes the conjugacy class of AiA_{i} in HH. Then we denote by 𝒜¯iψ𝒜iψ\bar{\mathcal{A}}_{i}^{\psi}\subset\mathcal{A}_{i}^{\psi} the subset corresponding to the conjugacy classes of non-cyclic subgroups. Define the \mathbb{Z}-stable number s(ψ)s\mathbb{Z}(\psi) of ψ\psi to be

s(ψ)=sup{k+1𝒜¯kψ}{}s\mathbb{Z}(\psi)=\sup{\{k+1\mid\bar{\mathcal{A}}_{k}^{\psi}\neq\emptyset\}}\in\mathbb{N}\cup\{\infty\}

where s(ψ)=s\mathbb{Z}(\psi)=\infty if 𝒜¯iψ\bar{\mathcal{A}}_{i}^{\psi}\neq\emptyset for all ii. In general, even 𝒜¯2ψ\bar{\mathcal{A}}_{2}^{\psi} may contain infinitely many conjugacy classes of subgroups. However, we show in Lemma 6.4 that if π1(X)π1(X1)ψ\pi_{1}(X)\cong\pi_{1}(X_{1})*_{\psi} is a one-relator splitting as in Theorem 4.13, then

[An]𝒜¯nψrr(An)rr(A)\sum_{[A_{n}]\in\bar{\mathcal{A}}_{n}^{\psi}}\operatorname{rr}(A_{n})\leq\operatorname{rr}(A)

where rr(A)=max{0,rk(A)1}\operatorname{rr}(A)=\max{\{0,\operatorname{rk}(A)-1\}} denotes the reduced rank of AA.

Definition 1.1.

A one-relator hierarchy XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X is a \mathbb{Z}-stable hierarchy if s(ψi)<s\mathbb{Z}(\psi_{i})<\infty for all i<Ni<N.

Our next result establishes an equivalence between quasi-convex one-relator hierarchies and \mathbb{Z}-stable one-relator hierarchies of hyperbolic one-relator groups.

{main}

Let XX be a one-relator complex and XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X a one-relator hierarchy. The following are equivalent:

  1. (1)

    XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X is a quasi-convex hierarchy and π1(X)\pi_{1}(X) is hyperbolic.

  2. (2)

    XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X is an acylindrical hierarchy.

  3. (3)

    XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X is a \mathbb{Z}-stable hierarchy and π1(X)\pi_{1}(X) contains no Baumslag–Solitar subgroups.

Moreover, if any of the above is satisfied, then π1(X)\pi_{1}(X) is virtually special and the image of π1(A)\pi_{1}(A) in π1(X)\pi_{1}(X) is quasi-convex for any connected subcomplex AXiA\subset X_{i}.

In [Wis21], Wise shows that Magnus–Moldavanskii hierarchies of one-relator groups with torsion are quasi-convex. We show that all one-relator hierarchies of one-relator complexes XX satisfying either of the following are \mathbb{Z}-stable:

  1. (1)

    π1(X)\pi_{1}(X) has torsion (by Corollary 6.7)

  2. (2)

    XX has negative immersions (by Corollary 6.9)

In the first case, since one-relator groups with torsion are hyperbolic [New68], we recover Wise’s result. In the second case, since fundamental groups of one-relator complexes with negative immersions are two-free, they cannot contain Baumslag–Solitar subgroups and so we prove Louder and Wilton’s conjecture.

By showing that hyperbolic groups with quasi-convex hierarchies are virtually special [Wis21], Wise also proved that one-relator groups with torsion are residually finite, settling an old conjecture of Baumslag [Bau67]. As a consequence of Wise’s work and Theorem 7.1, we may also establish virtual specialness, residual finiteness and linearity for one-relator groups with negative immersions.

Theorem 7.1 is the crux of the article and is easily applicable to concrete examples, as demonstrated by the following example.

Example 1.2.

Consider the following one-relator hierarchy of length one:

x,y,zz2yz2x2ψa,bb2a2b1aba2b2a2\langle x,y,z\mid z^{2}yz^{2}x^{-2}\rangle*_{\psi}\cong\langle a,b\mid b^{2}a^{2}b^{-1}aba^{2}b^{-2}a^{-2}\rangle

where ψ\psi is given by ψ(x)=y\psi(x)=y, ψ(y)=z\psi(y)=z. Since y=z2x2z2y=z^{-2}x^{2}z^{-2}, we see that x,y,z\langle x,y,z\rangle is a free group freely generated by xx and zz. Thus we have:

𝒜¯0ψ\displaystyle\bar{\mathcal{A}}_{0}^{\psi} ={[x,z]}\displaystyle=\{[\langle x,z\rangle]\}
𝒜¯1ψ\displaystyle\bar{\mathcal{A}}_{1}^{\psi} ={[x2,z]}\displaystyle=\{[\langle x^{2},z\rangle]\}
𝒜¯2ψ\displaystyle\bar{\mathcal{A}}_{2}^{\psi} ={[(z2x2z2)2,z]}\displaystyle=\{[\langle(z^{-2}x^{2}z^{-2})^{2},z\rangle]\}
𝒜¯3ψ\displaystyle\bar{\mathcal{A}}_{3}^{\psi} =\displaystyle=\emptyset

Hence, s(ψ)=3s\mathbb{Z}(\psi)=3 and this hierarchy is \mathbb{Z}-stable. Using a criterion for finding Baumslag–Solitar subgroups we develop in Subection 6.3, in Example 6.15 we show that

a,bb2a2b1aba2b2a2\langle a,b\mid b^{2}a^{2}b^{-1}aba^{2}b^{-2}a^{-2}\rangle

does not contain a Baumslag–Solitar subgroup. Thus, by Theorem 7.1, it is hyperbolic and virtually special.

The choice of group in Example 1.2 is motivated by [CH21] in which the authors verify hyperbolicity for one-relator groups with relator of length less than 17, using a combination of results from the literature and software in GAP and kbmag. This particular example does not satisfy any of the criteria the authors used: it is torsion-free, it is not small cancellation, it is not cyclically or conjugacy pinched, it does not satisfy the hypotheses of [IS98, Theorems 3 & 4] nor those of the hyperbolicity criterion in [BM22]. Moreover, it does not have negative immersions and does not split as an HNN-extension of a free group with a free factor edge group so that the results from [Mut21b, Mut21a] also do not apply.

1.4. Outline of the article

In Section 2, we introduce the necessary background and terminology on graphs and one-relator complexes. There we introduce the notion of a strongly inert graph immersion and use it to prove a key result, Theorem 2.11, bounding the sum of reduced ranks of intersections of certain subgroups in a one-relator group. This feeds into the proof of Theorem 7.1. In Section 3, we cover graphs of spaces and Bass–Serre theory. There we prove Proposition 3.2, a useful tool which will allow us to find HNN-splittings from cyclic covers of one-relator complexes as in Theorem 4.13. In Section 4 we introduce the notion of a \mathbb{Z}-domain, prove that \mathbb{Z}-domains of \mathbb{Z}-covers of finite CW-complexes exist and that minimal \mathbb{Z}-domains of \mathbb{Z}-covers of one-relator complexes are one-relator complexes. Combined with a complexity reduction argument, we then use this to establish Theorem 4.13. Section 5 is dedicated to showing that a hyperbolic one-relator group with a quasi-convex hierarchy has all of its Magnus subgroups quasi-convex. The proof relies on a careful analysis of normal forms induced by the hierarchy from Theorem 4.13. At this point of the article, it is possible to establish the equivalence of (1) and (2) from Theorem 7.1. In Section 6, \mathbb{Z}-stable hierarchies are introduced and some of their properties are proven. All hierarchies of one-relator groups with torsion and negative immersions are then shown to have such hierarchies. The section is concluded with Theorem 6.14 which shows the equivalence of (2) and (3) from Theorem 7.1 and provides a criterion for finding Baumslag–Solitar subgroups. Finally, Section 7 is dedicated to combining all the new tools from the article to prove our main results, Theorems 7.1 and 7.2.

Acknowledgements

We would like to thank Saul Schleimer and Henry Wilton for many stimulating conversations that helped improve the exposition of this article. We would also like to thank Lars Louder for his invaluable comments on Lemma 2.3. Finally, we would like to thank the anonymous referee for the many detailed and insightful comments which have enormously improved this article.

2. Graphs and one-relator complexes

A graph Γ\Gamma is a 1-dimensional CW-complex. We will write V(Γ)V(\Gamma) for the collection of 0-cells or vertices and E(Γ)E(\Gamma) for the collection of 1-cells or edges. We will usually assume Γ\Gamma to be oriented. An orientation will be induced by maps o:E(Γ)V(Γ)o:E(\Gamma)\to V(\Gamma) and t:E(Γ)V(Γ)t:E(\Gamma)\to V(\Gamma), the origin and target maps. For simplicity, we will write II to denote any connected graph whose vertices all have degree two, except for two vertices of degree one. Then S1S^{1} will denote a connected graph all of whose vertices have degree precisely two.

A map between graphs f:ΓΓf:\Gamma\to\Gamma^{\prime} is combinatorial if it sends each vertex to a vertex and each edge homeomorphically to an edge. A combinatorial map is an immersion if it is also locally injective. Combinatorial graph maps λ:IΓ\lambda:I\to\Gamma, λ:S1Γ\lambda:S^{1}\to\Gamma will be called paths and cycles respectively. The length of a combinatorial path λ:IΓ\lambda:I\to\Gamma is the number of edges in II and is denoted by |λ|\left\lvert\lambda\right\rvert. If λ:IX\lambda:I\to X is a path, we may often identify the vertices of II with the integers 0,1,,|λ|0,1,...,\left\lvert\lambda\right\rvert so that λ(i)\lambda(i) is the ithi^{\text{th}} vertex that λ\lambda traverses. We also put o(λ)=λ(0)o(\lambda)=\lambda(0) and t(λ)=λ(|λ|)t(\lambda)=\lambda(\left\lvert\lambda\right\rvert). We similarly define the length of a cycle λ:S1Γ\lambda:S^{1}\to\Gamma. A cycle λ:S1Γ\lambda:S^{1}\to\Gamma is primitive if λ\lambda does not factor through any non-trivial covering map S1S1S^{1}\to S^{1}. We will call it imprimitive otherwise. We will write deg(λ)\deg(\lambda) to denote the maximal degree of a covering map S1S1S^{1}\to S^{1} that λ\lambda factors through. Note that deg(λ)=1\deg(\lambda)=1 if and only if λ\lambda is primitive. We remark that this definition of primitivity should be compared with the definition of primitivity from the theory of combinatorics of words, not with the definition of primitivity in the theory of free groups. In particular, λ\lambda being imprimitive is not the same as Im(λ)\operatorname{Im}(\lambda_{*}) being imprimitive in π1(Γ)\pi_{1}(\Gamma).

The core of a graph Γ\Gamma is the subgraph consisting of the union of all the images of immersed cycles S1ΓS^{1}\looparrowright\Gamma and will be denoted by Core(Γ)\operatorname{Core}(\Gamma). Note that if Γ\Gamma is a forest, then Core(Γ)=\operatorname{Core}(\Gamma)=\emptyset. In particular Core(Γ)\operatorname{Core}(\Gamma) is unique.

2.1. Strongly inert graph immersions

If GG is a group and g,hGg,h\in G, we will adopt the usual convention that hg=g1hgh^{g}=g^{-1}\cdot h\cdot g. If H<GH<G, we will write [H][H] to denote the conjugacy class of HH in GG. More generally, If X,YGX,Y\subset G are subsets, then define the YY-conjugacy class of XX to be the following:

[X]Y={XyyY}[X]_{Y}=\{X^{y}\mid y\in Y\}

A subgroup H<GH<G is called inert if for every subgroup K<GK<G, we have rk(HK)rk(K)\operatorname{rk}(H\cap K)\leq\operatorname{rk}(K). This definition was first introduced in [DV96], motivated by the study of fixed subgroups of endomorphisms of free groups. More generally, as defined in [Iva18], we say that HH is strongly inert if for every subgroup K<GK<G, we have:

KgHrr(HKg)rr(K)\sum_{KgH}\operatorname{rr}(H\cap K^{g})\leq\operatorname{rr}(K)

Examples of strongly inert subgroups of free groups are:

  1. (1)

    subgroups of rank at most two [Tar92],

  2. (2)

    subgroups that are the fixed subgroup of an injective endomorphism [DV96, Theorem IV.5.5],

  3. (3)

    subgroups that are images of immersions of free groups, as defined in [Kap00].

The latter example follows by observing that the fibre product (in the sense of [Sta83]) of two rose graphs can contain at most a single vertex of valence greater than two.

Remark 2.1.

If H<GH<G is strongly inert, then applying the definition to K=HK=H, we get that rr(HHg)0\operatorname{rr}(H\cap H^{g})\leq 0 for all gHg\notin H. Thus, HH is cyclonormal.

Translating this to graphs, we make the following definition.

Definition 2.2.

A graph immersion γ:ΓΔ\gamma:\Gamma\looparrowright\Delta is strongly inert if for all graph immersions λ:ΛΔ\lambda:\Lambda\looparrowright\Delta, the following is satisfied:

χ(Core(Γ×ΔΛ))χ(Core(Λ)).\chi(\operatorname{Core}(\Gamma\times_{\Delta}\Lambda))\geq\chi(\operatorname{Core}(\Lambda)).

Let us briefly explain how this is a translation of strong inertia to graphs. If Δ\Delta is a (non-empty) connected core graph, we have that rr(π1(Δ))=χ(Δ)\operatorname{rr}(\pi_{1}(\Delta))=-\chi(\Delta). By work of Stallings [Sta83], if γ:ΓΔ\gamma\colon\Gamma\looparrowright\Delta and λ:ΛΔ\lambda\colon\Lambda\looparrowright\Delta are graph immersions, then the components of Core(Γ×ΔΛ)\operatorname{Core}(\Gamma\times_{\Delta}\Lambda) are in natural bijection with the non-trivial intersections γπ1(Γ)λπ1(Λ)g\gamma_{*}\pi_{1}(\Gamma)\cap\lambda_{*}\pi_{1}(\Lambda)^{g}, as gg varies over representatives for the double cosets λπ1(Λ)gγπ1(Γ)\lambda_{*}\pi_{1}(\Lambda)g\gamma_{*}\pi_{1}(\Gamma). As such, we have

χ(Core(Γ×ΔΛ))=λπ1(Λ)gγπ1(Γ)rr(γπ1(Γ)λπ1(Λ)g).-\chi(\operatorname{Core}(\Gamma\times_{\Delta}\Lambda))=\sum_{\lambda_{*}\pi_{1}(\Lambda)g\gamma_{*}\pi_{1}(\Gamma)}\operatorname{rr}(\gamma_{*}\pi_{1}(\Gamma)\cap\lambda_{*}\pi_{1}(\Lambda)^{g}).

The following lemma will produce more examples of strongly inert subgroups of free groups.

Lemma 2.3.

Let γ:ΓΔ\gamma:\Gamma\looparrowright\Delta be an immersion of finite graphs. If γπ1(Γ)F\gamma_{*}\pi_{1}(\Gamma)*F is an inert subgroup of π1(Δ)F\pi_{1}(\Delta)*F for all free groups FF, then γ\gamma is a strongly inert graph immersion.

Proof.

Suppose for a contradiction that there exists a graph immersion λ:ΛΔ\lambda:\Lambda\looparrowright\Delta such that:

χ(Core(Γ×ΔΛ))<χ(Core(Λ)).\chi(\operatorname{Core}(\Gamma\times_{\Delta}\Lambda))<\chi(\operatorname{Core}(\Lambda)).

Let Θ0,,ΘkCore(Γ×ΔΛ)\Theta_{0},...,\Theta_{k}\subset\operatorname{Core}(\Gamma\times_{\Delta}\Lambda) be the connected components. Choose vertices, (vi,wi)V(Θi)(v_{i},w_{i})\in V(\Theta_{i}) for each 0ik0\leqslant i\leqslant k. Let Δ\Delta^{\prime} be the graph with vertex set V(Δ)V(\Delta) and edge set E(Δ){e1,,ek}E(\Delta)\sqcup\{e_{1},\ldots,e_{k}\} where the origin and target of each eie_{i} is the basepoint. Similarly, we let Γ\Gamma^{\prime} be the graph with vertex set V(Γ)V(\Gamma) and edge set E(Γ){f1,,fk}E(\Gamma)\sqcup\{f_{1},\ldots,f_{k}\} where the origin and target of each fif_{i} is the basepoint. We then define a graph map γ:ΓΔ\gamma^{\prime}\colon\Gamma^{\prime}\looparrowright\Delta^{\prime} by γ(v)=γ(v)\gamma^{\prime}(v)=\gamma(v) for each vV(Γ)v\in V(\Gamma^{\prime}), γ(f)=γ(f)\gamma^{\prime}(f)=\gamma(f) if fE(Γ)f\in E(\Gamma) and γ(fi)=ei\gamma^{\prime}(f_{i})=e_{i} for each 1ik1\leqslant i\leqslant k. We have that π1(Δ)=π1(Δ)F\pi_{1}(\Delta^{\prime})=\pi_{1}(\Delta)*F where FF is a free group of rank kk and γπ1(Γ)=γπ1(Γ)F\gamma^{\prime}_{*}\pi_{1}(\Gamma^{\prime})=\gamma_{*}\pi_{1}(\Gamma)*F. Hence, by assumption, γπ1(Γ)\gamma^{\prime}_{*}\pi_{1}(\Gamma^{\prime}) is inert in π1(Δ)\pi_{1}(\Delta^{\prime}).

Now for each 1ik1\leqslant i\leqslant k, let gi:IiΓg_{i}:I_{i}\looparrowright\Gamma be any immersed path such that gig_{i} begins at vi1v_{i-1}, ends at viv_{i}, traverses fif_{i} precisely once and does not traverse fjf_{j} for any jij\neq i. Let Λ\Lambda^{\prime} be the graph obtained from Λ\Lambda by attaching segments IiI_{i} with origin wi1w_{i-1} and with target wiw_{i} for each 1ik1\leqslant i\leqslant k. Let λ:ΛΔ\lambda^{\prime}:\Lambda^{\prime}\to\Delta^{\prime} be the map obtained by extending λ\lambda by defining λIi=γgi\lambda^{\prime}\mid_{I_{i}}=\gamma\circ g_{i}. Finally, let λ′′:Λ′′Δ\lambda^{\prime\prime}:\Lambda^{\prime\prime}\looparrowright\Delta be the graph immersion obtained from λ\lambda^{\prime} by folding. By construction, for each ii there is only one edge in Λ\Lambda^{\prime} that maps to eie_{i}. Thus, each such edge does not get identified with any other edge under the folding map and so the folding map ΛΛ′′\Lambda^{\prime}\to\Lambda^{\prime\prime} is a homotopy equivalence, restricting to an isomorphism on ΛΛ\Lambda\subset\Lambda^{\prime}. Now not only will Core(Γ×ΔΛ′′)\operatorname{Core}(\Gamma^{\prime}\times_{\Delta^{\prime}}\Lambda^{\prime\prime}) contain Core(Γ×ΔΛ)\operatorname{Core}(\Gamma\times_{\Delta}\Lambda) as a subgraph, but it will also contain segments connecting Θi1\Theta_{i-1} to Θi\Theta_{i} for all 1ik1\leqslant i\leqslant k by construction. We have χ(Λ′′)=χ(Λ)k\chi(\Lambda^{\prime\prime})=\chi(\Lambda)-k, and so:

χ(Core(Γ×ΔΛ′′))\displaystyle\chi(\operatorname{Core}(\Gamma^{\prime}\times_{\Delta^{\prime}}\Lambda^{\prime\prime})) =χ(Core(Γ×ΔΛ))k\displaystyle=\chi(\operatorname{Core}(\Gamma^{\prime}\times_{\Delta^{\prime}}\Lambda))-k
<χ(Core(Λ′′)).\displaystyle<\chi(\operatorname{Core}(\Lambda^{\prime\prime})).

However, Core(Γ×ΔΛ′′)\operatorname{Core}(\Gamma^{\prime}\times_{\Delta^{\prime}}\Lambda^{\prime\prime}) is connected by construction, contradicting the fact that γπ1(Γ)\gamma^{\prime}_{*}\pi_{1}(\Gamma^{\prime}) was inert in π1(Δ)\pi_{1}(\Delta^{\prime}). ∎

2.2. One-relator complexes and Magnus subgraphs

For simplicity, we will mostly be restricting our attention to particular kinds of CW-complexes called combinatorial 22-complexes.

Definition 2.4.

A combinatorial 22-complex XX is a 22-dimensional CW-complex whose attaching maps are all immersions. We will usually write X=(Γ,λ)X=(\Gamma,\lambda) where Γ\Gamma is a graph and λ:𝕊=S1Γ\lambda:\mathbb{S}=\sqcup S^{1}\looparrowright\Gamma is an immersion of a disjoint union of cycles.

We will also restrict our maps to combinatorial maps.

Definition 2.5.

A combinatorial map of combinatorial 22-complexes f:YXf:Y\to X is a map that restricts to a combinatorial map of graphs fΓ:ΓYΓXf_{\Gamma}:\Gamma_{Y}\to\Gamma_{X} and induces a combinatorial map f𝕊:𝕊Y𝕊Xf_{\mathbb{S}}:\mathbb{S}_{Y}\to\mathbb{S}_{X} such that fΓλY=λXf𝕊f_{\Gamma}\circ\lambda_{Y}=\lambda_{X}\circ f_{\mathbb{S}}. We say that ff is an immersion if fΓf_{\Gamma} is an immersion and f𝕊f_{\mathbb{S}} restricts to a homeomorphism on each component.

Since we will always be assuming that our maps are combinatorial, we will often simply neglect to use the descriptor.

The main class of 22-complexes that we will be working with are one-relator complexes. That is, combinatorial 22-complexes of the form X=(Γ,λ)X=(\Gamma,\lambda) where λ:S1Γ\lambda:S^{1}\looparrowright\Gamma is an immersion of a single cycle. Denote by XλXX_{\lambda}\subset X the smallest subcomplex that is a one-relator complex. The following result is the classic Freiheitssatz of Magnus [Mag30].

Theorem 2.6.

Let X=(Γ,λ)X=(\Gamma,\lambda) be a one-relator complex. If ΛΓ\Lambda\subset\Gamma is a connected subgraph in which λ\lambda is not supported, then π1(Λ)π1(X)\pi_{1}(\Lambda)\to\pi_{1}(X) is injective.

We will call subgraphs of one-relator complexes that satisfy the hypothesis of Theorem 2.6, Magnus subgraphs. This is in analogy with Magnus subgroups: if GG is a one-relator group with one-relator presentation Σr\langle\Sigma\mid r\rangle, with rr cyclically reduced, then a Magnus subgroup for this presentation is a subgroup generated by a subset ΛΣ\Lambda\subset\Sigma such that rΛ<F(Σ)r\notin\langle\Lambda\rangle<F(\Sigma).

If XX is a one-relator complex and AXA\subset X is a Magnus subgraph, then π1(A)\pi_{1}(A) is a Magnus subgroup for some one-relator presentation of π1(X)\pi_{1}(X). We can see this by taking a spanning tree TΓT\subset\Gamma such that TAT\cap A is a spanning tree for AA. Then by contracting TT, we obtain a presentation complex for a one-relator presentation of π1(X)\pi_{1}(X) in which π1(A)\pi_{1}(A) is a Magnus subgroup.

