Hostname: page-component-54dcc4c588-trf7k Total loading time: 0 Render date: 2025-09-21T15:59:13.533Z Has data issue: false hasContentIssue false

Rigidity phenomena and the statistical properties of group actions on $\text {CAT}(0)$ cube complexes

Published online by Cambridge University Press:  19 September 2025

Stephen Cantrell
Affiliation:
Department of Mathematics, University of Warwick https://ror.org/01a77tt86 , Coventry, CV4 7AL, UK; E-mail: stephen.cantrell@warwick.co.uk
Eduardo Reyes*
Affiliation:
Facultad de Matemáticas, Pontificia Universidad Católica de Chile (PUC), Avenida Vicuña Mackenna 4860, Santiago, Chile
*
E-mail: eduardoreyes@uc.cl (corresponding author)

Abstract

We compare the marked length spectra of some pairs of proper and cocompact cubical actions of a nonvirtually cyclic group on $\mathrm {CAT}(0)$ cube complexes. The cubulations are required to be virtually co-special, have the same sets of convex-cocompact subgroups, and admit a contracting element. There are many groups for which these conditions are always fulfilled for any pair of cubulations, including nonelementary cubulable hyperbolic groups, many cubulable relatively hyperbolic groups, and many right-angled Artin and Coxeter groups.

For these pairs of cubulations, we study the Manhattan curve associated to their combinatorial metrics. We prove that this curve is analytic and convex, and a straight line if and only if the marked length spectra are homothetic. The same result holds if we consider invariant combinatorial metrics in which the lengths of the edges are not necessarily one. In addition, for their standard combinatorial metrics, we prove a large deviations theorem with shrinking intervals for their marked length spectra. We deduce the same result for pairs of word metrics on hyperbolic groups.

The main tool is the construction of a finite-state automaton that simultaneously encodes the marked length spectra of both cubulations in a coherent way, in analogy with results about (bi)combable functions on hyperbolic groups by Calegari-Fujiwara [14]. The existence of this automaton allows us to apply the machinery of thermodynamic formalism for suspension flows over subshifts of finite type, from which we deduce our results.

Information

Type
Dynamics
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1 Introduction

In this work we study rigidity phenomena and the statistical properties of group actions on $\mathrm {CAT}(0)$ cube complexes and the methods we use exploit the interplay between geometric group theory and dynamics. Group actions on $\mathrm {CAT}(0)$ cube complexes are nowadays a central object of study. Since the influential work of Sageev [Reference Sageev75], we have known that many groups admit proper and cubical actions on $\mathrm {CAT}(0)$ cube complexes (if in addition the actions are cocompact, in the sequel they are referred to as cubulations). The list includes small cancellation groups [Reference Martin and Steenbock60, Reference Wise85], many 3-manifold groups [Reference Bergeron and Wise3, Reference Hagen and Przytycki43, Reference Przytycki and Wise69, Reference Przytycki and Wise70, Reference Tidmore84], Coxeter groups [Reference Niblo and Reeves62], many Artin groups [Reference Charney and Davis21, Reference Godelle and Paris40], random groups at low density [Reference Odrzygóźdź63, Reference Ollivier and Wise64], 1-relator groups with torsion [Reference Lauer and Wise57, Reference Stucky83, Reference Wise86], hyperbolic free-by-cyclic groups [Reference Hagen and Wise44, Reference Hagen and Wise45], and so on. In particular, the fundamental groups of (compact) special cube complexes introduced by Haglund and Wise [Reference Haglund and Wise47] form a very rich class of convex-cocompact subgroups of right-angled Artin groups, and they played a key role in the resolution of the Virtual Haken and Virtual Fibering Conjectures [Reference Agol1, Reference Wise86].

In general, when nonempty, the space of geometric actions of a given group on $\mathrm {CAT}(0)$ cube complexes is quite large. For example, each filling multicurve on a closed hyperbolic surface is dual to a cubulation of its fundamental group [Reference Sageev75]. Similarly, cubulations for fundamental groups of cusped hyperbolic 3-manifolds can be obtained from their vast sets of (relatively) quasiconvex surface subgroups [Reference Bergeron and Wise3, Reference Cooper and Futer24, Reference Kahn and Markovic52]. Moreover, cubulations can be used to define deformation spaces, such as the classical Culler-Vogtmann outer space [Reference Culler and Vogtmann25] that encodes geometric actions of free groups on trees. This perspective has been extended to right-angled Artin groups, for which outer spaces have been constructed using cubulations with some particular special cube complexes as quotients [Reference Bregman, Charney and Vogtmann8, Reference Charney, Stambaugh and Vogtmann22].

Under some reasonable irreducibility assumptions, actions on $\mathrm {CAT}(0)$ cube complexes are marked length-spectrum rigid [Reference Beyrer and Fioravanti4, Reference Beyrer and Fioravanti5]. More precisely, let $\mathcal {X}$ be a cubulation of a group $\Gamma $ and let $\mathbf {conj}=\mathbf {conj}(\Gamma )$ denote the set of conjugacy classes of $\Gamma $ . The (stable) translation length of this action is the function $\ell _{\mathcal {X}}:\mathbf {conj} \rightarrow {\mathbb {R}}$ given by

$$\begin{align*}\ell_{\mathcal{X}}[g]=\lim_{n\to \infty}{\frac{d_{\mathcal{X}}(g^n x,x)}{n}},\end{align*}$$

where $d_{\mathcal {X}}$ denotes the combinatorial metric on the 1-skeleton of $\mathcal {X}$ and the limit above is independent of the representative $g\in [g]$ and the vertex $x\in \mathcal {X}$ .

For two cubulations $\mathcal {X},\mathcal {X}_\ast $ of $\Gamma $ , marked length-spectrum rigidity states that the equality of translation length functions $\ell _{\mathcal {X}}=\ell _{\mathcal {X}_\ast }$ implies the existence of a $\Gamma $ -equivariant cubical isometry from $\mathcal {X}$ onto $\mathcal {X}_\ast $ . Since in general the translation length functions $\ell _{\mathcal {X}}$ and $\ell _{\mathcal {X}_\ast }$ will not coincide, it is natural to ask about the behavior of $\ell _{\mathcal {X}_\ast }[g]$ when $\ell _{\mathcal {X}}[g]$ is large. The goal of this paper is to address this question for “compatible” pairs of virtually co-special cubulations, that is, those having quotients with a special cube complex as a finite cover. Such compatibility is described in Definition 5.1, and is guaranteed for any group in the following class.

Definition 1.1. Let $\mathfrak {G}$ be the class of nonvirtually cyclic groups $\Gamma $ satisfying the following:

  1. (1) $\Gamma $ admits a proper, cocompact and virtually co-special action on a $\mathrm {CAT}(0)$ cube complex.

  2. (2) The class of convex-cocompact subgroups of $\Gamma $ is the same with respect to any proper and cocompact action on a $\mathrm {CAT}(0)$ cube complex. That is, given any two proper, cocompact actions of $\Gamma $ on $\text {CAT}(0)$ cube complexes $\mathcal {X}, \mathcal {X}_\ast $ then the restricted action of a subgroup $H < \Gamma $ on $\mathcal {X}$ is convex-cocompact if and only if the restricted action of H on $\mathcal {X}_\ast $ is convex-cocompact.

  3. (3) Some (equivalently, any) proper and cocompact action of $\Gamma $ on a $\mathrm {CAT}(0)$ cube complex has a contracting element.

By [Reference Genevois39, Lemma 4.6], contracting elements for proper and cocompact actions on $\mathrm {CAT}(0)$ cube complexes are those having invariant geodesics that satisfy the conclusion of the Morse lemma. In particular, the notion of being a contracting element is independent of the cubulation of $\Gamma $ .

By Agol’s theorem [Reference Agol1, Theorem 1.1] and the characterization of convex-cocompact subgroups in terms of quasiconvexity [Reference Haglund and Wise47, Proposition 7.2], we see that every cubulable nonelementary hyperbolic group belongs to $\mathfrak {G}$ . Moreover, the class $\mathfrak {G}$ is closed under relative hyperbolicity, in the sense that a cubulable relatively hyperbolic group belongs to $\mathfrak {G}$ as long as its peripheral subgroups belong to $\mathfrak {G}$ . In particular, any $C'(1/6)$ small cancellation quotient of the free product of finitely many groups in $\mathfrak {G}$ belongs to $\mathfrak {G}$ [Reference Martin and Steenbock60]. However, this class is much larger since it contains some infinite families of right-angled Artin and Coxeter groups, most of them not being relatively hyperbolic with respect to any collection of proper subgroups. For instance, any right-angled Artin group with finite outer automorphism group belongs to $\mathfrak {G}$ . See Proposition 5.2 for the precise statement.

1.1 Manhattan curves

Let $\mathcal {X},\mathcal {X}_\ast $ be two cubulations of the group $\Gamma $ . We endow these cubulations with $\Gamma $ -invariant orthotope structures $\mathfrak {w},\mathfrak {w}_\ast $ respectively, consisting of (non-necessarily integer) positive lengths assigned to the hyperplanes which are invariant under the action of $\Gamma $ . This induces isometric actions of $\Gamma $ on the cuboid complexes $\mathcal {X}^{\mathfrak {w}}=(\mathcal {X},\mathfrak {w})$ and $\mathcal {X}_\ast ^{\mathfrak {w}_\ast }=(\mathcal {X}_\ast ,\mathfrak {w}_\ast )$ , see Subsection 2.2 for further details. The Manhattan curve for the pair $(\mathcal {X}^{\mathfrak {w}}, \mathcal {X}^{\mathfrak {w}_\ast }_\ast )$ is the boundary of the convex set

$$\begin{align*}\mathcal{C}_{\mathcal{X}^{\mathfrak{w}_\ast}_\ast/\mathcal{X}^{\mathfrak{w}}}:=\left\{ (a,b) \in {\mathbb{R}}^2 : \sum_{[g] \in \mathbf{conj}(\Gamma)} e^{-a\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g] - b\ell^{\mathfrak{w}}_{\mathcal{X}}[g]} < \infty \right\}, \end{align*}$$

where $\ell _{\mathcal {X}}^{\mathfrak {w}}$ and $\ell ^{\mathfrak {w}_\ast }_{\mathcal {X}_\ast }$ are the respective translation length functions of the actions of $\Gamma $ on the 1-skeleta of $\mathcal {X}^{\mathfrak {w}}$ and $\mathcal {X}_\ast ^{\mathfrak {w}_\ast }$ . We can parameterize this curve as $s \mapsto \theta _{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /{\mathcal {X}^{\mathfrak {w}}}}(s)$ , where for $s \in {\mathbb {R}}$ , $\theta _{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /{\mathcal {X}^{\mathfrak {w}}}}(s)$ is the abscissa of convergence of the series

$$\begin{align*}t \mapsto \sum_{[g] \in \mathbf{conj}(\Gamma)} e^{-t\ell^{\mathfrak{w}}_{\mathcal{X}}[g] - s\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]}. \end{align*}$$

By abuse of notation, we also call the parametrization $\theta _{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}$ the Manhattan curve of $(\mathcal {X}^{\mathfrak {w}},\mathcal {X}^{\mathfrak {w}_\ast }_\ast )$ .

Manhattan curves are useful tools for studying pairs of actions and are related to rigidity results, as well as recovering asymptotic invariants. They were introduced by Burger [Reference Burger12] for pairs of convex-cocompact representations of a group on rank 1 symmetric spaces, and later Sharp [Reference Sharp81] proved that they are analytic for pairs of cocompact Fuchsian representations. Sharp also extended these results for pairs of points in the outer space of a free group [Reference Sharp79, Reference Sharp80]. Recently, Manhattan curves have been studied for pairs of cusped Fuchsian representations [Reference Kao54, Reference Kao55], pairs of cusped quasi-Fuchsian representations [Reference Bray, Canary and Kao6], comparing quasi-Fuchsian representations with negatively curved metrics on surfaces [Reference Kao53], pairs of cusped Hitchin representations [Reference Bray, Canary, Kao and Martone7], and pairs of geometric actions on hyperbolic groups [Reference Cantrell and18, Reference Cantrell and Tanaka19].

In some sense, the Manhattan curve can be seen as a function that ‘interpolates’ between its pair of defining isometric actions. In particular, the regularity (i.e., differentiability) properties of the Manhattan curve somehow measure the “compatibility” of such actions. Moreover, when Manhattan curves are known to be analytic, then (as they are convex) they are either straight lines or strictly convex everywhere. This convexity characterization leads to length spectrum rigidity and other rigidity results for pairs of actions. See [Reference Cantrell and Tanaka19, Theorem 1] for some examples of such results. In addition, when Manhattan curves are known to be analytic, we immediately obtain precise large deviations principles comparing isometric actions. We consider such results in this work. Lastly, the $C^2$ -regularity of Manhattan curves can be use to construct pressure metrics, which recover the Weil-Petersson metric on Teichmüller space [Reference McMullen61] and generalize it to other geometric settings [Reference Aougab, Clay and Rieck2, Reference Bray, Canary and Kao6, Reference Bridgeman, Canary, Labourie and Sambarino9, Reference Kao55, Reference Pollicott and Sharp67].

Our first main theorem fits into the aforementioned results, and to the authors’ knowledge, is the first to address the analyticity of Manhattan curves outside the scope of relatively hyperbolic groups or representation theory.

Theorem 1.2. Let $\Gamma $ be a group in the class $\mathfrak {G}$ and let it act properly and cocompactly on the cuboid complexes $\mathcal {X}^{\mathfrak {w}}=(\mathcal {X},\mathfrak {w})$ and $\mathcal {X}_\ast ^{\mathfrak {w}_\ast }=(\mathcal {X}_\ast ,\mathfrak {w}_\ast )$ . Then the Manhattan curve $\theta _{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}:{\mathbb {R}} \rightarrow {\mathbb {R}}$ is convex, decreasing, and analytic. In addition, the following limit exists and equals $-\theta ^{\prime }_{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}(0)$ :

$$\begin{align*}\tau(\mathcal{X}^{\mathfrak{w}_\ast}_\ast/\mathcal{X}^{\mathfrak{w}}):= \lim_{T\to\infty} \frac{1}{\#\{[g]\in \mathbf{conj} \colon \ell^{\mathfrak{w}}_{\mathcal{X}}[g]< T\}} \sum_{\ell^{\mathfrak{w}}_{\mathcal{X}}[g]< T} \frac{\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]}{T}. \end{align*}$$

Moreover, we always have

$$\begin{align*}\tau(\mathcal{X}^{\mathfrak{w}_\ast}_\ast/\mathcal{X}^{\mathfrak{w}})\geq v_{\mathcal{X}^{\mathfrak{w}}}/v_{\mathcal{X}^{\mathfrak{w}_\ast}_\ast},\end{align*}$$

for $v_{\mathcal {X}^{\mathfrak {w}}},v_{\mathcal {X}^{\mathfrak {w}_\ast }_\ast }$ the corresponding exponential growth rates, and the following are equivalent:

  1. (1) $\theta _{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}$ is a straight line;

  2. (2) there exists $\Lambda>0$ such that $\ell ^{\mathfrak {w}}_{\mathcal {X}}[g] = \Lambda \ell ^{\mathfrak {w}_\ast }_{\mathcal {X}_\ast }[g]$ for all $[g] \in \mathbf {conj}(\Gamma )$ ; and

  3. (3) $\tau (\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}) = v_{\mathcal {X}^{\mathfrak {w}}}/v_{\mathcal {X}^{\mathfrak {w}_\ast }_\ast }$ .

Remark 1.3. In the result above, the group $\Gamma $ does not have to belong to $\mathfrak {G}$ as long as the triplet $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )$ belongs to the class $\mathfrak {X}$ in Definition 5.1. In particular, the action on $\mathcal {X}_\ast $ does not have to be proper. See Theorem 6.1 for the more general statement.

1.2 An automaton for pairs of cubulations

The main tool in the proof of Theorems 1.2 and 6.1 is the construction of a finite-state automaton that simultaneously encodes translation lengths for the actions on both $\mathcal {X}$ and $\mathcal {X}_\ast $ . Roughly speaking, an automaton is a finite directed graph $\mathcal {G}$ that encodes a group $\Gamma $ equipped with a finite generating set S. The edges of $\mathcal {G}$ are labeled by elements of S, so that finite paths in $\mathcal {G}$ correspond to group elements whose word length (with respect to S) equals the length of the corresponding path. Well-known examples include the Bowen-Series coding for Fuchsian groups [Reference Bowen and Series13], Cannon’s automatic structure for hyperbolic groups with arbitrary generating set [Reference Cannon15] (see Example 2.4) and Hermiller-Meier’s automatic structures for right-angled Artin and Coxeter groups with the standard generating sets [Reference Hermiller and Meier49]. Recently, the combinatorial structure of automata has been key to deduce strong counting results for some groups acting isometrically on $\delta $ -hyperbolic spaces. Examples include genericity of loxodromic elements [Reference Gekhtman, Taylor and Tiozzo37, Reference Gekhtman, Taylor and Tiozzo38] and central limit theorems [Reference Gekhtman, Taylor and Tiozzo36].

In principle, an automatic structure encodes a single length function associated to a group. However, for a hyperbolic group $\Gamma $ with two input generating sets $S,S_\ast $ , it is possible to enhance the automaton associated to S and equipping it with an extra (non-negative) integer-valued edge labeling. With respect to the new label, paths in the refined automaton record the word length with respect to $S_\ast $ of the group element associated to the path. This was achieved by Calegari and Fujiwara in [Reference Calegari and Fujiwara14], and having access to an automatic structure and labeling such as this means that one can apply powerful tools and techniques from thermodynamic formalism and symbolic dynamics to study pairs word metrics on hyperbolic groups. For example, this construction was used by Cantrell and Tanaka [Reference Cantrell and18] to show that Manhattan curves for pairs of word metrics on hyperbolic groups are analytic. Calegari-Fujiwara’s construction was the main inspiration for the construction of the automaton for pairs of actions on $\mathrm {CAT}(0)$ cube complexes, which we now proceed to describe.

In our setting, we start with a group $\Gamma $ in the class $\mathfrak {G}$ acting properly and cocompactly on the $\mathrm {CAT}(0)$ cube complexes $\mathcal {X}$ and $\mathcal {X}_\ast $ . The main result is Theorem 5.11, which is stated using the formalism of automatic structures (see Subsection 2.3). As the statement of this theorem is rather technical, we provide a simplified version that also incorporates Lemma 6.7, Lemma 6.11, and Lemma 6.12.

Theorem 1.4. Let $\Gamma $ be a group in the class $\mathfrak {G}$ and let it act properly and cocompactly on the $\mathrm {CAT}(0)$ cube complexes $\mathcal {X}$ and $\mathcal {X}_\ast $ . To the triplet $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )$ we associate the following data:

  1. i) A finite index subgroup $\overline {\Gamma } <\Gamma $ such that the quotient $\overline {\mathcal {X}}=\overline {\Gamma } \backslash \mathcal {X}$ is a special cube complex.

  2. ii) A finite directed graph $\mathcal {G}=\mathcal {G}(\Gamma ,\mathcal {X},\mathcal {X}_\ast )$ equipped with a labeling map $\pi $ that assigns to each edge of $\mathcal {G}$ an oriented hyperplane of $\overline {\mathcal {X}}$ .

  3. iii) An integer-valued functional $\psi $ on the edges of $\mathcal {G}$ .

From this data, any closed loop $\omega $ in $\mathcal {G}$ is assigned to a closed (combinatorial) geodesic $\overline {\gamma }_\omega $ in $\overline {\mathcal {X}}$ , and hence to a conjugacy class $[g_\omega ]\in \mathbf { conj}(\overline {\Gamma })$ , in such a way that:

  1. (1) If $\omega $ is determined by the sequence of edges $e_1,\dots ,e_n$ in $\mathcal {G}$ , then the loop $\overline {\gamma }_{\omega }$ consists of edges that are dual to the hyperplanes $\pi (e_1),\dots ,\pi (e_n)$ . In particular, the length of $\omega $ equals $\ell _{\mathcal {X}}[g_\omega ]$ .

  2. (2) If $\omega $ is as in (1), then $\psi (e_1)+\cdots +\psi (e_n)=\ell _{\mathcal {X}_\ast }[g_\omega ]$ .

  3. (3) The assignment $\omega \mapsto [g_\omega ]$ from the set of closed loops of $\mathcal {G}$ into $\mathbf {conj}(\overline {\Gamma })$ is

    • polynomial (in length)-to-one; and

    • has image with positive lower density with respect to the action of $\overline {\Gamma }$ on $\mathcal {X}$ .

We briefly sketch how Theorem 1.4 implies Theorem 1.2 in the case that $\mathcal {X}^{\mathfrak {w}}=\mathcal {X}$ and $\mathcal {X}_\ast ^{\mathfrak {w}_\ast }=\mathcal {X}_\ast $ . From the adjacency matrix of $\mathcal {G}$ we define a subshift of finite-type $(\Sigma ,\sigma )$ , whose periodic orbits correspond to loops in $\mathcal {G}$ and hence induce conjugacy classes in $\mathbf {conj}(\overline {\Gamma })$ . The periods of these periodic orbits correspond to $\ell _{\mathcal {X}}$ -translation lengths. The functional $\psi $ induces a potential $\Phi :\Sigma \rightarrow {\mathbb {Z}}$ that is constant on 2-cylinders, and whose Birkhoff sums correspond to $\ell _{\mathcal {X}_\ast }$ -translation lengths. Item (3) then allows us to describe the Manhattan curve $\theta _{\mathcal {X}_\ast /\mathcal {X}}$ in terms of pressure functions associated with $\Phi $ (see Proposition 6.8), and the theorems then follow by standard results in thermodynamic formalism.

1.3 Large deviations

For a pair $\mathcal {X},\mathcal {X}_\ast $ of cubulations of a group $\Gamma \in \mathfrak {G}$ , we also study large deviations for their translation lengths.

That is, we estimate the number of conjugacy classes $[g]$ for which

(1.1) $$ \begin{align} \left| \frac{\ell_{\mathcal{X}_\ast}[g]}{\ell_{\mathcal{X}}[g]} - \eta \right| < \epsilon \ \text{ for some }\eta \in {\mathbb{R}}\text{ and a small }\epsilon>0. \end{align} $$

We also study the set of conjugacy classes $[g]$ such that

(1.2) $$ \begin{align} |\ell_{\mathcal{X}_\ast}[g] - \eta \ell_{\mathcal{X}}[g]| < \epsilon \ \text{ for some }\eta \in {\mathbb{R}}\text{ and a small }\epsilon>0. \end{align} $$

This latter comparison is more delicate than the corresponding quotient comparison in (1.1) above, as (when $\ell _{\mathcal {X}}[g]$ is bounded away from $0$ ) (1.2) implies (1.1) but not vice versa.