If BXB\subset X is another Magnus subgraphs and ABA\cap B is connected, then, as above, we may obtain a one-relator presentation for π1(X)\pi_{1}(X) in which both π1(A)\pi_{1}(A) and π1(B)\pi_{1}(B) are Magnus subgroups. If ABA\cap B is not connected, then this is no longer the case. Nevertheless, by adding edges to A,BA,B so that ABA\cap B is connected, we see that π1(A)F\pi_{1}(A)*F and π1(B)F\pi_{1}(B)*F are Magnus subgroups for some one-relator presentation of π1(X)F\pi_{1}(X)*F, where FF is a finitely generated free group.

The interactions between Magnus subgroups of one-relator groups are well understood. The following theorems are the main results in [Col04] and [Col08] respectively.

Theorem 2.7.

Let X=(Γ,λ)X=(\Gamma,\lambda) be a one-relator complex and let A,BXA,B\subset X be Magnus subgraphs with ABA\cap B connected. Then one of the following holds:

  1. (1)

    π1(A)π1(B)=π1(AB)\pi_{1}(A)\cap\pi_{1}(B)=\pi_{1}(A\cap B),

  2. (2)

    π1(A)π1(B)=π1(AB)c\pi_{1}(A)\cap\pi_{1}(B)=\pi_{1}(A\cap B)*\langle c\rangle for some cπ1(X)c\in\pi_{1}(X).

If λ\lambda is imprimitive, then π1(A)π1(B)=π1(AB)\pi_{1}(A)\cap\pi_{1}(B)=\pi_{1}(A\cap B).

We say a pair of Magnus subgroups π1(A)\pi_{1}(A), π1(B)<π1(X)\pi_{1}(B)<\pi_{1}(X) have exceptional intersection if the latter situation occurs.

Theorem 2.8.

Let X=(Γ,λ)X=(\Gamma,\lambda) be a one-relator complex and let A,BXA,B\subset X be Magnus subgraphs with ABA\cap B connected. Then for any gπ1(X)g\in\pi_{1}(X), one of the following holds:

  1. (1)

    π1(A)π1(B)g=1\pi_{1}(A)\cap\pi_{1}(B)^{g}=1,

  2. (2)

    π1(A)π1(B)g=c\pi_{1}(A)\cap\pi_{1}(B)^{g}=\langle c\rangle for some cπ1(X)c\in\pi_{1}(X),

  3. (3)

    gπ1(B)π1(A)g\in\pi_{1}(B)\cdot\pi_{1}(A).

If λ\lambda is imprimitive, then either π1(A)π1(B)g=1\pi_{1}(A)\cap\pi_{1}(B)^{g}=1 or gπ1(B)π1(A)g\in\pi_{1}(B)\cdot\pi_{1}(A).

Remark 2.9.

Recall that the fundamental group of a one-relator complex X=(Γ,λ)X=(\Gamma,\lambda) has torsion if and only if λ\lambda is imprimitive by [KMS60].

We may, in some sense, strengthen Theorems 2.7 and 2.8 to incorporate intersections of subgroups of fundamental groups of Magnus subgraphs. First, we will need the following lemma.

Lemma 2.10.

Let X=(Γ,λ)X=(\Gamma,\lambda) be a one-relator complex and let A,BXA,B\subset X be Magnus subgraphs with ABA\cap B connected. If γ:ΓA\gamma:\Gamma\looparrowright A is a graph immersion such that γπ1(Γ)=π1(A)π1(B)\gamma_{*}\pi_{1}(\Gamma)=\pi_{1}(A)\cap\pi_{1}(B), then γ\gamma is a strongly inert graph immersion. In particular, π1(A)π1(B)\pi_{1}(A)\cap\pi_{1}(B) is strongly inert in π1(A)\pi_{1}(A).

Proof.

By Theorem 2.7, π1(A)π1(B)\pi_{1}(A)\cap\pi_{1}(B) is an echelon subgroup of π1(A)\pi_{1}(A) (see [Ros13, Definitions 3.1 & 3.3] for the definition of an echelon subgroup of a free group). If F,GF,G are free groups and J<FJ<F, K<GK<G are echelon subgroups, then by definition, JK<FGJ*K<F*G is an echelon subgroup. Thus, Lemma 2.3, combined with [Ros13, Theorem 3.6] implies that γ\gamma is a strongly inert graph immersion. ∎

Theorem 2.11.

Let X=(Γ,λ)X=(\Gamma,\lambda) be a one-relator complex and A,BXA,B\subset X be Magnus subgraphs with ABA\cap B connected. If Cπ1(A)C\leqslant\pi_{1}(A) and Dπ1(B)D\leqslant\pi_{1}(B) are finitely generated subgroups, with CC a strongly inert subgroup of π1(A)\pi_{1}(A), then the following is satisfied:

DgCgπ1(X)rr(CDg)=DgCgπ1(B)π1(A)rr(CDg)rr(D)\sum_{\begin{subarray}{c}DgC\\ g\in\pi_{1}(X)\end{subarray}}\operatorname{rr}(C\cap D^{g})=\sum_{\begin{subarray}{c}DgC\\ g\in\pi_{1}(B)\pi_{1}(A)\end{subarray}}\operatorname{rr}(C\cap D^{g})\leq\operatorname{rr}(D)
Proof.

The first equality follows from Theorem 2.8. Since Cπ1(A)C\leqslant\pi_{1}(A) and Dπ1(B)D\leqslant\pi_{1}(B), we have that

DgCgπ1(B)π1(A)rr(CDg)\displaystyle\sum_{\begin{subarray}{c}DgC\\ g\in\pi_{1}(B)\pi_{1}(A)\end{subarray}}\operatorname{rr}(C\cap D^{g}) DbaCbπ1(B),aπ1(A)rr(Ca1Db)\displaystyle\leqslant\sum_{\begin{subarray}{c}DbaC\\ b\in\pi_{1}(B),a\in\pi_{1}(A)\end{subarray}}\operatorname{rr}\left(C^{a^{-1}}\cap D^{b}\right)
=DbaCbπ1(B),aπ1(A)rr(Ca1ABDb)\displaystyle=\sum_{\begin{subarray}{c}DbaC\\ b\in\pi_{1}(B),a\in\pi_{1}(A)\end{subarray}}\operatorname{rr}\left(C^{a^{-1}}\cap A\cap B\cap D^{b}\right)
=DbaCbπ1(B),aπ1(A)rr(C((AB)Db)a).\displaystyle=\sum_{\begin{subarray}{c}DbaC\\ b\in\pi_{1}(B),a\in\pi_{1}(A)\end{subarray}}\operatorname{rr}\left(C\cap\left((A\cap B)\cap D^{b}\right)^{a}\right).

Next, let SS be a set of double coset representatives for D\π1(B)π1(A)/CD\backslash\pi_{1}(B)\pi_{1}(A)/C. For each gSg\in S, choose an element bπ1(B)b\in\pi_{1}(B) such that b1gπ1(A)b^{-1}g\in\pi_{1}(A). Denote by SBπ1(B)S_{B}\subset\pi_{1}(B) the set of such elements. Denote by SbS_{b} the set of distinct aπ1(A)a\in\pi_{1}(A) such that baSba\in S. Each element in SbS_{b} is in a distinct ((π1(A)π1(B))Db)\π1(A)/C((\pi_{1}(A)\cap\pi_{1}(B))\cap D^{b})\backslash\pi_{1}(A)/C double coset. Each element in SBS_{B} is in a distinct D\π1(B)/(π1(A)π1(B))D\backslash\pi_{1}(B)/(\pi_{1}(A)\cap\pi_{1}(B)) double coset. Then

DgCgπ1(B)π1(A)rr(CDg)\displaystyle\sum_{\begin{subarray}{c}DgC\\ g\in\pi_{1}(B)\pi_{1}(A)\end{subarray}}\operatorname{rr}\left(C\cap D^{g}\right) =bSBaSbrr(C((π1(A)π1(B))Db)a)\displaystyle=\sum_{b\in S_{B}}\sum_{a\in S_{b}}\operatorname{rr}\left(C\cap\left((\pi_{1}(A)\cap\pi_{1}(B))\cap D^{b}\right)^{a}\right)
bSBrr((π1(A)π1(B))Db)\displaystyle\leqslant\sum_{b\in S_{B}}\operatorname{rr}\left((\pi_{1}(A)\cap\pi_{1}(B))\cap D^{b}\right)
rr(D)\displaystyle\leqslant\operatorname{rr}(D)

where the first inequality follows from the fact that CC is strongly inert in π1(A)\pi_{1}(A) and the second inequality follows from the fact that π1(A)π1(B)\pi_{1}(A)\cap\pi_{1}(B) is strongly inert in π1(B)\pi_{1}(B) by Lemma 2.10. ∎

3. Graphs of spaces

Let Γ\Gamma be a connected graph. Let {Xv}vV(Γ)\{X_{v}\}_{v\in V(\Gamma)} and {Xe}eE(Γ)\{X_{e}\}_{e\in E(\Gamma)} be collections of connected CW-complexes. We call these the vertex spaces and edge spaces respectively. Let eE(Γ)e\in E(\Gamma) and o(e)=vo(e)=v_{-} and t(e)=v+t(e)=v_{+}; then let e±:XeXv±\partial^{\pm}_{e}:X_{e}\to X_{v_{\pm}} be π1\pi_{1}-injective combinatorial maps. This data determines a graph of spaces 𝒳=(Γ,{Xv},{Xe},{e±})\mathcal{X}=(\Gamma,\{X_{v}\},\{X_{e}\},\{\partial_{e}^{\pm}\}). The geometric realisation of 𝒳\mathcal{X} is defined as follows:

X𝒳=(vV(Γ)XveE(Γ)(Xe×[1,1]))/X_{\mathcal{X}}=\left(\bigsqcup_{v\in V(\Gamma)}X_{v}\sqcup\bigsqcup_{e\in E(\Gamma)}(X_{e}\times[-1,1])\right)/\sim

with (x,±1)e±(x)(x,\pm 1)\sim\partial^{\pm}_{e}(x) for each eE(Γ)e\in E(\Gamma). This space has a CW-complex structure in the obvious way. We will say a cell cX𝒳c\subset X_{\mathcal{X}} is horizontal if its attaching map is supported in a vertex space, vertical otherwise.

There is a natural vertical map

𝗏:X𝒳Γ\mathsf{v}:X_{\mathcal{X}}\to\Gamma

where XvX_{v} maps to vv and Xe×(1,1)X_{e}\times(-1,1) maps to the open edge ee in the obvious way. The following fact about the vertical map 𝗏\mathsf{v} is well known, see [Ser03].

Lemma 3.1.

The map:

𝗏:π1(X𝒳)π1(Γ)\mathsf{v}_{*}:\pi_{1}(X_{\mathcal{X}})\to\pi_{1}(\Gamma)

is surjective.

We may also define a horizontal map:

𝗁:X𝒳H𝒳\mathsf{h}:X_{\mathcal{X}}\to H_{\mathcal{X}}

as the quotient map given by the transitive closure of the relation which, for each eE(Γ)e\in E(\Gamma), i[1,1]i\in[-1,1] and xXex\in X_{e}, identifies (x,i)(x,i) with e+(x)\partial^{+}_{e}(x) and (x)\partial^{-}(x). A path p:IX𝒳p:I\to X_{\mathcal{X}} is vertical if 𝗁p\mathsf{h}\circ p is a constant path and horizontal if 𝗏p\mathsf{v}\circ p is a constant path.

In general, not much can be said about the horizontal map. However, with sufficient restrictions on the edge maps, we can show that it is a homotopy equivalence.

Proposition 3.2.

Let 𝒳=(Γ,{Xv},{Xe},{e±})\mathcal{X}=(\Gamma,\{X_{v}\},\{X_{e}\},\{\partial_{e}^{\pm}\}) be a graph of spaces. Suppose that e±\partial_{e}^{\pm} are given by inclusions of subcomplexes. Then the following hold:

  1. (1)

    𝗁:X𝒳H𝒳\mathsf{h}:X_{\mathcal{X}}\to H_{\mathcal{X}} is a homotopy equivalence if and only if 𝗁1(𝗁(c))\mathsf{h}^{-1}(\mathsf{h}(c)) is a tree for each 0-cell cX𝒳(0)c\in X_{\mathcal{X}}^{(0)}.

  2. (2)

    If 𝗁\mathsf{h} is a homotopy equivalence, then H𝒳H_{\mathcal{X}} has a CW-structure inherited from X𝒳X_{\mathcal{X}} and 𝗁Xv\mathsf{h}\mid X_{v} is an immersion for all vV(Γ)v\in V(\Gamma).

Proof.

Let cX𝒳c\subset X_{\mathcal{X}} be an open horizontal cell and denote by Γc=𝗁1(𝗁(c))\Gamma_{c}=\mathsf{h}^{-1}(\mathsf{h}(c)). If Γc\Gamma_{c} is not a tree for some 0-cell cc, then there is a loop in ΓcX𝒳\Gamma_{c}\subset X_{\mathcal{X}} that maps to a non-trivial loop in Γ\Gamma under the vertical map, but that maps to the trivial loop under 𝗁\mathsf{h}. Thus, 𝗁\mathsf{h} cannot be a homotopy equivalence. So now let us suppose that Γc\Gamma_{c} is a tree for all 0-cells cX𝒳(0)c\in X_{\mathcal{X}}^{(0)}.

We show by induction on nn that for each horizontal nn-cell cX𝒳c\subset X_{\mathcal{X}}, the preimage 𝗁1(𝗁(c))\mathsf{h}^{-1}(\mathsf{h}(c)) is homeomorphic to c×Γcc\times\Gamma_{c} where ΓcX𝒳\Gamma_{c}\subset X_{\mathcal{X}} is a tree containing only vertical edges and hence immersing into Γ\Gamma via the vertical map. The claim holds for n=0n=0 by the previous paragraph so now assume that n1n\geqslant 1 and the inductive hypothesis holds. If dX𝒳d\subset X_{\mathcal{X}} is an open cell of dimension strictly less than nn and such that the image of the attaching map for cc intersects dd, then the immersion ΓcΓ\Gamma_{c}\looparrowright\Gamma factors through the immersion ΓdΓ\Gamma_{d}\looparrowright\Gamma. Indeed, if cXvc\subset X_{v} lies in the image of e±\partial^{\pm}_{e}, then so must dXvd\subset X_{v}. Thus, the edge in Γc\Gamma_{c} corresponding to c×[1,1]Xe×[1,1]c\times[-1,1]\subset X_{e}\times[-1,1] maps to the edge in Γd\Gamma_{d} corresponding to d×[1,1]Xe×[1,1]d\times[-1,1]\subset X_{e}\times[-1,1]. Here we are using the fact that e±\partial^{\pm}_{e} is an inclusion of subcomplexes to naturally identify c,dXvc,d\subset X_{v} with the corresponding cells in XeX_{e}. Since Γd\Gamma_{d} is a tree by assumption, Γc\Gamma_{c} must also be a tree and, in particular, Γc\Gamma_{c} can be seen as a subtree of Γd\Gamma_{d}.

We may now define the cellular structure on H𝒳H_{\mathcal{X}}. Define an equivalence relation on the horizontal cells of X𝒳X_{\mathcal{X}} defined by c1c2c_{1}\sim c_{2} if c1𝗁1(𝗁(c2))c_{1}\subset\mathsf{h}^{-1}(\mathsf{h}(c_{2})) and c2𝗁1(𝗁(c1))c_{2}\subset\mathsf{h}^{-1}(\mathsf{h}(c_{1})). By the previous paragraph, each open cell c[c]c\in[c] maps homeomorphically to its image via 𝗁\mathsf{h} and each cell in a given equivalence class has the same image. Choosing a representative cc for each equivalence class [c][c], there is an open nn-cell 𝗁(c)H𝒳\mathsf{h}(c)\subset H_{\mathcal{X}} with attaching map given by composing the attaching map for cc with the horizontal map 𝗁\mathsf{h}. Since each preimage of an open nn-cell in H𝒳H_{\mathcal{X}} is contractible, a standard result now implies that 𝗁\mathsf{h} is a homotopy equivalence (for instance, see the proof of [Hat02, Proposition 0.17]).

With this cellular structure on H𝒳H_{\mathcal{X}}, we now show that 𝗁Xv\mathsf{h}\mid X_{v} is an immersion for each vV(Γ)v\in V(\Gamma). It is clearly an immersion on 0-skeleta. If it is an immersion on nn-skeleta, but not on (n+1)(n+1)-skeleta, let xXvx\in X_{v} be a point at which 𝗁Xv\mathsf{h}\mid X_{v} is not an immersion on the (n+1)(n+1)-skeleton. Since 𝗁\mathsf{h} restricts to homeomorphisms on each open cell in XvX_{v}, we must have that xx lies in some open cell cc of dimension nn or less. Since 𝗁Xv\mathsf{h}\mid X_{v} is an immersion on the nn-skeleton, there are two distinct open (n+1)(n+1)-cells c1,c2Xvc_{1},c_{2}\subset X_{v} that lie in the same equivalence class and such that the image of their attaching maps intersect cc. By definition of the tree Γc\Gamma_{c}, this would then yield a non-trivial loop in Γc\Gamma_{c} which is a contradiction. Hence Xv\mathsf{\mid}X_{v} is an immersion for all vV(Γ)v\in V(\Gamma). ∎

If 𝒳\mathcal{X} is a graph of spaces, then the universal cover X~𝒳X𝒳\tilde{X}_{\mathcal{X}}\to X_{\mathcal{X}} also has a graph of spaces structure where each vertex space is the universal cover of some vertex space of 𝒳\mathcal{X} and each edge space is the universal cover of some edge space of 𝒳\mathcal{X}. We will denote this by 𝒳~=(T,{X~v},{X~e},{~e±})\tilde{\mathcal{X}}=(T,\{\tilde{X}_{v}\},\{\tilde{X}_{e}\},\{\tilde{\partial}^{\pm}_{e}\}) where TT is the Bass–Serre tree of 𝒳\mathcal{X}. There is a natural covering action of π1(X𝒳)\pi_{1}(X_{\mathcal{X}}) on X~𝒳\tilde{X}_{\mathcal{X}} which pushes forward to an action on TT. Indeed, we have the following π1(X𝒳)\pi_{1}(X_{\mathcal{X}})-equivariant commuting diagram:

X~𝒳=X𝒳~{\tilde{X}_{\mathcal{X}}=X_{\tilde{\mathcal{X}}}}T{T}X𝒳{X_{\mathcal{X}}}Γ{\Gamma}

If X~\tilde{X} satisfies the hypothesis of Proposition 3.2, then we also have a π1(X𝒳)\pi_{1}(X_{\mathcal{X}})-equivariant commuting diagram:

H~𝒳=H𝒳~{\tilde{H}_{\mathcal{X}}=H_{\tilde{\mathcal{X}}}}X~𝒳{\tilde{X}_{\mathcal{X}}}T{T}H𝒳{H_{\mathcal{X}}}X𝒳{X_{\mathcal{X}}}Γ{\Gamma}

3.1. Hyperbolic graphs of spaces

Acylindrical actions were first defined by Sela in [Sel97].

Definition 3.3.

Let GG be a group acting on a tree TT. The action is kk-acylindrical if every subset of TT of diameter at least kk is pointwise stabilised by at most finitely many elements. We will say the action is acylindrical if there exists some constant k>0k>0 such that the action is kk-acylindrical.

By putting constraints on the geometry of the vertex and edge spaces of a graph of spaces 𝒳\mathcal{X}, we may deduce acylindricity of the action of π1(X𝒳)\pi_{1}(X_{\mathcal{X}}) on its Bass–Serre tree. The notion of hyperbolicity that we use is that of Gromov’s δ\delta-hyperbolicity. One direction of the following theorem is due to Bestvina–Feighn [BF92] (for the hyperbolicity statement) and Kapovich [Kap01, Theorem 1.2] (for the quasi-convexity statement), while the other direction follows from the fact that quasi-convex subgroups of hyperbolic groups have finite height, due to Gitik–Mitra–Rips–Sageev [GMRS98].

Theorem 3.4.

Let 𝒳=(Γ,{Xv},{Xe},{e±})\mathcal{X}=(\Gamma,\{X_{v}\},\{X_{e}\},\{\partial_{e}^{\pm}\}) be a graph of spaces such that X𝒳X_{\mathcal{X}} is compact, X~v\tilde{X}_{v} is hyperbolic for all vV(Γ)v\in V(\Gamma) and ~e±\tilde{\partial}_{e}^{\pm} is a quasi-isometric embedding for all eE(Γ)e\in E(\Gamma). Then π1(X𝒳)\pi_{1}(X_{\mathcal{X}}) acts acylindrically on the Bass-Serre tree TT, if and only if X~𝒳(1)\tilde{X}_{\mathcal{X}}^{(1)} is hyperbolic with the path metric and X~v(1)X~𝒳(1)\tilde{X}^{(1)}_{v}\to\tilde{X}^{(1)}_{\mathcal{X}} (or X~e(1)X~𝒳(1)\tilde{X}^{(1)}_{e}\to\tilde{X}^{(1)}_{\mathcal{X}}) is a quasi-isometric embedding for all vE(Γ)v\in E(\Gamma) (or for all eE(Γ)e\in E(\Gamma)).

4. \mathbb{Z}-domains and one-relator hierarchies

Let XX be a CW-complex. A \mathbb{Z}-cover is a connected covering map p:YXp\colon Y\to X with Deck(p)\operatorname{Deck}(p)\cong\mathbb{Z}. We now define \mathbb{Z}-domains. These are subcomplexes of \mathbb{Z}-covers of XX that allow us to construct a homotopy equivalence between XX and a graph of spaces in certain situations. We will prove that \mathbb{Z}-domains always exist for finite CW-complexes and use this, in combination with the Freiheitssatz, to establish our hierarchy result for one-relator groups.

Definition 4.1.

Let p:YXp\colon Y\to X be a \mathbb{Z}-cover of CW-complexes and denote by tt\in\mathbb{Z} a generator. A \mathbb{Z}-domain for pp is a subcomplex DYD\subset Y with the following properties:

  1. (1)

    D=Y\mathbb{Z}\cdot D=Y.

  2. (2)

    DtiDDtDtiDD\cap t^{i}\cdot D\subset D\cap t\cdot D\cap\ldots\cap t^{i}\cdot D for all i>0i>0.

  3. (3)

    DtDD\cap t\cdot D is connected and non-empty.

We will denote by D(p)\operatorname{\mathbb{Z}D}(p) the set of all \mathbb{Z}-domains. A minimal \mathbb{Z}-domain is a \mathbb{Z}-domain, minimal under the partial order of inclusion.

Example 4.2.

Let SgS_{g} be the orientable surface of genus g1g\geq 1 and γ:S1Sg\gamma:S^{1}\to S_{g} a non-separating simple closed curve. We may give SgS_{g} a CW structure so that γ\gamma is in the one-skeleton. The curve γ\gamma determines an epimorphism π1(Sg)\pi_{1}(S_{g})\to\mathbb{Z} via the intersection form. Let p:YSgp:Y\to S_{g} be the induced cyclic cover. Consider the subspace ZYZ\subset Y obtained by taking the closure of some component of p1(SgIm(γ))p^{-1}(S_{g}-\operatorname{Im}(\gamma)). Its translates cover YY and intersect in lifts of γ\gamma and so ZZ is a (minimal) \mathbb{Z}-domain for pp.