Both of these equations, (1.1) and (1.2), represent the conjugacy classes $[g]$ for which $\ell _{\mathcal {X}_\ast }[g]$ is approximately $\eta \ell _{\mathcal {X}}[g]$ . Therefore, understanding the growth rate/number of conjugacy classes satisfying these inequalities allows us to form a natural comparison between the actions of $\Gamma $ on $\mathcal {X}$ and $\mathcal {X}_\ast $ . These growth rates are also related to rigidity phenomena. See, for example, [Reference Cantrell and Reyes17, Theorem 1.4] that classifies word metrics on hyperbolic groups in terms of growth rates coming from such large deviations estimates. In fact, this current work was motivated by [Reference Cantrell and Reyes17, Theorem 4.1] and [Reference Cantrell and Reyes17, Remark 4.3] as we now explain. The result [Reference Cantrell and Reyes17, Theorem 4.1] shows that there is a precise large deviations theorem that compares certain ‘compatible’ word metrics on hyperbolic groups. Here ‘compatible’ is a condition regarding the exponential growth rates of the word metrics. In [Reference Cantrell and Reyes17, Remark 4.3] the authors asked if this compatibility condition was necessary. In this work we show that it is not and that we can deduce precise large deviations results for all pairs of word metrics on hyperbolic groups. We present and discuss these results below in Subsection 1.4.

Despite the fact that estimating the number of conjugacy classes satisfying (1.2) is significantly harder than studying the analogous question for (1.1), there are previous works that tackle this problem in other settings. For example, let $\Sigma $ be a closed surface with negative Euler characteristic and fundamental group $\Gamma $ , and suppose that $\mathfrak {g}$ and $\mathfrak {g}_\ast $ are two hyperbolic metrics on $\Sigma $ . These metrics induce isometric actions of $\Gamma $ on $\widetilde {\Sigma }$ with translation length functions $\ell _{\mathfrak {g}}$ and $\ell _{\mathfrak {g}_\ast }$ . A result of Schwartz and Sharp [Reference Schwartz and Sharp78] states that there is an interval $(\alpha , \beta ) \subset {\mathbb {R}}$ and constants $C, \lambda>0$ such that any $\eta \in (\alpha , \beta )$ satisfies

$$\begin{align*}\#\left\{ [g] \in \mathbf{conj}(\Gamma): \ell_{\mathfrak{g}}[g] < T : |\ell_{\mathfrak{g}_\ast}[g] - \eta \ell_{\mathfrak{g}}[g] | <\epsilon \right\} \sim \frac{Ce^{\lambda T}}{T^{3/2}} \end{align*}$$

as $T\to \infty $ for any fixed $\epsilon>0$ (here ‘ $\sim $ ’ represents that the quotients of the two quantities converge to $1$ as $T\to \infty $ ). Similar results are known to hold for surfaces of variable negative curvature by Dal’bo [Reference Dal’bo28], for Hitchin representations by Dai and Martone [Reference Dai and Martone27], for Green metrics by Cantrell [Reference Cantrell16] and for some pairs of points in outer space by Sharp [Reference Sharp79]. These asymptotics are often referred to as correlation results, and to prove them thermodynamic formalism is usually employed. To apply thermodynamic formalism one needs to know that the length spectra of the two considered metrics are not rationally related. That is, if $\ell _1$ , $\ell _2$ are the length spectra that we want to compare then we would need to know that there do not exist nonzero $a,b \in {\mathbb {R}}$ with $a\ell _1[g] + b\ell _2[g] \in {\mathbb {Z}}$ for all $[g] \in \mathbf {conj}(\Gamma ).$ This property is vital as it implies bounds on the operator norm of families of transfer operators, which are then used in the proof of the correlation asymptotic.

On the other hand, the length spectra of a pair of cubical actions on $\mathrm {CAT}(0)$ cube complexes are always rationally related. Indeed, after possibly performing one cubical barycentric subdivision, every cubical isometry of a $\mathrm {CAT}(0)$ cube complex either fixes a vertex or preserves a bi-infinite geodesic on which the isometry acts by nontrivial translations [Reference Haglund46]. In particular, the translation length function associated to any of these actions has image belonging to $\frac {1}{2}{\mathbb {Z}}$ .

However, by means of the automaton from Theorem 5.11 (Theorem 1.4), we are still able to estimate the number of conjugacy classes satisfying (1.2). As a consequence of Theorem 3.2 in the setting of subshifts of finite type, we can prove the following result that can be seen as large deviations with shrinking intervals.

Theorem 1.5. Let $\Gamma $ be a group in the class $\mathfrak {G}$ and let it act properly and cocompactly on the $\mathrm {CAT}(0)$ cube complexes $\mathcal {X}$ and $\mathcal {X}_\ast $ . Let $\mathbf {conj}'\subset \mathbf {conj}$ be the set of nontorsion conjugacy classes and consider the dilations

$$\begin{align*}\mathrm{Dil}(\mathcal{X}_\ast,\mathcal{X}) = \sup_{[g] \in \mathbf{conj}'} \frac{\ell_{\mathcal{X}_\ast}[g]}{\ell_{\mathcal{X}}[g]} \ \text{ and } \ \mathrm{Dil}(\mathcal{X},\mathcal{X}_\ast)^{-1} = \inf_{[g] \in \mathbf{conj}'} \frac{\ell_{\mathcal{X}_\ast}[g]}{\ell_{\mathcal{X}}[g]}. \end{align*}$$

Then there exists an analytic function

$$ \begin{align*}\mathcal{I}:[\mathrm{Dil}(\mathcal{X},\mathcal{X}_\ast)^{-1}, \mathrm{Dil}(\mathcal{X}_\ast,\mathcal{X})]\rightarrow {\mathbb{R}}\end{align*} $$

and $C>0$ such that for any $\eta \in (\mathrm {Dil}(\mathcal {X},\mathcal {X}_\ast )^{-1}, \mathrm {Dil}(\mathcal {X}_\ast ,\mathcal {X}))$ we have

(1.3) $$ \begin{align} 0 < \limsup_{T\to\infty} \frac{1}{T} \log \left(\#\left\{ [g] \in \mathbf{conj}: \ell_{\mathcal{X}}[g] < T, | \ell_{\mathcal{X}_\ast}[g] - \eta \ell_{\mathcal{X}}[g] | < \frac{C}{T} \right\} \right)= \mathcal{I}(\eta) \le v_{\mathcal{X}}. \end{align} $$

Furthermore, we have equality in the above inequality if and only if $\eta =\tau (\mathcal {X}_\ast /\mathcal {X})$ .

We have referred to this result as large deviations with shrinking intervals because it estimates the growth of the number of group elements satisfying (1.2) opposed to (1.1). It is desirable to prove this refinement, as it provides a much more precise comparison between the actions of $\Gamma $ on $\mathcal {X}$ and $\mathcal {X}_\ast $ .

Intuitively, the function $\mathcal {I}$ can be seen as measuring how similar the geometries actions on $\mathcal {X}$ and $\mathcal {X}_\ast $ are. That is, the closer $\mathcal {I}$ is to the constant function with value $v_{\mathcal {X}}$ , the more similar the length functions $v_{\mathcal {X}}\ell _{\mathcal {X}}$ and $v_{\mathcal {X}_\ast }\ell _{\mathcal {X}_\ast }$ are. We also note that the real analyticity of $\mathcal {I}$ is a useful property. Indeed, as shown in [Reference Cantrell and Reyes17], when $\mathcal {I}$ is analytic, it is possible to obtain rigidity results that compare metrics through the values taken by $\mathcal {I}$ .

Remark 1.6. As in Theorem 1.2, the conclusion above still holds for triplets $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )$ in the class $\mathfrak {X}$ ; see Theorem 6.2. However, for our arguments (particularly Theorem 3.2) it is crucial that the translation length functions belong to a lattice in ${\mathbb {R}}$ . We still expect Theorem 6.2 to hold for arbitrary cuboid complexes $\mathcal {X}^{\mathfrak {w}}$ and $\mathcal {X}^{\mathfrak {w}_\ast }_\ast $ , but we will not pursue this in this work.

As an application of Theorem 1.5 we deduce large deviations with shrinking intervals for the intersection of curves on hyperbolic surfaces. Let $\Sigma $ be a closed orientable surface of negative Euler characteristic and fundamental group $\Gamma $ . If $\alpha ,\beta $ are immersed closed oriented curves in $\Sigma $ , then the (geometric) intersection number is the minimal number $i_\Sigma (\alpha ,\beta )$ of intersections of closed curves in the free homotopy classes of $\alpha $ and $\beta $ . The function $i_\Sigma $ can be extended by bilinearity to weighted multicurves, which are finite sums of the form $\sum _j{\lambda _j\alpha _j}$ with $\alpha _j$ immersed oriented closed curves in $\Sigma $ and a set $(\lambda_j\geq 0)_j$ of weights. Any nontrivial element in $\mathbf {conj}(\Gamma )$ is represented by a unique free homotopy class of immersed oriented closed curves, so we can talk of the intersection number between a weighted multicurve in $\Sigma $ and a conjugacy class in $\mathbf {conj}(\Gamma )$ . For more details about the intersection number, see [Reference Farb and Margalit33].

A generating set S for $\Gamma $ is simple if there exists a point $p\in \Sigma $ such that elements of $S\subset \Gamma =\pi _1(\Sigma ,p)$ can be represented by simple loops that are pairwise nonhomotopic and disjoint except at the base point p. For example, the generating set for the standard presentation

$$ \begin{align*}\Gamma=\left<a_1,b_1,\dots,a_g,b_g\colon[a_1,b_1]\cdots [a_g,b_g]\right>\end{align*} $$

is simple. In [Reference Erlandsson31, Theorem 1.2], Erlandsson proved that the translation length function $\ell _S$ of the word metric of a simple generating set S can be recovered by pairing in the intersection number against a carefully chosen weighted multicurve $\alpha _S$ with weights in $\frac {1}{2}\mathbb {Z}$ . Up to scaling, this translation length function can also be recovered by looking at the $\mathrm {CAT}(0)$ cube complex dual to the multicurve $\alpha _S$ . Therefore, Theorem 6.2 applies and we obtain the following.

Corollary 1.7. Let $\Gamma $ be the fundamental group of the closed orientable hyperbolic surface $\Sigma $ and consider a simple generating set S of exponential growth rate $v_S$ . Let $\alpha $ be a nontrivial weighted multicurve on $\Sigma $ with integer weights, and define

$$\begin{align*}a_{\inf}:=\inf_{[g] \in \mathbf{conj}'} \frac{i_\Sigma(\alpha,[g])}{\ell_{S}[g]} \ \text{ and } \ a_{\sup} := \sup_{[g] \in \mathbf{conj}'} \frac{i_\Sigma(\alpha,[g])}{\ell_{S}[g]}. \end{align*}$$

Then there exists an analytic convex function $\mathcal {I}:[a_{\inf },a_{\sup }] \rightarrow {\mathbb {R}}$ and $C>0$ such that for any $\eta \in (a_{\inf },a_{\sup })$ we have

$$\begin{align*}0 < \limsup_{T\to\infty} \frac{1}{T} \log \left(\#\left\{ [g] \in \mathbf{conj}: \ell_S[g] < T, | i_\Sigma(\alpha,[g]) - \eta \ell_{S}[g] | < \frac{C}{T} \right\} \right)= \mathcal{I}(\eta) \le v_S. \end{align*}$$

As discussed above, we can see this corollary as providing a precise comparison between the intersection number (with $\alpha $ ) and word length of closed geodesics. In some sense, $\mathcal {I}$ is encoding how well the intersection number can be approximated by the simple word metric for S.

1.4 Word metrics on hyperbolic groups

Since the proof of Theorem 1.5 relies on Theorem 3.2 (a purely dynamical statement), the existence of an automaton encoding both actions on $\mathcal {X}$ and $\mathcal {X}_\ast $ , and the arithmeticity of the translation length functions, we can deduce large deviations with shrinking intervals for any pair of group actions fulfilling similar conditions. That is the case of word metrics on hyperbolic groups, and in fact, this was the author’s main motivation at the beginning of this project.

Let $\Gamma $ be a nonelementary hyperbolic group and let $S,S_\ast \subset \Gamma $ be finite generating sets with corresponding word metrics $d_S, d_{S_\ast }$ . By Cannon’s theorem [Reference Cannon15], for a total order on S, the language of lexicographically first geodesics in $\Gamma $ is regular, so it is parametrized by a finite-state automaton. As a consequence of [Reference Calegari and Fujiwara14, Lemma 3.8], Calegari and Fujiwara are able to modify this automaton (without modifying the parameterized language), and find an integer functional on the edges of the graph of the automaton so that its sum over paths recovers the $S_\ast $ -word length for the corresponding element in $\Gamma $ (this was our main motivation to construct the automaton in Theorem 5.11).

By studying the subshift of finite type associated to this automaton, Cantrell and Tanaka [Reference Cantrell and Tanaka19, Reference Cantrell and18] deduced analyticity of the Manhattan curve for $S,S_\ast $ as well as a large deviations principle. More precisely, let $\ell _S, \ell _{S_\ast }$ be the corresponding translation length functions and consider the dilations

$$\begin{align*}\mathrm{Dil}(S_\ast,S) = \sup_{[g] \in \mathbf{conj}'} \frac{\ell_{S_\ast}[g]}{\ell_{S}[g]} \ \text{ and } \ \mathrm{Dil}(S,S_\ast)^{-1} = \inf_{[g] \in \mathbf{conj}'} \frac{\ell_{S_\ast}[g]}{\ell_{S}[g].} \end{align*}$$

Then there exists a real analytic, concave function $\mathcal {I} :[ \mathrm {Dil}(S,S_\ast )^{-1}, \mathrm {Dil}(S_\ast ,S)] \to {\mathbb {R}}_{>0}$ such that for $\eta \in (\mathrm {Dil}(S,S_\ast )^{-1}, \mathrm {Dil}(S_\ast ,S))$ we have

(1.4) $$ \begin{align} \lim_{\epsilon \to 0^+} \limsup_{n\to\infty} \frac{1}{T} \log \left( \#\left\{[g] \in \mathbf{conj}(\Gamma) \colon \ell_S[g] < T , \left| \frac{\ell_{S_\ast}[g]}{\ell_S[g]} - \eta \right| < \epsilon \right\} \right)= \mathcal{I}(\eta). \end{align} $$

The rate function $\mathcal {I}$ is a Legendre transform constructed from the Manhattan curve for $S,S_\ast $ , see [Reference Cantrell and Tanaka19, Theorem 4.23]. By applying Theorem 3.2 to this subshift, we can improve this result and obtain a large deviations theorem with shrinking intervals.

Theorem 1.8. Let $\Gamma $ be a nonelementary hyperbolic group and consider two finite generating sets $S, S_\ast $ for $\Gamma $ with exponential growth rates $v_S,v_{S_\ast }$ , and $\mathcal {I}$ as above. Then there exists $C>0$ such that for any $\eta \in (\mathrm {Dil}(S,S_\ast )^{-1}, \mathrm {Dil}(S_\ast ,S))$ we have

$$\begin{align*}0 < \limsup_{T\to\infty} \frac{1}{T} \log \left(\#\left\{ [g] \in \mathbf{conj}: \ell_S[g] < T, | \ell_{S_\ast}[g] - \eta \ell_{S}[g] | < \frac{C}{T} \right\} \right)= \mathcal{I}(\eta) \le v_S. \end{align*}$$

Furthermore, we have equality in the above inequality if and only if

$$\begin{align*}\eta = \tau(S_\ast/S):=\lim_{T\to\infty} \frac{1}{\#\{[g]\in \mathbf{conj} \colon \ell_S[g]<T\}} \sum_{\ell_S[g] < T} \frac{\ell_{S_\ast}[g]}{T}. \end{align*}$$

This result implies that, after scaling a pair of word metrics by their exponential growth rates, there is always an exponentially growing set for which their translation lengths are close, that is, for any $\epsilon> 0$

(1.5) $$ \begin{align} 0 < \limsup_{T\to\infty} \frac{1}{T} \log \left(\#\left\{ [g] \in \mathbf{conj}: \ell_S[g] < T, | v_S\ell_S[g] - v_{S_\ast} \ell_{S_\ast}[g] | < \epsilon \right\}\right) \le v_S. \end{align} $$

This extends the recent work [Reference Cantrell and Reyes17, Theorem 4.1] where the authors proved a correlation result for pairs of word metrics under an additional rationality assumption on the exponential growth rates. This result also answers a question raised by the authors in [Reference Cantrell and Reyes17, Remark 4.3]. We also deduce the following corollary.

Corollary 1.9. Let $\Gamma $ be a nonelementary hyperbolic group, and let S and $S_\ast $ be two finite generating sets on $\Gamma $ . Then there exists $C>0$ such that for any $\eta \in [\mathrm {Dil}(S,S_\ast )^{-1}, \mathrm {Dil}(S_\ast ,S)]$ we can find an infinite sequence $(g_n)_{n\geq 1} \subset \Gamma $ such that

(1.6) $$ \begin{align} \left|\frac{\ell_{S_\ast}[g_n]}{\ell_{S}[g_n]} - \eta \right| \le \frac{C}{|g_n|_S^{2}}. \end{align} $$

If $\eta \in [\mathrm {Dil}(S,S_\ast )^{-1}, \mathrm {Dil}(S_\ast ,S)]$ is rational then there exists $g \in \Gamma $ such that

$$\begin{align*}\frac{\ell_{S_\ast}[g]}{\ell_{S}[g]} = \eta. \end{align*}$$

This result is concerned with understanding how well the values of the quotient of $\ell _{S_\ast }$ with $\ell _{S}$ can approximate a given $\eta \in [\mathrm {Dil}(S,S_\ast )^{-1}, \mathrm {Dil}(S_\ast ,S)]$ . It shows that the approximation rate is optimal. Indeed, it is well-known that the translation length function for word metrics takes values in a lattice $\frac {1}{N} \mathbb {Z}$ for some $N \in {\mathbb {N}}$ and therefore by Hurwitz’s Theorem [Reference Hurwitz51] we cannot find a sequence $g_n$ for which the convergence rate in (1.6) is faster.

In Subsection 4.2 we present an example for a pair of word metrics on a free group. In particular, we compute the limit supremum in (1.5) for this pair of word metrics.

Organization

The organization of the paper is as follows. In Section 2 we cover preliminary material about Manhattan curves, $\mathrm {CAT}(0)$ cube complexes and cubulable groups, finite-state automata, symbolic dynamics and suspension flows.

In Section 3 we prove Theorem 3.2, a large deviations result with shrinking intervals for lattice potentials on mixing subshifts of finite type that are constant on 2-cylinders. We apply this theorem to pairs of word metrics on hyperbolic groups in Section 4 and prove Theorem 1.8.

In Section 5 we prove Proposition 5.2 that describes large classes of groups included in $\mathfrak {G}$ . There we also prove Theorem 5.11, in which we construct a finite-state automaton for pairs of compatible actions on $\mathrm {CAT}(0)$ cube complexes. We use this automaton in Section 6 to prove Theorem 6.1 and Theorem 6.2, from which we deduce Theorem 1.2 and Theorem 1.5. In this section we also prove Theorem 1.4.

Finally, in the appendix we prove a Proposition A.1, a criterion for convex-cocompactness of subgroups of cubulable relatively hyperbolic groups, which may be of independent interest.

2 Preliminaries

2.1 Isometric group actions

Let $\Gamma $ be a finitely generated group acting by isometries on the metric space $(X,d_X)$ and let $x\in X$ be an arbitrary base point. The (stable) translation length of this action is the function $\ell _X:\mathbf { conj} \rightarrow {\mathbb {R}}$ given by

$$\begin{align*}\ell_X[g]=\lim_{n\to \infty}{\frac{d_X(g^nx,x)}{n}} \ \text{ for }g\in [g] \text{ in }\mathbf{conj}.\end{align*}$$

The exponential growth rate of this action is the quantity

$$\begin{align*}v_X:=\limsup_{T \to \infty}\frac{\log \left(\#\{g\in \Gamma \colon d_X(gx,x)<T\}\right)}{T}\in [0,+\infty].\end{align*}$$

As for the translation lengths, the exponential growth rate is independent of the chosen base point x. If X is geodesic (or more generally roughly geodesic) and the action of $\Gamma $ on X is proper and cocompact, then $v_X$ is finite. In some cases we can recover the exponential growth rate as the limit

$$\begin{align*}\limsup_{T\to \infty}{\frac{1}{T}\log \# \mathfrak{C}_X(T)}, \end{align*}$$

where each $T>0$ we denote

$$\begin{align*}\mathfrak{C}_X(T)=\{[g]\in \mathbf{conj} \colon \ell_X[g]<T\}.\end{align*}$$

This happens for example when $\Gamma $ is hyperbolic and the action on X is proper and cocompact.

Given two isometric actions of $\Gamma $ on the metric spaces X and $X_\ast $ , the Manhattan curve for the pair $(X, X_\ast )$ is the boundary of the convex set

$$\begin{align*}\mathcal{C}_{X_\ast/X}:=\left\{ (a,b) \in {\mathbb{R}}^2 : \sum_{[g] \in \mathbf{conj}} e^{-a\ell_{X_\ast}[g] - b\ell_{X}[g]} < \infty \right\}, \end{align*}$$

assuming it is nonempty. Equivalently, $\mathcal {C}_{X_\ast /X}$ is the set of points $(s,\theta _{X_\ast /X}(s))$ where $\theta _{X_\ast /{X}}(s)$ is the abscissa of convergence of the series

$$\begin{align*}t \mapsto \sum_{[g] \in \mathbf{conj}} e^{-t\ell_{X}[g] - s\ell_{X_\ast}[g]}. \end{align*}$$

By abuse of notation, $\theta _{X_\ast /X}$ is also called the Manhattan curve for $(X,X_\ast )$ .

2.2 $\mathrm {CAT}(0)$ cube complexes

For bibliography about $\mathrm {CAT}(0)$ cube complexes and groups acting on them, we refer the reader to [Reference Bridson and Haefliger10, Reference Sageev74]. A nonpositively curved (NPC) cube complex is a metric polyhedral complex in which all polyhedra are unit-length Euclidean cubes, and satisfies Gromov’s link condition: the link of each vertex is a flag complex. If this complex is simply connected we say that it is a $\mathrm {CAT}(0)$ cube complex.