Now let p:YXp:Y\to X be a \mathbb{Z}-cover, tDeck(p)t\in\mathbb{Z}\cong\operatorname{Deck}(p) a generator and DYD\subset Y a \mathbb{Z}-domain. Consider the following families of spaces:

{Di}i\{D_{i}\}_{i\in\mathbb{Z}}
{Ei}i\{E_{i}\}_{i\in\mathbb{Z}}

where DitiDD_{i}\cong t^{i}\cdot D and Eiti(DtD)E_{i}\cong t^{i}\cdot(D\cap t\cdot D). There are natural inclusion maps:

ιi:EiDi\iota_{i}^{-}\colon E_{i}\to D_{i}

and

ιi+:EiDi+1\iota_{i}^{+}\colon E_{i}\to D_{i+1}

given by the inclusions ti(DtD)tiDt^{i}\cdot(D\cap t\cdot D)\to t^{i}\cdot D and ti(DtD)ti+1Dt^{i}\cdot(D\cap t\cdot D)\to t^{i+1}\cdot D. With this data we define the space:

Θ(p,D)=(iDi)(iEi×[1,1])/\Theta(p,D)=\left(\bigsqcup_{i\in\mathbb{Z}}D_{i}\right)\sqcup\left(\bigsqcup_{i\in\mathbb{Z}}E_{i}\times[-1,1]\right)/\sim

where (x,±1)ιi±(x)(x,\pm 1)\sim\iota_{i}^{\pm}(x) for all ii\in\mathbb{Z} and xEix\in E_{i}. We also have an action of \mathbb{Z} on Θ(p,D)\Theta(p,D) induced by the action on {Di}i\{D_{i}\}_{i\in\mathbb{Z}}.

Proposition 4.3.

Let p:YXp\colon Y\to X be a \mathbb{Z}-cover with tDeck(p)t\in\operatorname{Deck}(p) a generator and let DYD\subset Y be a \mathbb{Z}-domain. If DtDD,tDD\cap t\cdot D\to D,t\cdot D are π1\pi_{1}-injective, then Θ(p,D)\Theta(p,D) and \Θ(p,D)\mathbb{Z}\backslash\Theta(p,D) are graphs of spaces and the induced maps

Θ(p,D)\displaystyle\Theta(p,D) Y\displaystyle\to Y
\Θ(p,D)\displaystyle\mathbb{Z}\backslash\Theta(p,D) X\displaystyle\to X

are homotopy equivalences factoring as horizontal maps composed with homeomorphisms. In particular, π1(X)\pi_{1}(X) splits as a HNN-extension with vertex group π1(D)\pi_{1}(D) and edge group π1(DtD)\pi_{1}(D\cap t\cdot D).

Proof.

Condition (3) of \mathbb{Z}-domains combined with π1\pi_{1}-injectivity of DtDD,tDD\cap t\cdot D\to D,t\cdot D implies that Θ(p,D)\Theta(p,D) and \Θ(p,D)\mathbb{Z}\backslash\Theta(p,D) are graphs of spaces. Condition (1) implies that the maps Θ(p,D)Y\Theta(p,D)\to Y and \Θ(p,D)X\mathbb{Z}\backslash\Theta(p,D)\to X are surjective and the definition of the horizontal map 𝗁\mathsf{h} shows that they factor through the horizontal maps. The preimages of points in YY under the map Θ(p,D)Y\Theta(p,D)\to Y are collections of intervals which can be identified with subintervals of \mathbb{R}, the underlying graph of the graph of spaces 𝒴\mathcal{Y} for Θ(p,D)=X𝒴\Theta(p,D)=X_{\mathcal{Y}}. The horizontal map Θ(p,D)=X𝒴H𝒴\Theta(p,D)=X_{\mathcal{Y}}\to H_{\mathcal{Y}} is the quotient map defined by identifying each connected component of these preimages to points. Condition (2) of \mathbb{Z}-domains implies these preimages are actually connected. Hence, the factored map H𝒴YH_{\mathcal{Y}}\to Y is actually a homeomorphism. Since \mathbb{Z} acts freely on YY, distinct points in a single point preimage in Θ(p,D)\Theta(p,D) lie in distinct \mathbb{Z}-orbits and so it follows from the commutativity of the diagram

Θ(p,D){{\Theta(p,D)}}\Θ(p,D){{\mathbb{Z}\backslash\Theta(p,D)}}Y{Y}X{X}

that preimages of points in XX under the map \Θ(p,D)X\mathbb{Z}\backslash\Theta(p,D)\to X are also connected subintervals of \mathbb{R}. As before, we obtain that the map \Θ(p,D)X\mathbb{Z}\backslash\Theta(p,D)\to X factors as the horizontal map composed with a homeomorphism. By Proposition 3.2, both horizontal maps are homotopy equivalences and so the induced maps are also homotopy equivalences. ∎

4.1. Existence of \mathbb{Z}-domains

Let us first fix some notation. Let p:YXp\colon Y\to X be a \mathbb{Z}-cover with tt a generator for Deck(p)\operatorname{Deck}(p). Choose some spanning tree TX(1)T\subset X^{(1)} and some orientation on X(1)X^{(1)}. This induces an identification of π1(X(1))\pi_{1}\left(X^{(1)}\right) with F(Σ)F(\Sigma), the free group generated by

Σ={e}eE(X(1))E(T).\Sigma=\{e\}_{e\in E(X^{(1)})\setminus E(T)}.

For any subset AΣA\subset\Sigma, define:

TA=T(eAe).T^{A}=T\cup\left(\bigcup_{e\in A}e\right).

Choose some lift of TT to YY and denote this by T0YT_{0}\subset Y. Then, since pp is regular, every lift of TT is obtained by translating T0T_{0} by an element of \mathbb{Z}. So we will denote by:

Ti=tiT0.T_{i}=t^{i}\cdot T_{0}.

For each eE(X(1))e\in E(X^{(1)}), denote by eie_{i} the lift of ee such that o(ei)Tio(e_{i})\in T_{i}. If eΣe\notin\Sigma, then t(ei)Tit(e_{i})\in T_{i}. If eΣe\in\Sigma, then t(ei)Ti+ι(e)t(e_{i})\in T_{i+\iota(e)} where

ι:Σπ1(X(1),x)\iota:\Sigma\to\pi_{1}(X^{(1)},x)\to\mathbb{Z}

is the obvious map.

If eΣe\in\Sigma and CC\subset\mathbb{Z}, define CeCC_{e}\subset C to be the subset consisting of the elements iCi\in C such that i+ι(e)Ci+\iota(e)\in C. Then if AΣA\subset\Sigma, we define the following subcomplex of the 1-skeleton of YY:

TCA=(iCTi)(eAiCeei)T_{C}^{A}=\left(\bigcup_{i\in C}T_{i}\right)\cup\left(\bigcup_{e\in A}\bigcup_{i\in C_{e}}e_{i}\right)

Examples of such subcomplexes can be seen in Figures 1 and 2. We also write

KA\displaystyle K_{A} =ι(A)<\displaystyle=\langle\iota(A)\rangle<\mathbb{Z}
kA\displaystyle k_{A} =[:KA].\displaystyle=[\mathbb{Z}:K_{A}].
Remark 4.4.

The subgroup KAK_{A} acts on each component of TAT^{A}_{\mathbb{Z}} and the quantity kAk_{A} is precisely the number of connected components of TAT_{\mathbb{Z}}^{A}. In fact, if we contract all the lifts of TT in Y(1)Y^{(1)}, then the resulting graph is the Cayley graph for \mathbb{Z} over the generating set ι(Σ)\iota(\Sigma). So the connected components of TAT_{\mathbb{Z}}^{A} correspond to KAK_{A} cosets in \mathbb{Z}.

We will call a subset CC\subset\mathbb{Z} connected if for all triples of integers ijki\leqslant j\leqslant k, if i,kCi,k\in C then jCj\in C.

Proposition 4.5.

Let p:YXp:Y\to X be a \mathbb{Z}-cover. For any subset AΣA\subset\Sigma and any connected subset CC\subset\mathbb{Z}, the following hold:

  1. (1)

    If kA=k_{A}=\infty, then TCAT_{C}^{A} consists of |C||C| connected components.

  2. (2)

    If kA<k_{A}<\infty, then there exists some constant k=k(A)0k=k(A)\geq 0 such that TCAT_{C}^{A} consists of kAk_{A} connected components whenever |C|k|C|\geq k.

Proof.

For (1), note that we have that KA={0}K_{A}=\{0\}, hence each edge eje_{j} with eAe\in A has both endpoints in TjT_{j}. So for each jCj\in C, the subcomplexes TjATCAT_{j}^{A}\subset T_{C}^{A} are all pairwise disjoint and cover TCAT_{C}^{A}.

Now assume that kA<k_{A}<\infty. Let AAA^{\prime}\subset A be any finite subset such that gcd(ι(A))=gcd(ι(A))=kA\gcd(\iota(A^{\prime}))=\gcd(\iota(A))=k_{A}. For any pair of connected subsets CDC\subset D\subset\mathbb{Z} of size larger than the largest element of ι(A)\iota(A^{\prime}) in absolute value, the inclusion TCATDAT_{C}^{A^{\prime}}\hookrightarrow T_{D}^{A^{\prime}} is surjective on components. Hence, for large enough CC, the number of components must stabilise at kAk_{A} by Remark 4.4. By definition of TCAT_{C}^{A}, we also have that TCAT_{C}^{A} has kAk_{A} connected components for all connected CC\subset\mathbb{Z} such that |C|k=k(A)|C|\geqslant k=k(A^{\prime}). This establishes (2). ∎

Proposition 4.6.

Every \mathbb{Z}-cover of a finite CW-complex has a finite \mathbb{Z}-domain.

Proof.

Let XX be a finite CW-complex and p:YXp:Y\to X a \mathbb{Z}-cover. We prove this by induction on nn-skeleta of XX. Denote by pn:Y(n)X(n)p_{n}:Y^{(n)}\to X^{(n)} the restriction of pp to the nn-skeleton of YY. Since XX is finite, |ι(Σ)||\iota(\Sigma)| is finite. As kΣ=1k_{\Sigma}=1, we may apply Proposition 4.5 and obtain an integer k(Σ)0k(\Sigma)\geq 0 such that TCΣT_{C}^{\Sigma} is connected for all |C|k|C|\geq k. So if we choose CC so that |C||C| is greater than kk and greater than the maximal size of an element in ι(Σ)\iota(\Sigma) in absolute value, then TCΣT_{C}^{\Sigma} satisfies conditions (1) and (2) of the definition of \mathbb{Z}-domains. By possibly enlarging CC by one element, we may ensure that (3) is also satisfied and hence that TCΣT_{C}^{\Sigma} is a \mathbb{Z}-domain for p1:Y(1)X(1)p_{1}:Y^{(1)}\to X^{(1)}.

Now suppose we have a \mathbb{Z}-domain Dn1D_{n-1} for pn1:Y(n1)X(n1)p_{n-1}:Y^{(n-1)}\to X^{(n-1)}. Then for any connected subset CC\subset\mathbb{Z}, the subcomplex jCtjDn1\bigcup_{j\in C}t^{j}\cdot D_{n-1} is also a \mathbb{Z}-domain for pn1p_{n-1}. The first and last properties are clear, whereas the second property holds because

(jCtjDn1)ti(jCtjDn1)=m+ijMtjDn1\left(\bigcup_{j\in C}t^{j}\cdot D_{n-1}\right)\cap t^{i}\cdot\left(\bigcup_{j\in C}t^{j}\cdot D_{n-1}\right)=\bigcup_{m+i\leqslant j\leqslant M}t^{j}\cdot D_{n-1}

if i|C|i\leqslant|C|, where m=min{C}m=\min{\{C\}} and M=min{C}M=\min{\{C\}}, and

(jCtjDn1)ti(jCtjDn1)\displaystyle\left(\bigcup_{j\in C}t^{j}\cdot D_{n-1}\right)\cap t^{i}\cdot\left(\bigcup_{j\in C}t^{j}\cdot D_{n-1}\right) =tMDn1tiDn1\displaystyle=t^{M}\cdot D_{n-1}\cap t^{i}\cdot D_{n-1}
tMDn1Dn1M+1tiDn1\displaystyle\subseteq t^{M}\cdot D_{n-1}\cap D_{n-1}^{M+1}\cap\ldots\cap t^{i}\cdot D_{n-1}

otherwise, where the last inclusion uses the fact that Dn1D_{n-1} was a \mathbb{Z}-domain for pn1p_{n-1}. We may choose CC large enough so that jCtjDn1\bigcup_{j\in C}t^{j}\cdot D_{n-1} contains a lift of each attaching map of nn-cells in XX. Then the full subcomplex of Y(n)Y^{(n)} containing jCtjDn1\bigcup_{j\in C}t^{j}\cdot D_{n-1} is a \mathbb{Z}-domain for pnp_{n}. ∎

4.2. Graphs and one-relator complexes

If we restrict our attention to graphs, the topology of minimal \mathbb{Z}-domains is much simpler.

Lemma 4.7.

Let Γ\Gamma be a finite graph and let p:YΓp:Y\to\Gamma be a \mathbb{Z}-cover. If DD(p)D\in\operatorname{\mathbb{Z}D}(p) is a minimal \mathbb{Z}-domain, then

χ(D)\displaystyle\chi(D) =χ(Γ)+1,\displaystyle=\chi(\Gamma)+1,
χ(DtD)\displaystyle\chi(D\cap t\cdot D) =1.\displaystyle=1.
Proof.

Up to contracting a spanning tree, we may assume that Γ\Gamma has a single vertex and nn edges. Let DD(p)D\in\operatorname{\mathbb{Z}D}(p) be a minimal \mathbb{Z}-domain. We have χ(DtD)=χ(D)+n1\chi(D\cap t\cdot D)=\chi(D)+n-1 since DtDD\cap t\cdot D is obtained from DD by removing the lowest vertex and each of the nn lowest edges. We claim that DtDD\cap t\cdot D is a tree. If not, then there would be an embedded cycle S1DtDS^{1}\hookrightarrow D\cap t\cdot D. By possibly replacing this cycle with a translate, we may assume that it traverses an edge eDe\subset D such that tet\cdot e does not lie in DD. Then removing this edge will leave us with a smaller \mathbb{Z}-domain, contradicting minimality. Thus, we must have that DtDD\cap t\cdot D is actually a tree, otherwise we could remove some edge. Thus χ(D)=χ(Γ)+1\chi(D)=\chi(\Gamma)+1 and χ(DtD)=1\chi(D\cap t\cdot D)=1. ∎

As a consequence of Lemma 4.7 and Proposition 4.3, we can see that minimal \mathbb{Z}-domains of \mathbb{Z}-covers of graphs Γ\Gamma correspond to certain free product decompositions of π1(Γ)\pi_{1}(\Gamma).

Example 4.8.

Let Γ\Gamma be a graph with a single vertex and two edges, so χ(Γ)=1\chi(\Gamma)=-1. The spanning tree is the unique vertex vΓv\in\Gamma. Let Σ={a,b}\Sigma=\{a,b\} be the two edges and let ϕ:F(a,b)\phi:F(a,b)\to\mathbb{Z} be the homomorphism that sends aa to 55 and bb to 33. Then let p:YΓp:Y\to\Gamma be the corresponding \mathbb{Z}-cover. Let C={0,1,,6}C=\{0,1,...,6\}, then TCΣT_{C}^{\Sigma} can be seen in Figure 1. The numbers ii in the diagram correspond to TiT_{i}, the lower black edges are lifts of aa and the upper red edges are lifts of bb edges. This is not quite a \mathbb{Z}-domain as its intersection with a translate is disconnected. However, we can see that TC7ΣT_{C\cup 7}^{\Sigma} is a genuine \mathbb{Z}-domain, see Figure 2. In fact, it is the unique minimal \mathbb{Z}-domain for pp. The unique primitive cycle that factors through TC7ΣT_{C\cup 7}^{\Sigma}, represents the unique conjugacy class of primitive element in ker(ϕ)\ker(\phi).

Refer to caption
Figure 1. Not quite a \mathbb{Z}-domain.
Refer to caption
Figure 2. A minimal \mathbb{Z}-domain.

Another special case is that of one-relator complexes. The following two propositions are the foundations for our one-relator hierarchies.

Proposition 4.9.

Let XX be a finite one-relator complex, let p:YXp\colon Y\to X be a \mathbb{Z}-cover and let ZD(p)Z\subset\operatorname{\mathbb{Z}D}(p) be a minimal \mathbb{Z}-domain. Then ZZ is a finite one-relator complex, such that either ZtZZ\cap t\cdot Z is a tree, or t1ZZtZt^{-1}\cdot Z\cap Z\cap t\cdot Z is connected.

Proof.

Let X=(Γ,λ)X=(\Gamma,\lambda) be a finite one-relator complex and p:YXp:Y\to X a \mathbb{Z}-cover. Let ZYZ\subset Y be a minimal \mathbb{Z}-domain for pp. There is a minimal \mathbb{Z}-domain DY(1)D\subset Y^{(1)} for the \mathbb{Z}-cover Y(1)X(1)Y^{(1)}\to X^{(1)} contained in ZZ. By Lemma 4.7, DtDD\cap t\cdot D is a tree. Just as in the proof of Proposition 4.6, we have that iCtiD\bigcup_{i\in C}t^{i}\cdot D is also a \mathbb{Z}-domain for all connected subsets CC\subset\mathbb{Z}. Now let C={0,,m}C=\{0,\ldots,m\} be of smallest size such that a lift of λ\lambda is supported in iCtiD\bigcup_{i\in C}t^{i}\cdot D. Since DtDD\cap t\cdot D is a tree, at most one lift λ~\tilde{\lambda} can be supported in this subgraph. Now the union of this subgraph with the two-cell attached along λ~\tilde{\lambda} is a \mathbb{Z}-domain which must contain a translate of ZZ by minimality. Thus, ZZ is a one-relator complex. By construction, either C={0}C=\{0\} and so ZtZ=DtDZ\cap t\cdot Z=D\cap t\cdot D and so is a tree, or |C|>1|C|>1 and so ZtZZ\cap t\cdot Z is a \mathbb{Z}-domain for Y(1)X(1)Y^{(1)}\to X^{(1)} and so t1ZZtZt^{-1}\cdot Z\cap Z\cap t\cdot Z is connected. ∎

Remark 4.10.

More generally, by [How87, Lemma 2], if XX is a staggered two-complex and p:YXp:Y\to X a cyclic cover, then every \mathbb{Z}-domain DD(p)D\in\operatorname{\mathbb{Z}D}(p) is also a staggered two-complex. Furthermore, by [HW01, Corollary 6.2] and Proposition 4.3, each DD(p)D\in\operatorname{\mathbb{Z}D}(p) induces a HNN-splitting of π1(X)\pi_{1}(X) over π1(D)\pi_{1}(D).

4.3. One-relator hierarchies

Define the complexity of a one-relator complex X=(Γ,λ)X=(\Gamma,\lambda) to be the following quantity:

c(X)=(|λ|deg(λ)|Xλ(0)|,χ(X)).c(X)=\left(\frac{\left\lvert\lambda\right\rvert}{\deg(\lambda)}-\left\lvert X_{\lambda}^{(0)}\right\rvert,-\chi(X)\right).

We endow the complexity of one-relator complexes with the dictionary order so that (q,r)<(s,t)(q,r)<(s,t) if q<sq<s or q=sq=s and r<tr<t. Note that the first component of c(X)c(X) is a slight modification of the notion of complexity used by Howie in [How81].

Proposition 4.11.

Let XX be a finite one-relator complex and p:YXp:Y\to X a cyclic cover. If DD(p)D\in\operatorname{\mathbb{Z}D}(p) is a minimal \mathbb{Z}-domain, then:

c(D)<c(X).c(D)<c(X).
Proof.

It is clear that if c(D)=(q,r)c(D)=(q,r) and c(X)=(s,t)c(X)=(s,t), we cannot have q>sq>s. Suppose that we have q=sq=s, then we must have that XλX_{\lambda} actually lifts to all \mathbb{Z}-domains. But then this implies that XλX_{\lambda} is a subcomplex of DD. So applying Lemma 4.7 to the induced \mathbb{Z}-cover of the graph X/XλX/X_{\lambda}, we see that χ(D)=χ(X)+1\chi(D)=\chi(X)+1 when DD is minimal. Thus, r<tr<t and c(D)<c(X)c(D)<c(X) when DD is minimal. ∎

By Proposition 4.9, if XX is a one-relator complex and p:YXp:Y\to X is a cyclic cover, there exists a one-relator complex X1D(p)X_{1}\in\operatorname{\mathbb{Z}D}(p). Repeating this, we obtain a sequence of immersions of one-relator complexes XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X where XiX_{i} is a one-relator \mathbb{Z}-domain of a \mathbb{Z}-cover of Xi1X_{i-1} for each ii. We will call this a one-relator tower. Note that each immersion is a tower map in the sense of [How81]. If XNX_{N} does not admit any \mathbb{Z}-covers, we will call this a maximal one-relator tower.

Proposition 4.12.

Every finite one-relator complex X=(Γ,λ)X=(\Gamma,\lambda) has a maximal one-relator tower XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X.

Proof.

The proof is by induction on c(X)c(X), the base case is c(X)=(0,0)c(X)=(0,0). We have c(X)=(0,0)c(X)=(0,0) precisely when ΓS1\Gamma\simeq S^{1}. This is the only case where XX does not admit any \mathbb{Z}-cover. This is because if c(X)>(0,0)c(X)>(0,0), then χ(X)0\chi(X)\leq 0 and so rk(H1(X,))1\operatorname{rk}(H_{1}(X,\mathbb{Z}))\geq 1. Hence, the base case holds. The inductive step is simply Proposition 4.11. ∎

Now we are ready to prove our version of the Magnus–Moldavanskii hierarchy for one-relator groups.

Theorem 4.13.

Let XX be a finite one-relator complex. There exists a finite sequence of immersions of one-relator complexes:

XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X

such that π1(Xi)π1(Xi+1)ψi\pi_{1}(X_{i})\cong\pi_{1}(X_{i+1})*_{\psi_{i}} where ψi\psi_{i} is induced by an identification of Magnus subgraphs, and such that π1(XN)\pi_{1}(X_{N}) is finite cyclic.

Proof.

By Proposition 4.12 there is a maximal one-relator tower XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X. By Theorem 2.6, the inclusions XitXiXi,tXiX_{i}\cap t\cdot X_{i}\hookrightarrow X_{i},t\cdot X_{i} are all injective on π1\pi_{1}. Thus, by Proposition 4.3, π1(Xi)π1(Xi+1)ψi\pi_{1}(X_{i})\cong\pi_{1}(X_{i+1})*_{\psi_{i}} for some isomorphisms ψi\psi_{i} induced by an identification of Magnus subgraphs. ∎

We will call a splitting π1(Xi)π1(Xi+1)ψ\pi_{1}(X_{i})\cong\pi_{1}(X_{i+1})*_{\psi} as in Theorem 4.13 a one-relator splitting. Then the associated Magnus subgraphs inducing ψ\psi are A=t1Xi+1Xi+1A=t^{-1}\cdot X_{i+1}\cap X_{i+1} and B=Xi+1tXi+1B=X_{i+1}\cap t\cdot X_{i+1}.