Let $\mathcal {X}$ be an NPC cube complex. Consider the minimal equivalence relations on the set of edges (resp. oriented edges) of $\mathcal {X}$ such that the edges $e=\{v,w\}$ and $e=\{v',w'\}$ (resp. oriented edges $v\xrightarrow {e} w$ and $v'\xrightarrow {e'} w'$ ) are in the same equivalence class if $v,w,v',w'$ span a square in $\mathcal {X}$ (resp. $v,w,v'w'$ span a square in $\mathcal {X}$ with $v,v'$ adjacent and $w,w'$ adjacent). Equivalence classes of these equivalence relations are called hyperplanes (resp. oriented hyperplanes), and we let $\mathbb {H}(\mathcal {X})$ denote the set of hyperplanes of $\mathcal {X}$ . If the equivalence class of an (oriented) edge e is the (oriented) hyperplane $\mathfrak {h}$ , then we say that e dual to $\mathfrak {h}$ . A hyperplane is 2-sided if it corresponds to exactly two oriented hyperplanes; otherwise it is 1-sided.

Suppose now that $\mathcal {X}$ is a $\mathrm {CAT}(0)$ cube complex, in which case all hyperplanes are 2-sided. A combinatorial path in $\mathcal {X}$ is a sequence $\gamma =(\gamma _0,\dots ,\gamma _n)$ of vertices in $\mathcal {X}$ such that $\gamma _i$ is adjacent to $\gamma _{i+1}$ for $i=0,\dots ,n-1$ . In that case, we say that $\gamma _0$ is the initial vertex of $\gamma $ and $\gamma _n$ is its final vertex. The length of $\gamma =(\gamma _0,\dots ,\gamma _n)$ is defined as n. This path is often seen as a continuous path by also considering the edges $e_{i+1}$ joining each $\gamma _i$ with $\gamma _{i+1}$ . Such a path is geodesic if no two distinct edges $e_i$ are dual to the same hyperplane. The combinatorial metric on $\mathcal {X}$ is the graph metric $d_{\mathcal {X}}$ on its 1-skeleton $\mathcal {X}^1$ so that each edge has length 1. It follows that a combinatorial path is geodesic if and only if it is geodesic for the metric $d_{\mathcal {X}}$ .

A hyperplane $\mathfrak {h}$ in the $\mathrm {CAT}(0)$ cube complex $\mathcal {X}$ separates two vertices of $\mathcal {X}$ if some (any) combinatorial path connecting these vertices has an edge dual to $\mathfrak {h}$ . It follows that the combinatorial distance of any two vertices in $\mathcal {X}$ equals the number of hyperplanes separating them. Also, a hyperplane $\mathfrak {h}$ determines the equivalence relation of “not being separated by $\mathfrak {h}$ ” on the set of vertices of $\mathcal {X}$ . This equivalence relation has exactly 2 equivalence classes $\{\mathfrak {h}^-,\mathfrak {h}^+\}$ , which are the halfspaces determined by $\mathfrak {h}$ . A subcomplex of $\mathcal {X}$ is convex if its vertex set is the intersection of halfspaces. Equivalently, $Z\subset \mathcal {X}$ is convex if any combinatorial geodesic joining points in $Z^0$ is contained in $Z^0$ .

An orthotope structure on the $\mathrm {CAT}(0)$ cube complex $\mathcal {X}$ is a function

$$ \begin{align*}\mathfrak{w}: \mathbb{H}(\mathcal{X}) \rightarrow {\mathbb{R}}_{>0},\end{align*} $$

and the pair $\mathcal {X}^{\mathfrak {w}}=(\mathcal {X},\mathfrak {w})$ is called a cuboid complex. An orthotope structure induces a metric $d_{\mathcal {X}}^{\mathfrak {w}}$ on $\mathcal {X}^1$ by declaring each edge e to have length $\mathfrak {w}(\mathfrak {h})$ for $\mathfrak {h}$ the hyperplane dual to e. In this way, for any two vertices $x,y\in \mathcal {X}^0$ we have

$$\begin{align*}d_{\mathcal{X}}^{\mathfrak{w}}(x,y)=\sum_{\mathfrak{h}\in \mathbb{H}(x|y)}{\mathfrak{w}(\mathfrak{h})},\end{align*}$$

where $\mathbb {H}(x|y)\subset \mathbb {H}(\mathcal {X})$ is the collection of hyperplanes separating x and y. Note that if $\mathfrak {w}$ is the constant function equal to 1, then $d_{\mathcal {X}}^{\mathfrak {w}}$ is just the standard combinatorial metric $d_{\mathcal {X}}$ .

Remark 2.1. It is clear that a geodesic in $\mathcal {X}$ with respect to $d_{\mathcal {X}}$ is also geodesic with respect to $d_{\mathcal {X}}^{\mathfrak {w}}$ for any orthotope structure $\mathfrak {w}$ .

Now let $\Gamma $ be a group acting on the $\mathrm {CAT}(0)$ cube complex $\mathcal {X}$ . We always assume that the action is cubical, meaning that it preserves the cube complex structure. Under this assumption, the action is isometric on $\mathcal {X}^1$ with the combinatorial metric $d_{\mathcal {X}}$ . Similarly, $\ell _{\mathcal {X}}$ always denotes the stable translation length of $\Gamma $ with respect to the action on $(\mathcal {X}^1,d_{\mathcal {X}})$ . If the action of $\Gamma $ is proper and cocompact, we say that $\mathcal {X}$ is a cubulation of $\Gamma $ .

The action of $\Gamma $ on $\mathcal {X}$ induces a natural action on $\mathbb {H}(\mathcal {X})$ . If $\mathfrak {w}$ is a $\Gamma $ -invariant orthotope structure on $\mathcal {X}$ in the sense that $\mathfrak {w}(\mathfrak {h})=\mathfrak {w}(g\mathfrak {h})$ for all $g\in \Gamma $ and $\mathfrak {h}\in \mathbb {H}(\mathcal {X})$ , then we say that $\Gamma $ acts on the cuboid complex $(\mathcal {X},\mathfrak {w})$ . In that case the action of $\Gamma $ on $(\mathcal {X}^1,d_{\mathcal {X}}^{\mathfrak {w}})$ is also by isometries. We let $\ell _{\mathcal {X}}^{\mathfrak {w}}$ denote the stable translation length of $\Gamma $ for its action on $(\mathcal {X}^1,d_{\mathcal {X}}^{\mathfrak {w}})$ .

By a hyperplane stabilizer we mean a subgroup of $\Gamma $ consisting of group elements g such that $g \mathfrak {h}=\mathfrak {h}$ for some fixed hyperplane $\mathfrak {h}\in \mathbb {H}(\mathcal {X})$ . We denote the hyperplane stabilizer of $\mathfrak {h}$ by $\Gamma _{\mathfrak {h}}$ . A hyperplane $\mathfrak {h}$ is essential for the action of $\Gamma $ if for any vertex x of $\mathcal {X}$ , the halfspaces $\mathfrak {h}^\pm $ contain elements in the $\Gamma $ -orbit of x arbitrarily far from $\mathfrak {h}^{\mp }$ . The action of $\Gamma $ on $\mathcal {X}$ is essential if every hyperplane is essential.

If $\mathcal {X}$ is a cubulation of $\Gamma $ , a subgroup $H<\Gamma $ is called convex-cocompact with respect to $\mathcal {X}$ if there exists a convex subcomplex $Z\subset \mathcal {X}$ that is H-invariant and so that the action of H on Z is cocompact. Such a subcomplex Z is called a convex core for H. Note that the hyperplane stabilizer of any hyperplane $\mathfrak {h}$ is convex-cocompact since it acts cocompactly on the (convex subcomplex spanned by the) set of vertices in edges dual to $\mathfrak {h}$ [Reference Haglund and Wise47, Lemma 13.4].

If $\mathcal {X}$ is a $\mathrm {CAT}(0)$ cube complex and $\mathbb {W}\subset \mathbb {H}(\mathcal {X})$ is any collection of hyperplanes, in [Reference Caprace and Sageev20, Section 2.3] Caprace and Sageev introduced the restriction quotient, which is a $\mathrm {CAT}(0)$ cube complex $\mathcal {X}(\mathbb {W})$ equipped with a surjective cellular map $\phi :\mathcal {X} \rightarrow \mathcal {X}(\mathbb {W})$ satisfying the following: an edge in $\mathcal {X}$ is collapsed to a single vertex under $\phi $ if and only if it is dual to a hyperplane not in $\mathbb {W}$ . The projection $\phi $ induces a natural bijection between $\mathbb {W}$ and $\mathbb {H}(\mathcal {X}(\mathbb {W}))$ , and hence any orthotope structure $\mathfrak {w}$ on $\mathcal {X}$ induces an orthotope structure $\phi _\ast (\mathfrak {w})$ on $\mathcal {X}(\mathbb {W})$ .

Remark 2.2. Note that (pre)images of convex subcomplexes under restriction quotients remain convex. In particular, images of geodesic paths remain geodesic, although some subpaths are allowed to collapse to points. Also note that if $\Gamma $ acts on $\mathcal {X}$ and $\mathbb {W}$ is $\Gamma $ -invariant, then there is a natural action of $\Gamma $ on $\mathcal {X}(\mathbb {W})$ .

A cubulation $\mathcal {X}$ of $\Gamma $ is co-special if the quotient $\overline {\mathcal {X}}=\Gamma \backslash \mathcal {X}$ is a special cube complex in the sense of Haglund-Wise [Reference Haglund and Wise47]. Equivalently, $\mathcal {X}$ is co-special if $\Gamma $ injects into a right-angled Artin group $A_G$ inducing a $\Gamma $ -equivariant isometric embedding of $\mathcal {X}$ into $R_G$ as a convex subcomplex, where $R_G$ is the universal cover of the Salvetti complex $\overline {R}_G$ associated to the graph G. Among other properties of special cube complexes, hyperplanes are 2-sided and embedded, and they do not self-osculate [Reference Haglund and Wise47, Def. 3.2]. In particular, different oriented edges with the same initial vertex are dual to different oriented hyperplanes. The cubulation $\mathcal {X}$ of $\Gamma $ is virtually co-special if there exists a finite-index subgroup $\overline {\Gamma } <\Gamma $ such that the action of $\overline {\Gamma }$ on $\mathcal {X}$ is co-special.

Fundamental groups of compact special cube complexes are residually finite, and more generally, their convex-cocompact subgroups are separable [Reference Haglund and Wise47, Corollary 7.9]. For our purposes, co-special cubulations will be used to construct finite-state automata parameterizing combinatorial geodesics, as we explain in Example 2.5 in the next subsection.

2.3 Finite-state automata

For references on automatic structures and some of their connections with group theory, see [Reference Epstein, Cannon, Holt, Levy, Paterson and Thurston32]. For the relation between automatic structures and special cube complexes, we refer the reader to [Reference Li and Wise58, Section 5]. Let S be a finite set and let $S^\ast $ denote the set of finite words over the alphabet S. If $w=h_1\cdots h_n$ and $w'=h^{\prime }_1\cdots h_m'$ are words in $S^\ast $ , then its concatenation is the word $ww':=h_1\cdots h_nh^{\prime }_1\cdots h_m'$ . The length of a word is the number of letters in S composing it. We let the empty set correspond to the unique word of length 0 in $S^\ast $ . A language over S is any subset of words in $S^\ast $ .

A (finite-state) automaton over S is a tuple

$$\begin{align*}\mathcal{A}=(\mathcal{G},\pi,I,F),\end{align*}$$

where $\mathcal {G}=(V,E)$ is a finite directed graph, $\pi :E\rightarrow S$ is a labeling function and $I,F\subset V$ are nonempty sets of initial and final states.

Remark 2.3. This convention differs from the standard definition of automaton (e.g., [Reference Li and Wise58, Definition 5.2]), where it is required for I to consist of a single vertex. This difference in convention does not significantly change the discussion, but it will be useful in Theorem 5.11 when we construct an automaton for which we have no control on the number of initial states.

By a path in $\mathcal {G}$ we mean a sequence $\omega $ of (always directed) edges $e_1,\dots ,e_n$ in E such that the final vertex $v_i$ of $e_i$ is the initial vertex of $e_{i+1}$ for $i=1,\dots , n-1$ . If $v_0$ is the initial vertex of $e_1$ , we denote this path $\omega $ either by $\omega =(\xrightarrow { e_1} \cdots \xrightarrow {e_n})$ or $\omega =(v_0\xrightarrow { e_1} \cdots \xrightarrow {e_n} v_n)$ depending on the emphasis we want to give to the vertices. If there is no ambiguity on the edges, we can also denote this path by $(v_0 \rightarrow \cdots \rightarrow v_n)$ . The length of a path is the number of edges that determine it. Note that paths of length 1 correspond to the edges in E. We also allow the degenerate case of paths of length 0, which are the vertices in V.

If $\omega , \omega '$ are paths in $\mathcal {G}$ such that the final vertex of $\omega $ is the initial vertex of $\omega '$ , then the concatenation $\omega \omega '$ is the path in $\mathcal {G}$ defined in the expected way. Similarly we define the concatenation of any finite number of paths.

A word w in $S^\ast $ is represented by the path $\omega =(v_0\xrightarrow { e_1} v_1\cdots \xrightarrow {e_n} v_n)$ in $\mathcal {G}$ if $w=\pi (\omega ):=\pi (e_1)\cdots \pi (e_n)$ . If in addition $v_0\in I$ and $v_n\in F$ , we say that w accepted by $\mathcal {A}$ and that $\omega $ is admissible. It is clear that the word represented by a concatenation $\omega \omega '$ is the concatenation of the words represented by $\omega $ and $\omega '$ . Let $L=L_{\mathcal {A}}$ be the language consisting of the words accepted by $\mathcal {A}$ . In this case we say that L is parametrized by $\mathcal {A}$ .

The automaton $\mathcal {A}$ is deterministic if any two distinct edges in $\mathcal {G}$ with the same initial vertex have different labels. In that case, for any $w\in L_{\mathcal {A}}$ and any initial state $v\in I$ there exists at most one path in $\mathcal {G}$ representing w and starting at v. The automaton is pruned if any vertex in $\mathcal {G}$ is the final vertex of a path starting at an initial state.

Example 2.4 (Automatic structures on hyperbolic groups)

Let $\Gamma $ be a hyperbolic group and consider a finite set $S\subset \Gamma $ generating $\Gamma $ as a semi-group and with word length $|\cdot |_S$ . We consider S as our alphabet, so that there is a natural evaluation map $\text {ev}:S^\ast \rightarrow \Gamma $ . For a fixed total order on S we obtain the lexicographic order $\prec $ on $S^*$ . Let $L=L_S$ be language of lexicographically first geodesics. That is, for each $g \in \Gamma $ , a word $w \in S^\ast $ with $\text {ev}(w)=g$ is in L if and only if w has length $|g|_S$ and $w \prec w'$ for all other $w' \in S^\ast $ with $\text {ev}(w')=g$ and of length $|g|_S$ . Cannon [Reference Cannon15] showed that the language L defined above is regular, in the sense that $L=L_{\mathcal {A}}$ for $\mathcal {A}=(\mathcal {G}=(V,E),\pi ,\{\ast \},V)$ a deterministic finite-state automaton over S. In particular, the evaluation map gives us a length-preserving bijection from L onto $\Gamma $ .

Consider now another finite generating subset $S_\ast \subset \Gamma $ . In [Reference Calegari and Fujiwara14, Lemma 3.8], Calegari and Fujiwara constructed a new deterministic automaton $\mathcal {A}'=(\mathcal {G}'=(V',E'),\pi ',\{\ast '\},V')$ over S parameterizing $L_S$ and an integer-valued function $\phi :V' \rightarrow {\mathbb {Z}}$ such that for any word $w\in L_S$ represented by the path $\omega =(\ast '\xrightarrow { e_1} \cdots \xrightarrow {e_n} v_n)$ in $\mathcal {G}'$ we have

$$\begin{align*}|\text{ev}(w)|_{S_\ast}=\sum_{i=1}^{n}\phi(v_i).\end{align*}$$

We note that it is equivalent to define the labeling $\phi $ on the directed edge set $E'$ instead.

We will see in the following subsection that the automatic structure discussed above gives rise to a dynamical system $(\Sigma ,\sigma )$ called a subshift of finite type. The labeling $\phi $ above corresponds to a real-valued function on $\Sigma $ which satisfies the Markovian property of being ‘constant on $2$ -cylinders’: see Subsection 2.4.

Example 2.5 (Automatic structures on special cube complexes)

For this example we follow Sections 5.2 and 5.3 in [Reference Li and Wise58]. Let $\overline {\mathcal {Z}}$ be a compact special cube complex with universal cover $\mathcal {Z}$ and fix a base vertex $o \in \mathcal {Z}$ . Let $S_{\overline {\mathcal {Z}}}$ be the set of the oriented hyperplanes of $\overline {\mathcal {Z}}$ . In [Reference Li and Wise58], Li and Wise constructed a deterministic pruned finite-state automaton

$$\begin{align*}\mathcal{A}_{\overline{\mathcal{Z}}}=(\mathcal{G}_{\overline{\mathcal{Z}}}=(V,E),\pi,\{\ast\},V)\end{align*}$$

over $S_{\overline {\mathcal {Z}}}$ and parameterizing a language $L_{\overline {\mathcal {Z}}}$ . This language describes geodesics in $\mathcal {Z}$ based at o in the following sense. Given any word $w \in L_{\overline {\mathcal {Z}}}$ represented by the (necessarily unique) admissible path $\omega $ there exists a geodesic path $\gamma _{w}$ in $\mathcal {Z}$ starting at o and ending at the vertex $\tau _{\mathcal {Z}}(w)$ such that:

  • if $w=\mathfrak {h}_1\cdots \mathfrak {h}_n\in L_{\overline {\mathcal {Z}}}$ and $\omega =(\ast '\xrightarrow { e_1}\cdots \xrightarrow {e_n} v_n)$ , then $\gamma _w=(o=x_0,\dots ,x_n)$ is a geodesic of length n;

  • for each $1 \leq i \leq n$ we have $\pi (e_i)=\mathfrak {h}_i$ and the oriented hyperplane in $\mathcal {Z}$ dual to the oriented edge from $x_{i-1}$ to $x_{i}$ maps to $\mathfrak {h}_i$ under the quotient $\mathcal {Z} \rightarrow \overline {\mathcal {Z}}$ ; and,

  • the map $\tau _{\mathcal {Z}}: L_{\overline {\mathcal {Z}}} \rightarrow \mathcal {Z}^0$ is a bijection.

The automaton $\mathcal {A}_{\overline {\mathcal {Z}}}$ being deterministic implies that the geodesic $\gamma _w$ is uniquely determined by w.

Remark 2.6. The language constructed in [Reference Li and Wise58] actually depends on an injection of $\Gamma =\pi _1(\overline {\mathcal {Z}})$ into a right-angled Artin group $A_G$ inducing a $\Gamma $ -equivariant isometric embedding of $\mathcal {Z}$ into $R_G$ as a convex subcomplex, where $R_G$ is the universal cover of the Salvetti complex $\overline {R}_G$ associated to G. In that case, the language obtained is over the alphabet of oriented hyperplanes of $\overline {R}_G$ . The language $L_{\overline {\mathcal {Z}}}$ described above is a particular case of this construction, when we consider the local isometric immersion $\overline {\mathcal {Z}} \rightarrow \overline {R}_{G_{\overline {\mathcal {Z}}}}$ for $G_{\overline {\mathcal {Z}}}$ being the crossing graph of $\overline {\mathcal {Z}}$ . For this immersion there is a natural bijection between $S_{\overline {\mathcal {Z}}}$ and the set of oriented hyperplanes in $\overline {R}_{G_{\overline {\mathcal {Z}}}}$ , see for instance [Reference Haglund and Wise47, Lemma 4.1].

2.4 Symbolic dynamics

In this subsection we introduce the preliminary material we need from symbolic dynamics. See Chapter $1$ of [Reference Parry and Pollicott66] for more information regarding the basic definitions we now present. Let A be a $k \times k$ matrix with entries $0$ or $1$ . This matrix is said to be aperiodic if there exists $N \ge 1$ such that all of the entries of $A^N$ are strictly positive. We say that A is irreducible if for any $i,j \in \{1, \ldots , k \}$ there exists $n \ge 1$ such that $(A^n)_{i,j}$ (i.e., the $(i,j)$ th entry of $A^n$ ) is strictly positive.

The (one-sided) subshift of finite type $\Sigma _A$ associated to A is the set of infinite sequences

$$\begin{align*}\Sigma_A = \left\{(x_n)_{n=0}^{\infty} : x_n \in \{1, \ldots, k\} \text{ and } A_{x_n, x_{n+1}} =1 \text{ for all } n \ge 0\right\}. \end{align*}$$

These infinite sequences can be seen as infinite paths in a directed graph $\mathcal {G}_A$ with vertices labeled $1, \ldots , k$ and a directed edge from vertex i to j if and only if $A_{i,j} =1$ . We will therefore refer to the numbers $1, \ldots , k$ as the states of $\Sigma _A$ . We equip $\Sigma _A$ with the shift map $\sigma : \Sigma _A \to \Sigma _A$ defined by

$$\begin{align*}\sigma((x_n)_{n=0}^\infty) = (x_{n+1})_{n=0}^\infty \end{align*}$$

to obtain a dynamical system $(\Sigma _A, \sigma )$ .

Consider a finite ordered string $x_0, \ldots , x_{m-1} \in \{1, \ldots , k \}$ where $A_{x_j, x_{j+1}} =1$ for each $j=0, \ldots , m-2$ . The cylinder set associated to this string is the subset of $\Sigma _A$ given by

$$\begin{align*}[x_0, \ldots, x_{m-1}] := \left\{ (y_n)_{n=0}^\infty \in \Sigma_A : y_j = x_j \text{ for } j=0, \ldots, m-1 \right\}. \end{align*}$$

We endow $\Sigma _A$ with a topology by declaring the set of all cylinder sets to be an open basis.

The system $(\Sigma _A, \sigma )$ is said to be mixing if for any two open sets $U, V \subset \Sigma _A$ there is $N \ge 1$ such that $\sigma ^n(U) \cap V \neq \emptyset $ for all $n \ge N$ . We say that $(\Sigma _A, \sigma )$ is transitive if for any two open sets $U,V \subset \Sigma _A$ there exists $n \ge 1$ such that $\sigma ^n(U) \cap V \neq \emptyset .$ We have that $(\Sigma _A, \sigma )$ is mixing if and only if A aperiodic and $(\Sigma _A, \sigma )$ is transitive if and only if A is irreducible. We will often suppress the dependence of A in the notation for a subshift and will write $(\Sigma , \sigma )$ .

Example 2.7. Let $\Gamma $ be a hyperbolic group equipped with finite generating set S. Consider a corresponding automatic structure $\mathcal {A}=(\mathcal {G}=(V,E),\pi ,\{\ast \},V)$ as discussed in Example 2.4. Suppose we have labeled the vertices in V by $1$ to k where k is the cardinality of V. Then the graph $\mathcal {G}$ is encoded by a $k \times k$ transition matrix A where the $(i,j)$ th entry of A is $1$ if there is a directed edge joining vertex i to j and is $0$ otherwise. This matrix gives a subshift $(\Sigma _A,\sigma )$ that encodes $(\Gamma , S)$ . A subshift obtained in this way is never transitive (as $\ast $ only has outgoing edges) and it is not known whether, after removing $\ast $ , it is always possible to find a connected graph $\mathcal {G}$ representing a given pair $(\Gamma ,S)$ . In general it is possible to decompose the graph $\mathcal {G}$ into connected components (i.e., maximal connected subgraphs). If the the transition matrices for these subgraphs are $\mathcal {C}_1,\ldots , \mathcal {C}_m$ then the subshifts $\Sigma _{\mathcal {C}_j}$ are each transitive. We call a connected component $\mathcal {C}$ maximal if the number of paths in $\mathcal {C}$ consisting of n edges grows like $\lambda ^n$ , where $\lambda ^n$ is the growth rate of the n spheres in the Cayley graph $\text {Cay}(\Gamma , S)$ , that is, the growth rate of the number of paths of length n in $\mathcal {C}$ is as large as possible.

Throughout the rest of the section $(\Sigma , \sigma )$ will be a mixing subshift of finite type, and consider a function (which we will, at some points, refer to as a potential) $\psi :\Sigma \to {\mathbb {R}}$ . We say that $\psi $ is constant on $2$ -cylinders if $\psi $ is constant on each set of the form $[x_0, x_1]$ where $x_0, x_1 \in \{1, \ldots , k \}$ and $A_{x_0, x_1} = 1$ . The assumption that $\psi $ is constant on $2$ -cylinders guarantees that $\psi $ has Markovian behaviour: the value that $\psi $ takes at $x \in \Sigma $ depends only on the initial cylinder that x belongs to. That is, $\psi (x)$ does not depend on the future cylinders that x visits under the iterates of $\sigma $ . Although this is a restrictive condition for a general function on $\Sigma $ , it is not restrictive for our purposes. This is because we are interested in understanding the growth rate properties of functions that have lattice image: their image lies in $\alpha \mathbb {Z}$ for some $\alpha \in \mathbb {R}$ (as discussed in Subsection 1.3). Hölder continuous functions on $\Sigma $ that have a lattice image have the property that their image only depends on finitely many symbols of the input. Then after relabeling the subshift $\Sigma $ (i.e., moving to a topologically conjugate subshift) we can assume that the function is in fact constant on $2$ -cylinders. See for example the proof of Proposition 5.1 in [Reference Parry and Pollicott66] for an example of this argument. To summarize, functions that are constant on $2$ -cylinders naturally arise when studying discrete geometries. Lastly, it is worth mentioning that such functions have particularly nice properties: see for example Lemma 3.4 below.

For each $n\ge 1$ , the nth Birkhoff sum of $\psi $ is the function

$$\begin{align*}\psi^n : \Sigma \to {\mathbb{R}} \ \text{ such that } \ \psi^n(x) := \psi(x) + \psi(\sigma(x)) + \cdots + \psi(\sigma^{n-1}(x)). \end{align*}$$

A point $x \in \Sigma $ is said to be periodic if $\sigma ^n(x) = x$ for some $n \ge 1$ . Such an n is called a period of x. Note that a periodic point has infinitely many periods. Given a periodic point we will assume that it comes with with a choice of period (which may not be its least period) which we will label $|x|$ (so that $\sigma ^{|x|}(x) = x$ ).

Example 2.8. Consider the automaton $\mathcal {A}'=(\mathcal {G}'=(V',E'),\pi ',\{\ast '\},V')$ and the integer-valued function $\phi :V' \rightarrow {\mathbb {Z}}$ introduced in Example 2.4. Then, as discussed in Example 2.7, $\mathcal {A}'$ gives rise to a subshift of finite type $(\Sigma ,\sigma ).$ Furthermore, the labeling $\phi $ defines a function $f: \Sigma \to {\mathbb {Z}}$ by

$$\begin{align*}f(x) = \phi(v_{x_0}), \end{align*}$$

where $x = (x_n)_{n=0}^\infty $ and $v_{x_0} \in V$ is the vertex corresponding to the symbol $x_0$ . The function f is constant on $2$ -cylinders and furthermore if $x = (x_n)_{n=0}^\infty \in \Sigma $ then

$$\begin{align*}f^n(x) = |g_{x,n}|_{S_\ast} \end{align*}$$

where $g_{x,n} \in \Gamma $ is the group element obtained from multiplying the first n labelings in the infinite path corresponding to x, that is, if the first n edges in the path corresponding to x are $e_1,\ldots , e_n$ then $f^n(x) = |\pi (e_1)\cdots \pi (e_n)|_{S_\ast }$ . We note that, when $x \in \Sigma $ satisfies that $\sigma ^n(x) = x$ then

$$\begin{align*}mf^n(x) = f^{mn}(x) = |g_{x,mn}|_{S_\ast} = |g_{x,n}^m|_{S_\ast} \end{align*}$$

and so

$$\begin{align*}f^n(x) = \lim_{m\to\infty} |g_{x,n}^m|_{S_\ast} / m = \ell_{S_\ast}[g_{x,n}]. \end{align*}$$

Two functions $\psi , \varphi : \Sigma \to {\mathbb {R}}$ , which we are assuming to be constant on $2$ -cylinders, are said to be cohomologous if there exists a continuous function $u: \Sigma \to {\mathbb {R}}$ such that $\psi (x) = \varphi (x) + u(\sigma (x)) - u(x)$ for all $x \in \Sigma $ . By Livsic’s Theorem [Reference Parry and Pollicott66, Proposition 3.7], $\psi $ and $\varphi $ are cohomologous if and only if $\psi ^n(x) = \varphi ^n(x)$ whenever $\sigma ^n(x) = x.$

The variational principle states that there is a unique $\sigma $ -invariant Borel probability measure on $(\Sigma , \sigma )$ that achieves the supremum

$$\begin{align*}\text{P}(\psi) := \sup_{\mu\in \mathcal{M}_\sigma}\left\{ h_\mu(\sigma) + \int \psi \ d\mu \right\}, \end{align*}$$

where $\mathcal {M}_\sigma $ is the collection of all $\sigma $ -invariant Borel probability measures on $\Sigma $ and $h_\mu (\sigma )$ denotes the (metric) entropy of $\sigma $ with respect to the measure $\mu $ [Reference Parry and Pollicott66, Theorem 3.5]. The quantity $\text {P}(\psi )$ is referred to as the pressure of $\psi $ and the measure attaining the supremum is called the equilibrium state of $\psi $ . When $\psi $ is a constant function, the measure achieving the supremum for the pressure of $\psi $ is the measure of maximal entropy. Furthermore the topological entropy $h=h(\sigma )$ of $(\Sigma ,\sigma )$ is given by $h = \text {P}(0)$ .

Consider the quantities

$$ \begin{align*} \alpha_{\min} := \inf_{\mu \in \mathcal{M}_\sigma}\int_\Sigma \psi \ d\mu \ \ \text{ and } \ \ \alpha_{\max} := \sup_{\mu \in \mathcal{M}_\sigma} \int_\Sigma \psi \ d\mu. \end{align*} $$

The large deviations principle implies (since functions that are constant on $2$ -cylinders are Hölder) that there exists a real analytic, concave function $\mathcal {L}(\psi , \cdot ) : {\mathbb {R}} \to {\mathbb {R}}_{>0}\cup \{\infty \}$ such that, for any nonempty sets $U \subset V \subset {\mathbb {R}}$ with U open and V closed we have

(2.1) $$ \begin{align} -\inf_{s\in U} \mathcal{L}(\psi, s) &\le \liminf_{n\to\infty} \frac{1}{n} \log \mu \left( x \in \Sigma : \frac{\psi^n(x)}{n} \in U \right) \nonumber \\ &\le \limsup_{n\to\infty} \frac{1}{n} \log \mu \left( x \in \Sigma : \frac{\psi^n(x)}{n} \in V \right) \le -\inf_{s\in V} \mathcal{L}(\psi,s). \end{align} $$

See [Reference Kifer56, Theorem in Section 2.1] for this result. The same result holds, with the same rate function $\mathcal {L}(\psi ,\cdot )$ , when we replace the sets

$$\begin{align*}\mu \left( x \in \Sigma : \left| \frac{\psi^n(x)}{n} - \eta \right| < \epsilon \right) \end{align*}$$

with the sequence of (normalized) cardinalities

$$\begin{align*}\frac{1}{\#\{x \in \Sigma : \sigma^n(x) = x \}}\#\left\{x \in \Sigma: \sigma^n(x) = x \text{ and } \left| \frac{\psi^n(x)}{n} - \eta \right| < \epsilon \right\}. \end{align*}$$

This is a well-known result that follows from the same proof as that of [Reference Kifer56, Theorem in Section 2.1]. The function $\mathcal {L}(\psi ,\cdot )$ is the Legendre transform of $t \mapsto \text {P}(t\psi ) - h$ . That is,

(2.2) $$ \begin{align} - \mathcal{L}(\psi,s) = \inf_{t\in {\mathbb{R}}} (\text{P}(t\psi) - h - ts ). \end{align} $$

Furthermore, $\mathcal {L}(\psi ,\cdot )$ is finite on $[\alpha _{\min }, \alpha _{\max }]$ and is infinite otherwise. An alternative characterization for $\mathcal {L}$ is the following:

$$\begin{align*}-\mathcal{L}(\psi,\eta)= \sup\left\{ h_\mu(\sigma) : \mu \in\mathcal{M}_\sigma \text{ and} \int \psi \ d\mu = \eta\right\} - h. \end{align*}$$

A function $\psi : \Sigma \rightarrow {\mathbb {R}}$ is lattice if there are $a,b \in {\mathbb {R}}$ satisfying

$$\begin{align*}\left\{ \psi^n(x) + an : x\in\Sigma \text{ and } \sigma^n(x)=x \text{ for some } n\ge 1 \right\} \subset b{\mathbb{Z}}. \end{align*}$$

If this is not the case then we say that $\psi $ is nonlattice.

Remark 2.9. Suppose that $\psi $ is lattice. Then $\psi $ is cohomologous to a function of the form $a + b \varphi $ where $a,b \in {\mathbb {R}}$ and $\varphi : \Sigma \to {\mathbb {Z}}$ [Reference Parry and Pollicott66, Proposition 5.2]. When this is the case, the large deviations behaviour of $\psi $ and $\varphi $ over periodic orbits is the same, since $\psi ^n(x) = an + b\varphi ^n(x)$ when $\sigma ^n(x) =x$ .

2.5 Suspension flows

In this subsection we define suspension flows of subshifts of finite type. See Chapters $1$ to $6$ of [Reference Parry and Pollicott66] for more details on the results stated in this subsection. Let $\Sigma _A$ be a transitive subshift of finite type and $r: \Sigma _A \to {\mathbb {R}}_{>0}$ a function that is constant on $2$ -cylinders. We note that in [Reference Parry and Pollicott66, Chapter 6] suspension flows are considered over mixing subshifts, however the same proofs (with some minor modifications) work when the subshift is transitive. We define the suspension flow of $\Sigma _A^r$ to be the space

$$\begin{align*}\Sigma_A^r = \{(x,t) \in \Sigma_A \times {\mathbb{R}}_{\ge 0} : 0 \le t \le r(x) \} / \sim \end{align*}$$

where $(x,t) \sim (r(x),0)$ , equipped with the flow $\sigma ^r=(\sigma ^r_t)_{t>0}$ so that $\sigma _t^r$ sends $(x,s)$ to $(x,s+t)$ for $s \in {\mathbb {R}}$ . There is a natural metric on $\Sigma _A^r$ which can be constructed as in [Reference Parry and Pollicott66]. We will not present the construction of this metric here as it is a little technical.

For a Hölder continuous function $\Phi : \Sigma _A^r \to {\mathbb {R}}$ we can define its pressure as

$$\begin{align*}\text{P}_{\sigma^r}(\Phi) = \sup_{m \in \mathcal{M}_{\sigma^r}} \left\{ h_m(\sigma^r) + \int_{\Sigma_A^r} \Phi \ dm \right\}, \end{align*}$$

where $\mathcal {M}_{\sigma ^r}$ is the space of $\sigma ^r_t$ - invariant Borel probability measures on $\Sigma _A^r$ and $h_m(\sigma ^r)$ is the entropy of the time-one map $\sigma ^r_1$ for the measure m [Reference Parry and Pollicott66, Section 6].

Remark 2.10. For such $\Phi $ , the pressure function $s \mapsto \text {P}_{\sigma ^r}(s\Phi )$ is real analytic. This is a well-known result that follows from Proposition 6.1 in [Reference Parry and Pollicott66] and the implicit function theorem.

Let $\delta _r>0$ be the unique number such that $\text {P}(-\delta _r r) =0$ and write $\mu _{-\delta _r r}$ for the equilibrium state of $-\delta _r r$ on $\Sigma _A$ . The measure of maximal entropy for $\Sigma _A^r$ is (locally) given by

$$\begin{align*}\frac{\mu_{-\delta_r r} \times \text{Leb}}{\int r \ d\mu_{-\delta_r r}}, \end{align*}$$

where $\text {Leb}$ represents the Lebesgue measure along $\mathbb {R}_{\ge 0}$ ([Reference Parry and Pollicott66, Proposition 6.1]). That is, up to normalization, the measure of maximal entropy is the measure that acts as Lebesgue along the fibers of the suspension and as $\mu _{-\delta _r r}$ on the base. If we write m for the measure of maximal entropy then we have that

$$\begin{align*}\left. \frac{d}{ds}\right|{}_{s=0} \text{P}_{\sigma^r}(s\Phi) = \int \Phi \ dm \end{align*}$$

(see [Reference Parry and Pollicott66, Proposition 4.10]).

For $T>0$ we will write $P(\Sigma _A^r, T)$ for the collection of periodic orbits of $\sigma ^r$ of length less than T. Given $R>0$ we will write $P(\Sigma _A^r, R , T)$ for the collection of periodic orbits of length between $T -R$ and $T+R$ .

Given a Hölder continuous function $\Phi : \Sigma _A^r \to {\mathbb {R}}$ it is a standard result that for any $R>0$ sufficiently large

$$\begin{align*}\lim_{T\to\infty} \frac{1}{T} \log \left(\sum_{\tau \in P(\Sigma_A^r, R, T)} e^{-s \int_\tau \Phi }\right) = \text{P}_{\sigma^r}(-s\Phi) \end{align*}$$

for any $ s\in {\mathbb {R}}$ (see for example [Reference Parry and Pollicott66, Proposition 5.10]).

Lastly we recall that two functions $\Phi , \Psi : \Sigma _A^r \to {\mathbb {R}}$ are cohomologous if $\Phi - \Psi = u'$ where $u : \Sigma _A^r \to {\mathbb {R}}$ is continuously differentiable (along flow lines) and

$$\begin{align*}u'(x) = \lim_{t\to0} \frac{u(\sigma^r_t(x)) - u(x)}{t}. \end{align*}$$

Further the function $s \mapsto \text {P}_{\sigma ^r}(s\Phi )$ is a straight line if and only if $\Phi $ is cohomologous to a constant function.

3 Large deviations

In this section we discuss large deviations with shrinking intervals for potentials on mixing subshifts of finite type. The main result of the section is Theorem 3.2, and it will be used in the proof of Theorems 1.8 and 6.2 in subsequent sections.

Suppose that $(\Sigma ,\sigma )$ is a mixing subshift of finite type with $k \times k$ transition matrix A, and let $\mu $ denote its measure of maximal entropy. Also, let M be the least number such that $A^M$ has strictly positive entries. The large deviations principle (2.1) from Subsection 2.4 implies that there is a real analytic, concave function $\mathcal {L}(\psi , \cdot ): [\alpha _{\min }, \alpha _{\max }] \to {\mathbb {R}}_{>0}$ such that the following holds: for any $\eta \in (\alpha _{\min }, \alpha _{\max })$

(3.1) $$ \begin{align} \lim_{\epsilon \to 0^-} \limsup_{n\to\infty} \frac{1}{n} \log \mu \left( x \in \Sigma : \left| \frac{\psi^n(x)}{n} - \eta \right| < \epsilon \right) = - \mathcal{L}(\psi,\eta). \end{align} $$

Instead of taking two limits as above, first with respect to n and then with respect to $\epsilon $ , it is natural to ask the following.

Question 3.1. How quickly can a sequence $\delta _n$ decay to $0$ as $n\to \infty $ so that we have

(3.2) $$ \begin{align} \lim_{n\to\infty} \frac{1}{n} \log \mu\left( x \in \Sigma : \left| \frac{\psi^n(x)}{n} - \eta \right| < \delta_n \right) = - \mathcal{L}(\psi,\eta) \end{align} $$

for each $\eta \in (\alpha _{\min }, \alpha _{\max })$ ?

We refer to this problem as large deviations with shrinking intervals. We can ask the same question when the limit in (3.2) is replaced with the limit supremum.

Large deviations with shrinking intervals are best understood for functions $\psi : \Sigma \to \mathbb {R}$ that are nonlattice. For example, when $\psi $ is nonlattice the local central limit theorem [Reference Guivarc’h and Hardy42] implies that (3.2) holds when $\delta _n^{-1} = O(n)$ . In [Reference Pollicott and Sharp68] Pollicott and Sharp improved this result under an additional assumption. They showed that if $\psi $ satisfies a non-Diophantine condition then there exist $\kappa>0$ such that (3.2) holds when $\delta _n^{-1} = O(n^{1+\kappa })$ .

When $\psi $ is lattice, large deviations with shrinking intervals are not as well-understood. The aim of this section is to study (3.2) for functions $\psi $ that are constant on $2$ -cylinders and are lattice.

We now state our large deviations theorem with shrinking intervals. Write $\mathcal {L}(\psi , \cdot ): [\alpha _{\min }, \alpha _{\max }] \to {\mathbb {R}}_{>0}$ be the function introduced above.

Theorem 3.2. Suppose that $(\Sigma ,\sigma )$ is a mixing subshift of finite type and that $\psi : \Sigma \to {\mathbb {R}}$ is a function that is constant on $2$ -cylinders. Then there exists $C>0$ such that for any $\eta \in (\alpha _{\min }, \alpha _{\max })$

(3.3) $$ \begin{align} \limsup_{n\to\infty} \frac{1}{n} \log\left( \#\left\{ x \in \Sigma: \sigma^n(x) = x \text{ and } \left| \frac{\psi^n(x)}{n} - \eta \right| < \frac{C}{n^2} \right\} \right) = h -\mathcal{L}(\psi,\eta) \end{align} $$

where h is the topological entropy of $(\Sigma ,\sigma )$ . Furthermore we can take

$$\begin{align*}C = \frac{4M^2(1+k^2)^2(\alpha_{\max} - \alpha_{\min}) }{\sqrt{5}}. \end{align*}$$

In the case that $\psi $ takes values in ${\mathbb {Z}}$ then there exists $\eta \in (\alpha _{\min }, \alpha _{\max })$ and $\epsilon>0$ such that

$$\begin{align*}\limsup_{n\to\infty} \frac{1}{n} \log\left( \#\left\{ x \in \Sigma: \sigma^n(x) = x \text{ and } \left| \frac{\psi^n(x)}{n} - \eta \right| < \frac{\epsilon}{n^2} \right\} \right) = 0. \end{align*}$$

Remark 3.3. i) This theorem also holds if we only assume that $(\Sigma ,\sigma )$ is transitive (opposed to mixing). Indeed, if $(\Sigma ,\sigma )$ is transitive we can find an integer $p \ge 1$ such that $(\Sigma ,\sigma ^p)$ decomposes into p disjoint $\sigma ^p$ -invariant sets. These $\sigma ^p$ -invariant sets are mixing subshifts of finite type when equipped with $\sigma ^p$ . We can then apply the mixing version of our theorem above to these subshifts to deduce the transitive version. ii) Our result significantly improves the decay rate implied by the local limit theorem under the nonlattice assumption [Reference Rousseau-Egele73, Theoreme 5]. Furthermore, the case that $\psi $ takes values in a lattice shows that the decay rates obtained in the first part of Theorem 3.2 are optimal.

For the rest of the section we note the correspondence between periodic orbits of $(\Sigma ,\sigma )$ and cycles (i.e., closed paths) in the adjacency graph $\mathcal {G}_A$ . We say that a cycle is simple if it does not visit any state more than once.

To prove the above result we start with the following observation.

Lemma 3.4. Suppose that $(\Sigma ,\sigma )$ is a mixing subshift of finite type and that $\psi : \Sigma \to {\mathbb {R}}$ is a function that is constant on $2$ -cylinders. Then

$$\begin{align*}\alpha_{\min} = \inf_{\sigma^n(x) = x} \frac{\psi^n(x)}{n} \ \text{ and } \ \alpha_{\max} = \sup_{\sigma^n(x) = x} \frac{\psi^n(x)}{n}, \end{align*}$$

and furthermore there exist periodic orbits $\overline {x},\overline {y}$ that achieve these values, that is, $\psi ^{|\overline {x}|}(\overline {x}) = |\overline {x}|\alpha _{\min }, \psi ^{|\overline {y}|}(\overline {y}) = |\overline {y}|\alpha _{\max }.$ Here the infimum and supremum are over all the periodic orbits.

Proof. The first statement follows from a result of Sigmund [Reference Sigmund82, Theorem 1] which states that the set of probability measures supported on periodic orbits is dense in $\mathcal {M}_\sigma $ (equipped with the weak- $\ast $ topology). For the furthermore statement note that each periodic orbit can be written as a disjoint union of simple cycles. It follows easily that the above infimum and supremum are attained by simple cycles (and powers of them).

We now want to use the orbits $\overline {x}, \overline {y}$ from the lemma above to construct periodic orbits along which the average value of $\psi $ approximates a given $\eta \in (\alpha _{\min },\alpha _{\max })$ (see Proposition 3.8). We begin with the following observation.

Lemma 3.5. Take an interval $(s,t) \subset {\mathbb {R}}$ and a number $\eta \in (s,t)$ . Then there are infinitely many $n \ge 1$ for which there exist integers $0 \le a,b \le n$ with $a + b = n$ and such that

$$\begin{align*}\left|\frac{as + bt}{n} - \eta \right| \le \frac{t-s}{\sqrt{5} \, n^2}. \end{align*}$$

Proof. Note that it suffices to prove this result when $s = 0$ , $t=1$ . The general result then follows by shifting and rescaling the interval $(0,1)$ into $(s,t)$ . When $s = 0$ , $t=1$ the result follows from the well-known Hurwitz’s Theorem [Reference Hurwitz51] from Diophantine approximation.