Remark 4.14.

By Proposition 4.9, each π1(Xi+1)\pi_{1}(X_{i+1}) from Theorem 4.13 has a one-relator presentation so that ψi\psi_{i} actually identifies two Magnus subgroups for this presentation.

We will call a one-relator tower XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X a hierarchy of length NN if π1(XN)\pi_{1}(X_{N}) splits as a free product of cyclic groups. Denote by h(X)h(X) the hierarchy length of XX, that is, the smallest integer NN such that a hierarchy for XX of length NN exists. By Theorem 4.13 this is well-defined for all one-relator complexes. We extend this definition also to one-relator presentations Σw\langle\Sigma\mid w\rangle by saying that the hierarchy length of a one-relator presentation is the hierarchy length of its presentation complex. A hierarchy XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X is of minimal length if N=h(X)N=h(X).

The hierarchy length of a one-relator complex is not preserved under homotopy equivalence. This is illustrated by the following examples.

Example 4.15.

Let ±1p,q\pm 1\neq p,q\in\mathbb{Z} be a pair of coprime integers. Denote by prp,q(x,y)\operatorname{pr}_{p,q}(x,y) any cyclically reduced element in the unique conjugacy class of primitive elements in ab1(p,q)\operatorname{ab}^{-1}(p,q) where ab:F(x,y)2\operatorname{ab}:F(x,y)\to\mathbb{Z}^{2} is the abelianisation map. Then the Baumslag–Solitar group BS(p,q)\operatorname{BS}(p,q) has two one-relator presentations:

BS(p,q)a,t|t1apt=aqb,s|prp,q(b,s1bs)\operatorname{BS}(p,q)\cong\langle a,t|t^{-1}a^{p}t=a^{q}\rangle\cong\langle b,s|\operatorname{pr}_{p,-q}(b,s^{-1}bs)\rangle

such that the hierarchy length of the first presentation is 2, but of the second is 1. Since presentation complexes of one-relator groups without torsion are aspherical by Lyndon’s identity theorem[Lyn50], the two associated presentation complexes are homotopy equivalent.

Note that the words t1aptaqt^{-1}a^{p}ta^{-q} and prp,q(a,t1at)\operatorname{pr}_{p,-q}(a,t^{-1}at) are not equivalent under the action of Aut(F(a,t))\operatorname{Aut}(F(a,t)) if |p|,|q|2\left\lvert p\right\rvert,\left\lvert q\right\rvert\geq 2. Similarly for t1aptaqt^{-1}a^{p}ta^{-q} and prp,q(a,t1at)1\operatorname{pr}_{p,-q}(a,t^{-1}at)^{-1}.

We may do the same with the Baumslag–Gersten groups:

BG(p,q)a,t|t1atapt1a1t=aqb,s|prp,q(b,s1bsbs1b1s)BG(p,q)\cong\langle a,t|t^{-1}ata^{p}t^{-1}a^{-1}t=a^{q}\rangle\cong\langle b,s|\operatorname{pr}_{p,-q}(b,s^{-1}bsbs^{-1}b^{-1}s)\rangle

The hierarchy length of the first presentation is three, but of the second is two.

5. Normal forms and quasi-convex one-relator hierarchies

5.1. Normal forms

Computable normal forms for one-relator complexes can be derived from the original Magnus hierarchy [Mag30]. Following the same idea, we build normal forms for universal covers of one-relator complexes. The geometry of these normal forms will then allow us to prove our main theorem.

Let XX be a combinatorial 22-complex. We define I(X)I(X) to be the set of all combinatorial immersed paths. If cI(X)c\in I(X), then we denote by c¯\bar{c} the reverse path. If a,bI(X)a,b\in I(X) are paths with t(a)=o(b)t(a)=o(b), we will write aba*b for their concatenation.

Definition 5.1.

A normal form for XX is a map η:X(0)×X(0)I(X)\eta:X^{(0)}\times X^{(0)}\to I(X) where o(η(p,q))=po(\eta(p,q))=p and t(η(p,q))=qt(\eta(p,q))=q and η(p,p)\eta(p,p) is the constant path, for all p,qX(0)p,q\in X^{(0)}. We say that η\eta is prefix-closed if for every p,qX(0)p,q\in X^{(0)} and i[0,|η(p,q)|]i\in[0,\left\lvert\eta(p,q)\right\rvert], we have that η(p,r)=η(p,q)[0,i]\eta(p,r)=\eta(p,q)\mid_{[0,i]} where r=(η(p,q))(i)r=(\eta(p,q))(i).

We will be particularly interested in certain kinds of normal forms, defined below, the first being quasi-geodesic normal forms and the second being normal forms relative to a given subcomplex. The motivation for these definitions is the following: if XX is a two-complex, ZXZ\subset X a π1\pi_{1}-injective subcomplex such that the universal cover X~\tilde{X} admits quasi-geodesic normal forms relative to a copy of the universal cover Z~X~\tilde{Z}\subset\tilde{X}, then π1(Z)\pi_{1}(Z) will be undistorted in π1(X)\pi_{1}(X).

Definition 5.2.

Let K>0K>0. A normal form η:X(0)×X(0)I(X)\eta:X^{(0)}\times X^{(0)}\to I(X) is KK-quasi-geodesic if every path in Im(η)\operatorname{Im}(\eta) is a KK-quasi-geodesic. We will call η\eta quasi-geodesic if it is KK-quasi-geodesic for some K>0K>0.

Definition 5.3.

Suppose ZXZ\subset X is a subcomplex. A normal form η:X(0)×X(0)I(X)\eta:X^{(0)}\times X^{(0)}\to I(X) is a normal form relative to ZZ if, for any p,qp,q contained in the same connected component of ZZ and rXr\in X, the following hold:

  1. (1)

    η(p,q)\eta(p,q) is supported in ZZ,

  2. (2)

    η(p,r)η¯(q,r)\eta(p,r)*\bar{\eta}(q,r) is supported in ZZ after removing backtracking.

By definition, any normal form for XX is a normal form relative to X(0)X^{(0)}.

The general strategy for building our normal forms for universal covers of one-relator complexes will be to use one-relator hierarchies and induction. In order to do so, we must first discuss normal forms for graphs of spaces.

5.2. Graph of spaces normal forms

For simplicity, we only discuss normal forms for graphs of spaces with underlying graph consisting of a single vertex and single edge. However, the construction is the same for any graph.

Let 𝒳=(Γ,{X},{C},±)\mathcal{X}=(\Gamma,\{X\},\{C\},\partial^{\pm}) be a graph of spaces where Γ\Gamma consists of a single vertex and a single edge, XX and CC are combinatorial two-complexes and ±\partial^{\pm} are inclusions of subcomplexes. Denote by A=Im()A=\operatorname{Im}(\partial^{-}) and B=Im(+)B=\operatorname{Im}(\partial^{+}). We may orient the vertical edges so that they go from the subcomplex AA to the subcomplex BB. When it is clear from context which path we mean, we will abuse notation and write tkt^{k} for all k0k\geq 0, to denote a path that follows kk vertical edges consecutively, respecting this orientation. We will write tkt^{-k} for all k0k\geq 0, for the path that follows kk vertical edges in the opposite direction. Since ±\partial^{\pm} are embeddings, these paths are uniquely defined given an initial vertex.

The universal cover X~𝒳X𝒳\tilde{X}_{\mathcal{X}}\to X_{\mathcal{X}} also has a graph of spaces structure with underlying graph the Bass–Serre tree of the associated splitting, the vertex spaces are copies of the universal cover X~\tilde{X}, the edge spaces are copies of the universal cover C~\tilde{C} and the edge maps are π1(X)\pi_{1}(X)-translates of lifts ~±:C~X~\tilde{\partial}^{\pm}:\tilde{C}\to\tilde{X} of the maps ±\partial^{\pm}. We will denote by 𝒳~\tilde{\mathcal{X}} the underlying graph of spaces so that X𝒳~=X~𝒳X_{\tilde{\mathcal{X}}}=\tilde{X}_{\mathcal{X}}. Every path cI(X𝒳~)c\in I(X_{\tilde{\mathcal{X}}}) can be uniquely factorised:

c=c0tϵ1c1tϵncnc=c_{0}*t^{\epsilon_{1}}*c_{1}*...*t^{\epsilon_{n}}*c_{n}

with ϵi=±1\epsilon_{i}=\pm 1 and where each cic_{i} is (a possibly empty path) supported in some copy of X~\tilde{X}. We say cc is reduced if there are no two subpaths ci,cjc_{i},c_{j} that are both supported in the same copy of X~\tilde{X}. Hence that no subpath citϵi+1tϵjcjc_{i}*t^{\epsilon_{i+1}}*...*t^{\epsilon_{j}}*c_{j} has both endpoints in the same copy of X~\tilde{X}. In other words, the path vcv\circ c is immersed, where v:X~𝒳Tv:\tilde{X}_{\mathcal{X}}\to T is the vertical map to the Bass–Serre tree TT.

A vertical square in X~𝒳\tilde{X}_{\mathcal{X}} is a two-cell with boundary path etϵftϵe*t^{\epsilon}*f*t^{-\epsilon} where ee and ff are edges in two different copies of X~\tilde{X}. We will call the paths etϵe*t^{\epsilon}, tϵft^{\epsilon}*f, ftϵf*t^{-\epsilon}, tϵet^{-\epsilon}*e and their inverses corners of this square. We may pair up two corners α\alpha and β\beta if their concatenation αβ¯\alpha*\bar{\beta} forms the boundary of a square. In this way, each corner exhibits a vertical homotopy to the opposing corner it is paired up with. For instance, there is a vertical homotopy through the square between tϵf¯t^{\epsilon}*\bar{f} and etϵe*t^{\epsilon}.

Now let

ηA~:X~(0)×X~(0)I(X~)\eta_{\tilde{A}}:\tilde{X}^{(0)}\times\tilde{X}^{(0)}\to I(\tilde{X})
ηB~:X~(0)×X~(0)I(X~)\eta_{\tilde{B}}:\tilde{X}^{(0)}\times\tilde{X}^{(0)}\to I(\tilde{X})

be π1(X)\pi_{1}(X)-equivariant normal forms relative to A~\tilde{A} and B~\tilde{B} respectively. Since these are π1(X)\pi_{1}(X)-equivariant, it is not important which lift of A~\tilde{A} and B~\tilde{B} we choose. Let

ηX~:X~(0)×X~(0)I(X~)\eta_{\tilde{X}}:\tilde{X}^{(0)}\times\tilde{X}^{(0)}\to I(\tilde{X})

be a π1(X)\pi_{1}(X)-equivariant normal form for X~\tilde{X}. From this data, we may define normal forms for X𝒳~X_{\tilde{\mathcal{X}}} as follows. We say a normal form

η:X𝒳~(0)×X𝒳~(0)I(X𝒳~)\eta:X_{\tilde{\mathcal{X}}}^{(0)}\times X_{\tilde{\mathcal{X}}}^{(0)}\to I(X_{\tilde{\mathcal{X}}})

is a graph of spaces normal form induced by ({ηX~},{ηA~,ηB~})(\{\eta_{\tilde{X}}\},\{\eta_{\tilde{A}},\eta_{\tilde{B}}\}) if the following holds for all c=c0tϵ1c1tϵncnIm(η)c=c_{0}*t^{\epsilon_{1}}*c_{1}*...*t^{\epsilon_{n}}*c_{n}\in\operatorname{Im}(\eta):

  1. (1)

    c0=ηX~(o(c0),t(c0))c_{0}=\eta_{\tilde{X}}(o(c_{0}),t(c_{0})).

  2. (2)

    If ϵi=1\epsilon_{i}=1 then ci=ηB~(o(ci),t(ci))c_{i}=\eta_{\tilde{B}}(o(c_{i}),t(c_{i})). Furthermore, if ee is the first edge cic_{i} traverses, then tet*e does not form a corner of a vertical square.

  3. (3)

    If ϵi=1\epsilon_{i}=-1 then ci=ηA~(o(ci),t(ci))c_{i}=\eta_{\tilde{A}}(o(c_{i}),t(c_{i})). Furthermore, if ee is the first edge cic_{i} traverses, then t1et^{-1}*e does not form a corner of a vertical square.

Our definition of graph of spaces normal form and the following theorem are the topological translations of the algebraic definition and the normal form theorem [LS01, p. 181].

Theorem 5.4.

If ηA~,ηB~\eta_{\tilde{A}},\eta_{\tilde{B}} and ηX~\eta_{\tilde{X}} are as above, there is a unique graph of spaces normal form η\eta induced by ({ηX~},{ηA~,ηB~})(\{\eta_{\tilde{X}}\},\{\eta_{\tilde{A}},\eta_{\tilde{B}}\}). Moreover, the following are satisfied:

  1. (1)

    η\eta is π1(X𝒳)\pi_{1}(X_{\mathcal{X}})-equivariant,

  2. (2)

    the action of π1(X)\pi_{1}(X) on η\eta and ηX~\eta_{\tilde{X}} coincide,

  3. (3)

    if ηX~\eta_{\tilde{X}}, ηA~\eta_{\tilde{A}} and ηB~\eta_{\tilde{B}} are prefix-closed, then so is η\eta.

Before moving on, we shall provide a brief sketch of the proof of Theorem 5.4.

Given existence and uniqueness of graph of spaces normal forms, the fact that they must be prefix closed provided ηX~,ηA~\eta_{\tilde{X}},\eta_{\tilde{A}} and ηB~\eta_{\tilde{B}} are can be shown directly from the definition of graph of spaces normal forms. Indeed, any prefix of a path satisfying the three conditions, will also satisfy the three conditions. Similarly for the equivariance claims.

The existence of normal forms can be shown by choosing paths c=c0tϵ1c1tϵncnc=c_{0}*t^{\epsilon_{1}}*c_{1}*\ldots*t^{\epsilon_{n}}*c_{n} between each pair of points and putting them into normal form as follows. If n=0n=0, then η(o(c),t(c))=ηX~(o(c),t(c))\eta(o(c),t(c))=\eta_{\tilde{X}}(o(c),t(c)). If n1n\geqslant 1 and ϵn=1\epsilon_{n}=1, then replace cnc_{n} with ηB~(o(cn),t(cn))=bncn\eta_{\tilde{B}}(o(c_{n}),t(c_{n}))=b_{n}*c_{n}^{\prime}, where bnb_{n} is the largest prefix contained in the same copy of B~\tilde{B} that o(cn)o(c_{n}) was in. Then perform a sequence of vertical homotopies through vertical squares, replacing tϵnbnt^{\epsilon_{n}}*b_{n} with antϵna_{n}*t^{\epsilon_{n}}, where ana_{n} lies in the same copy of A~\tilde{A} that o(tϵn)o(t^{\epsilon_{n}}) was in. Then η(o(c),t(c))\eta(o(c),t(c)) can be defined inductively on nn by setting η(o(c),t(c))=η(o(c),t(c))cn\eta(o(c),t(c))=\eta(o(c^{\prime}),t(c^{\prime}))*c_{n}^{\prime}, where c=c0tϵ1c1tϵn1cn1anc^{\prime}=c_{0}*t^{\epsilon_{1}}*c_{1}*\ldots*t^{\epsilon_{n-1}}*c_{n-1}*a_{n}, after possibly removing backtracking. The case where ϵn=1\epsilon_{n}=-1 is handled similarly. The fact that these are graph of spaces normal forms is by construction. This procedure will be important for the proof of Theorem 5.7.

Now suppose that c=c0tϵ1tϵncnc=c_{0}*t^{\epsilon_{1}}*\ldots*t^{\epsilon_{n}}*c_{n} and c=c0tη1tηmcmc^{\prime}=c_{0}^{\prime}*t^{\eta_{1}}*\ldots*t^{\eta_{m}}*c^{\prime}_{m} are two paths in normal form with the same origin and target vertices. Since the underlying graph of the graph of spaces 𝒳~\tilde{\mathcal{X}} is a tree, the Bass–Serre tree of the graph of groups induced by 𝒳\mathcal{X}, it follows that n=mn=m and ϵi=ηi\epsilon_{i}=\eta_{i} for all ii. Moreover, it also follows that cnc¯nc_{n}*\bar{c}_{n}^{\prime} must be equal if n=0n=0 or path homotopic into a copy of B~\tilde{B} if ϵn=1\epsilon_{n}=1 or A~\tilde{A} if ϵn=1\epsilon_{n}=-1. In the second case, since cn=ηB~(o(cn),t(cn))c_{n}=\eta_{\tilde{B}}(o(c_{n}),t(c_{n})) and cn=ηB~(o(cn),t(cn))c_{n}^{\prime}=\eta_{\tilde{B}}(o(c_{n}^{\prime}),t(c_{n}^{\prime})), the definition of relative normal forms tells us that cnc¯nc_{n}*\bar{c}_{n}^{\prime} lies in B~\tilde{B} after removing backtracking. Hence cn=cnc_{n}=c_{n}^{\prime} otherwise the first edge of cnc_{n} or cnc_{n}^{\prime} would form the corner of a vertical square together with tϵnt^{\epsilon_{n}}. The first case is handled similarly and so cc and cc^{\prime} are equal as paths.

5.3. Quasi-geodesic normal forms for graphs of hyperbolic spaces

The definition of quasi-geodesics may be generalised in the following way. Let XX be a metric space and f:00f:\mathbb{R}_{\geq 0}\to\mathbb{R}_{\geq 0} be a monotonic increasing function. A path c:IXc:I\to X, parametrised by arc length, is an ff-quasi-geodesic if, for all 0pq|c|0\leq p\leq q\leq\left\lvert c\right\rvert, the following is satisfied:

qpf(d(c(p),c(q)))d(c(p),c(q))+f(d(c(p),c(q)))q-p\leq f(d(c(p),c(q)))\cdot d(c(p),c(q))+f(d(c(p),c(q)))

If ff is bounded above by a constant KK, then cc is simply a KK-quasi-geodesic.

Theorem 5.5.

Let XX be a geodesic hyperbolic metric space and f:00f:\mathbb{R}_{\geq 0}\to\mathbb{R}_{\geq 0} a subexponential function. There is a constant K(f)K(f) such that all ff-quasi-geodesics are K(f)K(f)-quasi-geodesics.

Proof.

Follows from [Gro87, Corollary 7.1.B]. Alternatively, a slight modification of the proof of the Morse lemma, see for example [ABC+91, Proposition 3.3], replacing quasi-geodesics with ff-quasi-geodesics, yields that there is a constant K(f)K^{\prime}(f) such that ff-quasi-geodesics and geodesics lie in the KK^{\prime}-neighbourhoods of each other. So now let γ\gamma be a geodesic and let cc be an ff-quasi-geodesic with the same endpoints. For each positive integer i|γ|i\leqslant|\gamma|, there is some ji|c|j_{i}\leqslant|c| such that d(γ(i),c(ji))Kd(\gamma(i),c(j_{i}))\leqslant K^{\prime} and so d(c(ji),c(ji+1))2K+1d(c(j_{i}),c(j_{i+1}))\leqslant 2K^{\prime}+1. Hence, we have ji+1jif(2K+1)(2K+1)+f(2K+1)j_{i+1}-j_{i}\leqslant f(2K^{\prime}+1)\cdot(2K^{\prime}+1)+f(2K^{\prime}+1) for all i|γ|i\leqslant|\gamma|. Thus, setting K(f)=f(2K+1)(2K+2)K(f)=f(2K^{\prime}+1)\cdot(2K^{\prime}+2) yields that |c|K(f)d(o(c),t(c))|c|\leqslant K(f)\cdot d(o(c),t(c)) and thus, the result. ∎

It follows from Theorem 5.5 that there is a gap in the spectrum of possible distortion functions of subgroups of hyperbolic groups. This is known by work of Gromov [Gro87] and appears as [Kap01, Proposition 2.1].

Corollary 5.6.

Subexponentially distorted subgroups of hyperbolic groups are undistorted and hence, quasi-convex.

The following result can be thought of as a strengthening of [Kap01, Corollary 1.1]. We make use of Theorem 5.5.

Theorem 5.7.

Let 𝒳=(Γ,{X},{C},{±})\mathcal{X}=(\Gamma,\{X\},\{C\},\{\partial^{\pm}\}) be a graph of spaces with Γ\Gamma consisting of a single vertex and edge. Let ηA~:X~(0)×X~(0)I(X~)\eta_{\tilde{A}}:\tilde{X}^{(0)}\times\tilde{X}^{(0)}\to I(\tilde{X}), ηB~:X~(0)×X~(0)I(X~)\eta_{\tilde{B}}:\tilde{X}^{(0)}\times\tilde{X}^{(0)}\to I(\tilde{X}) and ηX~:X~(0)×X~(0)I(X~)\eta_{\tilde{X}}:\tilde{X}^{(0)}\times\tilde{X}^{(0)}\to I(\tilde{X}) be π1(X)\pi_{1}(X)-equivariant normal forms, with ηA~\eta_{\tilde{A}} relative to A~\tilde{A} and ηB~\eta_{\tilde{B}} relative to B~\tilde{B}. Suppose that the following are satisfied:

  1. (1)

    X~\tilde{X} is hyperbolic,

  2. (2)

    π1(X𝒳)\pi_{1}(X_{\mathcal{X}}) acts acylindrically on the Bass-Serre tree TT,

  3. (3)

    ηA~\eta_{\tilde{A}}, ηB~\eta_{\tilde{B}} and ηX~\eta_{\tilde{X}} are prefix-closed quasi-geodesic normal forms.

Then the graph of spaces normal form induced by ({ηX~},{ηA~,ηB~})\left(\{\eta_{\tilde{X}}\},\{\eta_{\tilde{A}},\eta_{\tilde{B}}\}\right):

η:X~𝒳(0)×X~𝒳(0)I(X~𝒳)\eta:\tilde{X}^{(0)}_{\mathcal{X}}\times\tilde{X}^{(0)}_{\mathcal{X}}\to I(\tilde{X}_{\mathcal{X}})

is prefix-closed and quasi-geodesic.

Under the same hypothesis, Kapovich showed that for each pair of points in X𝒳X_{\mathcal{X}}, there exists a reduced quasi-geodesic path connecting them. Theorem 5.7 shows that graph of spaces normal forms provide us with such paths. This fact will be key for the proof of our main theorems.

Proof of Theorem 5.7.

The fact that η\eta is prefix-closed is Theorem 5.4. By Theorem 3.4, we have that X~𝒳\tilde{X}_{\mathcal{X}} is δ\delta-hyperbolic and X~X𝒳~\tilde{X}\hookrightarrow X_{\tilde{\mathcal{X}}} is a quasi-isometric embedding. Thus, there is a constant K0K\geq 0 such that the images of ηA~\eta_{\tilde{A}}, ηB~\eta_{\tilde{B}} and ηX~\eta_{\tilde{X}} are KK-quasi-geodesic in X~𝒳\tilde{X}_{\mathcal{X}}. Now let x,yX~𝒳(0)x,y\in\tilde{X}_{\mathcal{X}}^{(0)} be any two points and γ\gamma a geodesic connecting them through the 1-skeleton. We may factorise this geodesic

γ=γ0tϵ1γ1tϵkγk\gamma=\gamma_{0}*t^{\epsilon_{1}}*\gamma_{1}*...*t^{\epsilon_{k}}*\gamma_{k}

such that ϵi=±1\epsilon_{i}=\pm 1, each γi\gamma_{i} is path homotopic into a copy of X~\tilde{X} and there is no subpath γitϵi+1γj\gamma_{i}*t^{\epsilon_{i+1}}*...*\gamma_{j} with i<ji<j that is path homotopic into some copy of X~\tilde{X}. If ρI(X~𝒳)\rho\in I(\tilde{X}_{\mathcal{X}}), we will write X~ρ\tilde{X}_{\rho} to denote the copy of X~\tilde{X} in X~𝒳\tilde{X}_{\mathcal{X}} that ρ\rho ends in. When it makes sense, we will do the same for A~\tilde{A} and B~\tilde{B}. This is well defined since all the π1(X𝒳)\pi_{1}(X_{\mathcal{X}}) translates of A~\tilde{A} are disjoint or equal and all the π1(X𝒳)\pi_{1}(X_{\mathcal{X}}) translates of B~\tilde{B} are disjoint or equal.

Let pX~γ(0)p\in\tilde{X}^{(0)}_{\gamma}, let α\alpha be a geodesic connecting yy with pp and let β\beta be a geodesic connecting o(γk)o(\gamma_{k}) with pp. Denoting by M=M(K,δ)M=M(K,\delta) the maximal Hausdorff distance of all ηX~\eta_{\tilde{X}}, ηA~\eta_{\tilde{A}} and ηB~\eta_{\tilde{B}} normal forms from their respective geodesics in X~𝒳\tilde{X}_{\mathcal{X}} and by L=δ+ML=\delta+M, we claim that

|η(x,p)|K(k+1)(kL+|γ|+|α|+k+1).\left\lvert\eta(x,p)\right\rvert\leq K(k+1)(kL+|\gamma|+|\alpha|+k+1).

The proof of the claim is by induction on kk. First suppose that k=0k=0. We will write \simeq to mean path homotopic. Since βγkα\beta\simeq\gamma_{k}*\alpha and o(β),t(β)X~γo(\beta),t(\beta)\in\tilde{X}_{\gamma}, we get that η(x,p)=ηX~(x,p)\eta(x,p)=\eta_{\tilde{X}}(x,p) is actually a KK-quasi-geodesic path. This establishes the base case.

Now suppose it is true for all i<ki<k. Suppose ϵk=1\epsilon_{k}=-1, the other case is the same. Let β=ηA~(o(γk),p)=ac\beta^{\prime}=\eta_{\tilde{A}}(o(\gamma_{k}),p)=a*c where aA~γγ¯ka\subset\tilde{A}_{\gamma*\bar{\gamma}_{k}} and the first edge of cc is not in A~γγ¯k\tilde{A}_{\gamma*\bar{\gamma}_{k}}. Since aa is supported in a copy of A~\tilde{A}, the path t1at^{-1}*a may be homotoped through vertical squares to a path bt1b*t^{-1}, where bb is supported in B~γγ¯kt\tilde{B}_{\gamma*\bar{\gamma}_{k}*t}. Let α\alpha^{\prime} be a geodesic connecting o(b)o(b) to t(b)t(b). Since ac=ηA~(o(a),t(c))a*c=\eta_{\tilde{A}}(o(a),t(c)), the definition of relative normal forms implies that c=ηA~(o(c),t(c))c=\eta_{\tilde{A}}(o(c),t(c)). Let t(α)=pt(\alpha^{\prime})=p^{\prime}. By uniqueness and the definition of graph of space normal forms, we have

(1) η(x,p)\displaystyle\eta(x,p) =η(x,p)t1c.\displaystyle=\eta(x,p^{\prime})*t^{-1}*c.

Letting γ=γ0tϵ1γk1\gamma^{\prime}=\gamma_{0}*t^{\epsilon_{1}}*...*\gamma_{k-1}, the inductive hypothesis applied to γ\gamma^{\prime} and α\alpha^{\prime} yields that

(2) |η(x,p)|\displaystyle\left\lvert\eta(x,p^{\prime})\right\rvert Kk((k1)L+|γ|+|α|+k).\displaystyle\leq Kk((k-1)L+|\gamma^{\prime}|+|\alpha^{\prime}|+k).

Applying (2) to (1), we obtain

(3) |η(x,q)|\displaystyle\left\lvert\eta(x,q)\right\rvert Kk((k1)L+|γ|+|α|+k)+|c|+1\displaystyle\leq Kk((k-1)L+\left\lvert\gamma^{\prime}\right\rvert+\left\lvert\alpha^{\prime}\right\rvert+k)+\left\lvert c\right\rvert+1

By assumption, β\beta^{\prime} is a KK-quasi-geodesic. Hence there is a point qβq\in\beta such that d(t(a),q)Md(t(a),q)\leq M. But then since βαγk\beta\cup\alpha\cup\gamma_{k} forms a geodesic triangle, there is a point zαγkz\in\alpha\cup\gamma_{k} such that d(t(a),z)Ld(t(a),z)\leq L. We may now divide into two subcases: either zαz\in\alpha or zγkz\in\gamma_{k}.

Refer to caption
Figure 3. Case 1: zαz\in\alpha.

First suppose zz is a vertex traversed by α\alpha, see Figure 3. It follows that

d(o(a),t(a))\displaystyle d(o(a),t(a)) |γk|+L+|α|\displaystyle\leq\left\lvert\gamma_{k}\right\rvert+L+\left\lvert\alpha\right\rvert
d(o(c),t(c))\displaystyle d(o(c),t(c)) L+|α|\displaystyle\leq L+\left\lvert\alpha\right\rvert

and hence, since cc is a KK-quasi-geodesic, we have

(4) |c|\displaystyle\left\lvert c\right\rvert K(L+|α|)+K.\displaystyle\leq K(L+\left\lvert\alpha\right\rvert)+K.

We also have

(5) |α|\displaystyle\left\lvert\alpha^{\prime}\right\rvert d(o(a),t(a))+2|γk|+L+|α|+2.\displaystyle\leq d(o(a),t(a))+2\leq\left\lvert\gamma_{k}\right\rvert+L+\left\lvert\alpha\right\rvert+2.

In order, applying (4), (5) and the fact that |γ|=|γ|+1+|γk|\left\lvert\gamma\right\rvert=\left\lvert\gamma^{\prime}\right\rvert+1+\left\lvert\gamma_{k}\right\rvert to (3), we obtain

|η(x,q)|\displaystyle\left\lvert\eta(x,q)\right\rvert Kk((k1)L+|γ|+|α|+k)+K(L+|α|)+K+1\displaystyle\leq Kk((k-1)L+\left\lvert\gamma^{\prime}\right\rvert+\left\lvert\alpha^{\prime}\right\rvert+k)+K(L+\left\lvert\alpha\right\rvert)+K+1
Kk((k1)L+|γ|+|γk|+L+|α|+k+2)+K(L+|α|)+K+1\displaystyle\leq Kk((k-1)L+\left\lvert\gamma^{\prime}\right\rvert+\left\lvert\gamma_{k}\right\rvert+L+\left\lvert\alpha\right\rvert+k+2)+K(L+\left\lvert\alpha\right\rvert)+K+1
Kk(kL+|γ|+|α|+k+1)+K(L+|α|)+K+1\displaystyle\leq Kk(kL+\left\lvert\gamma\right\rvert+\left\lvert\alpha\right\rvert+k+1)+K(L+\left\lvert\alpha\right\rvert)+K+1
Kk(kL+|γ|+|α|+k+1)+K(L+|α|+2)\displaystyle\leq Kk(kL+\left\lvert\gamma\right\rvert+\left\lvert\alpha\right\rvert+k+1)+K(L+\left\lvert\alpha\right\rvert+2)
K(k+1)(kL+|γ|+|α|+k+1).\displaystyle\leq K(k+1)(kL+\left\lvert\gamma\right\rvert+\left\lvert\alpha\right\rvert+k+1).
Refer to caption
Figure 4. Case 2: zγkz\in\gamma_{k}.

Now we deal with the other case, see Figure 4. If zz is a vertex traversed by γk\gamma_{k}, then

d(o(a),t(a))\displaystyle d(o(a),t(a)) d(o(γk),z)+L\displaystyle\leq d(o(\gamma_{k}),z)+L
d(o(c),t(c))\displaystyle d(o(c),t(c)) L+d(z,y)+|α|\displaystyle\leq L+d(z,y)+\left\lvert\alpha\right\rvert

and hence, since cc is a KK-quasi-geodesic, we have

(6) |c|\displaystyle\left\lvert c\right\rvert K(L+d(z,y)+|α|)+K.\displaystyle\leq K(L+d(z,y)+\left\lvert\alpha\right\rvert)+K.

We have

(7) |α|\displaystyle\left\lvert\alpha^{\prime}\right\rvert d(o(a),t(a))+2d(o(γk),z)+L+2.\displaystyle\leq d(o(a),t(a))+2\leq d(o(\gamma_{k}),z)+L+2.

In order, applying (6) and (7) to (2) we obtain

|η(x,p)|\displaystyle\left\lvert\eta(x,p)\right\rvert\leq Kk((k1)L+|γ|+|α|+k)\displaystyle Kk((k-1)L+\left\lvert\gamma^{\prime}\right\rvert+\left\lvert\alpha^{\prime}\right\rvert+k)
+K(L+d(z,q)+|α|)+K+1\displaystyle+K(L+d(z,q)+\left\lvert\alpha\right\rvert)+K+1
\displaystyle\leq Kk((k1)L+|γ|+d(o(γk),z)+L+k+2)\displaystyle Kk((k-1)L+\left\lvert\gamma^{\prime}\right\rvert+d(o(\gamma_{k}),z)+L+k+2)
+K(L+d(z,q)+|α|)+K+1\displaystyle+K(L+d(z,q)+\left\lvert\alpha\right\rvert)+K+1
\displaystyle\leq Kk(kL+|γ|+d(o(γk),z)+k+2)\displaystyle Kk(kL+\left\lvert\gamma^{\prime}\right\rvert+d(o(\gamma_{k}),z)+k+2)
+K(L+d(z,q)+|α|+2)\displaystyle+K(L+d(z,q)+\left\lvert\alpha\right\rvert+2)

and then using the fact that |γ|=|γ|+1+d(o(γk),z)+d(z,q)|\gamma|=|\gamma^{\prime}|+1+d(o(\gamma_{k}),z)+d(z,q), we obtain

|η(x,p)|\displaystyle\left\lvert\eta(x,p)\right\rvert\leq Kk(kL+|γ|+|α|+k+1)+K(L+|γ|+|α|+k+1)\displaystyle Kk(kL+\left\lvert\gamma\right\rvert+\left\lvert\alpha\right\rvert+k+1)+K(L+\left\lvert\gamma\right\rvert+\left\lvert\alpha\right\rvert+k+1)
\displaystyle\leq K(k+1)(kL+|γ|+|α|+k+1)\displaystyle K(k+1)(kL+\left\lvert\gamma\right\rvert+\left\lvert\alpha\right\rvert+k+1)

This concludes the proof of the claim.

Consider the polynomial function f:f:\mathbb{N}\to\mathbb{R} given by

f(n)=K(n+1)2(L+2).f(n)=K(n+1)^{2}(L+2).

We have shown that for all x,yX~𝒳(0)x,y\in\tilde{X}_{\mathcal{X}}^{(0)}, we have |η(x,y)|f(d(x,y))\left\lvert\eta(x,y)\right\rvert\leq f(d(x,y)). We now want to show that the graph of spaces normal forms are actually ff-quasi-geodesics.

Let h=h0tϵ1h1tϵkhkIm(η)h=h_{0}*t^{\epsilon_{1}}*h_{1}*...*t^{\epsilon_{k}}*h_{k}\in\operatorname{Im}(\eta) be a normal form and let h=hitϵi+1hi+1tϵjhjh^{\prime}=h_{i}^{\prime}*t^{\epsilon_{i+1}}*h_{i+1}*...*t^{\epsilon_{j}}*h_{j}^{\prime} be a subpath of hh. Then we have that

ηX~(o(hi),t(hi))tϵi+1hi+1tϵjhj\eta_{\tilde{X}}(o(h_{i}^{\prime}),t(h_{i}^{\prime}))*t^{\epsilon_{i+1}}*h_{i+1}*...*t^{\epsilon_{j}}*h_{j}^{\prime}

satisfies the three conditions of the definition of a graph of space normal form, where here we are using the fact that ηA~\eta_{\tilde{A}} and ηB~\eta_{\tilde{B}} are prefix closed to see that hjh_{j}^{\prime} is a normal form. Combining this with the uniqueness of graph of space normal forms we have

η(o(h),t(h))=ηX~(o(hi),t(hi))tϵi+1hi+1tϵjhj.\eta(o(h^{\prime}),t(h^{\prime}))=\eta_{\tilde{X}}(o(h_{i}^{\prime}),t(h_{i}^{\prime}))*t^{\epsilon_{i+1}}*h_{i+1}*...*t^{\epsilon_{j}}*h_{j}^{\prime}.

We showed that:

|η(t(hi),t(h))|f(d(t(hi),t(h)))\left\lvert\eta(t(h^{\prime}_{i}),t(h^{\prime}))\right\rvert\leq f(d(t(h_{i}^{\prime}),t(h^{\prime})))

and so we know that

|h|\displaystyle|h^{\prime}| =|hi|+|η(t(hi),t(h))|\displaystyle=\left\lvert h_{i}^{\prime}\right\rvert+\left\lvert\eta(t(h_{i}^{\prime}),t(h^{\prime}))\right\rvert
Kd(o(hi),t(hi))+K+|η(t(hi),t(h))|\displaystyle\leq Kd(o(h_{i}^{\prime}),t(h_{i}^{\prime}))+K+\left\lvert\eta(t(h_{i}^{\prime}),t(h^{\prime}))\right\rvert
Kd(o(hi),t(hi))+K+f(d(t(hi),t(h)))\displaystyle\leq Kd(o(h_{i}^{\prime}),t(h_{i}^{\prime}))+K+f(d(t(h_{i}^{\prime}),t(h^{\prime})))
f(d(o(hi),t(hi)))+f(d(t(hi),t(h)))\displaystyle\leq f(d(o(h_{i}^{\prime}),t(h_{i}^{\prime})))+f(d(t(h_{i}^{\prime}),t(h^{\prime})))
f(d(o(h),t(h))).\displaystyle\leq f(d(o(h^{\prime}),t(h^{\prime}))).

By Theorem 5.5, it follows that there is some constant K=K(f)K^{\prime}=K^{\prime}(f) such that η\eta is a KK^{\prime}-quasi-geodesic normal form. ∎

Theorem 5.7 can be used to show that certain subgroups of a hyperbolic acylindrical graph of hyperbolic groups are quasi-convex. In the section that follows, we are going to do precisely this to deduce that Magnus subgroups of one-relator groups with a quasi-convex hierarchy are quasi-convex. The main idea we employ is to use quasi-geodesic normal forms in vertex spaces to build quasi-geodesic normal forms relative to given subgroups. Other conditions of a different flavour already exist in the literature, see [BW13] and the references therein. These results cannot be applied directly as the edge groups in the Magnus splittings are not necessarily malnormal and are usually very far from being Noetherian. Nevertheless, an alternative approach could involve understanding the splittings of Magnus subgroups of one-relator groups induced by the Magnus hierarchy and then following the approach of Dahmani from [Dah03] for showing local (relative) quasi-convexity of limit groups. We opt for our more direct approach which may be of independent interest.

5.4. Quasi-convex Magnus subgraphs

Let us first define what a quasi-convex one-relator hierarchy is. This definition is a reformulation of Wise’s [Wis21] notion.

Definition 5.8.

A one-relator tower (respectively, hierarchy) XNX1X0X_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0} is a quasi-convex tower (respectively, hierarchy) if π1(Ai+1),π1(Bi+1)\pi_{1}(A_{i+1}),\pi_{1}(B_{i+1}) are quasi-isometrically embedded in π1(Xi)\pi_{1}(X_{i}) for all ii, where Ai+1,Bi+1Xi+1A_{i+1},B_{i+1}\subset X_{i+1} are the associated Magnus subgraphs.

In order to prove that one-relator groups with torsion have quasi-convex (one-relator) hierarchies, Wise showed in [Wis21, Lemma 19.8] that their Magnus subgroups are quasi-convex. When the torsion assumption is dropped, this is certainly no longer true. However, under additional hypotheses, we may recover quasi-convexity of Magnus subgroups. The only results specifically from the theory of one-relator groups we shall need are the Freiheitssatz and our hierarchy results.

Theorem 5.9.

Let X=(Γ,λ)X=(\Gamma,\lambda) be a finite one-relator complex, p:YXp:Y\to X a cyclic cover and ZD(p)Z\in\operatorname{\mathbb{Z}D}(p) a finite one-relator \mathbb{Z}-domain. Suppose further that the following hold:

  1. (1)

    π1(X)\pi_{1}(X) is hyperbolic.

  2. (2)

    For all connected Magnus subgraphs WZW\subset Z, π1(W)\pi_{1}(W) is quasi-convex in π1(Z)\pi_{1}(Z).

  3. (3)

    π1(Z)\pi_{1}(Z) is quasi-convex in π1(X)\pi_{1}(X).

Then for all connected Magnus subgraphs CXC\subset X, π1(C)\pi_{1}(C) is quasi-convex in π1(X)\pi_{1}(X).

Proof.

Let CXC\subset X be a connected Magnus subgraph. Denote by Z=(Λ,λ~)Z=(\Lambda,\tilde{\lambda}). Since quasi-convexity is transitive, we may assume that E(C)=E(Γ){f}E(C)=E(\Gamma)-\{f\} for some edge ff. Note that CC is connected and so ff is non separating.

For each edge eE(X)e\in E(X), choose a lift e0e_{0} in YY as in Subsection 4.1 and denote by ei=tie0e_{i}=t^{i}\cdot e_{0} for each ii\in\mathbb{Z}. Denote by mem_{e} the smallest integer such that emee_{m_{e}} is traversed by λ~\tilde{\lambda}, and by MeM_{e} the largest. We have two cases to consider.

Case 1. For all fiE(Λ)f_{i}\in E(\Lambda), we have that mfiMfm_{f}\leq i\leq M_{f} and fif_{i} is non separating in Λ\Lambda.

The action of Deck(p)\operatorname{Deck}(p) on YY, combined with the Freiheitssatz, gives us a graph of spaces 𝒳=(S1,{Z},{ZtZ},{±})\mathcal{X}=(S^{1},\{Z\},\{Z\cap t\cdot Z\},\{\partial^{\pm}\}) as in Proposition 4.3. We moreover have a map (the horizontal map):

𝗁:X𝒳X\mathsf{h}:X_{\mathcal{X}}\to X

that is a homotopy equivalence obtained by collapsing the edge space onto the vertex space as in Proposition 3.2. This lifts to a map in the universal covers:

𝗁~:X𝒳~X~.\tilde{\mathsf{h}}:X_{\tilde{\mathcal{X}}}\to\tilde{X}.

Denote by A,BZA,B\subset Z the Magnus subgraphs that are the images of ±\partial^{\pm}. That is, A=t1(ZtZ)A=t^{-1}\cdot(Z\cap t\cdot Z) and B=ZtZB=Z\cap t\cdot Z.

Denote by A=ΛfMfA^{\prime}=\Lambda-f_{M_{f}} and B=ΛfmfB^{\prime}=\Lambda-f_{m_{f}}. These are connected since fmff_{m_{f}} and fMff_{M_{f}} are non separating and they are Magnus subgraphs since both fmff_{m_{f}} and fMff_{M_{f}} are traversed by λ~\tilde{\lambda}. In particular, p1(C)ZABp^{-1}(C)\cap Z\subset A^{\prime}\cap B^{\prime}. Denote by Z~\tilde{Z} the universal cover of ZZ. By the Freiheitssatz, the universal covers of AAA\subset A^{\prime} and BBB\subset B^{\prime} are trees, including into Z~\tilde{Z}. Denote by A~A~\tilde{A}\subset\tilde{A}^{\prime} and B~B~\tilde{B}\subset\tilde{B}^{\prime} subgraphs of Z~\tilde{Z} corresponding to universal covers of AAA\subset A^{\prime} and BBB\subset B^{\prime} respectively. Since π1(A)\pi_{1}(A^{\prime}) and π1(B)\pi_{1}(B^{\prime}) are quasi-convex subgroups of π1(Z)\pi_{1}(Z), there are prefix-closed, quasi-geodesic, π1(Z)\pi_{1}(Z)-equivariant normal forms ηA~\eta_{\tilde{A}^{\prime}}, ηB~\eta_{\tilde{B}^{\prime}} for Z~\tilde{Z}, relative to A~\tilde{A}^{\prime} and B~\tilde{B}^{\prime} respectively. Recall that our normal forms are immersed paths and so ηA~\eta_{\tilde{A}^{\prime}} and ηB~\eta_{\tilde{B}^{\prime}} necessarily restrict to geodesic normal forms on A~\tilde{A}^{\prime} and B~\tilde{B}^{\prime}. Moreover, they necessarily agree on intersections of translates of A~\tilde{A}^{\prime} with translates of B~\tilde{B}^{\prime}. Since A~\tilde{A} and B~\tilde{B} are subtrees of A~\tilde{A}^{\prime} and B~\tilde{B}^{\prime} respectively, these are also normal forms relative to A~\tilde{A} and B~\tilde{B}. Hence, we may apply Theorem 5.4 to obtain unique prefix-closed π1(X𝒳)\pi_{1}(X_{\mathcal{X}})-equivariant graph of space normal forms η\eta for X𝒳~X_{\tilde{\mathcal{X}}}, induced by ({ηA~},{ηA~,ηB~})\left(\{\eta_{\tilde{A}^{\prime}}\},\{\eta_{\tilde{A}^{\prime}},\eta_{\tilde{B}^{\prime}}\}\right). By Theorems 3.4 and 5.7, η\eta is also quasi-geodesic.

Now let c:IC~X~c:I\looparrowright\tilde{C}\subset\tilde{X} be an immersed path and let c:IX𝒳~c^{\prime}:I\looparrowright X_{\tilde{\mathcal{X}}} be a path such that 𝗁c\mathsf{h}\circ c^{\prime} is path homotopic to cc via a sequence of collapses of segments in the domain which mapped to vertical edges in X𝒳~X_{\tilde{\mathcal{X}}} under cc^{\prime}. Since X𝒳X_{\mathcal{X}} is finite, there is a constant kk such that preimages of vertices in X~\tilde{X} under 𝗁\mathsf{h} are segments of length at most kk. In particular, |c|k|c||c^{\prime}|\leqslant k|c|. After possibly performing finitely many homotopies through vertical squares (which do not alter the composition with 𝗁\mathsf{h}), we may assume that c=c0tϵ1c1tϵncnc^{\prime}=c_{0}*t^{\epsilon_{1}}*c_{1}*...*t^{\epsilon_{n}}*c_{n} has the property that for all ii, if ee is the first edge that cic_{i} traverses, then tϵiet^{\epsilon_{i}}*e does not form the corner of a vertical square. Since p1(C)ZABp^{-1}(C)\cap Z\subset A^{\prime}\cap B^{\prime}, we see that each cic_{i} is both in normal form with respect to ηA~\eta_{\tilde{A}^{\prime}} and ηB~\eta_{\tilde{B}^{\prime}} by the remarks in the previous paragraph. Hence, cc^{\prime} is in normal form with respect to η\eta. Since η\eta was a quasi-geodesic normal form, it follows that π1(C)\pi_{1}(C) is quasi-convex in π1(X)\pi_{1}(X).