We also need the following lemma. Suppose that the simple periodic orbit $\overline {x}$ realizing $\alpha _{\max }$ has initial state i and that the initial state for $\overline {y}$ realizing $\alpha _{\min }$ is j (from Lemma 3.4). Further assume that we repeat $\overline {x}$ and $\overline {y}$ by each other’s periods so that they both have period l satisfying $1<l \leq k^2$ .

Lemma 3.6. Suppose that $(\Sigma ,\sigma )$ is a mixing subshift of finite type and that $\psi : \Sigma \to {\mathbb {R}}$ is a function that is constant on $2$ -cylinders such that $\alpha _{\min }<\alpha _{\max }$ . Then we can find periodic orbits $x, y$ both with period at most $2M(1+k^2)$ such that the initial state for x is i, the initial state for y is j and we have that

$$\begin{align*}\alpha_{\min} < \frac{\psi^{|x|}(x)}{|x|} < \frac{\psi^{|y|}(y)}{|y|} < \alpha_{\max}. \end{align*}$$

Proof. Consider the start state i and find a path p from i to j of length M and a path q from j to i of length M. By composing the path p with r (to be chosen later) repeats of a single cycle of $\overline {y}$ and q we obtain a periodic orbit x of period $2M + lr$ . Further we have that

$$\begin{align*}\frac{\psi^{|x|}(x)}{|x|} \le \frac{2M\alpha_{\max} + lr\alpha_{\min}}{2M + lr} \end{align*}$$

and x starts at i. Similarly we find y starting at j with $|y| = 2M + lr$ and

$$\begin{align*}\frac{\psi^{|y|}(y)}{|y|}\ge \frac{2M\alpha_{\min} + lr\alpha_{\max}}{2M + lr}. \end{align*}$$

Now, as long as r is chosen so that

$$\begin{align*}2M\alpha_{\min} + lr\alpha_{\max}> 2M\alpha_{\max} + lr\alpha_{\min} \end{align*}$$

$x,y$ will satisfy the final inequality in the lemma. Note that this inequality is satisfied for $ r =2M$ in which case x and y have periods $2M+2Ml \le 2M(1+k^2)$ as required.

We also require the following.

Lemma 3.7. Suppose that $(\Sigma ,\sigma )$ is a mixing subshift of finite type and that $\psi : \Sigma \to {\mathbb {R}}$ is a function that is constant on $2$ -cylinders. Suppose that there exist periodic orbits $x,y$ both with period l and same initial state such that $lA = \psi ^l(x) < \psi ^l(y) = lB$ . Then there exists $\overline {C}>0$ such that for any $\eta \in (A, B)$ we can find an infinite sequence of periodic orbits $x_n \in \Sigma $ such that

$$\begin{align*}\left| \frac{\psi^{|x_n|}(x_n)}{|x_n|} - \eta \right| \le \frac{\overline{C}}{|x_n|^2}. \end{align*}$$

Furthermore we can take

$$\begin{align*}\overline{C} = \frac{(\alpha_{\max} - \alpha_{\min})l^2}{\sqrt{5}}. \end{align*}$$

Proof. By Lemma 3.5 there exist infinitely many $n \ge 1$ such that the following holds. There are integers $n_1, n_2\geq 0$ with $n_1 + n_2 = n$ and such that

$$\begin{align*}\left| \frac{n_1 A + n_2 B}{n_1+n_2} - \eta \right| \le \frac{B-A}{\sqrt{5} n^2} \le \frac{(\alpha_{\max} - \alpha_{\min}) l^2}{\sqrt{5} (l n)^2}. \end{align*}$$

Since x and y have the same initial vertex, we can form a new periodic orbit by composing $n_1$ copies of x followed by $n_2$ copies of y. This creates a periodic orbit z of orbit length $nl$ with the property that

$$\begin{align*}\frac{\psi^{nl}(z)}{nl} = \frac{n_1 A + n_2 B}{n_1+n_2} \ \text{ and so } \ \left| \frac{\psi^{nl}(z)}{nl} - \eta \right| \leq \frac{(\alpha_{\max} - \alpha_{\min})l^2}{\sqrt{5} |z|^2}. \end{align*}$$

Since we can run this construction for infinitely many n, the result follows.

To obtain uniformity over $\eta $ , that is, to show the existence of C in Theorem 3.2, we need to upgrade Lemma 3.7 using Lemma 3.6.

Proposition 3.8. Suppose that $(\Sigma ,\sigma )$ is a mixing subshift of finite type and that $\psi : \Sigma \to {\mathbb {R}}$ is a function that is constant on $2$ -cylinders. Then there exists $C>0$ such that for any $\eta \in (\alpha _{\min }, \alpha _{\max })$ there exists an infinite sequence of periodic orbits $x_n \in \Sigma $ such that

$$\begin{align*}\left| \frac{\psi^{|x_n|}(x_n)}{|x_n|} - \eta \right| \le \frac{C}{|x_n|^2}. \end{align*}$$

Furthermore we can take

$$\begin{align*}C = \frac{4M^2 (1+k^2)^2 (\alpha_{\max} - \alpha_{\min})}{\sqrt{5}}. \end{align*}$$

Proof. We can assume that $\psi $ is not cohomologous to a constant function (otherwise the conclusion is clear), so that $\alpha _{\min }<\alpha _{\max }$ .

We split the interval $(\alpha _{\min }, \alpha _{\max }) = I_1 \cup I_2$ into the two (nondisjoint) intervals

$$\begin{align*}I_1 = \left(\alpha_{\min}, \frac{\psi^{|y|}(y)}{|y|}\right) \ \text{ and } \ I_2 = \left(\frac{\psi^{|x|}(x)}{|x|}, \alpha_{\max}\right) \end{align*}$$

where $x,y$ are the orbits constructed in Lemma 3.6. In particular, both $|x|$ and $|y|$ are bounded above by $2M(1+k^2)$ . We can now apply Lemma 3.7 to both of the intervals $I_1$ and $I_2$ to deduce the result.

Definition 3.9. Let $\psi ,\Sigma $ be as above. Suppose $w \in (\alpha _{\min }, \alpha _{\max })$ is chosen so that there exists x with $\sigma ^n(x) = x$ and $\psi ^n(x) = nw$ . We define $d(\psi ,w)$ to be the greatest common divisor of all numbers $n \ge 1$ such that

$$\begin{align*}\#\left\{ x\in\Sigma : \sigma^n(x) = x, \psi^n(x) = wn \right\}> 0. \end{align*}$$

If $\#\left \{ x\in \Sigma : \sigma ^n(x) = x, \psi ^n(x) = wn \right \} =0$ for all n then we set $d(\psi ,w) = 0$ .

Note that $d(\psi , w) = 0$ for all but countably many values of $w.$ This is because the values for which $d(\psi , w)> 0$ are contained in the rational span of the values that the averaged Birkhoff sum of $\psi $ attains on simple cycles. We now state the key result of Marcus and Tuncel that we need to prove large deviations with shrinking intervals. Recall that $\mathcal {L}$ represents the large deviations rate function introduced in Subsection 2.4 (see (2.1)).

Theorem 3.10 (Theorem 14 [Reference Marcus and Tuncel59])

For any $\xi>0$ and any closed subset $W \subset ( \alpha _{\min }, \alpha _{\max }) $ there exist $r, N \in {\mathbb {Z}}_{\ge 0}$ and $\delta> 0$ such that, for any $n \ge N$ with $d(\psi ,w)|n$ ,

(3.4) $$ \begin{align} \#\left\{ x\in\Sigma : \sigma^n(x) = x, \psi^n(x) = wn \right\} \ge \delta n^{-r} e^{n(h-(\mathcal{L}(\psi, w)) - \xi)} \end{align} $$

for each $w\in W$ satisfying $d(\psi ,w)>0$ .

To make use of this result we need to control the values of $d(\psi ,w)$ as w takes values in a shrinking interval.

Lemma 3.11. There exists $C>0$ such that for any $\eta \in (\alpha _{\min }, \alpha _{\max })$ we can find a sequence $w_n \in (\alpha _{\min }, \alpha _{\max })$ and a sequence $x_n \in \Sigma $ of periodic orbits such that

$$\begin{align*}w_n \in \left[ \eta - \frac{C}{|x_n|^2}, \eta +\frac{C}{|x_n|^2} \right] \ \text{ with } \ \frac{\psi^{|x_n|}(x_n)}{|x_n|} = w_n \end{align*}$$

and $d(\psi ,w_n)\, \big | \, |x_n|$ .

Proof. By Proposition 3.8 there exist $C>0, N\ge 1$ depending only on $\psi , \Sigma $ such that for any ${\eta \in (\alpha _{\min }, \alpha _{\max })}$ and $n \ge N$ there is a sequence of periodic orbits $x_n \in \Sigma $ with periods $|x_n|$ (i.e., $\sigma ^{|x_n|}(x_n) = x_n$ ) such that

$$\begin{align*}\left| \frac{\psi^{|x_n|}(x_n)}{|x_n|} - \eta \right| \le \frac{C}{|x_n|^2}. \end{align*}$$

We define $w_n = \frac {\psi ^n(x_n)}{n}$ and note that $d(\psi , w_n) \, \big | \, |x_n|$ if and only if

$$\begin{align*}\#\left\{ x \in \Sigma: \sigma^{|x_n|}(x) = x, \frac{\psi^{|x_n|}(x_n)}{|x_n|} = w_n \right\}> 0. \end{align*}$$

Hence we are done.

For a sequence $(\delta _n)_{n\geq 1}$ of positive numbers and $n \ge 1$ we define

$$\begin{align*}F_n(\eta, \delta_n) = \left\{x\in \Sigma : \sigma^n(x) = x, \left|\frac{\psi^n(x)}{n} - \eta\right| < \delta_n \right\} \end{align*}$$

for $\eta \in (\alpha _{\min }, \alpha _{\max })$ . Intuitively, this set contains the periodic orbits of period n along which the average value of $\psi $ approximates $\eta $ up to an error less than $\delta _n$ . Theorem 3.2 is concerned with the exponential growth rate of the cardinality of $F_n(\eta , Cn^{-2})$ for some constant $C>0$ as $n\to \infty $ . We therefore want to obtain bounds on $\#F_n(\eta , Cn^{-2})$ which we achieve in the following proposition.

Proposition 3.12. Suppose that $(\Sigma , \sigma )$ is a mixing subshift of finite type and that $\psi : \Sigma \to {\mathbb {R}}$ is a function that is constant on $2$ -cylinders. Then there exists $C>0$ such that for any $\eta \in (\alpha _{\min }, \alpha _{\max })$ and $ \xi> 0$ there exist $\delta , r, M>0$ and a sequence of integers $n_l$ with $n_l \to \infty $ as $l \to \infty $ such that

$$\begin{align*}\#F_{n_l}(\eta, C n_l^{-2}) \ge \delta n_l^{-r} e^{n_l ((h-\mathcal{L}(\psi,\eta)) -\xi) } \end{align*}$$

for all $l \ge 1.$

Remark 3.13. For $\eta $ with $d(\psi , \eta )> 0$ this result follows immediately from estimates due to Marcus and Tuncel [Reference Marcus and Tuncel59]. The main strength of the above estimates are that they hold for all values of $\eta \in (\alpha _{\min }, \alpha _{\max })$ .

Proof of Proposition 3.12

We use Proposition 3.8 to find a sequence of periodic orbits $(x_n)_n$ such that

$$\begin{align*}|w_n - \eta| \le \frac{C}{|x_n|^2} \ \text{ where } \ w_n = \frac{\psi^{|x_n|}(x_n)}{|x_n|} \end{align*}$$

for each $n \ge 1$ . By (3.4), for any $\xi>0$ and for all n sufficiently large we have that

(3.5) $$ \begin{align} \#\left\{ x \in \Sigma: \sigma^{|x_n|}(x) = x \text{ and } \left| \frac{\psi^{|x_n|}(x_n)}{|x_n|} - \eta \right| < \frac{C}{|x_n|^2} \right\} \ge \delta |x_n|^{-r} e^{|x_n|((h-\mathcal{L}(\psi,w_n)) - \xi)} \end{align} $$

for some $\delta , r>0$ . Now, by analyticity of $\mathcal {L}$ , there is $\widetilde {C}>0$ such that

$$\begin{align*}|\mathcal{L}(\psi,\eta) - \mathcal{L}(\psi,w_n)| \le \widetilde{C}|\eta - w_n| = O(|x_n|^{-2}) \end{align*}$$

(where the implied error constant is independent of n). Substituting this into the right hand side of (3.5) provides the required bound concluding the proof.

Proof of Theorem 3.2

Note that the large deviations principle where we count over periodic orbits (see the discusion after (2.1)) implies that

$$\begin{align*}\limsup_{n\to\infty} \frac{1}{n} \log \left(\#\left\{ x\in\Sigma : \sigma^n(x) = x, \left| \frac{\psi^n(x)}{n} - \eta\right| < \frac{C'}{n^2} \right\} \right)\le h-\mathcal{L}(\psi, \eta) \end{align*}$$

for any $C'>0$ and $\eta \in (\alpha _{\min }, \alpha _{\max })$ . Proposition 3.12 also implies that for $C> 0$ (as in the proposition)

(3.6) $$ \begin{align} \limsup_{n\to\infty} \frac{1}{n} \log \left(\#\left\{ x\in\Sigma_{\mathcal{C}} : \sigma^n(x) = x, \left| \frac{\psi^n(x)}{n} - \eta\right| < \frac{C}{n^2} \right\} \right)\ge h-\mathcal{L}(\psi, \eta) \end{align} $$

for any $\eta \in (\alpha _{\min }, \alpha _{\max })$ . This proves the identity (3.3), and the estimate for C follows from Proposition 3.8.

We now finish by proving the final statement of the theorem which assumes that $\psi $ takes values in $\mathbb {Z}$ . When this is the case, the set $ \left \{ \frac {\psi ^n(x)}{n} : \sigma ^n(x) = x \right \}$ contains rational numbers all with denominator at most n. By [Reference Hurwitz51, Satz II] there is $\eta \in (\alpha _{\min }, \alpha _{\max })$ such that if $\epsilon> 0$ is sufficiently small then for all but finitely many values of n,

$$\begin{align*}\left|\frac{\psi^n(x)}{n} - \eta \right|> \frac{\epsilon}{n^2} \ \text{ for all }x\in \Sigma\text{ with }\sigma^n(x) = x. \end{align*}$$

Hence $F_n(\eta , \epsilon n^{-2})$ is empty for all n sufficiently large.

4 Large deviations for pairs of word metrics

In this section we prove Theorem 1.8.

Throughout the proof, we will follow the same terminology and notation that we established above. Before presenting the proof, we restate the theorem for the reader’s convenience.

Theorem 4.1. Let $\Gamma $ be a nonelementary hyperbolic group and consider two finite generating sets $S, S_\ast $ for $\Gamma $ with exponential growth rates $v_S,v_{S_\ast }$ . Then there exist $C>0$ and a real analytic, concave function $\mathcal {I} :[ \mathrm {Dil}(S,S_\ast )^{-1}, \mathrm {Dil}(S_\ast ,S)] \to {\mathbb {R}}_{>0}$ such that for any $\eta \in (\mathrm {Dil}(S,S_\ast )^{-1}, \mathrm {Dil}(S_\ast ,S))$ we have

$$\begin{align*}0 < \limsup_{T\to\infty} \frac{1}{T} \log \left(\#\left\{ [g] \in \mathbf{conj}: \ell_S[g] < T, | \ell_{S_\ast}[g] - \eta \ell_{S}[g] | < \frac{C}{T} \right\} \right)= \mathcal{I}(\eta) \le v_S. \end{align*}$$

Furthermore, we have equality in the above inequality if and only if

$$\begin{align*}\eta = \tau(S_\ast/S):=\lim_{T\to\infty} \frac{1}{\#\{[g]\in \mathbf{conj} \colon \ell_S[g]<T\}} \sum_{\ell_S[g] < T} \frac{\ell_{S_\ast}[g]}{T}. \end{align*}$$

Proof. Without loss of generality assume that $d_S, d_{S_{\ast }}$ are not roughly similar (i.e., there does not exist $\tau , C>0$ such that $|d_S(g,h) - \tau d_{S_\ast }(g,h)| < C$ for all $g,h \in \Gamma $ ). As discussed in Example 2.8, by Lemma 3.8 and Example 3.9 in [Reference Calegari and Fujiwara14] we can find a Cannon coding for $(\Gamma , S)$ with corresponding shift space $(\Sigma ,\sigma )$ and a constant on $2$ -cylinders function $\psi : \Sigma \to {\mathbb {Z}}$ satisfying the following: if $z = (z_0, z_1,\ldots , z_{n-1},z_0, \ldots )\in \Sigma $ is a periodic orbit of period n, then the nth Birkhoff sum of $\psi $ on z outputs the value $\ell _{S_\ast }[g]$ , where $g\in \Gamma $ is the group element obtained by multiplying the labelings in the finite path $z_0,z_1, \ldots , z_{n-1}$ .

Fix a maximal component $\Sigma _{\mathcal {C}}$ in $\Sigma $ (as in Example 2.7). We know that the function $\psi $ satisfies a large deviations principle (as discussed at (3.1)) over the periodic orbits on $\Sigma _{\mathcal {C}}$ , as discussed after (2.1). We set $\mathcal {I}=h-\mathcal {L}$ , where h is the topological entropy of $(\Sigma ,\sigma )$ and $\mathcal {L}=\mathcal {L}(\psi ,\cdot )$ is the Legendre transform (as defined in (2.2) above). By [Reference Cantrell16, Lemma 3.3] we know that $\mathcal {I}$ is also the Legendre transform of the Manhattan curve for $d_S, d_{S_\ast }$ . In particular,

$$\begin{align*}\alpha_{\min}=\inf_{\sigma^n(x) = x} \frac{\psi^n(x)}{n} = \mathrm{Dil}(S,S_\ast)^{-1} \ \text{ and } \ \alpha_{\max}=\sup_{\sigma^n(x) = x} \frac{\psi^n(x)}{n} =\mathrm{Dil}(S_\ast,S) \end{align*}$$

where the infimum/supremum is taken over all periodic orbits in $\Sigma _{\mathcal {C}}$ . Now, by Theorem 3.2 we obtain the same limits with shrinking intervals, that is, $\eta \in (\alpha _{\min }, \alpha _{\max })$

$$\begin{align*}\limsup_{n\to\infty} \frac{1}{n} \log\left( \#\left\{ x \in \Sigma: \sigma^n(x) = x \text{ and } \left| \frac{\psi^n(x)}{n} - \eta \right| \le \frac{C}{n^2} \right\} \right) = \mathcal{I}(\eta) \end{align*}$$

for some C independent of $\eta $ . It is possible that the system $(\Sigma _{\mathcal {C}}, \sigma )$ is not mixing but is instead transitive. This is no issue as explained in Remark 3.3. Now note that each periodic orbit has a corresponding conjugacy class as described above: if $z = (z_0, z_1,\ldots , z_{n-1},z_0, \ldots )\in \Sigma $ is a periodic orbit of period n then we associate to it the conjugacy class $[g]$ with $\ell _{S}[g] = n$ where g is the group element obtained by multiplying the labelings along the path $z_0, z_1, \ldots , z_{n-1}$ . Furthermore [Reference Cantrell and Reyes17, Lemma 4.2] says that the periodic orbits of period n overcount the number of conjugacy classes by at most a linear factor in n. Hence we must have

$$\begin{align*}\limsup_{n\to\infty} \frac{1}{n} \log \left(\#\left\{ [g] \in \mathbf{conj}: \ell_S[g] = n, | \ell_{S_\ast}[g] - \eta \ell_{S}[g] | < \epsilon \right\}\right) = \mathcal{I}(\eta) \le v_S \end{align*}$$

for any $\epsilon> 0$ and $\eta \in (\alpha _{\min }, \alpha _{\max })$ (note that $h=v_S$ ). By [Reference Cantrell and Tanaka19, Theorem 1.1] we have that $\mathcal {I}$ is maximized (and equals $v_S$ ) when

$$\begin{align*}\eta = \lim_{T\to\infty} \frac{1}{\#\{[g]\in \mathbf{conj} \colon \ell_S[g]<T\}} \sum_{\ell_S[g] < T} \frac{\ell_{S_\ast}[g]}{T}. \end{align*}$$

This concludes the proof.

As a consequence we deduce Corollary 1.9.

Proof of Corollary 1.9

The first part of the theorem, when $\eta $ is irrational, follows as a direct corollary of Theorem 1.8. To deduce the final statement when $\eta $ is rational it suffices to show that, when $(\Sigma , \sigma )$ is a mixing subshift of finite type and $\psi : \Sigma \to {\mathbb {Z}}$ is a function that is constant of $2$ -cylinders, then the following holds: if $p/q \in (\alpha _{\min }, \alpha _{\max }) \cap {\mathbb {Q}}$ then there exists a periodic orbit $x \in \Sigma $ such that $\psi ^{|x|}(x) = |x|p/q.$ It is an easy exercise to verify this and so we leave it to the reader.

Example 4.2. In this example we show how to apply Theorem 1.8 to a pair of word metrics on a free group. We calculate the exact value of $\mathcal {I}$ (evaluated at a natural value) and provide explicit constants which determine how similar the length spectra of the two word metrics are.