Case 2. There is some fiE(Λ)f_{i}\in E(\Lambda) such that either i<mfi<m_{f}, or i>Mfi>M_{f}, or fif_{i} is separating in Λ\Lambda.

Let XX^{\prime} be the one-relator complex obtained from XX by adding a single edge, dd. Thus, we have π1(X)=π1(X)x\pi_{1}(X^{\prime})=\pi_{1}(X)*\langle x\rangle. If ϕ:π1(X)\phi:\pi_{1}(X)\to\mathbb{Z} is the epimorphism inducing pp, denote by ϕ:π1(X)\phi^{\prime}:\pi_{1}(X^{\prime})\to\mathbb{Z} the epimorphism such that ϕπ1(X)=ϕ\phi^{\prime}\mid\pi_{1}(X)=\phi and ϕ(x)=1\phi^{\prime}(x)=1. Let Z′′ZZ^{\prime\prime}\subset Z be the subcomplex obtained from ZZ by removing each fif_{i} with i<mfi<m_{f} or i>Mfi>M_{f}. If p:YXp^{\prime}:Y^{\prime}\to X^{\prime} is the cyclic cover induced by ϕ\phi^{\prime}, then Z′′Z^{\prime\prime} lifts to YY^{\prime}. Moreover, for appropriately chosen integers klk\leq l, we see that

(Λ,λ~)=Z=Z′′(i=kldi)(\Lambda^{\prime},\tilde{\lambda}^{\prime})=Z^{\prime}=Z^{\prime\prime}\cup\left(\bigcup_{i=k}^{l}d_{i}\right)

is a one-relator \mathbb{Z}-domain for pp^{\prime}. By construction, for all fiE(Λ)f_{i}\in E(\Lambda^{\prime}), we have that mfiMfm_{f}\leq i\leq M_{f} and that fif_{i} is non separating in Λ\Lambda^{\prime}. Moreover

π1(Z)=i=0klπ1(Ai)xi\pi_{1}(Z^{\prime})=\bigast_{i=0}^{k-l}\pi_{1}(A_{i})^{x^{i}}

where each AiA_{i} is a (possible empty) connected subcomplex of ZZ. By hypothesis, π1(A)\pi_{1}(A^{\prime}) is quasi-convex in π1(X)\pi_{1}(X^{\prime}) for all connected subgraphs AZA^{\prime}\subset Z^{\prime}. By the proof of the first case, we see that π1(C)\pi_{1}(C) is quasi-convex in π1(X)\pi_{1}(X^{\prime}). But then π1(C)\pi_{1}(C) must also be quasi-convex in π1(X)\pi_{1}(X) and we are done.∎

Corollary 5.10.

Let X=(Γ,λ)X=(\Gamma,\lambda) be a one-relator complex. Suppose that π1(X)\pi_{1}(X) is hyperbolic and XX admits a quasi-convex one-relator hierarchy. Then π1(A)\pi_{1}(A) is quasi-convex in π1(X)\pi_{1}(X) for all Magnus subgraphs AXA\subset X.

Proof.

The proof is by induction on hierarchy length. The base case is clear. Theorems 3.4 and 5.9 handle the inductive step. ∎

At this point we remark that we have essentially proved the equivalence between (1) and (2) from Theorem 7.1. Nevertheless, we postpone the complete proof to Section 7.

6. \mathbb{Z}-stable HNN-extensions of one-relator groups

Let us repeat the definition of the \mathbb{Z}-stable number s(ψ)s\mathbb{Z}(\psi) from the introduction. If ψ:AB\psi:A\to B is an isomorphism of two subgroups A,B<HA,B<H, inductively define

𝒜0ψ={[H]},,𝒜i+1ψ={[ψ(AAi)]}Ai[Ai]𝒜iψ,\mathcal{A}^{\psi}_{0}=\{[H]\},\quad\ldots,\quad\mathcal{A}^{\psi}_{i+1}=\{[\psi(A\cap A_{i})]\}_{A_{i}\in[A_{i}]\in\mathcal{A}^{\psi}_{i}},\quad\ldots

Then we denote by 𝒜¯iψ𝒜iψ\bar{\mathcal{A}}_{i}^{\psi}\subset\mathcal{A}_{i}^{\psi} the subset corresponding to the conjugacy classes of non-cyclic subgroups. Define the \mathbb{Z}-stable number s(ψ)s\mathbb{Z}(\psi) of ψ\psi as

s(ψ)=sup{k+1𝒜¯kψ}{}s\mathbb{Z}(\psi)=\sup{\{k+1\mid\bar{\mathcal{A}}_{k}^{\psi}\neq\emptyset\}}\in\mathbb{N}\cup\{\infty\}

where s(ψ)=s\mathbb{Z}(\psi)=\infty if 𝒜¯iψ\bar{\mathcal{A}}_{i}^{\psi}\neq\emptyset for all ii. We say that HψH*_{\psi} is \mathbb{Z}-stable if s(ψ)<s\mathbb{Z}(\psi)<\infty.

For the purposes of this section, unless otherwise stated, we will always assume that HH is a finitely generated one-relator group and that A,B<HA,B<H are two strongly inert subgroups of Magnus subgroups for some one-relator presentation of HH. For the main results of this article, the reader should have in mind the case in which AA and BB are Magnus subgroups themselves. Let ψ:AB\psi:A\to B be an isomorphism. Consider the HNN-extension of HH over ψ\psi:

GHψH,ttat1=ψ(a),aAG\cong H*_{\psi}\cong\langle H,t\mid tat^{-1}=\psi(a),\forall a\in A\rangle

We will call this an inertial one-relator extension.

Remark 6.1.

Recall that if XX is a one-relator complex, then a one-relator splitting of π1(X)\pi_{1}(X) is a HNN-splitting of π1(X)\pi_{1}(X) as in Theorem 4.13. By Remark 4.14, every one-relator splitting π1(Z)ψ\pi_{1}(Z)*_{\psi} of π1(X)\pi_{1}(X) is an inertial one-relator extension.

Remark 6.2.

If HH is a one-relator group with the trivial relation, HH itself is a Magnus subgroup of HH. Thus, HNN-extensions of free groups with finitely generated strongly inert edge groups are inertial one-relator extensions.

Remark 6.3.

For simplicity, we focus only on HNN-extensions. However, with minor modifications, most of the results in this section hold for graphs of groups {Γ,{Gv},{Ge},{e±}}\{\Gamma,\{G_{v}\},\{G_{e}\},\{\partial_{e}^{\pm}\}\} satisfying the following:

  1. (1)

    Γ\Gamma is a connected graph with χ(Γ)0\chi(\Gamma)\geq 0,

  2. (2)

    for each vV(Γ)v\in V(\Gamma), GvG_{v} has a one-relator presentation so that each adjacent edge group is a strongly inert subgroup of a Magnus subgroup of GvG_{v}.

This class of groups includes groups with staggered presentations.

Let TT denote the Bass–Serre tree associated with the inertial one-relator extension HψH*_{\psi}. See [Ser03] for the relevant notions in Bass–Serre theory. We will identify each vertex of TT with a left coset of HH. Denote by

𝒮n\displaystyle\mathcal{S}^{n} ={SS a geodesic segment of length n with an endpoint at H}\displaystyle=\{S\mid S\text{ a geodesic segment of length $n$ with an endpoint at $H$}\}
𝒮\displaystyle\mathcal{S} =i𝒮i\displaystyle=\bigcup_{i}\mathcal{S}^{i}

Each edge in TT has an orientation induced by ψ\psi. Denote by 𝒮+n𝒮n\mathcal{S}^{n}_{+}\subset\mathcal{S}^{n} the geodesic segments which only consist of edges pointing away from HH and by 𝒮n𝒮n\mathcal{S}^{n}_{-}\subset\mathcal{S}^{n} those pointing towards HH. The elements [An]𝒜¯nψ[A_{n}]\in\bar{\mathcal{A}}_{n}^{\psi} correspond to stabilisers of segments S𝒮nS\in\mathcal{S}_{-}^{n} such that rr(Stab(S))0\operatorname{rr}(\operatorname{Stab}(S))\neq 0.

For inertial one-relator extensions, the ranks of stabilisers of elements in 𝒮n\mathcal{S}^{n} are bounded in a strong sense.

Lemma 6.4.

For all n1n\geq 1, the following holds:

HSS𝒮nrr(Stab(S))\displaystyle\sum_{\begin{subarray}{c}H\cdot S\\ S\in\mathcal{S}^{n}\end{subarray}}\operatorname{rr}(\operatorname{Stab}(S)) =HSS𝒮nrr(Stab(S))+HSS𝒮+nrr(Stab(S))\displaystyle=\sum_{\begin{subarray}{c}H\cdot S\\ S\in\mathcal{S}_{-}^{n}\end{subarray}}\operatorname{rr}(\operatorname{Stab}(S))+\sum_{\begin{subarray}{c}H\cdot S\\ S\in\mathcal{S}_{+}^{n}\end{subarray}}\operatorname{rr}(\operatorname{Stab}(S))
HSS𝒮±nrr(Stab(S))\displaystyle\sum_{\begin{subarray}{c}H\cdot S\\ S\in\mathcal{S}_{\pm}^{n}\end{subarray}}\operatorname{rr}(\operatorname{Stab}(S)) rr(A).\displaystyle\leq\operatorname{rr}(A).
Proof.

Let S𝒮nS\in\mathcal{S}^{n}, then Stab(S)=HHc1\operatorname{Stab}(S)=H\cap H^{c^{-1}} where cc is equal to a reduced word of the form:

c=c1tϵ1cntϵnc=c_{1}t^{\epsilon_{1}}\ldots c_{n}t^{\epsilon_{n}}

with ϵi=±1\epsilon_{i}=\pm 1 and ciHc_{i}\in H. The segment SS is the geodesic segment with vertices H,c1tϵ1H,,cHH,c_{1}t^{\epsilon_{1}}H,\ldots,cH. By assumption we have that AA is a strongly inert subgroup of a Magnus subgroup A<HA^{\prime}<H. We have that rr(AAa)=0\operatorname{rr}(A\cap A^{a})=0 for all aAAa\in A^{\prime}-A by Remark 2.1 (similarly for BB). Combining this with Theorem 2.8, if there is some ii such that ϵi=ϵi+1\epsilon_{i}=-\epsilon_{i+1}, then Stab(S)\operatorname{Stab}(S) is either cyclic or trivial since our word was reduced. Denote by W+1nW_{+1}^{n} the set of reduced words of the form tc2tcnttc_{2}t\ldots c_{n}t and by W1nW_{-1}^{n} the set of reduced words of the form t1c2t1cnt1t^{-1}c_{2}t^{-1}\ldots c_{n}t^{-1}. Let T±1nT_{\pm 1}^{n} be a set of representatives for the set of double cosets H\W±1n/HH\backslash W_{\pm 1}^{n}/H. Then we have

HSS𝒮nrr(Stab(S))=cT+1nrr(HHc1)+cT1nrr(HHc1)\sum_{\begin{subarray}{c}H\cdot S\\ S\in\mathcal{S}^{n}\end{subarray}}\operatorname{rr}(\operatorname{Stab}(S))=\sum_{c\in T_{+1}^{n}}\operatorname{rr}\left(H\cap H^{c^{-1}}\right)+\sum_{c\in T_{-1}^{n}}\operatorname{rr}\left(H\cap H^{c^{-1}}\right)

which yields the equality from the statement. By symmetry, it suffices to only bound one of the sums above to establish the inequality. We may assume that each word in T+1nT_{+1}^{n} is a prefix of a word in T+1n+1T_{+1}^{n+1}. If cT+1nc\in T_{+1}^{n}, denote by TcT_{c} the set of elements cn+1Hc_{n+1}\in H such that ccn+1tT+1n+1cc_{n+1}t\in T_{+1}^{n+1}. We first claim that each element cn+1Tcc_{n+1}\in T_{c} is in a distinct (HHc)cn+1B(H\cap H^{c})c_{n+1}B double coset. Let cn+1,cn+1Tcc_{n+1},c_{n+1}^{\prime}\in T_{c} be such that cn+1=hcn+1bc_{n+1}^{\prime}=hc_{n+1}b where h=c1hcHHch=c^{-1}h^{\prime}c\in H\cap H^{c} and bBb\in B. Then we have that ccn+1t=hccn+1tψ1(b)Hccn+1tHcc_{n+1}^{\prime}t=h^{\prime}cc_{n+1}t\psi^{-1}(b)\in Hcc_{n+1}tH, a contradiction. We next claim that

cT+1nrr(HHc1)=cT+1nrr(HHc)rr(A).\sum_{c\in T_{+1}^{n}}\operatorname{rr}\left(H\cap H^{c^{-1}}\right)=\sum_{c\in T_{+1}^{n}}\operatorname{rr}(H\cap H^{c})\leqslant\operatorname{rr}(A).

The equality is obvious, we now show the inequality. The proof of the claim proceeds by induction with the base case being clear as HHt1=AH\cap H^{t^{-1}}=A. So let us now assume the inductive hypothesis. We have

cT+1n+1rr(HHc)\displaystyle\sum_{c\in T_{+1}^{n+1}}\operatorname{rr}(H\cap H^{c}) =cT+1ncn+1Tcrr(HHccn+1t)\displaystyle=\sum_{c\in T_{+1}^{n}}\sum_{c_{n+1}\in T_{c}}\operatorname{rr}\left(H\cap H^{cc_{n+1}t}\right)
cT+1n(HHc)cn+1Bcn+1Hrr(Ht1HHccn+1)\displaystyle\leqslant\sum_{c\in T_{+1}^{n}}\sum_{\begin{subarray}{c}(H\cap H^{c})c_{n+1}B\\ c_{n+1}\in H\end{subarray}}\operatorname{rr}\left(H^{t^{-1}}\cap H\cap H^{cc_{n+1}}\right)
=cT+1n(HHc)cn+1Bcn+1Hrr(B(HHc)cn+1)\displaystyle=\sum_{c\in T_{+1}^{n}}\sum_{\begin{subarray}{c}(H\cap H^{c})c_{n+1}B\\ c_{n+1}\in H\end{subarray}}\operatorname{rr}\left(B\cap(H\cap H^{c})^{c_{n+1}}\right)
cT+1nrr(HHc)\displaystyle\leqslant\sum_{c\in T^{n}_{+1}}\operatorname{rr}(H\cap H^{c})
rr(A)\displaystyle\leqslant\operatorname{rr}(A)

where the first inequality follows from our previous claim, the second inequality follows from Theorem 2.11 and where the final inequality follows from the inductive hypothesis. The lemma now readily follows. ∎

The following proposition is key in our proof of \mathbb{Z}-stability of one-relator hierarchies of one-relator groups with negative immersions. The main consequence is that if s(ψ)=s\mathbb{Z}(\psi)=\infty, then there exists some element gGg\in G acting hyperbolically on TT such that rr(HHgn)0\operatorname{rr}(H\cap H^{g^{n}})\neq 0 for all nn.

Proposition 6.5.

If s(ψ)=s\mathbb{Z}(\psi)=\infty, then there are 1nrr(A)1\leq n\leq\operatorname{rr}(A) many HH-orbits of biinfinite geodesics STS\subset T such that the following holds:

  1. (1)

    SS contains the vertex HH.

  2. (2)

    Every finite subset of SS has non-cyclic, non-trivial stabiliser.

Moreover, for every such biinfinite geodesic, the following holds:

  1. (1)

    Every edge in SS is directed in the same direction.

  2. (2)

    There exists an element gGg\in G acting hyperbolically on TT with translation length at most rr(A)\operatorname{rr}(A) such that gS=Sg\cdot S=S.

Proof.

The fact that every such geodesic must consist of edges directed in the same direction follows from the equality in Lemma 6.4. The fact that there are 1nrr(A)1\leq n\leq\operatorname{rr}(A) many such HH-orbits of geodesics follows by applying the pigeonhole principle and the inequality from Lemma 6.4. So now let S1,,SnS_{1},...,S_{n} be a collection of HH-orbit representatives of such biinfinite geodesics in TT. Identify each vertex of SiS_{i} with an integer so that HH is the vertex associated with 0. Let gi,jGg_{i,j}\in G be an element such that gi,jHg_{i,j}H is the jthj^{\text{th}} vertex in SiS_{i}. Then gi,j1Sig_{i,j}^{-1}\cdot S_{i} must be in the same HH-orbit of some SmS_{m}. But then by the pigeonhole principle, Si,gi,11Si,,gi,rr(A)1SiS_{i},g_{i,1}^{-1}\cdot S_{i},...,g_{i,\operatorname{rr}(A)}^{-1}\cdot S_{i} must contain two biinfinite geodesics in the same HH-orbit. Suppose that h1gi,k1Si=h2gi,l1Si=Slh_{1}g_{i,k}^{-1}\cdot S_{i}=h_{2}g_{i,l}^{-1}\cdot S_{i}=S_{l} for some h1,h2Hh_{1},h_{2}\in H and 1k<lrr(A)1\leqslant k<l\leqslant\operatorname{rr}(A). Then we have

Si=h2gi,l(h1gi,k)1SiS_{i}=h_{2}g_{i,l}(h_{1}g_{i,k})^{-1}\cdot S_{i}

where h2gi,l(h1gi,k)1h_{2}g_{i,l}(h_{1}g_{i,k})^{-1} acts hyperbolically on TT with translation length at most rr(A)\operatorname{rr}(A). ∎

6.1. One-relator groups with torsion

The following lemma allows one to easily establish \mathbb{Z}-stability of a one-relator splitting in certain cases: if the edge groups of a one-relator splitting have no exceptional intersection, then the splitting is \mathbb{Z}-stable.

Lemma 6.6.

Let XX be a one-relator complex and let π1(X)π1(X1)ψ\pi_{1}(X)\cong\pi_{1}(X_{1})*_{\psi} be a one-relator splitting where A,BX1A,B\subset X_{1} are the associated Magnus subgraphs. Let A,BX1A^{\prime},B^{\prime}\subset X_{1} be Magnus subgraphs such that AAA\subset A^{\prime}, BBB\subset B^{\prime} and ABA^{\prime}\cap B^{\prime} is connected. If π1(A)π1(B)=π1(AB)\pi_{1}(A^{\prime})\cap\pi_{1}(B^{\prime})=\pi_{1}(A^{\prime}\cap B^{\prime}), then s(ψ)<s\mathbb{Z}(\psi)<\infty.

Proof.

Let YXY\to X be the cyclic cover containing X1X_{1} and let tt be a generator of the deck group. Let γ:ΓA\gamma:\Gamma\looparrowright A be a graph immersion with χ(Γ)1\chi(\Gamma)\leq-1. By assumption and Theorem 2.8, γ\gamma is homotopic in X1X_{1} to a graph immersion γ1:ΓB=tA\gamma_{1}:\Gamma\looparrowright B=t\cdot A if and and only if γ\gamma is homotopic in AA to a graph immersion γ1:Γ1AB\gamma_{1}:\Gamma_{1}\looparrowright A\cap B. If s(ψ)2s\mathbb{Z}(\psi)\geq 2, there is a graph immersion γ\gamma as above such that γ1:Γ1ABtA\gamma_{1}:\Gamma_{1}\looparrowright A\cap B\hookrightarrow t\cdot A is homotopic in tX1t\cdot X_{1} to a graph immersion γ2:Γ2AtAt2A1(AB)\gamma_{2}:\Gamma_{2}\looparrowright A\cap t\cdot A\cap t^{2}\cdot A\hookrightarrow 1\cdot(A\cap B). Carrying on this argument, we see that if s(ψ)ns\mathbb{Z}(\psi)\geq n, then there exists a graph immersion γ:ΓA\gamma:\Gamma\looparrowright A such that χ(Γ)1\chi(\Gamma)\leq-1 and such that γ\gamma is homotopic to a graph immersion γn:ΓnAtAtnA\gamma_{n}:\Gamma_{n}\looparrowright A\cap t\cdot A\cap...\cap t^{n}\cdot A. However, clearly for large enough nn we have that AtAtnA=A\cap t\cdot A\cap...\cap t^{n}\cdot A=\emptyset. ∎

The condition from Lemma 6.6 always holds for one-relator splittings of one-relator groups with torsion by Remark 6.1 and Theorem 2.7 (see Remark 2.9), yielding the following.

Corollary 6.7.

All one-relator splittings of one-relator groups with torsion are \mathbb{Z}-stable.

6.2. One-relator groups with negative immersions

The aim of this subsection is to show that every one-relator splitting of a one-relator group with negative immersions is \mathbb{Z}-stable. In order to do this, we show that certain constraints on the subgroups of HψH*_{\psi} that are not conjugate into HH imply \mathbb{Z}-stability.

Theorem 6.8.

If every descending chain of non-cyclic, freely indecomposable proper subgroups of bounded rank

G=H0>H1>>Hn>G=H_{0}>H_{1}>...>H_{n}>...

is either finite or eventually conjugate into HH, then s(ψ)<s\mathbb{Z}(\psi)<\infty.

Proof.

Suppose for contradiction that s(ψ)=s\mathbb{Z}(\psi)=\infty. Let SS be a biinfinite geodesic as in Proposition 6.5. Let gGg\in G be the element acting by translations on SS. Since every edge in SS points in the same direction, we have that ϕ(g)0\phi(g)\neq 0. Hence, by definition of SS, HHgnH\cap H^{g^{n}} is non-trivial and non-cyclic for all nn.

Denote by ϕ:GG/H\phi:G\to G/\langle\langle H\rangle\rangle\cong\mathbb{Z}. Consider the descending chain of subgroups

G=H,g>H,g2>>H,g2n>G=\langle H,g\rangle>\langle H,g^{2}\rangle>...>\langle H,g^{2^{n}}\rangle>...

where we denote by Hn=H,g2nH_{n}=\langle H,g^{2^{n}}\rangle. Since rk(Hn)rk(H)+1\operatorname{rk}(H_{n})\leq\operatorname{rk}(H)+1, the ranks are bounded. The chain must be proper since ϕ(Hi)=ϕ(g)2iϕ(g)2j=ϕ(Hj)\phi(H_{i})=\langle\phi(g)2^{i}\rangle\neq\langle\phi(g)2^{j}\rangle=\phi(H_{j}) for all iji\neq j.

Each HiH_{i} has a Grushko decomposition

Hi=F(Xi)Ji,1Ji,niH_{i}=F(X_{i})*J_{i,1}*...*J_{i,n_{i}}

that is unique up to permutation and conjugation of factors and where each Ji,jJ_{i,j} is non-cyclic and freely indecomposable. By the Kurosh subgroup theorem, each Ji,jJ_{i,j} is a conjugate of a subgroup of some Ji1,lJ_{i-1,l}. Furthermore, we have that |Xi|+2nirk(H)+1\left\lvert X_{i}\right\rvert+2n_{i}\leq\operatorname{rk}(H)+1.

Now the claim is that for some integer m0m\geq 0 and for all i>mi>m, each Ji,jJ_{i,j} is either a conjugate of some Ji1,lJ_{i-1,l} or is conjugate into HH. If this was not the case, then there would be infinite sequences of integers i0,i1,i2,i_{0},i_{1},i_{2},... and j0,j1,j2,j_{0},j_{1},j_{2},... and elements g1,g2,Gg_{1},g_{2},...\in G such that

Ji0,j0>Ji1,j1g1>Ji2,j2g2>J_{i_{0},j_{0}}>J_{i_{1},j_{1}}^{g_{1}}>J_{i_{2},j_{2}}^{g_{2}}>...

where the inclusions are all proper and no Jik,jkJ_{i_{k},j_{k}} is conjugate into HH. But this contradicts the hypothesis and so the claim is proven.

So now denote by Ji=Ji,1Ji,niJ_{i}=J_{i,1}*...*J_{i,n_{i}}. For all imi\geq m, we have that ϕ(Ji)=l\phi(J_{i})=\langle l\rangle for some fixed ll. But now since ϕ(Hi)=ϕ(g)2i\phi(H_{i})=\langle\phi(g)2^{i}\rangle, this forces Ji<ker(ϕ)J_{i}<\ker(\phi). Thus, the induced homomorphism ϕHm\phi\mid_{H_{m}} factors through the projection HmHm/JmF(Xm)H_{m}\to H_{m}/\langle\langle J_{m}\rangle\rangle\cong F(X_{m}). But now ϕ(g2m)\phi\left(g^{2^{m}}\right) is a generator of ϕ(Hm)\phi(H_{m}), so there is some primitive element xmF(Xm)x_{m}\in F(X_{m}) that maps to a generator of ϕ(Hm)\phi(H_{m}) and such that xm=kmg2mx_{m}=k_{m}g^{2^{m}} for some kmker(ϕHm)k_{m}\in\ker(\phi\mid_{H_{m}}). Since xmx_{m} is primitive in F(Xm)F(X_{m}), we have:

Hm=xmKm=kmg2mKmH_{m}=\langle x_{m}\rangle*K_{m}=\left\langle k_{m}g^{2^{m}}\right\rangle*K_{m}

for some subgroup Km<HmK_{m}<H_{m} such that Km=ker(ϕ)\langle\langle K_{m}\rangle\rangle=\ker(\phi).

Now let TmT_{m} be the Bass–Serre tree associated with the HNN-extension xmKm\langle x_{m}\rangle*K_{m} with trivial edge groups. Each vertex is stabilised by a conjugate of KmK_{m} and each edge has trivial stabiliser. Since H<HmH<H_{m}, H<ker(ϕ)H<\ker(\phi) and HH is finitely generated, it follows that HH is a subgroup of a free product of finitely many conjugates of KmK_{m}. Since g2mg^{2^{m}} acts hyperbolically on TmT_{m} and TmT_{m} has trivial edge stabilisers, it follows that

HHg2nm=1H\cap H^{g^{2^{nm}}}=1

for all nn sufficiently large. But then this contradicts the assumption that s(ψ)=s\mathbb{Z}(\psi)=\infty. ∎

Corollary 6.9.

All one-relator splittings of one-relator groups with negative immersions are \mathbb{Z}-stable.

Proof.

The result follows from a result of Louder–Wilton [LW24, Theorem B] and Theorem 6.8. ∎

6.3. The graph of cyclic stabilisers and Baumslag–Solitar subgroups

To any inertial one-relator extension HψH*_{\psi}, we now describe how to associate a unique graph of cyclic groups 𝒢\mathcal{G}. This graph of cyclic groups will be called the graph of cyclic stabilisers of HψH*_{\psi} and will encode relations between cyclic stabilisers of segments leading out of HH. Moreover, certain Baumslag–Solitar subgroups of HψH*_{\psi} can be read off from 𝒢\mathcal{G}.

Denote by

𝒜\displaystyle\mathcal{A} ={[a]AAStab(S)a,S𝒮, a not a proper power in A}\displaystyle=\{[\langle a\rangle]_{A}\mid A\cap\operatorname{Stab}(S)\leqslant\langle a\rangle,S\in\mathcal{S},\text{ $a$ not a proper power in $A$}\}
\displaystyle\mathcal{B} ={[b]BBStab(S)b,S𝒮, b not a proper power in B}\displaystyle=\{[\langle b\rangle]_{B}\mid B\cap\operatorname{Stab}(S)\leqslant\langle b\rangle,S\in\mathcal{S},\text{ $b$ not a proper power in $B$}\}

In other words, 𝒜\mathcal{A} (respectively \mathcal{B}) is the set of conjugacy classes in AA (respectively in BB) of maximal cyclic subgroups of AA (respectively in BB) which contain a stabiliser Stab(S)\operatorname{Stab}(S) for some S𝒮S\in\mathcal{S}.

We now define the graph of cyclic stabilisers 𝒢=(Γ,{cv},{ce},{e±})\mathcal{G}=(\Gamma,\{\langle c_{v}\rangle\},\{\langle c_{e}\rangle\},\{\partial_{e}^{\pm}\}) as follows. Identify V(Γ)V(\Gamma) with the set 𝒜\mathcal{A}\sqcup\mathcal{B}. Choose any map ν:V(Γ)H\nu:V(\Gamma)\to H sending an element [a]A𝒜[\langle a\rangle]_{A}\in\mathcal{A} to a generator of a representative aAa\in A and [b]B[\langle b\rangle]_{B}\in\mathcal{B} to a generator of a representative bBb\in B. There are two types of edges: HH-edges and tt-edges.

For each pair of vertices v,wV(Γ)v,w\in V(\Gamma) and each double coset ν(v)hν(w)\langle\nu(v)\rangle h\langle\nu(w)\rangle, such that hHh\in H or hAt1BBtAh\in At^{-1}B\sqcup BtA and

ν(v)ν(w)h1,\langle\nu(v)\rangle\cap\langle\nu(w)\rangle^{h}\neq 1,

there is an edge ee connecting vv and ww. We further assume that if v,w𝒜v,w\in\mathcal{A} or v,wv,w\in\mathcal{B}, then hν(v)ν(w)h\notin\langle\nu(v)\rangle\langle\nu(w)\rangle. Then the boundary maps e±\partial_{e}^{\pm} are induced by the monomorphisms

ν(v)ν(w)h\displaystyle\langle\nu(v)\rangle\cap\langle\nu(w)\rangle^{h} ν(v)\displaystyle\hookrightarrow\langle\nu(v)\rangle
ν(v)h1ν(w)\displaystyle\langle\nu(v)\rangle^{h^{-1}}\cap\langle\nu(w)\rangle ν(w)\displaystyle\hookrightarrow\langle\nu(w)\rangle

If hHh\in H, then we say that ee is an HH-edge and if hAt1BBtAh\in At^{-1}B\sqcup BtA, we say that ee is a tt-edge. All edges between vertices in 𝒜\mathcal{A} and \mathcal{B} are oriented towards \mathcal{B} and a choice of orientation is made for the remaining edges.

Remark 6.10.

The isomorphism class of 𝒢\mathcal{G} as a graph of groups depends only on ψ\psi. This follows from the construction and justifies 𝒢\mathcal{G} being called the graph of cyclic stabilisers of HψH*_{\psi}.

Choose any map ξ:E(Γ)G\xi:E(\Gamma)\to G such that ξ(e)ν(o(e))hν(t(e))\xi(e)\in\langle\nu(o(e))\rangle h\langle\nu(t(e))\rangle where ν(o(e))hν(t(e))\langle\nu(o(e))\rangle h\langle\nu(t(e))\rangle is the double coset associated with ee. For all vV(Γ)v\in V(\Gamma), we may define a group homomorphism

μ:π1(𝒢,v)Hψ\mu:\pi_{1}(\mathcal{G},v)\to H*_{\psi}

induced by the choices ν,ξ\nu,\xi as follows.

Each element of π1(𝒢,v)\pi_{1}(\mathcal{G},v) can be represented by words of the form

cv0i0e1ϵ1cv1i1enϵncvninc_{v_{0}}^{i_{0}}e^{\epsilon_{1}}_{1}c_{v_{1}}^{i_{1}}...e^{\epsilon_{n}}_{n}c_{v_{n}}^{i_{n}}

where v0,vn=vv_{0},v_{n}=v, eiE(Γ)e_{i}\in E(\Gamma), ϵi=±1\epsilon_{i}=\pm 1, and t(ei)=vi=o(ei+1)t(e_{i})=v_{i}=o(e_{i+1}). Then we define

μ(cv0i0e1ϵ1cv1i1enϵncvnin)=ν(v0)i0ξ(e1)ϵ1ν(v1)i1ξ(en)ϵnν(vn)in\mu(c_{v_{0}}^{i_{0}}e^{\epsilon_{1}}_{1}c_{v_{1}}^{i_{1}}...e^{\epsilon_{n}}_{n}c_{v_{n}}^{i_{n}})=\nu(v_{0})^{i_{0}}\xi(e_{1})^{\epsilon_{1}}\nu(v_{1})^{i_{1}}...\xi(e_{n})^{\epsilon_{n}}\nu(v_{n})^{i_{n}}

One can check that this is well defined and depends only on ν,ξ\nu,\xi. We will call μ\mu the homomorphism induced by ν,ξ\nu,\xi.

A path γ:IΓ\gamma:I\to\Gamma is alternating if it does not traverse two HH-edges or tt-edges in a row. An HH-path is a path γ:IΓ\gamma:I\to\Gamma that only traverses HH-edges. We say a word e1ϵ1cv1i1enϵncvnine^{\epsilon_{1}}_{1}c_{v_{1}}^{i_{1}}...e^{\epsilon_{n}}_{n}c_{v_{n}}^{i_{n}} is an alternating word if e1ϵ1e2ϵ2enϵne^{\epsilon_{1}}_{1}*e^{\epsilon_{2}}_{2}*...*e^{\epsilon_{n}}_{n} is an alternating path. Recall that a word cv0i0e1ϵ1cv1i1enϵncvninc_{v_{0}}^{i_{0}}e^{\epsilon_{1}}_{1}c_{v_{1}}^{i_{1}}...e^{\epsilon_{n}}_{n}c_{v_{n}}^{i_{n}} is cyclically reduced if all of its cyclic permutations ejϵjcvjijenϵncvnin+i0ej1ϵj1cvj1ij1e^{\epsilon_{j}}_{j}c_{v_{j}}^{i_{j}}...e^{\epsilon_{n}}_{n}c_{v_{n}}^{i_{n}+i_{0}}...e^{\epsilon_{j-1}}_{j-1}c_{v_{j-1}}^{i_{j-1}} are also reduced. A word is cyclically alternating if all of its cyclic permutations are also alternating.

Lemma 6.11.

If c=e1ϵ1cv1i1enϵncvninc=e^{\epsilon_{1}}_{1}c_{v_{1}}^{i_{1}}...e^{\epsilon_{n}}_{n}c_{v_{n}}^{i_{n}} is a cyclically reduced and cyclically alternating word, then μ(c)\mu(c) is a cyclically reduced word. In particular, if n2n\geq 2, then μ(cvn),μ(c)\langle\mu(c_{v_{n}}),\mu(c)\rangle contains a Baumslag–Solitar subgroup, not conjugate into HH.

Proof.

If n=1n=1 the result is clear, so suppose that n2n\geq 2. We may assume that e1e_{1} is a tt-edge. Taking indices modulo nn, we have that μ(c)\mu(c) is not cyclically reduced if and only if ϵ2j1=ϵ2j+1=1\epsilon_{2j-1}=-\epsilon_{2j+1}=1 for some jj and

ν(v2j1)i2j1ξ(e2j)ϵ2jν(v2j)i2jA\nu(v_{2j-1})^{i_{2j-1}}\xi(e_{2j})^{\epsilon_{2j}}\nu(v_{2j})^{i_{2j}}\in A

or ϵ2j1=ϵ2j+1=1\epsilon_{2j-1}=-\epsilon_{2j+1}=-1 for some jj and

ν(v2j1)i2j1ξ(e2j)ϵ2jν(v2j)i2jB.\nu(v_{2j-1})^{i_{2j-1}}\xi(e_{2j})^{\epsilon_{2j}}\nu(v_{2j})^{i_{2j}}\in B.

In the first case, ν(v2j1)i2j1ξ(e2j)ϵ2jν(v2j)i2jA\nu(v_{2j-1})^{i_{2j-1}}\xi(e_{2j})^{\epsilon_{2j}}\nu(v_{2j})^{i_{2j}}\in A if and only if ξ(e2j)ϵ2jA\xi(e_{2j})^{\epsilon_{2j}}\in A. But this would then imply that v2j1=v2jv_{2j-1}=v_{2j}. Since AA is free and ν(v2j)\nu(v_{2j}) generates a maximal cyclic subgroup, it follows that ξ(e2j)ϵ2jν(v2j)\xi(e_{2j})^{\epsilon_{2j}}\in\langle\nu(v_{2j})\rangle, a contradiction. The same argument is valid for the other case. Thus, when n2n\geq 2, μ(c)\mu(c) acts hyperbolically on TT.

Now, by definition of 𝒢\mathcal{G}, there are integers k,lk,l, such that

μ(c)1μ(cvn)kμ(c)=μ(cvn)l.\mu(c)^{-1}\mu(c_{v_{n}})^{k}\mu(c)=\mu(c_{v_{n}})^{l}.

We may assume that kk and ll are smallest possible. If k=±1k=\pm 1 or l=±1l=\pm 1, then μ(c),μ(cvn)BS(1,±l)\langle\mu(c),\mu(c_{v_{n}})\rangle\cong\operatorname{BS}(1,\pm l) or BS(1,±k)\operatorname{BS}(1,\pm k) respectively. Since μ(c)\mu(c) acts hyperbolically on TT, it follows that this Baumslag–Solitar subgroup is not conjugate into HH. If k,l±1k,l\neq\pm 1, since μ(c)\mu(c) is cyclically reduced, we have that [μ(c),μ(cvn)][\mu(c),\mu(c_{v_{n}})] is also cyclically reduced. Thus, μ(cvn)l,[μ(c),μ(cvn)]2\langle\mu(c_{v_{n}})^{l},[\mu(c),\mu(c_{v_{n}})]\rangle\cong\mathbb{Z}^{2}. As before, this copy of 2\mathbb{Z}^{2} cannot be conjugate into HH. ∎

Example 6.12.

Consider the following one-relator splitting:

a,b,tta2t1bab1tat1bab1\displaystyle\langle a,b,t\mid ta^{2}t^{-1}bab^{-1}tat^{-1}bab^{-1}\rangle x,y,z,ttxt1=y,y2zxz1yzxz1\displaystyle\cong\langle x,y,z,t\mid txt^{-1}=y,y^{2}zxz^{-1}yzxz^{-1}\rangle
x,y,zy2zxz1yzxz1ψ\displaystyle\cong\langle x,y,z\mid y^{2}zxz^{-1}yzxz^{-1}\rangle*_{\psi}

This one-relator splitting is also a one-relator hierarchy of length one as y2zxz1yzxz1y^{2}zxz^{-1}yzxz^{-1} is a primitive element of F(x,y,z)F(x,y,z).

Since the edge groups of this splitting are cyclic, it follows that 𝒢\mathcal{G} has two vertices, one corresponding to [x]x[\langle x\rangle]_{\langle x\rangle} and the other to [y]y[\langle y\rangle]_{\langle y\rangle} and a tt-edge connecting them. Since y2zxz1yzxz1y^{2}zxz^{-1}yzxz^{-1} is a primitive word over y,zxz1y,zxz^{-1}, we see that zx2z1=y3zx^{2}z^{-1}=y^{-3}. Hence there is an edge connecting [x]x[\langle x\rangle]_{\langle x\rangle} with [y]y[\langle y\rangle]_{\langle y\rangle} corresponding to conjugation by zz, and where the edge monomorphisms are given by multiplication by 22 and 3-3 respectively. See Figure 5 for the graph of cyclic groups 𝒢\mathcal{G}.

We see that π1(𝒢)\pi_{1}(\mathcal{G}) has a cyclically alternating and cyclically reduced word zt1zt^{-1} and so our one-relator group contains a Baumslag–Solitar subgroup by Lemma 6.11. More explicitly, we have

x,zt1=a,bt1BS(2,3).\langle x,zt^{-1}\rangle=\langle a,bt^{-1}\rangle\cong\operatorname{BS}(2,-3).
Refer to caption
Figure 5. The graph of cyclic stabilisers 𝒢\mathcal{G}

We will denote by 𝒢k\mathcal{G}_{k} the full subgraph of groups of 𝒢\mathcal{G} on the vertices corresponding to maximal cyclic subgroups containing some Stab(S)\operatorname{Stab}(S) where S𝒮iS\in\mathcal{S}^{i} for some iki\leq k.

Lemma 6.13.

Let HψH*_{\psi} be an inertial one-relator extension with HH hyperbolic and with both edge groups quasi-convex in HH. Then the graph of cyclic stabilisers 𝒢\mathcal{G} is locally finite. Moreover, if s(ψ)<s\mathbb{Z}(\psi)<\infty, then 𝒢=𝒢s(ψ)\mathcal{G}=\mathcal{G}_{s\mathbb{Z}(\psi)} is finite.

Proof.

We first show that 𝒢\mathcal{G} is locally finite. Each vertex has at most one tt-edge connected to it, so it suffices to only bound the number of HH-edges connected to any given vertex. Such a bound follows from [KMW17, Proposition 6.7] (alternatively, one could use a slight modification of [GMRS98]).

Since 𝒢\mathcal{G} is locally finite, it follows from the construction that 𝒢k\mathcal{G}_{k} is finite for all kk. We claim that 𝒢=𝒢k\mathcal{G}=\mathcal{G}_{k} for all ks(ψ)k\geqslant s\mathbb{Z}(\psi). Denoting by

𝒮n=𝒮i<n𝒮i\mathcal{S}^{\geqslant n}=\mathcal{S}-\bigcup_{i<n}\mathcal{S}^{i}

we have that Stab(S)\operatorname{Stab}(S) is either trivial or infinite cyclic for all S𝒮s(ψ)S\in\mathcal{S}^{\geqslant s\mathbb{Z}(\psi)} by Lemma 6.4. Hence, for all S𝒮s(ψ)+1S\in\mathcal{S}^{\geqslant s\mathbb{Z}(\psi)+1}, there is some S𝒮s(ψ)S^{\prime}\in\mathcal{S}^{s\mathbb{Z}(\psi)} such that Stab(S)<Stab(S)\operatorname{Stab}(S)<\operatorname{Stab}(S^{\prime}). The result follows. ∎

Lemma 6.11 told us that certain Baumslag–Solitar subgroups of HψH*_{\psi} could be read off from its graph of cyclic stabilisers. Although we may not find all such subgroups in this way, we now show that under certain conditions, if HψH*_{\psi} does contain Baumslag–Solitar subgroups, then Lemma 6.11 will always produce a witness.

Theorem 6.14.

Let HψH*_{\psi} be an inertial one-relator extension. Suppose that s(ψ)<s\mathbb{Z}(\psi)<\infty, that HH is hyperbolic and that AA and BB are quasi-convex in HH. The following are equivalent:

  1. (1)

    GG acts acylindrically on TT.

  2. (2)

    GG does not contain any Baumslag–Solitar subgroups.

  3. (3)

    π1(𝒢)\pi_{1}(\mathcal{G}) admits no cyclically alternating and cyclically reduced word.

Proof.

We prove that (3) implies (1), that (1) implies (2) and that (2) implies (3) by proving the contrapositive statements.

Let 𝒢\mathcal{G} be the graph of cyclic stabilisers of ψ\psi and let ν:V(Γ)H\nu:V(\Gamma)\to H be a choice of representatives. Suppose that GG does not act acylindrically on TT. Then for any nn, there exists a sequence of geodesic segments S1S2SnS_{1}\subset S_{2}\subset...\subset S_{n} with Si𝒮iS_{i}\in\mathcal{S}^{i}, each with infinite stabiliser. Let giGg_{i}\in G be an element such that the endpoints of SiS_{i} are HH and giHg_{i}H. For all is(ψ)i\geq s\mathbb{Z}(\psi), by Lemma 6.4, Stab(Si)\operatorname{Stab}(S_{i}) is cyclic. By definition, each subgroup Stab(gi1Si)\operatorname{Stab}(g_{i}^{-1}\cdot S_{i}) is HH-conjugate into some ν(v)\langle\nu(v)\rangle where vV(Γ)v\in V(\Gamma). Since 𝒢\mathcal{G} is finite by Lemma 6.13, by the pigeonhole principle there are three integers s(ψ)<i<j<ks\mathbb{Z}(\psi)<i<j<k such that Stab(gi1Si)\operatorname{Stab}(g_{i}^{-1}S_{i}), Stab(gj1Si)\operatorname{Stab}(g_{j}^{-1}S_{i}) and Stab(gk1Si)\operatorname{Stab}(g_{k}^{-1}S_{i}) are HH-conjugate into some ν(v)\langle\nu(v)\rangle. Thus, there are elements h1,h2,h3,h4h_{1},h_{2},h_{3},h_{4} such that:

ν(v)gjh2ν(v)gih1\displaystyle\langle\nu(v)\rangle^{g_{j}h_{2}}\cap\langle\nu(v)\rangle^{g_{i}h_{1}} 1,\displaystyle\neq 1,
ν(v)gkh4ν(v)gih3\displaystyle\langle\nu(v)\rangle^{g_{k}h_{4}}\cap\langle\nu(v)\rangle^{g_{i}h_{3}} 1.\displaystyle\neq 1.

In particular, if f=gih1h21gj1f=g_{i}h_{1}h_{2}^{-1}g_{j}^{-1} and g=gih3h41gk1g=g_{i}h_{3}h_{4}^{-1}g_{k}^{-1}, then

ν(v)ν(v)f\displaystyle\langle\nu(v)\rangle\cap\langle\nu(v)\rangle^{f} 1,\displaystyle\neq 1,
ν(v)ν(v)g\displaystyle\langle\nu(v)\rangle\cap\langle\nu(v)\rangle^{g} 1,\displaystyle\neq 1,
ν(v)ν(v)fg\displaystyle\langle\nu(v)\rangle\cap\langle\nu(v)\rangle^{fg} 1.\displaystyle\neq 1.

Suppose that both ff and gg act elliptically on TT. If they do not both fix a common vertex, then fgfg acts hyperbolically on TT. So suppose that they both fix a common vertex uu. Now the geodesic connecting uu with SkS_{k} must meet SkS_{k} at the midpoint between giHg_{i}H and gjHg_{j}H and the midpoint between giHg_{i}H and gkHg_{k}H. Since j<kj<k, this is not possible and so we may assume that one of ff, gg or fgfg acts hyperbolically on TT. The cyclic reduction of ff, gg or fgfg provides us with a cyclically alternating and cylically reduced word in π1(𝒢)\pi_{1}(\mathcal{G}). Thus, (3) implies (1).

Now suppose that GG contains a Baumslag–Solitar subgroup J<GJ<G. Since HH is hyperbolic and so cannot contain a Baumslag–Solitar subgroup, JJ cannot act elliptically on TT. It follows from [MO15, Theorem 2.1] that JJ cannot act acylindrically on TT. Hence, GG does not act acylindrically on TT and (1) implies (2).

Finally, Lemma 6.11 shows that (2) implies (3). ∎

Example 6.15.