Let $F_2 = \langle a, b \rangle $ be the free group on two generators and consider the generating sets

$$\begin{align*}S = \{a,b,a^{-1}, b^{-1}\} \ \text{ and } \ S_\ast = \{a,b,ab,a^{-1}, b^{-1}, (ab)^{-1} \}. \end{align*}$$

The corresponding word metrics have exponential growth rates $v_{S_\ast }= \log (4)$ and $v_S = \log (3)$ . The Manhattan curve $\theta _{S/S_\ast }$ was computed in [Reference Cantrell and Tanaka19] and is given by

$$\begin{align*}\theta_{S/S_\ast}(t) = \log\left( \frac{1}{2} e^{-t} \left(e^{-t} + \sqrt{e^{-t}(e^{-t}+8)} + 4\right)\right). \end{align*}$$

From the definition we see that $\mathcal {I}(v_{S_\ast }/v_S) = \log (4) - \Lambda /\log (3)$ where $\Lambda $ is the constant

$$\begin{align*}\Lambda = \sup_{t\in{\mathbb{R}}} \left\{\log(4) - \frac{\log(4)}{\log(3)} t - \theta_{S/S_\ast}(t) \right\}. \end{align*}$$

This can be computed (using, say Wolfram Alpha) to be

$$\begin{align*}\log(16) \log\left(\log\left(\frac{4}{3}\right)\right) + \log(9) \log\left(\frac{\log\left(\frac{3}{2}\right)}{\log(\frac{4}{3})}\right) + \log(4) \left(\log(2) - \log\left(\log\left(\frac{3}{2}\right) \log(2)\right)\right). \end{align*}$$

Theorem 1.8 then implies that there is $C>0$ such that

$$ \begin{align*} \limsup_{T\to\infty} \frac{1}{T} \log \left(\#\left\{ [g] \in \mathbf{conj}: \ell_{S_\ast}[g] < T, |v_S\ell_S[g] - v_{S_\ast}\ell_{S_\ast}[g]| \le \frac{C}{T} \right\}\right) &= \log(4) - \frac{\Lambda}{\log(3)} \\ &\approx 1.3679878759 \end{align*} $$

and similarly we also have

$$ \begin{align*} \limsup_{T\to\infty} \frac{1}{T} \log \left(\#\left\{ [g] \in \mathbf{conj}: \ell_{S}[g] < T, |v_S\ell_S[g] - v_{S_\ast}\ell_{S_\ast}[g]| \le \frac{C}{T} \right\}\right) &= \log(3) - \frac{\Lambda}{\log(4)}\\ &\approx 1.0841047424. \end{align*} $$

In this case we can set $k = 6, M=2, \alpha _{\max } = 2, \alpha _{\min } =1$ by [Reference Cantrell and Tanaka19]. Also, it was computed in [Reference Cantrell and Tanaka19] that $ -\theta ^{\prime }_{S/S_\ast }(0) = 4/3$ . We therefore have that

$$\begin{align*}\frac{4M^2(1+k^2)^2(\alpha_{\max} - \alpha_{\min}) }{\sqrt{5}} = \frac{4(2^2)(1+6^2)(2-1)}{\sqrt{5}} = \frac{592}{\sqrt{5}} \approx 264.7 \le 300 \end{align*}$$

and

$$\begin{align*}\limsup_{T\to\infty} \frac{1}{T} \log \left(\#\left\{ [g] \in \mathbf{conj}: \ell_{S_\ast}[g] < T, |3\ell_S[g] - 4 \ell_{S_\ast}[g]| \le \frac{300}{T} \right\}\right) = v_{S_\ast}. \end{align*}$$

Therefore the length spectra of $3d_S$ and $4d_{S_\ast }$ are within distance $300$ on a set of full exponential growth rate for $d_{S_\ast }$ . In particular, for any $\eta \in (1,2)$ we can find an infinite sequence of conjugacy classes $[g_n] \in \mathbf {conj}(\Gamma )$ such that

$$\begin{align*}\left| \frac{\ell_{S}[g_n]}{\ell_{S_\ast}[g_n]} - \eta\right| \le \frac{300}{|g_n|_{S_\ast}^2}. \end{align*}$$

5 Encoding cubulations via finite-state automata

In this section we study further the class $\mathfrak {G}$ defined in the introduction. Recall that a group belongs to $\mathfrak {G}$ if it is not virtually cyclic, it admits a virtually co-special cubulation with a contracting element, and its set of convex-cocompact subgroups is independent of the cubulation. See Subsection 2.2 for the notions of virtual co-specialness and convex-cocompact subgroups. We prove Proposition 5.2, which provides plenty of examples of groups in this class. Then we construct a finite-state automaton that encodes a pair of cubulations of a group in $\mathfrak {G}$ , our main result being Theorem 5.11, that implies most of the claims in Theorem 1.4 from the introduction. This will be done in the greater generality of the class $\mathfrak {X}$ of pairs of compatible cubulations. Theorem 5.11 is also key to prove our main results in Section 6.

5.1 The classes $\mathfrak {G}$ and $\mathfrak {X}$

Throughout this and the next section we will work with the following notion of compatibility of pairs of group actions on $\mathrm {CAT}(0)$ cube complexes.

Definition 5.1. Let $\mathfrak {X}$ be the class of triplets $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )$ , where $\Gamma $ is a nonvirtually cyclic group acting cocompactly on the $\mathrm {CAT}(0)$ cube complexes $\mathcal {X},\mathcal {X}_\ast $ and satisfying:

  1. (1) the action on $\mathcal {X}$ is proper and virtually co-special;

  2. (2) every hyperplane stabilizer for the action on $\mathcal {X}_\ast $ is convex-cocompact for the action on $\mathcal {X}$ ; and,

  3. (3) the action of $\Gamma $ on $\mathcal {X}$ has a contracting element.

Note that in the definition above we do not require the action on $\mathcal {X}_\ast $ to be proper.

Since hyperplane stabilizers are always convex-cocompact and being a contracting element does not depend on the cubulation [Reference Genevois39, Lemma 4.6], it follows that $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )\in \mathfrak {X}$ whenever $\Gamma \in \mathfrak {G}$ and $\mathcal {X},\mathcal {X}_\ast $ are cubulations of $\Gamma $ .

If $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )\in \mathfrak {X}$ , we should think of $\mathcal {X}$ and $\mathcal {X}_\ast $ as “compatible” cubulations of $\Gamma $ with well-behaved counting properties. Compatibility is guaranteed by (2), by means of Proposition 5.7 that allows us to combine $\mathcal {X}$ and $\mathcal {X}_\ast $ into a new cubulation of $\Gamma $ . Then (1) implies that the new cubulation is also virtually co-special, allowing us to use the automaton from Example 2.5. This automaton is upgraded in Theorem 5.11 to an automaton that “remembers” both $\mathcal {X}$ and $\mathcal {X}_\ast $ . Then (3) is used to relate counting computations over this automaton to counting properties of the length functions $\ell _{\mathcal {X}},\ell _{\mathcal {X}_\ast }$ over conjugacy classes of $\Gamma $ . This counting is done in Section 6.

The main result of this subsection is the following proposition, which tells us that the class $\mathfrak {G}$ is much bigger than the class of cubulable hyperbolic groups.

Proposition 5.2. The following classes of groups are contained in $\mathfrak {G}$ .

  1. i) Hyperbolic cubulable groups that are nonelementary.

  2. ii) Right-angled Artin groups of the form $\Gamma =A_G$ , where G is a finite graph with more than one vertex, that is not a join and such that $\mathrm {st}(v)$ is not contained in $\mathrm {st}(w)$ for every pair of distinct vertices $v,w$ of G.

  3. iii) Right-angled Coxeter groups that are not virtually cyclic or direct products and are of the form $\Gamma =W_G$ , where G is finite and does not have any loose squares.

Moreover, if $\Gamma $ is cubulable and hyperbolic relative to groups belonging to $\mathfrak {G}$ , then $\Gamma $ belongs to $\mathfrak {G}$ .

For $\Gamma =W_G$ a right-angled Coxeter group, a loose square is a full subgraph $\Delta \subset G$ that is a square, and such that for every maximal subgraph $\Lambda \subset G$ with $W_\Lambda $ virtually abelian, either $\Delta \subset \Lambda $ or $\Delta \cap \Lambda $ generates a finite subgroup of $\Gamma $ .

Many nonrelatively hyperbolic right-angled Artin and Coxeter groups satisfy assumptions ii) and iii) of Proposition 5.2. Indeed, a RAAG is nontrivially relatively hyperbolic if and only if it is a nontrivial free product, and hence RAAGs with underlying graphs n-agons for $n\geq 5$ are not relatively hyperbolic and belong to $\mathfrak {G}$ .

For the proof of Proposition 5.2 we require two results, the first one being an observation about virtual specialness.

Lemma 5.3. Let $\mathcal {X},\mathcal {X}_\ast $ be cubulations of the group $\Gamma $ and assume that the action on $\mathcal {X}$ is virtually co-special. If every hyperplane stabilizer for the action on $\mathcal {X}_\ast $ is convex-cocompact with respect to $\mathcal {X}$ , then the action of $\Gamma $ on $\mathcal {X}_\ast $ is virtually co-special.

Proof. If $\mathcal {X}$ and $\mathcal {X}_\ast $ satisfy the assumptions of the lemma, then all the double cosets of the hyperplane stabilizers for the action of $\Gamma $ on $\mathcal {X}_\ast $ are separable by [Reference Reyes71, Theorem A.1]. Then the action of on $\mathcal {X}_\ast $ is virtually co-special by the double-cosets criterion [Reference Haglund and Wise47, Theorem 9.19].

The second result we need is a criterion of convex-cocompactness for subgroups of cubulable relatively hyperbolic groups, which may be of independent interest and whose proof is postponed to the appendix.

Proposition 5.4. Let $\Gamma $ be a relatively hyperbolic group acting properly and cocompactly on the $\mathrm {CAT}(0)$ cube complex $\mathcal {X}$ . Then the following are equivalent for a subgroup $H< \Gamma $ .

  1. (1) H is convex-cocompact for the action of $\Gamma $ on $\mathcal {X}$ .

  2. (2) H is relatively quasiconvex and $H\cap P$ is convex-cocompact for the action of $\Gamma $ on $\mathcal {X}$ for any maximal parabolic subgroup $P<\Gamma $ .

Proof of Proposition 5.2

By Agol’s Theorem [Reference Agol1, Theorem 1.1] every cubulation of a hyperbolic group is virtually co-special. Moreover, the class of convex-cocompact subgroups for any such cubulation coincides with the class of quasiconvex subgroups [Reference Haglund and Wise47, Proposition 7.2]. Since any loxodromic element in a hyperbolic group is contracting, this solves the proposition for groups in class i).

Now, let $\Gamma $ be a group belonging to the class ii) (respectively iii)). In [Reference Fioravanti, Levcovitz and Sageev35, Theorem A] (resp. [Reference Fioravanti, Levcovitz and Sageev35, Corollary C]) it was proven that $\Gamma $ has a unique cubical coarse median structure. By [Reference Fioravanti, Levcovitz and Sageev35, Theorem 2.15] this uniqueness result is equivalent to the class of convex-cocompact subgroups being independent of the cubulation. Since finitely generated right-angled Artin (resp. Coxeter) groups are virtually special, $\Gamma $ satisfies Items (1) and (2) of Definition 1.1. The existence of a contracting element for $\Gamma $ for a geometric group action on a $\mathrm {CAT}(0)$ cube complex follows from [Reference Yang88, Section 5.2] and the references therein.

To prove the moreover statement, let $\Gamma $ be a group that is hyperbolic relative to groups belonging to $\mathfrak {G}$ and let $\mathcal {X},\mathcal {X}_\ast $ be two cubulations of $\Gamma $ . If $P<\Gamma $ is a maximal parabolic subgroup, then by [Reference Sageev and Wise76, Theorem 1.1] it has convex cores $Z_P\subset \mathcal {X}$ and $(Z_P)_\ast \subset \mathcal {X}_\ast $ . Also, since P belongs to $\mathfrak {G}$ , by Lemma 5.3 the action of P on $Z_P$ is virtually co-special. As this holds for every maximal parabolic subgroup, the action of $\Gamma $ on $\mathcal {X}$ is virtually co-special either by [Reference Groves and Manning41, Theorem A] or [Reference Reyes71, Theorem 1.2].

Consider a group $H<\Gamma $ that is convex-cocompact for the action of $\Gamma $ on $\mathcal {X}$ . By Proposition 5.4, H is relatively quasiconvex and $H\cap P$ is a convex-cocompact subgroup for any maximal parabolic subgroup $P<\Gamma $ . This implies that $H\cap P<P$ is convex-cocompact for the action of P on $Z_P$ , see, for example, Lemmas 2.14 and 2.15 in [Reference Reyes71]. But each such P belongs to $\mathfrak {G}$ , so $H\cap P$ is also convex-cocompact for the action of P on $(Z_P)_\ast $ , implying that $H\cap P$ is convex-cocompact for the action of $\Gamma $ on $\mathcal {X}_\ast $ . By Proposition 5.4 we deduce that H is convex-cocompact for the action of $\Gamma $ on $\mathcal {X}_\ast $ , so that the actions on $\mathcal {X}$ and $\mathcal {X}_\ast $ have the same sets of convex-cocompact subgroups. Since any contracting element in a maximal parabolic subgroup is contracting for $\Gamma $ , we have proven that $\Gamma $ belongs to $\mathfrak {G}$ .

Example 5.5. Let M be a cusped hyperbolic 3-manifold with cusps $V_1,\dots ,V_r\subset M$ . We affirm that $\Gamma =\pi _1(M)$ does not belong to $\mathfrak {G}$ . For each $i=1,\dots ,r$ choose distinct slopes $\alpha _i,\beta _i$ for the cusp $V_i$ . Equivalently, each pair $\alpha _i,\beta _i$ represents a pair of cyclic subgroups (up to commensurability) such that their union generates a finite-index subgroup of $\pi _1(V_i)$ . By [Reference Cooper and Futer24, Corollary 1.3], there exists a cubulation $\mathcal {X}$ of $\Gamma $ such that each $\alpha _i$ and $\beta _i$ represents a convex-cocompact subgroup. Since any cubulation of ${\mathbb {Z}}^2$ has a subgroup that is not convex-cocompact (this follows for instance from [Reference Wise and Woodhouse87, Theorem 3.6]), for $\Gamma $ as above we can produce two cubulations $\mathcal {X},\mathcal {X}_\ast $ not satisfying (2) in Definition 1.1.

However, the data of the slopes that are convex-cocompact in cubulations of $\Gamma $ are the only obstruction for them determining triplets in $\mathfrak {X}$ . That is, if $\mathcal {X},\mathcal {X}_\ast $ are two cubulations of $\Gamma $ , then $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )\in \mathfrak {X}$ if and only if they have the same pairs of convex-cocompact slopes for each cusp subgroup. The proof of this follows the same lines as the proof of Proposition 5.2.

Example 5.6. There exist groups $\Gamma $ not necessarily belonging to $\mathfrak {G}$ for which we still can find essentially distinct cubulations $\mathcal {X},\mathcal {X}_\ast $ such that $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )\in \mathfrak {X}$ .

As an example, let $\Gamma $ be a finite graph product of finite groups and let $\mathcal {X}$ be the graph-product complex with the standard action by $\Gamma $ . If $\phi \in {Aut}(\Gamma )$ is any automorphism, then we let $\mathcal {X}_\ast $ be the cubulation obtained by precomposing by $\phi $ the action of $\Gamma $ on $\mathcal {X}$ . Then, as long as $\Gamma $ has a contracting element, $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )\in \mathfrak {X}$ by combining Theorem D (1) and Theorem 2.15 in [Reference Fioravanti, Levcovitz and Sageev35]. We note that there are many right-angled Coxeter groups with large outer automorphism groups [Reference Sale and Susse77].

If instead we apply [Reference Fioravanti, Levcovitz and Sageev35, Theorem D (2)], the same conclusion holds for $\Gamma $ any Coxeter group with a contracting element, with $\mathcal {X}$ being its Niblo-Reeves cubulation, and $\mathcal {X}_\ast $ obtained from $\mathcal {X}$ after twisting by an automorphism of $\Gamma $ .

5.2 Constructing the appropriate automaton

In this subsection we construct a finite-state automaton for a triplet $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )$ in $\mathfrak {X}$ , our main result being Theorem 5.11. This is the automaton we will use to prove our main results for the length functions $\ell _{\mathcal {X}}^{\mathfrak {w}}, \ell ^{\mathfrak {w}_\ast }_{\mathcal {X}_\ast }$ in Section 6, and it implies most of Theorem 1.4. This automaton can be thought of as a cubical analog of the automaton for pairs of word metrics on hyperbolic groups constructed by Calegari-Fujiwara in [Reference Calegari and Fujiwara14, Lemma 3.8] and explained in Example 2.4.

Our starting point in the construction is the automaton for special cube complexes highlighted in Example 2.5. To use this automaton, our first step is the construction of a (virtually co-special) cubulation for $\Gamma $ that simultaneously encodes the actions on $\mathcal {X}$ and $\mathcal {X}_\ast $ . This is the content of the next proposition.

Proposition 5.7. Given $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )\in \mathfrak {X}$ there exists a proper, cocompact, and essential action of $\Gamma $ on a $\mathrm {CAT}(0)$ cube complex $\mathcal {Z}$ , $\Gamma $ -invariant subsets $\mathbb {W},\mathbb {W}_\ast \subset \mathbb {H}(\mathcal {Z})$ such that $\mathbb {W} \cup \mathbb {W}_\ast =\mathbb {H}(\mathcal {Z})$ , and a finite index subgroup $\overline {\Gamma }<\Gamma $ satisfying the following.

  • Let $\hat {\mathcal {X}}=\mathcal {Z}(\mathbb {W})$ and $\hat {\mathcal {X}}_\ast =\mathcal {Z}(\mathbb {W}_\ast )$ . Then $\hat {\mathcal {X}}$ and $\hat {\mathcal {X}}_\ast $ embed $\Gamma $ -equivariantly as convex subcomplexes of $\mathcal {X}$ and $\mathcal {X}_\ast $ respectively. In particular, the triplet $(\Gamma ,\hat {\mathcal {X}},\hat {\mathcal {X}}_\ast )$ belongs to $\mathfrak {X}$ and we have the equalities $\ell _{\mathcal {X}}= \ell _{\hat {\mathcal {X}}}$ and $\ell _{\mathcal {X}_\ast }=\ell _{\hat {\mathcal {X}}_\ast }$ .

  • The action of $\overline {\Gamma }$ on both $\mathcal {Z}$ and $\hat {\mathcal {X}}$ is co-special.

Remark 5.8. As we will see in the proof, the first conclusion of the proposition above only uses Item (2) in Definition 5.1. Item (1) is only used to find the finite index subgroup $\overline {\Gamma }$ satisfying the second conclusion.

The proof of Proposition 5.7 uses the formalism of median algebras, from which refer the reader to [Reference Fioravanti34, Subsection 2.1.5]. We require the following lemma, which is a slight generalization of the implication $(3) \Rightarrow (1)$ in [Reference Fioravanti34, Proposition 7.9].

Lemma 5.9. Let $\Gamma $ act on the $\mathrm {CAT}(0)$ cube complexes $\mathcal {X},\mathcal {X}_\ast $ so that action on $\mathcal {X}$ is proper and cocompact, and the action on $\mathcal {X}_\ast $ is essential and has only finitely many orbits of hyperplanes. If every hyperplane stabilizer for the action of $\Gamma $ on $\mathcal {X}_\ast $ is convex-cocompact for the action on $\mathcal {X}$ , then for any finite subset $F\subset \mathcal {X}^0\times \mathcal {X}_\ast ^0$ the median algebra generated by the $\Gamma $ -translates of F is $\Gamma $ -cofinite.

If $\Gamma $ acts on a $\mathrm {CAT}(0)$ cube complex and $\mathfrak {h}$ is a hyperplane, recall that $\Gamma _{\mathfrak {h}}$ denotes the hyperplane stabilizer of $\mathfrak {h}$ .

Proof. Without loss of generality, suppose that $F=P\times P_\ast $ for $P,P_\ast $ some finite sets of vertices. The main idea in the proof of $(3) \Rightarrow (1)$ in [Reference Fioravanti34, Proposition 7.9] is the construction, for each hyperplane $\frak h \in \mathbb {H}(\mathcal {X}_\ast )$ with halfspaces $\mathfrak {h}^+$ and $\mathfrak {h}^-$ , of a partition $C(\mathfrak {h}^-)\sqcup C(\mathfrak {h}) \sqcup C(\mathfrak {h}^+)$ of $\mathcal {X}$ such that:

  • $C(\mathfrak {h})$ is a $\Gamma _{\mathfrak {h}}$ -invariant convex subcomplex of $\mathcal {X}$ and the action of $\Gamma _{\mathfrak {h}}$ on $C(\mathfrak {h})$ is cocompact;

  • $C(\mathfrak {h}^+)$ and $C(\mathfrak {h}^-)$ are $\Gamma _{\mathfrak {h}}$ -invariant unions of connected components of $\mathcal {X} \backslash C(\mathfrak {h})$ ;

  • if $g\in \Gamma $ , $y\in P_\ast $ and $gy \in \mathfrak {h}^+$ , then $gx\in C(\mathfrak {h}^+)\cup C(\mathfrak {h})$ for all $x\in P$ ;

  • if $g\in \Gamma $ , $y\in P_\ast $ and $gy\in \mathfrak {h}^-$ , then $gx\in C(\mathfrak {h}^-)\cup C(\mathfrak {h})$ for all $x\in P$ ; and,

  • for any $g \in \Gamma $ we have $C(g\mathfrak {h})=gC(\mathfrak {h})$ and $C(g\mathfrak {h}^+)=gC(\mathfrak {h}^+)$ .

Let $M\subset \mathcal {X}^0 \times \mathcal {X}_\ast ^0$ be the median algebra generated by F, so that any wall of M is induced by a hyperplane in $\mathcal {X}$ or $\mathcal {X}_\ast $ . Consider two transverse walls $\mathfrak {v},\mathfrak {w}$ of M with $\mathfrak {v}$ induced by $\mathfrak {h} \in \mathbb {H}(\mathcal {X}_\ast )$ and $\mathfrak {w}$ induced by $\mathfrak {l}\in \mathbb {H}(\mathcal {X}) \sqcup \mathbb {H}(\mathcal {X}_\ast )$ . As in the proof of $(3) \Rightarrow (1)$ in [Reference Fioravanti34, Proposition 7.9], we can verify that $\mathfrak {l} \in \mathbb {H}(\mathcal {X})$ implies $C(\mathfrak {h})\cap \mathfrak {l} \neq \emptyset $ , and $\mathfrak {l} \in \mathbb {H}(\mathcal {X}_\ast )$ implies $C(\mathfrak {h})\cap C(\mathfrak {l}) \neq \emptyset $ . Then we can argue as in the proof of [Reference Fioravanti34, Proposition 7.9] to conclude that M is the 0-skeleton of a $\mathrm {CAT}(0)$ cube complex on which the induced action of $\Gamma $ is cocompact, implying that M is $\Gamma $ -cofinite.

On the other hand, the construction of the partitions $C(\mathfrak {h}^-)\sqcup C(\mathfrak {h}) \sqcup C(\mathfrak {h}^+)$ is possible by the following slight generalization of [Reference Fioravanti34, Lemma 7.11]: if $H < \Gamma $ is convex-cocompact for the action on $\mathcal {X}$ and $A\subset \Gamma $ is an H-almost invariant set (see [Reference Fioravanti34, Lemma 7.11]), then there exists a partition $\mathcal {X}=C_- \sqcup C_0 \sqcup C_+$ such that:

  • $C_0$ is an H-invariant convex subcomplex and the action of H on $C_0$ is cocompact;

  • $C_-$ and $C_+$ are H-invariant unions of connected components of $X \backslash C_0$ ; and,

  • $A\cdot P \subset C_0 \cup C_+$ and $(\Gamma \backslash A) \cdot P \subset C_0 \cup C_-$ .

The proof of this generalization follows from the expected modifications of the proof of [Reference Fioravanti34, Lemma 7.11] and is left to the reader. From this, for a halfspace $\mathfrak {h}^+$ of $\mathcal {X}_\ast $ with bounding hyperplane $\mathfrak {h}$ in a complete set of representatives of $\Gamma $ -orbits of hyperplanes, we consider $(C(\mathfrak {h}^+),C(\mathfrak {h}),C(\mathfrak {h}^-))=(C_+,C_0,C_-)$ for $A=\{g\in \Gamma \colon gy\in \mathfrak {h}^+ \text { for some }y\in P_\ast \}$ . We note that A is a $\Gamma _{\mathfrak {h}}$ -almost invariant set by [Reference Fioravanti34, Remark 7.12]. The proof of the lemma concludes after extending this construction $\Gamma $ -equivariantly.

Proof of Proposition 5.7

By [Reference Caprace and Sageev20, Proposition 3.5], let $\hat {\mathcal {X}}$ and $\hat {\mathcal {X}}_\ast $ be the $\Gamma $ -essential cores of $\mathcal {X}$ and $\mathcal {X}_\ast $ respectively, which $\Gamma $ -equivariantly embed in $\mathcal {X}$ and $\mathcal {X}_\ast $ as convex subcomplexes.

We claim that there exists a cubulation $\mathcal {Z}'$ of $\Gamma $ and $\Gamma $ -invariant subsets $\mathbb {W},\mathbb {W}_\ast \subset \mathbb {H}(\mathcal {Z}')$ so that $\hat {\mathcal {X}}$ and $\hat {\mathcal {X}}_\ast $ are $\Gamma $ -equivariantly isometric to $\mathcal {Z}'(\mathbb {W})$ and $\mathcal {Z}'(\mathbb {W}_\ast )$ . Under the additional assumption that the action on $\mathcal {X}_\ast $ is proper, this is the content of the implication $(3) \Rightarrow (4)$ in [Reference Fioravanti, Levcovitz and Sageev35, Theorem 2.17], so we now explain how to use Lemma 5.9 the prove the general case.

Let $P\subset \hat {\mathcal {X}}^0$ and $P_\ast \subset \hat {\mathcal {X}}_\ast ^0$ be the vertex sets of compact connected subcomplexes $K,K_\ast $ such that $\Gamma \cdot K=\hat {\mathcal {X}}^0$ and $\Gamma \cdot K_\ast =\hat {\mathcal {X}}_\ast ^0$ . By Lemma 5.9, the median algebra $M\subset \hat {\mathcal {X}}^0\times \hat {\mathcal {X}}_\ast ^0$ generated by the $\Gamma $ -translates of $P\times P_\ast $ is $\Gamma $ -cofinite. By Chepoi–Roller duality [Reference Chepoi23, Reference Roller72], M is the 0-skeleton of a $\mathrm {CAT}(0)$ cube complex $\mathcal {Z}'$ equipped with a proper and cocompact cubical action of $\Gamma $ . The restriction to M of the natural projection $\hat {\mathcal {X}}^0 \times \hat {\mathcal {X}}_\ast ^0 \rightarrow \hat {\mathcal {X}}^0$ then induces a $\Gamma $ -equivariant cubical map $\mathcal {Z}' \rightarrow \hat {\mathcal {X}}$ that is surjective because M contains $K \times K_\ast $ . This map is also a median morphism on the 0-skeleton, hence a restriction quotient by the discussion preceding the proof of Theorem 2.17 in [Reference Fioravanti, Levcovitz and Sageev35]. The same argument gives us a restriction quotient $\mathcal {Z}' \rightarrow \hat {\mathcal {X}}_\ast $ .

To finish the proof of the first assertion, note that the action of $\Gamma $ on $\mathcal {Z}'$ may not be essential, so instead we consider the projection quotient $\mathcal {Z}=\mathcal {Z}'(\mathbb {W}\cup \mathbb {W}_\ast )$ , which is essential and cocompact since the action of $\Gamma $ on both $\hat {\mathcal {X}}$ and $\hat {\mathcal {X}}_\ast $ is essential and the action on $\mathcal {Z}'$ is cocompact. The complex $\mathcal {Z}$ still $\Gamma $ -equivariantly projects onto $\hat {\mathcal {X}}$ and $\hat {\mathcal {X}}_\ast $ , so the action of $\Gamma $ on $\mathcal {Z}$ is proper because the action of $\Gamma $ on $\hat {\mathcal {X}}$ is proper.

To prove the second assertion, note that by [Reference Fioravanti, Levcovitz and Sageev35, Theorem 2.17] the actions of $\Gamma $ on $\hat {\mathcal {X}}$ and $\mathcal {Z}$ have the same sets of convex-cocompact subgroups, so any hyperplane stabilizer for the action of $\Gamma $ on $\mathcal {Z}$ will be convex-cocompact with respect to $\hat {\mathcal {X}}$ . But the action of $\Gamma $ on $\hat {\mathcal {X}}$ is virtually co-special (it is a convex core for $\Gamma $ acting on $\mathcal {X}$ ), so $\mathcal {Z}$ is virtually co-special by Lemma 5.3. Finally, co-specialness is preserved under taking finite-index subgroups, so we can choose $\overline {\Gamma }$ so that both quotients $\overline {\Gamma } \backslash \mathcal {Z}$ and $\overline {\Gamma } \backslash \hat {\mathcal {X}}$ are special.

By virtue of the proposition above, throughout the rest of the section we will work under the following convention.

Convention 5.10. $\Gamma $ is a nonvirtually cyclic group acting properly, cocompactly and essentially on the $\mathrm {CAT}(0)$ cube complex $\mathcal {Z}$ , and $\mathbb {W}\subset \mathbb {H}(\mathcal {Z})$ is a $\Gamma $ -invariant subset such that the action of $\Gamma $ on $\mathcal {X}=\mathcal {Z}(\mathbb {W})$ is proper, cocompact and essential. Let $\phi :\mathcal {Z} \rightarrow \mathcal {X}$ be the restriction quotient.

Also, let $\overline {\Gamma }<\Gamma $ be a finite index subgroup such that the quotients $\overline {\mathcal {Z}}=\overline {\Gamma } \backslash \mathcal {Z}$ and $\overline {\mathcal {X}}=\overline {\Gamma } \backslash \mathcal {X}$ are special cube complexes. We fix a base vertex $\widetilde o\in \mathcal {Z}$ and set $o=\phi (\widetilde o)\in \mathcal {X}$ .

Let $S_{\overline {\mathcal {Z}}}$ and $S_{\overline {\mathcal {X}}}$ be the set of oriented hyperplanes in $\overline {\mathcal {Z}}$ and $\overline {\mathcal {X}}$ respectively. By specialness all the hyperplanes in $\overline {\mathcal {Z}}$ and $\overline {\mathcal {X}}$ are 2-sided, so there exist two orientations for each hyperplane. Since each hyperplane in $S_{\overline {\mathcal {X}}}$ corresponds to a $\overline {\Gamma }$ -orbit of oriented hyperplanes in $\mathbb {W}\subset \mathbb {H}(\mathcal {Z})$ , there is a natural injection of $S_{\overline {\mathcal {X}}}$ into $S_{\overline {\mathcal {Z}}}$ , so often we will consider $S_{\overline {\mathcal {X}}}$ as a subset of $S_{\overline {\mathcal {Z}}}$ . The label of an oriented hyperplane in $\mathcal {Z}$ (resp. $\mathcal {X}$ ) is its projection in $\overline {\mathcal {Z}}$ (resp. $\overline {\mathcal {X}}$ ).

We say that a word $w=\mathfrak {h}_1\cdots \mathfrak {h}_n$ in $(S_{\overline {\mathcal {Z}}})^\ast $ is represented by a (combinatorial) path $\gamma =(\gamma _0,\dots ,\gamma _n)$ in $\mathcal {Z}$ if for each i the oriented hyperplane $\mathfrak {h}_i$ is the image in $\overline {\mathcal {Z}}$ of the oriented hyperplane dual to the edge from $\gamma _{i-1}$ to $\gamma _i$ . Similarly, we define when a word in $(S_{\overline {\mathcal {X}}})^\ast $ is represented by a path in $\mathcal {X}$ . A consequence of specialness is that if a word is represented by two paths with the same initial vertex, then the paths must coincide.

The next theorem gives us a finite-state automaton over $S_{\overline {\mathcal {X}}}$ that keeps track of the action of $\overline {\Gamma }$ on both $\mathcal {X}$ and $\mathcal {Z}$ .

Theorem 5.11. Let $\Gamma ,\mathcal {Z}$ and $\mathcal {X}$ satisfy Convention 5.10. Then there exists a language $L=L_{\overline {\Gamma },\phi }\subset (S_{\overline {\mathcal {X}}})^\ast $ parametrized by the pruned finite-state automaton

$$\begin{align*}\mathcal{A}_{\overline{\Gamma},\phi}=(\mathcal{G}_{\phi}=(V_{\phi},E_{\phi}),\pi_{\phi},I_{\phi},V_{\phi})\end{align*}$$

satisfying the following.

  1. (1) There exists $C\geq 1$ depending only on $\overline {\Gamma }$ and $\phi :\mathcal {Z} \rightarrow \mathcal {X}$ such that any $w\in L_{\overline {\Gamma },\phi }$ is represented by at most C paths in $\mathcal {G}_{\phi }$ starting at an initial state.

  2. (2) Every $w\in L_{\overline {\Gamma },\phi }$ is represented by a unique combinatorial geodesic $\gamma _w\subset \mathcal {X}$ starting at the vertex o. We let $\tau _{\mathcal {X}}(w)$ denote the final vertex of $\gamma _w$ .

  3. (3) The map $\tau _{\mathcal {X}}:L_{\overline {\Gamma },\phi }\rightarrow \mathcal {X}^0$ is a bijection.

Moreover, there exist maps $\Psi :V_\phi \rightarrow (S_{\overline {\mathcal {Z}}} \backslash S_{\overline {\mathcal {X}}})^\ast $ and $\Xi :V_\phi \rightarrow \overline {\mathcal {X}}^0$ satisfying the following.

  1. (4) If $w\in L_{\overline {\Gamma },\phi }$ is represented by the path $\omega =(v_0\xrightarrow {e_1} v_1\cdots \xrightarrow {e_n} v_n)$ in $\mathcal {G}_\phi $ starting at an initial state, then the concatenation

    (5.1) $$ \begin{align} \alpha(\omega):=\Psi(v_0)\pi_\phi(e_1)\Psi(v_1) \cdots \pi_\phi(e_n)\in (S_{\overline{\mathcal{Z}}})^* \end{align} $$
    can be represented by a unique geodesic path $\widetilde \gamma _{\alpha (\omega )}$ in $\mathcal {Z}$ with initial vertex $\widetilde o$ and final vertex $\tau _{\mathcal {Z}}(\alpha (\omega ))$ , so that $\phi ( \tau _{\mathcal {Z}}(\alpha (\omega )))=\tau _{\mathcal {X}}(w)$ .
  2. (5) If $w\in L_{\overline {\Gamma },\phi }$ is represented by the path $\omega =(v_0\rightarrow \cdots \rightarrow v_n)$ in $\mathcal {G}_\phi $ starting at an initial state, then the path $\gamma _w=(\gamma _0,\dots ,\gamma _n)$ in $\mathcal {X}$ projects to $(\Xi (v_0),\dots ,\Xi (v_n))$ in $\overline {\mathcal {X}}$ .

Recall from Subsection 2.3 that an automaton with underlying graph $\mathcal {G}$ is deterministic if any two edges of $\mathcal {G}$ with the same initial vertex have different labels, and that the automaton is pruned if any vertex in $\mathcal {G}$ is the final vertex of an admissible path.

Remark 5.12. i) The notation $L_{\overline {\Gamma },\phi }$ and $\mathcal {A}_{\Gamma ,\phi }$ in the theorem above is chosen to emphasize the dependence on $\overline {\Gamma },\mathcal {X}$ and $\mathcal {Z}$ , but also on the restriction quotient map $\phi :\mathcal {Z} \rightarrow \mathcal {X}$ . We have suppressed some notation for simplicity, but this is also the case for the data involved in the definition of $\mathcal {A}_{\overline {\Gamma },\phi }$ , as well as for $\tau _{\mathcal {X}},\tau _{\mathcal {Z}}, \Psi , \Xi $ and $\alpha $ . For simplicity, often we will use the simplified notation $L_\phi =L_{\overline {\Gamma },\phi }$ and $\mathcal {A}_{\phi }=\mathcal {A}_{\overline {\Gamma },\phi }$ . ii) The specialness of ${\overline {\mathcal {X}}}$ and ${\overline {\mathcal {Z}}}$ is used at several steps in the construction of the automaton $\mathcal {A}_{\overline {\Gamma },\phi }$ . Crucially, specialness of $\overline {\mathcal {Z}}$ allows us to use the automaton constructed by Li and Wise in [Reference Li and Wise58] (see Example 2.5). This automaton does not remember the quotient $\Gamma \backslash \mathcal {Z}$ covered by ${\overline {\mathcal {Z}}}$ , and in particular, the automaton $\mathcal {A}_{\overline {\Gamma },\phi }$ does not remember $\Gamma $ . iii). We note that Convention 5.10, and hence the construction of $\mathcal {A}_{\overline {\Gamma },\phi }$ does not assume that $\Gamma $ has a contracting element. The full strength of Item (3) in Definition 5.1 is used in Section 6 when we prove our counting theorems (compare with Convention 6.3). For these counting results, passing to the finite index subgroup $\overline {\Gamma }$ is not an issue, as it suffices for the automaton to see a subset of $\Gamma $ having positive lower density, see Lemma 6.11.

We start the construction of $\mathcal {A}_{\overline {\Gamma },\phi }$ by considering a regular language $L_{\overline {\mathcal {Z}}} \subset (S_{\overline {\mathcal {Z}}})^\ast $ parameterized by the (pruned and deterministic) automaton

$$\begin{align*}\mathcal{A}_{\overline{\mathcal{Z}}}=(\mathcal{G}_{\overline{\mathcal{Z}}}=(V_{\overline{\mathcal{Z}}},E_{\overline{\mathcal{Z}}}),\pi_{\overline{\mathcal{Z}}},\{\ast_{\overline{\mathcal{Z}}}\},V_{\overline{\mathcal{Z}}})\end{align*}$$

over $S_{\overline {\mathcal {Z}}}$ , constructed by Li and Wise in [Reference Li and Wise58] and discussed in Example 2.5. This automaton satisfies:

  • every $\widetilde w\in L_{\overline {\mathcal {Z}}}$ is represented by a unique geodesic path $\widetilde {\gamma }_{\widetilde w}$ in $\mathcal {Z}$ starting at $\widetilde o$ and ending at the vertex $\tau _{\mathcal {Z}}(\widetilde w)$ ; and,

  • the map $\tau _{\overline {\mathcal {Z}}}: L_{{\overline {\mathcal {Z}}}} \rightarrow \mathcal {Z}^0$ is a bijection.

We use the automaton $\mathcal {A}_{\overline {\mathcal {Z}}}$ to produce a language $L_\phi =L_{\overline {\Gamma },\phi }$ over the alphabet $S_{\overline {\mathcal {X}}}$ . We do this by first constructing an automaton

$$ \begin{align*}\hat{\mathcal{A}}_\phi=(\hat{\mathcal{G}}_\phi=(\hat{V}_\phi,\hat{E}_\phi),\hat{\pi}_\phi,\hat{I}_\phi,\hat{V}_\phi)\end{align*} $$

as follows.

Definition 5.13. Let $\hat {V}_\phi $ be the set of finite directed paths (possibly of length 0) in $\mathcal {G}_{\overline {\mathcal {Z}}}$ of the form $\omega =(v_0\xrightarrow {e_1}\cdots \xrightarrow {e_n} v_n)$ and satisfying:

  • $\pi _{\overline {\mathcal {Z}}}(e_i)\in S_{\overline {\mathcal {Z}}} \backslash S_{\overline {\mathcal {X}}}$ for every $1\leq i\leq n$ ;

  • either $v_0=\ast _{\overline {\mathcal {Z}}}$ or there exists an edge in $E_{\overline {\mathcal {Z}}}$ with label in $S_{\overline {\mathcal {X}}}$ and final vertex $v_0$ ; and,

  • either there exists an edge in $E_{\overline {\mathcal {Z}}}$ with label in $S_{\overline {\mathcal {X}}}$ and initial vertex $v_n$ or there are no edges in $E_{\overline {\mathcal {Z}}}$ with initial vertex $v_n$ .

We consider an edge $\hat {e}$ from the vertex $\omega $ to the vertex $\omega '$ in $\hat {V}_\phi $ if there exists an edge $e\in E_{\overline {\mathcal {Z}}}$ with $\pi _{\overline {\mathcal {Z}}}(e)\in S_{\overline {\mathcal {X}}}$ and such that the concatenation $\omega e \omega '$ is a path in $\mathcal {G}_{\overline {\mathcal {Z}}}$ . We define $\hat {\pi }_\phi (\hat {e}):=\pi _{\overline {\mathcal {Z}}}(e)$ and let $\hat {E}_\phi $ be the set of all of the edges defined in this way. Finally, a vertex of $\hat {V}_\phi $ is an initial state if its initial vertex (as a path in $\mathcal {G}_{\overline {\mathcal {Z}}}$ ) is $\ast _{\overline {\mathcal {Z}}}$ . We let $\hat {I}_\phi $ be the set of all the initial states.

Lemma 5.14. The set $\hat {V}_\phi $ is finite and nonempty. Therefore, $\hat {\mathcal {A}}_\phi $ defines a pruned finite-state automaton over $S_{\overline {\mathcal {X}}}$ .

Proof. The set $S_{\overline {\mathcal {X}}}$ is nonempty because $\Gamma $ is nonelementary and $\mathcal {X}$ is essential, and since $\mathcal {A}_{\overline {\mathcal {Z}}}$ is pruned we can find a vertex in $\hat {I}_\phi $ , so that $\hat {V}_\phi $ is nonempty.

To show finiteness, let M be the maximum cardinality of a preimage $\phi ^{-1}(x)\cap \mathcal {Z}^0$ among $x\in \mathcal {X}^0$ , which is finite since $\phi $ is $\Gamma $ -invariant and the action of $\Gamma $ on $\mathcal {Z}$ is cocompact. If $\omega =(v_0\xrightarrow {e_1}\cdots \xrightarrow {e_n} v_n)$ is a vertex in $\hat {V}_\phi $ , the fact that $\mathcal {A}_{\overline {\mathcal {Z}}}$ is pruned implies the existence of a geodesic path $\widetilde \gamma $ in $\mathcal {Z}$ representing the word $\pi _{\overline {\mathcal {Z}}}(e_1)\cdots \pi _{\overline {\mathcal {Z}}}(e_n)\in (S_{\overline {\mathcal {Z}}} \backslash S_{\overline {\mathcal {X}}})^\ast $ . Since $\phi $ collapses the hyperplanes not belonging to $\mathbb {W}$ , the image $\phi (\widetilde \gamma )$ consists of a single point, implying that $n+1 \leq M$ . We conclude that every vertex in $\hat {V}_\phi $ has uniformly bounded length as a path in $\mathcal {G}_{\overline {\mathcal {Z}}}$ , so $\hat {V}_\phi $ is finite because $\mathcal {G}_{\overline {\mathcal {Z}}}$ is.

Finally, $\mathcal {A}_{\overline {\mathcal {Z}}}$ being pruned implies that $\hat {\mathcal {A}}_\phi $ is pruned, and by construction this automaton is over the alphabet $S_{\overline {\mathcal {X}}}$ .

Definition 5.15. We let $L_\phi =L_{\overline {\Gamma },\phi } \subset (S_{\overline {\mathcal {X}}})^*$ be the language parametrized by $\hat {\mathcal {A}}_\phi $ .

If $\hat {\omega } =(\omega _0\xrightarrow {\hat {e}_1} \cdots \xrightarrow {\hat {e}_n} \omega _n)$ is a path in $\hat {\mathcal {G}}_\phi $ , then the concatenation

(5.2) $$ \begin{align} c(\hat{\omega}):=\omega_0e_1\cdots \omega_{n-1}e_n \end{align} $$

is a path in $\mathcal {G}_{\overline {\mathcal {Z}}}$ . Let $\hat \alpha (\hat {\omega })\in (S_{\overline {\mathcal {Z}}})^\ast $ be the word represented by $c(\hat {\omega })$ .

Lemma 5.16.

  1. (1) If $\hat {\omega },\hat {\omega }'$ are paths in $\hat {\mathcal {G}}_\phi $ starting at an initial state and representing the same word $w\in L_\phi $ , then $\phi (\tau _{\mathcal {Z}}(\hat {\alpha } (\hat {\omega })))=\phi (\tau _{\mathcal {Z}}(\hat {\alpha } (\hat {\omega }')))\in \mathcal {X}^0$ . We denote this vertex by $\tau _{\mathcal {X}}(w)$ .

  2. (2) Any w in $L_\phi $ is represented by a unique geodesic path $\gamma _w$ in $\mathcal {X}$ starting at o and ending at $\tau _{\mathcal {X}}(w)$ .

  3. (3) The map $\tau _{\mathcal {X}}:L_\phi \rightarrow \mathcal {X}^0$ is a bijection.

Proof. By induction on the length of w we will prove simultaneously assertion (1) and that $d_{\mathcal {X}}(o,\tau _{\mathcal {X}}(w))$ equals the length of w. Suppose that w has length n and is represented by the paths

$$\begin{align*}\hat{\omega} =(\omega_0\xrightarrow{\hat{e}_1} \cdots \xrightarrow{\hat{e}_n} \omega_n) \text{ and } \hat{\omega}' =(\omega_0'\xrightarrow{\hat{e}_1'} \cdots \xrightarrow{\hat{e}_n'} \omega_n')\end{align*}$$

in $\hat {\mathcal {G}}_\phi $ . If $n=0$ then $\hat {\alpha }(\hat {\omega })=\omega _0$ has no letters in $S_{\overline {\mathcal {X}}}$ , and hence the projection of $\widetilde \gamma _{\hat {\alpha }(\hat {\omega })}$ under $\phi $ consists of a single vertex, which must be o. As the same happens for $\omega '$ , this solves the base case.