Consider Example 1.2 from the introduction:

Hψ=x,y,zz2yz2x2ψH*_{\psi}=\langle x,y,z\mid z^{2}yz^{2}x^{-2}\rangle*_{\psi}

where ψ:AB\psi:A\to B is given by ψ(x)=y\psi(x)=y, ψ(y)=z\psi(y)=z. We showed that HψH*_{\psi} is \mathbb{Z}-stable and that the vertex group was freely generated by x,zx,z. We see that the edge groups are generated by x,z2x2z2x,z^{-2}x^{2}z^{-2} and x2,zx^{2},z respectively. Every subgroup of the form ABgA\cap B^{g}, AAgA\cap A^{g} or BBgB\cap B^{g} is either trivial, or conjugate to AA, BB or one of the following:

x,z2x2z2x2,z\displaystyle\langle x,z^{-2}x^{2}z^{-2}\rangle\cap\langle x^{2},z\rangle =x2,z2x2z2,\displaystyle=\langle x^{2},z^{-2}x^{2}z^{-2}\rangle,
x,z2x2z2x2,zx\displaystyle\langle x,z^{-2}x^{2}z^{-2}\rangle\cap\langle x^{2},z\rangle^{x} =x2,\displaystyle=\langle x^{2}\rangle,
x2,zx2,zx\displaystyle\langle x^{2},z\rangle\cap\langle x^{2},z\rangle^{x} =x2.\displaystyle=\langle x^{2}\rangle.

By applying ψ\psi to xx and ψ1\psi^{-1} to x2=z2yz2x^{2}=z^{2}yz^{2}, we see that the vertices in Γ\Gamma corresponding to stabilisers of elements in 𝒮2\mathcal{S}^{2} are the \mathcal{B}-vertex [y]B[\langle y\rangle]_{B} and the 𝒜\mathcal{A}-vertices [y]A[\langle y\rangle]_{A} and [y2xy2]A[\langle y^{2}xy^{2}\rangle]_{A}. The graph of cyclic stabilisers can be seen in Figure 6. Since 𝒢\mathcal{G} does not admit any cyclically alternating and cyclically reduced words, by Theorem 6.14 we see that HψH*_{\psi} does not contain any Baumslag–Solitar subgroups.

Refer to caption
Figure 6. The graph of cyclic stabilisers 𝒢\mathcal{G}
Remark 6.16.

Algorithmic problems relating to this section are handled in detail in the author’s thesis [Lin22]. Combining work of Howie [How05], Stallings [Sta83] and Kharlampovich–Miasnikov–Weil [KMW17], there it is shown that given as input an inertial one-relator extension HψH*_{\psi} with the assumptions of Theorem 6.14, the \mathbb{Z}-stable number and the graph of cyclic stabilisers are computable. Thus, one can effectively decide whether HψH*_{\psi} contains a Baumslag–Solitar subgroup or not. The problem of deciding whether an arbitrary one-relator group contains a Baumslag–Solitar subgroup is still open.

7. Main results and further questions

We are finally ready to prove our main results. A one-relator tower (respectively, hierarchy) XNX1X0X_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0} is an acylindrical tower (respectively, hierarchy) if π1(Xi)\pi_{1}(X_{i}) acts acylindrically on the Bass-Serre tree associated with the one-relator splitting π1(Xi)=π1(Xi+1)ψi\pi_{1}(X_{i})=\pi_{1}(X_{i+1})*_{\psi_{i}} for all ii. It is a \mathbb{Z}-stable tower (respectively, hierarchy) if s(ψi)<s\mathbb{Z}(\psi_{i})<\infty for all ii.

Theorem 7.1.

Let XX be a one-relator complex and XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X a one-relator hierarchy. The following are equivalent:

  1. (1)

    XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X is a quasi-convex hierarchy and π1(X)\pi_{1}(X) is hyperbolic.

  2. (2)

    XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X is an acylindrical hierarchy.

  3. (3)

    XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X is a \mathbb{Z}-stable hierarchy and π1(X)\pi_{1}(X) contains no Baumslag–Solitar subgroups.

Moreover, if any of the above is satisfied, then π1(X)\pi_{1}(X) is virtually special and the image of π1(A)\pi_{1}(A) in π1(X)\pi_{1}(X) is quasi-convex for any connected subcomplex AXiA\subset X_{i}.

Proof.

The proof is by induction. The base case is clear. By the inductive hypothesis, we may assume that π1(X1)\pi_{1}(X_{1}) is hyperbolic, virtually special and that π1(A)<π1(X1)\pi_{1}(A)<\pi_{1}(X_{1}) is quasi-convex for any connected full subcomplex AXiA\subset X_{i} with i1i\geqslant 1. The equivalence between (1) and (2) now follows from Theorem 3.4. The equivalence between (2) and (3) is Theorem 6.14.

Finally, assuming π1(X)\pi_{1}(X) is hyperbolic and the hierarchy is quasi-convex, π1(X)\pi_{1}(X) is virtually special by [Wis21, Theorem 13.3] and the image of π1(A)\pi_{1}(A) in π1(X)\pi_{1}(X) is quasi-convex for any connected subcomplex AXiA\subset X_{i} by Theorem 5.9 and induction. ∎

We now prove a stronger form of [LW22, Conjecture 1.9].

Theorem 7.2.

Let XX be a one-relator complex with negative immersions. Then π1(X)\pi_{1}(X) is hyperbolic, virtually special and all of its one-relator hierarchies XNX1X0=XX_{N}\looparrowright...\looparrowright X_{1}\looparrowright X_{0}=X are quasi-convex hierarchies.

Proof.

Since π1(X)\pi_{1}(X) does not contain any Baumslag–Solitar subgroups [LW22, Corollary 1.8], the result follows from Corollary 6.9 and Theorem 7.1. ∎

Corollary 7.3.

The isomorphism problem for one-relator groups with negative immersions is decidable within the class of one-relator groups.

Proof.

Let GG be a one-relator group with negative immersions and let HH be any one-relator group. By [LW22, Theorem 1.3 & Lemma 6.4], there is an algorithm to decide whether HH has negative immersions or not. If it does not, then it is not isomorphic to GG. If it does, then both GG and HH are hyperbolic by Theorem 7.2 and so isomorphism between the two can be decided by [DG08]. ∎

We may now solve a problem of Baumslag’s [Bau86, Problem 4].

Corollary 7.4.

Parafree one-relator groups have negative immersions. In particular, they are hyperbolic, virtually special and their isomorphism problem is decidable.

Proof.

By [Bau69, Theorem 4.2] and [LW22, Theorems 1.3 & 1.5], parafree one-relator groups have negative immersions. Now the result follows from Theorems 7.2 and 7.3. ∎

Corollary 7.5.

One-relator groups with negative immersions are residually finite and linear.

Corollary 7.6.

Let XX be a one-relator complex with negative immersions. Then every finitely generated subgroup of π1(X)\pi_{1}(X) is hyperbolic.

Proof.

Follows from [Ger96, Corollary 7.8] and [LW24, Theorem A]. ∎

Extending any of these results to all one-relator groups would require a better understanding of one-relator hierarchies that are not \mathbb{Z}-stable. Therefore, we ask the following question.

Problem 7.7.

Characterise non \mathbb{Z}-stable one-relator hierarchies.

Any one-relator group satisfying the hypothesis of Brown’s criterion (see [Bro87]), either has a one-relator splitting that is not \mathbb{Z}-stable, or is isomorphic to a Baumslag–Solitar group BS(1,k)\operatorname{BS}(1,k). More generally, if π1(X)π1(X1)ψ\pi_{1}(X)\cong\pi_{1}(X_{1})*_{\psi} is a one-relator splitting and there is some gπ1(X)g\in\pi_{1}(X) acting hyperbolically on its Bass–Serre tree and some An[An]𝒜¯nψA_{n}\in[A_{n}]\in\bar{\mathcal{A}}_{n}^{\psi} such that Ang<AnA^{g}_{n}<A_{n}, then An,g\langle A_{n},g\rangle splits as an ascending HNN-extension of a finitely generated free group. We ask whether this is the only situation that can occur.

Question 7.8.

Let XX be a one-relator complex and π1(X)π1(X1)ψ\pi_{1}(X)\cong\pi_{1}(X_{1})*_{\psi} a one-relator splitting that is not \mathbb{Z}-stable. Is there some n0n\geq 0 such that one of the following holds

𝒜¯nψ\displaystyle\bar{\mathcal{A}}_{n}^{\psi} =𝒜¯n+iψ\displaystyle=\bar{\mathcal{A}}_{n+i}^{\psi}
𝒜¯nψ1\displaystyle\bar{\mathcal{A}}_{n}^{\psi^{-1}} =𝒜¯n+iψ1\displaystyle=\bar{\mathcal{A}}_{n+i}^{\psi^{-1}}

for all i0i\geq 0?

If so, does there exist an immersion of one-relator complexes ZXZ\looparrowright X that does not factor through X1XX_{1}\looparrowright X and such that π1(Z)\pi_{1}(Z) has non-empty BNS invariant?

References

  • [ABC+91] J. M. Alonso, T. Brady, D. Cooper, V. Ferlini, M. Lustig, M. Mihalik, M. Shapiro, and H. Short, Notes on word hyperbolic groups, Group theory from a geometrical viewpoint (Trieste, 1990), World Sci. Publ., River Edge, NJ, 1991, Edited by Short, pp. 3–63. MR 1170363
  • [AG99] D. J. Allcock and S. M. Gersten, A homological characterization of hyperbolic groups, Invent. Math. 135 (1999), no. 3, 723–742. MR 1669272
  • [Bau67] Gilbert Baumslag, Residually finite one-relator groups, Bull. Amer. Math. Soc. 73 (1967), 618–620. MR 212078
  • [Bau69] by same author, Groups with the same lower central sequence as a relatively free group. II. Properties, Trans. Amer. Math. Soc. 142 (1969), 507–538. MR 245653
  • [Bau86] by same author, A survey of groups with a single defining relation, Proceedings of groups—St. Andrews 1985, London Math. Soc. Lecture Note Ser., vol. 121, Cambridge Univ. Press, Cambridge, 1986, pp. 30–58. MR 896500
  • [BC06] Gilbert Baumslag and Sean Cleary, Parafree one-relator groups, J. Group Theory 9 (2006), no. 2, 191–201. MR 2220574
  • [BCH04] Gilbert Baumslag, Sean Cleary, and George Havas, Experimenting with infinite groups. I, Experiment. Math. 13 (2004), no. 4, 495–502. MR 2118274
  • [Bes04] Mladen Bestvina, Questions in geometric group theory, 2004, https://www.math.utah.edu/ bestvina/eprints/questions-updated.pdf.
  • [BF92] Mladen Bestvina and Mark Feighn, A combination theorem for negatively curved groups, J. Differential Geom. 35 (1992), no. 1, 85–101. MR 1152226
  • [BFR19] Gilbert Baumslag, Benjamin Fine, and Gerhard Rosenberger, One-relator groups: an overview, Groups St Andrews 2017 in Birmingham, London Math. Soc. Lecture Note Ser., vol. 455, Cambridge Univ. Press, Cambridge, 2019, pp. 119–157. MR 3931411
  • [BM22] Martín Axel Blufstein and Elías Minian, Strictly systolic angled complexes and hyperbolicity of one-relator groups, Algebr. Geom. Topol. 22 (2022), no. 3, 1159–1175. MR 4474781
  • [Bra99] Noel Brady, Branched coverings of cubical complexes and subgroups of hyperbolic groups, J. London Math. Soc. (2) 60 (1999), no. 2, 461–480. MR 1724853
  • [Bro87] Kenneth S. Brown, Trees, valuations, and the Bieri-Neumann-Strebel invariant, Invent. Math. 90 (1987), no. 3, 479–504. MR 914847
  • [BW13] Hadi Bigdely and Daniel T. Wise, Quasiconvexity and relatively hyperbolic groups that split, Michigan Math. J. 62 (2013), no. 2, 387–406. MR 3079269
  • [CH21] Christopher H. Cashen and Charlotte Hoffmann, Short, Highly Imprimitive Words Yield Hyperbolic One-Relator Groups, Experimental Mathematics 0 (2021), no. 0, 1–10.
  • [Che21] Haimiao Chen, Solving the isomorphism problems for two families of parafree groups, J. Algebra 585 (2021), 616–636. MR 4282650
  • [CM82] Bruce Chandler and Wilhelm Magnus, The history of combinatorial group theory, Studies in the History of Mathematics and Physical Sciences, vol. 9, Springer-Verlag, New York, 1982, A case study in the history of ideas. MR 680777
  • [Col04] Donald J. Collins, Intersections of Magnus subgroups of one-relator groups, Groups: topological, combinatorial and arithmetic aspects, London Math. Soc. Lecture Note Ser., vol. 311, Cambridge Univ. Press, Cambridge, 2004, pp. 255–296. MR 2073350
  • [Col08] by same author, Intersections of conjugates of Magnus subgroups of one-relator groups, The Zieschang Gedenkschrift, Geom. Topol. Monogr., vol. 14, Geom. Topol. Publ., Coventry, 2008, pp. 135–171. MR 2484702
  • [Dah03] François Dahmani, Combination of convergence groups, Geom. Topol. 7 (2003), 933–963. MR 2026551
  • [DG08] François Dahmani and Daniel Groves, The isomorphism problem for toral relatively hyperbolic groups, Publ. Math. Inst. Hautes Études Sci. (2008), no. 107, 211–290. MR 2434694
  • [DV96] Warren Dicks and Enric Ventura, The group fixed by a family of injective endomorphisms of a free group, Contemporary Mathematics, vol. 195, American Mathematical Society, Providence, RI, 1996. MR 1385923
  • [FRMW21] Benjamin Fine, Gerhard Rosenberger, Anja Moldenhauer, and Leonard Wienke, Topics in infinite group theory—Nielsen methods, covering spaces, and hyperbolic groups, De Gruyter STEM, De Gruyter, Berlin, [2021] ©2021. MR 4368862
  • [FRS97] Benjamin Fine, Gerhard Rosenberger, and Michael Stille, The isomorphism problem for a class of para-free groups, Proc. Edinburgh Math. Soc. (2) 40 (1997), no. 3, 541–549. MR 1475915
  • [Ger92] Stephen M. Gersten, Problems on automatic groups, Algorithms and classification in combinatorial group theory (Berkeley, CA, 1989), Math. Sci. Res. Inst. Publ., vol. 23, Springer, New York, 1992, pp. 225–232. MR 1230636
  • [Ger96] by same author, Subgroups of word hyperbolic groups in dimension 22, J. London Math. Soc. (2) 54 (1996), no. 2, 261–283. MR 1405055
  • [GKL21] Giles Gardam, Dawid Kielak, and Alan D. Logan, Algebraically hyperbolic groups, 2021, arXiv:2112.01331.
  • [GMRS98] Rita Gitik, Mahan Mitra, Eliyahu Rips, and Michah Sageev, Widths of subgroups, Transactions of the American Mathematical Society 350 (1998), no. 1, 321–329.
  • [Gro87] Mikhael Gromov, Hyperbolic groups, Essays in group theory, Math. Sci. Res. Inst. Publ., vol. 8, Springer, New York, 1987, pp. 75–263. MR 919829
  • [Gro93] M. Gromov, Asymptotic invariants of infinite groups, Geometric group theory, Vol. 2 (Sussex, 1991), London Math. Soc. Lecture Note Ser., vol. 182, Cambridge Univ. Press, Cambridge, 1993, pp. 1–295. MR 1253544
  • [GW19] Giles Gardam and Daniel J. Woodhouse, The geometry of one-relator groups satisfying a polynomial isoperimetric inequality, Proc. Amer. Math. Soc. 147 (2019), no. 1, 125–129. MR 3876736
  • [Hat02] Allen Hatcher, Algebraic topology, Cambridge University Press, Cambridge, 2002. MR 1867354
  • [HK17] Do Viet Hung and Vu The Khoi, Applications of the Alexander ideals to the isomorphism problem for families of groups, Proc. Edinb. Math. Soc. (2) 60 (2017), no. 1, 177–185. MR 3589847
  • [HK20] by same author, Twisted Alexander ideals and the isomorphism problem for a family of parafree groups, Proc. Edinb. Math. Soc. (2) 63 (2020), no. 3, 780–806. MR 4163871
  • [How81] James Howie, On pairs of 22-complexes and systems of equations over groups, J. Reine Angew. Math. 324 (1981), 165–174. MR 614523
  • [How87] by same author, How to generalize one-relator group theory, Combinatorial group theory and topology (Alta, Utah, 1984), Ann. of Math. Stud., vol. 111, Princeton Univ. Press, Princeton, NJ, 1987, pp. 53–78. MR 895609
  • [How05] by same author, Magnus intersections in one-relator products, Michigan Math. J. 53 (2005), no. 3, 597–623. MR 2207211
  • [HW01] Geoffrey Christopher Hruska and Daniel T. Wise, Towers, ladders and the B. B. Newman spelling theorem, J. Aust. Math. Soc. 71 (2001), no. 1, 53–69. MR 1840493
  • [HW16] Joseph Helfer and Daniel T. Wise, Counting cycles in labeled graphs: the nonpositive immersion property for one-relator groups, Int. Math. Res. Not. IMRN (2016), no. 9, 2813–2827. MR 3519130
  • [IMM23] Giovanni Italiano, Bruno Martelli, and Matteo Migliorini, Hyperbolic 5-manifolds that fiber over S1S^{1}, Invent. Math. 231 (2023), no. 1, 1–38. MR 4526820
  • [IS98] Sergei V. Ivanov and Paul E. Schupp, On the hyperbolicity of small cancellation groups and one-relator groups, Trans. Amer. Math. Soc. 350 (1998), no. 5, 1851–1894. MR 1401522
  • [Iva18] Sergei V. Ivanov, The intersection of subgroups in free groups and linear programming, Math. Ann. 370 (2018), no. 3-4, 1909–1940. MR 3770185
  • [Kap00] Ilya Kapovich, Mapping tori of endomorphisms of free groups, Comm. Algebra 28 (2000), no. 6, 2895–2917. MR 1757436
  • [Kap01] by same author, The combination theorem and quasiconvexity, Internat. J. Algebra Comput. 11 (2001), no. 2, 185–216. MR 1829050
  • [KMS60] Abraham Karrass, Wilhelm Magnus, and Donald Solitar, Elements of finite order in groups with a single defining relation, Comm. Pure Appl. Math. 13 (1960), 57–66. MR 124384
  • [KMW17] Olga Kharlampovich, Alexei Miasnikov, and Pascal Weil, Stallings graphs for quasi-convex subgroups, J. Algebra 488 (2017), 442–483. MR 3680926
  • [Lin22] Marco Linton, One-relator hierarchies, Ph.D. thesis, University of Warwick, UK, 2022.
  • [LL94] Robert H. Lewis and Sal Liriano, Isomorphism classes and derived series of certain almost-free groups, Experiment. Math. 3 (1994), no. 3, 255–258. MR 1329373
  • [LS01] Roger C. Lyndon and Paul E. Schupp, Combinatorial group theory, Classics in Mathematics, Springer-Verlag, Berlin, 2001, Reprint of the 1977 edition. MR 1812024
  • [LW17] Larsen Louder and Henry Wilton, Stackings and the WW-cycles conjecture, Canad. Math. Bull. 60 (2017), no. 3, 604–612. MR 3679733
  • [LW22] by same author, Negative immersions for one-relator groups, Duke Math. J. 171 (2022), no. 3, 547–594. MR 4382976
  • [LW24] by same author, Uniform negative immersions and the coherence of one-relator groups, Invent. Math. 236 (2024), no. 2, 673–712. MR 4728240
  • [Lyn50] Roger C. Lyndon, Cohomology theory of groups with a single defining relation, Ann. of Math. (2) 52 (1950), 650–665. MR 47046
  • [Mag30] Wilhelm Magnus, Über diskontinuierliche Gruppen mit einer definierenden Relation. (Der Freiheitssatz), J. Reine Angew. Math. 163 (1930), 141–165. MR 1581238
  • [Mas06] Joseph D. Masters, Heegaard splittings and 1-relator groups, unpublished paper, 2006.
  • [MKS66] Wilhelm Magnus, Abraham Karrass, and Donald Solitar, Combinatorial group theory: Presentations of groups in terms of generators and relations, Interscience Publishers [John Wiley & Sons], New York-London-Sydney, 1966. MR 0207802
  • [MO15] Ashot Minasyan and Denis Osin, Acylindrical hyperbolicity of groups acting on trees, Math. Ann. 362 (2015), no. 3-4, 1055–1105. MR 3368093
  • [Mol67] D. I. Moldavanskiĭ, Certain subgroups of groups with one defining relation, Sibirsk. Mat. Ž. 8 (1967), 1370–1384. MR 0220810
  • [Mut21a] Jean Pierre Mutanguha, Constructing stable images, preprint on webpage at mutanguha.com/pdfs/relimmalgo.pdf, 2021.
  • [Mut21b] Jean Pierre Mutanguha, The dynamics and geometry of free group endomorphisms, Adv. Math. 384 (2021), Paper No. 107714, 60. MR 4237417
  • [MUW11] Alexei Myasnikov, Alexander Ushakov, and Dong Wook Won, The word problem in the Baumslag group with a non-elementary Dehn function is polynomial time decidable, J. Algebra 345 (2011), 324–342. MR 2842068
  • [New68] B. B. Newman, Some results on one-relator groups, Bull. Amer. Math. Soc. 74 (1968), 568–571. MR 222152
  • [Ros13] Amnon Rosenmann, On the intersection of subgroups in free groups: echelon subgroups are inert, Groups Complex. Cryptol. 5 (2013), no. 2, 211–221. MR 3245107
  • [Sel97] Z. Sela, Acylindrical accessibility for groups, Invent. Math. 129 (1997), no. 3, 527–565. MR 1465334
  • [Ser03] 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
  • [Sta83] John R. Stallings, Topology of finite graphs, Invent. Math. 71 (1983), no. 3, 551–565. MR 695906
  • [Tar92] Gábor Tardos, On the intersection of subgroups of a free group, Invent. Math. 108 (1992), no. 1, 29–36. MR 1156384
  • [Wis03] Daniel T. Wise, Nonpositive immersions, sectional curvature, and subgroup properties, Electron. Res. Announc. Amer. Math. Soc. 9 (2003), 1–9. MR 1988866
  • [Wis04] D. T. Wise, Sectional curvature, compact cores, and local quasiconvexity, Geom. Funct. Anal. 14 (2004), no. 2, 433–468. MR 2062762
  • [Wis05] Daniel T. Wise, Approximating flats by periodic flats in CAT(0) square complexes, Canad. J. Math. 57 (2005), no. 2, 416–448. MR 2124924
  • [Wis14] by same author, Cubular tubular groups, Trans. Amer. Math. Soc. 366 (2014), no. 10, 5503–5521. MR 3240932
  • [Wis20] by same author, An invitation to coherent groups, What’s next?—the mathematical legacy of William P. Thurston, Ann. of Math. Stud., vol. 205, Princeton Univ. Press, Princeton, NJ, 2020, pp. 326–414. MR 4205645
  • [Wis21] by same author, The structure of groups with a quasiconvex hierarchy, Annals of Mathematics Studies, vol. 209, Princeton University Press, Princeton, NJ, [2021] ©2021. MR 4298722