If $n\geq 1$ we consider $\hat {\xi }=(\omega _0\xrightarrow {\hat {e}_1} \cdots \xrightarrow {\hat {e}_{n-1}} \omega _{n-1})$ and $\hat {\xi }'=(\omega _0'\xrightarrow {\hat {e}_1'} \cdots \xrightarrow {\hat {e}_{n-1}'} \omega _{n-1}')$ . Our inductive assumption implies that $\phi (\tau _{\mathcal {Z}}(\hat {\alpha } (\hat {\xi })))=\phi (\tau _{\mathcal {Z}}(\hat {\alpha } (\hat {\xi }')))=:x$ , so that no hyperplane with label in $S_{\overline {\mathcal {X}}}$ separates $\tau _{\mathcal {Z}}(\hat {\alpha } (\hat {\xi }))$ and $\tau _{\mathcal {Z}}(\hat {\alpha } (\hat {\xi }'))$ . Since $\overline {\mathcal {X}}$ is special there exists at most one edge in $\mathcal {X}$ with initial vertex x and dual to a hyperplane with label $\mathfrak {h}=\hat {\pi }_\phi (\hat {e}_{n})= \hat {\pi }_\phi (\hat {e}_{n}') \in S_{\overline {\mathcal {X}}}$ . But $\phi $ is injective on $\mathbb {W}$ , so that the hyperplane labeled $\mathfrak {h}$ that separates $\tau _{\mathcal {Z}}(\hat {\alpha } (\hat {\xi }))$ and $\tau _{\mathcal {Z}}(\hat {\alpha } (\hat {\omega }))$ in $\mathcal {Z}$ is the same as the hyperplane labeled $\mathfrak {h}$ that separates $\tau _{\mathcal {Z}}(\hat {\alpha } (\hat {\xi }'))$ and $\tau _{\mathcal {Z}}(\hat {\alpha } (\hat {\omega }'))$ . Since this is the only hyperplane with a label in $S_{\overline {\mathcal {X}}}$ that separates these pairs, we conclude that every hyperplane separating $\tau _{\mathcal {Z}}(\hat {\alpha } (\hat {\omega }))$ and $\tau _{\mathcal {Z}}(\hat {\alpha } (\hat {\omega }'))$ has a label in $S_{\overline {\mathcal {Z}}} \backslash S_{\overline {\mathcal {X}}}$ , which gives us $\phi (\tau _{\mathcal {Z}}(\hat {\alpha } (\hat {\omega })))=\phi (\tau _{\mathcal {Z}}(\hat {\alpha } (\hat {\omega }')))$ . Moreover, if $w'\in L_\phi $ is the word represented by $\hat {\xi }$ , then by induction we have $d_{\mathcal {X}}(o,\tau _{\mathcal {X}}(w'))=n-1$ and $d_{\mathcal {X}}(\tau _{\mathcal {X}}(w'),\tau _{\mathcal {X}}(\xi ))=1$ . But all these points belong to the projection under $\phi $ of the geodesic $\widetilde \gamma _{\hat {\alpha }(\hat {\omega })}$ , so Remark 2.2 implies that $d_{\mathcal {X}}(o,\tau _{\mathcal {X}}(w))=n$ and concludes the proof by induction, proving (1).

It is not hard to see that if $w=\mathfrak {h}_1\cdots \mathfrak {h}_n\in L_\phi $ then

$$ \begin{align*}\gamma_w:=(o,\tau_{\mathcal{X}}(\mathfrak{h}_1),\tau_{\mathcal{X}}(\mathfrak{h}_1\mathfrak{h}_2),\dots,\tau_{\mathcal{X}}(\mathfrak{h}_1\cdots \mathfrak{h}_n))\end{align*} $$

is the unique geodesic representing w in $\mathcal {X}$ and starting at o, which settles (2).

To prove (3), injectivity can be deduced by induction on the length of words in $L_\phi $ combined with the fact that no two distinct edges with the same initial vertex in $\mathcal {X}$ are dual to hyperplanes with the same label in $S_{\overline {\mathcal {X}}}$ . This last statement is true by specialness of $\overline {\mathcal {X}}$ .

To prove surjectivity, let $x\in \mathcal {X}^0$ and consider $\widetilde w\in L_{{\overline {\mathcal {Z}}}}$ such that $\phi (\tau _{\mathcal {Z}}(\widetilde w))=x$ . Such an $\widetilde w$ exists because both $\tau _{\mathcal {Z}}$ and $\phi $ are surjective. We write $\widetilde w=w_0e_1\cdots e_nw_n,$ where each $e_i$ is a letter in $S_{\overline {\mathcal {X}}}$ and each $w_i$ is a (possibly empty) word in $(S_{\overline {\mathcal {Z}}} \backslash S_{\overline {\mathcal {X}}})^\ast $ . Then $\widetilde w':=w_0e_1\cdots w_{n-1}e_n$ equals $\hat {\alpha } (\hat {\omega })$ for some path $\hat {\omega }$ in $\hat {\mathcal {G}}_\phi $ representing the word $w\in L_\phi $ , for which $\tau _{\mathcal {X}}(w)=\phi (\tau _{\mathcal {Z}}(\widetilde w'))=\phi (\tau _{\mathcal {Z}}(\widetilde w))=x$ .

Lemma 5.17. There exists $C\geq 1$ such that every $w\in L_\phi $ is represented by at most C paths in $\hat {\mathcal {G}}_\phi $ starting at an initial vertex.

Proof. Since $\tau _{\mathcal {Z}}$ is injective and $\phi $ is uniformly finite-to-one when restricted to vertices, it is enough to prove that the assignment $\hat {\omega } \mapsto \hat {\alpha }(\hat {\omega })$ from the paths in $\hat {\mathcal {G}}_\phi $ starting at an initial vertex into $L_{{\overline {\mathcal {Z}}}}$ is uniformly finite-to one. To show this, note that such $\hat {\omega }$ is completely determined by its concatenation $c(\hat {\omega })$ in $\mathcal {G}_{\overline {\mathcal {Z}}}$ (defined in (5.2)) and its final vertex in $\hat {V}_\phi $ , and that $c(\hat {\omega })$ is completely determined by $\hat {\alpha } (\hat {\omega })$ since $\mathcal {A}_{\overline {\mathcal {Z}}}$ is deterministic. The lemma then follows from Lemma 5.14.

Proof of Theorem 5.11

First we note that $\overline {\mathcal {X}}^1$ can be seen as the finite-state automaton

$$ \begin{align*}\overline{\mathcal{X}}^1=((\overline{\mathcal{X}}^0,E(\overline{\mathcal{X}})),{pr}_{{\overline{\mathcal{X}}}},\{\overline{o}\}, \overline{\mathcal{X}}^0)\end{align*} $$

over $S_{\overline {\mathcal {X}}}$ , where $E(\overline {\mathcal {X}})$ is the set of oriented edges of $\overline {\mathcal {X}}$ , $\text {pr}_{{\overline {\mathcal {X}}}}$ labels each directed edge of $\overline {\mathcal {X}}$ with its corresponding dual oriented hyperplane, and $\overline {o}$ is the image of o under the quotient $\mathcal {X} \rightarrow \overline {\mathcal {X}}$ . This automaton is deterministic since $\overline {\mathcal {X}}$ is special.

Let $\mathcal {A}_\phi =\mathcal {A}_{\overline {\Gamma },\phi }=(\mathcal {G}_\phi =(V_\phi ,E_\phi ),\pi _\phi ,I_\phi ,V_\phi )$ be the fiber product of $\hat {\mathcal {A}}_\phi $ and $\overline {\mathcal {X}}^{1}$ . That is, in $\hat {V}_\phi \times \overline {\mathcal {X}}^0$ consider a directed edge from $(\hat {\omega }, \overline {x})$ to $(\hat {\omega }',\overline {x} ')$ if there exists an edge $\hat {e}$ from $\omega $ to $\omega '$ in $\hat {\mathcal {G}}_\phi $ such that $\hat {\pi }_\phi ( \hat {e})$ is the oriented hyperplane dual to an edge from $\overline {x}$ to $\overline {x}'$ in $\overline {\mathcal {X}}$ (so that $\overline {x}, \overline {x}'$ must be adjacent). By abuse of notation we will call this edge e and define $\pi _\phi (e):=\hat \pi _\phi (\hat {e})$ . Let $I_\phi =\hat {I}_\phi \times \{\overline {o}\}$ be the set of initial states of $\mathcal {G}_\phi $ and let $V_\phi \subset \hat {V}_\phi \times \overline {\mathcal {X}}^0$ be the set of all the vertices in some directed path in $\hat {V}_\phi \times \overline {\mathcal {X}}^0$ starting at an initial state. Let $E_\phi $ be the set of all the directed edges between vertices in $V_\phi $ as defined above. Clearly $\mathcal {A}_\phi $ is pruned and finite.

There exists a label-preserving map $\mathfrak {p}$ from the set of paths in $\mathcal {G}_\phi $ starting at an initial state into the set of paths in $\hat {\mathcal {G}}_\phi $ starting at an initial state, which sends the path $((\omega _0,\overline {x}_0)\xrightarrow {e_1} \cdots \xrightarrow {e_n} (\omega _n,\overline {x}_n ))$ to the path $\hat {\omega }=(\omega _0\xrightarrow {\hat {e}_1} \cdots \xrightarrow {\hat {e}_n} \omega _n)$ . The map $\mathfrak {p}$ is a bijection since the sequence $\overline {x}_0,\dots ,\overline {x}_n$ is the image of $\gamma _w$ under $\mathcal {X} \rightarrow \overline {\mathcal {X}}$ , where w is the word represented by $\hat {\omega }$ . In particular, the language parametrized by $\mathcal {A}_\phi $ is precisely $L_\phi $ , so Lemma 5.17 implies Item (1). Also, Items (2) and (3) follow from Lemma 5.16.

For Item (4) we consider the map $\Psi :V_\phi \rightarrow (S_{\overline {\mathcal {Z}}} \backslash S_{\overline {\mathcal {X}}})^\ast $ that sends the vertex $(\omega ,\overline {x})$ to the word represented by $\omega $ , seen as a path in $\mathcal {G}_{\overline {\mathcal {Z}}}$ . From the definition of $\hat {\mathcal {G}}_\phi $ it is clear that for the path $\omega =(v_0\xrightarrow {e_1} \cdots \xrightarrow {e_n}v_n)$ in $\mathcal {G}_\phi $ starting at an initial state, the concatenation

$$ \begin{align*}\alpha(\omega):=\hat{\alpha}(\mathfrak{p}(\omega))=\Psi(v_0)\pi_\phi(e_1)\Psi(v_1) \cdots \pi_\phi(e_n)\end{align*} $$

belongs to $L_{{\overline {\mathcal {Z}}}}$ , so it is represented by the geodesic path $\widetilde \gamma _{\alpha (\omega )}$ in $\mathcal {Z}$ starting at $\widetilde o$ . The word w represented by $\omega $ is also represented by $\mathfrak {p}(\omega )$ , so Lemma 5.16 implies that $\tau _{\mathcal {X}}(w)=\phi (\tau _{\mathcal {Z}}(\alpha (\omega )))$ . $\mathcal {A}_{\overline {\mathcal {Z}}}$ being deterministic implies that $\widetilde \gamma _{\alpha (\omega )}$ is the unique geodesic in $\mathcal {Z}$ starting at $\widetilde o$ and representing w.

Finally, we define $\Xi :V_\phi \rightarrow \overline {\mathcal {X}}^0$ as the coordinate projection $\Xi (\omega , \overline {x})=\overline {x}$ , and the same argument for the proof that $\mathfrak {p}$ is a bijection implies Item (5).

6 Proof of the main theorems

In this section we prove our main results about pairs of actions on $\mathrm {CAT}(0)$ cube complexes from the introduction. Theorem 1.2 and Theorem 1.5 are consequences of more general statements, given by Theorems 6.1 and 6.2 respectively. The strategy is to use the automaton from Theorem 5.11 to define an appropriate suspension flow, and then use Proposition 6.8 to relate the Manhattan curves with pressure functions for potentials on this suspension (at this point we have enough formalism to deduce Theorem 1.4). This relation will allow us to use the tools from symbolic dynamics and thermodynamic formalism discussed in Section 3 to deduce our main results.

The following are the main theorems of the section, and they are proven in Subsection 6.2. For their statements, we interpret the quantities $\mathrm {Dil}(\mathcal {X},\mathcal {X}_\ast )^{-1}$ and $v_{\mathcal {X}^{\mathfrak {w}}}/v_{\mathcal {X}^{\mathfrak {w}_\ast }_\ast }$ as zero if the action of $\Gamma $ on $\mathcal {X}_\ast $ is not proper.

Theorem 6.1. Let $(\Gamma ,\mathcal {X},\mathcal {X}_\ast ) \in \mathfrak {X}$ and let $\mathfrak {w},\mathfrak {w}_\ast $ be $\Gamma $ -invariant orthotope structures on $\mathcal {X}$ and $\mathcal {X}_\ast $ respectively. Then the Manhattan curve $\theta _{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}:{\mathbb {R}} \rightarrow {\mathbb {R}}$ is convex, decreasing, and analytic. In addition, the following limit exists and equals $-\theta ^{\prime }_{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}(0)$ :

$$\begin{align*}\tau(\mathcal{X}^{\mathfrak{w}_\ast}_\ast/\mathcal{X}^{\mathfrak{w}}):= \lim_{T\to\infty} \frac{1}{\# \mathfrak{C}_{\mathcal{X}^{\mathfrak{w}}}(T)} \sum_{[g]\in \mathfrak{C}_{\mathcal{X}^{\mathfrak{w}}}(T)} \frac{\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]}{\ell^{\mathfrak{w}}_{\mathcal{X}}[g]}. \end{align*}$$

Moreover, we always have

$$\begin{align*}\tau(\mathcal{X}^{\mathfrak{w}_\ast}_\ast/\mathcal{X}^{\mathfrak{w}})\geq v_{\mathcal{X}^{\mathfrak{w}}}/v_{\mathcal{X}^{\mathfrak{w}_\ast}_\ast}. \end{align*}$$

If the action of $\Gamma $ on $\mathcal {X}_\ast $ is proper then the following are equivalent:

  1. (1) $\theta _{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}$ is a straight line;

  2. (2) there exists $\Lambda>0$ such that $\ell ^{\mathfrak {w}}_{\mathcal {X}}[g] = \Lambda \ell ^{\mathfrak {w}_\ast }_{\mathcal {X}_\ast }[g]$ for all $[g] \in \mathbf {conj}(\Gamma )$ ; and,

  3. (3) $\tau (\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}) = v_{\mathcal {X}^{\mathfrak {w}}}/v_{\mathcal {X}^{\mathfrak {w}_\ast }_\ast }$ .

Theorem 6.2. Let $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )\in \mathfrak {X}$ . Then there exists an analytic function

$$ \begin{align*}\mathcal{I}:[\mathrm{Dil}(\mathcal{X},\mathcal{X}_\ast)^{-1}, \mathrm{Dil}(\mathcal{X}_\ast,\mathcal{X})]\rightarrow {\mathbb{R}}\end{align*} $$

and $C>0$ such that for any $\eta \in (\mathrm {Dil}(\mathcal {X},\mathcal {X}_\ast )^{-1}, \mathrm {Dil}(\mathcal {X}_\ast ,\mathcal {X}))$ we have

(6.1) $$ \begin{align} 0 < \limsup_{T\to\infty} \frac{1}{T} \log\left( \#\left\{ [g] \in \mathfrak{C}_{\mathcal{X}}(T): | \ell_{\mathcal{X}_\ast}[g] - \eta \ell_{\mathcal{X}}[g] | < \frac{C}{T} \right\}\right) = \mathcal{I}(\eta) \le v_{\mathcal{X}}. \end{align} $$

Furthermore, we have equality in the above inequality if and only if $\eta =\tau (\mathcal {X}_*/\mathcal {X})$ .

As we noted in the previous section, $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )\in \mathfrak {X}$ whenever $\Gamma \in \mathfrak {G}$ and $\mathcal {X},\mathcal {X}_\ast $ are cubulations of $\Gamma $ . From this we see that Theorems 6.1 and 6.2 imply Theorems 1.2 and 1.5 from the introduction.

The outline to prove these theorems is as follows. Given $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )\in \mathfrak {X}$ and corresponding orthotope structures $\mathfrak {w},\mathfrak {w}_\ast $ , our goal is to relate the Manhattan curve of $(\mathcal {X}^{\mathfrak {w}},\mathcal {X}^{\mathfrak {w}_\ast }_\ast )$ to a pressure function for a potential on a suspension flow (Proposition 6.8), which is done in Subsection 6.1. The suspension flow and the potential are constructed using the automaton $\mathcal {A}_{\overline {\Gamma },\phi }$ from Theorem 5.11, and the proof of Proposition 6.8 relies on showing that any closed orbit in the suspension flow has associated a conjugacy class in $\Gamma $ in an “almost bijective” way (Lemma 6.12 and Lemma 6.13). These two lemmas are the last pieces we need to prove Theorem 1.4. Then in Subsection 6.2, and with Proposition 6.8 at our disposal, we can deduce a standard large deviations principle for the pair $(\mathcal {X}^{\mathfrak {w}},\mathcal {X}_\ast ^{\mathfrak {w}_\ast })$ (Corollary 6.16). Combining this with tools from thermodynamic formalism, we prove Theorem 6.1. Finally, we apply Theorem 3.2 to deduce a large deviations theorem with shrinking intervals for $(\mathcal {X},\mathcal {X}_\ast )$ , which is Theorem 6.2.

6.1 Manhattan curves for pairs of cubulations

In this subsection we use the finite-state automaton given by Theorem 5.11 to describe the Manhattan geodesics for a pair of cubulations in terms of pressure functions. The main result is Proposition 6.8, which will allow us to use thermodynamic formalism to prove our main results in the next subsection. For this section we keep the notation from the previous section and assume the following.

Convention 6.3. Let $(\Gamma ,\mathcal {X},\mathcal {Z})$ be a triplet satisfying Convention 5.10. Consider a nonempty $\Gamma $ -invariant subset $\mathbb {W}_\ast \subset \mathbb {H}(\mathcal {Z})$ such that $\mathbb {W} \cup \mathbb {W}_\ast =\mathbb {H}(\mathcal {Z})$ and set $\mathcal {X}_\ast =\mathcal {Z}(\mathbb {W}_\ast )$ . Then the action of $\Gamma $ on $\mathcal {X}_\ast $ is cocompact, but not necessarily proper, and we further assume that this action is essential. Let $\mathfrak {w}$ and $\mathfrak {w}_\ast $ be $\Gamma $ -invariant orthotope structures on $\mathcal {X}$ and $\mathcal {X}_\ast $ respectively. We also require the action of $\Gamma $ on $\mathcal {X}$ to have a contracting element, so that $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )\in \mathfrak {X}$ .

Let $S_{\overline {\mathcal {X}}_\ast }\subset S_{\overline {\mathcal {Z}}}$ be the set of all the oriented hyperplanes in $\overline {\mathcal {Z}}$ whose lifts to $\mathcal {Z}$ correspond to hyperplanes in $\mathbb {W}_\ast $ , and let $\phi _\ast :\mathcal {Z} \rightarrow \mathcal {X}_\ast $ be the projection quotient with $o_\ast =\phi _\ast (\widetilde o)$ .

Since the structures $\mathfrak {w},\mathfrak {w}_\ast $ are $\Gamma $ -invariant, there exist natural weighting maps $\overline {\mathfrak {w}}, \overline {\mathfrak {w}}_\ast :S_{{\overline {\mathcal {Z}}}} \rightarrow {\mathbb {R}}$ defined according to:

$$ \begin{align*} \overline{\mathfrak{w}}(\mathfrak{h}):=\begin{cases} \mathfrak{w}(\widetilde{\mathfrak{h}}) & \text{ if }\mathfrak{h} \in S_{{\overline{\mathcal{X}}}} \text{ and } \widetilde{\mathfrak{h}}\in \mathbb{H}(\mathcal{X}) \text{ is oriented and projects to }\mathfrak{h} \text{ under } \mathcal{X} \rightarrow \overline{\mathcal{X}}, \\ 0 & \text{ if }\mathfrak{h} \in S_{{\overline{\mathcal{Z}}}}\backslash S_{{\overline{\mathcal{X}}}}, \end{cases} \end{align*} $$

and

$$ \begin{align*} \overline{\mathfrak{w}}_\ast(\mathfrak{h}):=\begin{cases} \mathfrak{w}_\ast(\widetilde{\mathfrak{h}}) & \text{ if }\mathfrak{h} \in S_{{\overline{\mathcal{X}}}_\ast} \text{ and } \widetilde{\mathfrak{h}}\in \mathbb{H}(\mathcal{X}) \text{ is oriented and projects to }\mathfrak{h} \text{ under } \mathcal{Z} \rightarrow \overline{\mathcal{Z}}, \\ 0 & \text{ if }\mathfrak{h} \in S_{{\overline{\mathcal{Z}}}}\backslash S_{{\overline{\mathcal{X}}}_\ast}. \end{cases} \end{align*} $$

By abuse of notation we extend these weightings to $\overline {\mathfrak {w}}, \overline {\mathfrak {w}}_\ast : (S_{\overline {\mathcal {Z}}})^\ast \rightarrow {\mathbb {R}}$ by declaring the empty word to have weights 0 and assigning

$$\begin{align*}\overline{\mathfrak{w}} (\mathfrak{h}_1\cdots \mathfrak{h}_n)= \overline{\mathfrak{w}}(\mathfrak{h}_1)+\cdots + \overline{\mathfrak{w}}(\mathfrak{h}_n) \ \text{ and } \ \overline{\mathfrak{w}}_\ast(\mathfrak{h}_1\cdots \mathfrak{h}_n)= \overline{\mathfrak{w}}_\ast(\mathfrak{h}_1)+\cdots + \overline{\mathfrak{w}}_\ast(\mathfrak{h}_n) \end{align*}$$

for a word $\mathfrak {h}_1\cdots \mathfrak {h}_n\in (S_{\overline {\mathcal {Z}}})^\ast $ of positive length.

We consider the automaton $\mathcal {A}_{\overline {\Gamma },\phi }$ given by Theorem 5.11, and we let $\Sigma ^\times $ be the set of all the finite directed paths in the underlying graph $\mathcal {G}_\phi $ of $\mathcal {A}_{\overline {\Gamma },\phi }$ . We can see $\Sigma ^\times $ as a set of finite sequences of edges in $E_\phi $ . Similarly, let $\Sigma \subset (E_\phi )^{{\mathbb {N}}}$ be the set of infinite sequences $(e_i)_{i\geq 1}$ such that $(e_i)_{1\leq i\leq k}\in \Sigma ^\times $ for all k. We also let $\sigma :\Sigma \rightarrow \Sigma $ denote the shift map $\sigma ((e_i)_{i\geq 1})=(e_{i+1})_{i\geq 1}$ . For each n we let $P_n(\Sigma ^\times )\subset \Sigma ^\times $ be the subset of all the closed paths of length n, and set $P_{\leq n}(\Sigma ^\times )=\bigcup _{j\leq n}{P_j(\Sigma ^\times )}$ and $P(\Sigma ^\times )=\bigcup _{j}{P_j(\Sigma ^\times )}$ . We will often identify the set $\operatorname {\mathrm {Fix}}_n(\Sigma )$ of sequences $\omega \in \Sigma $ satisfying $\sigma ^n(\omega )=\omega $ with $P_n(\Sigma ^\times )$ via the truncation $\omega =(e_1,\dots ,e_n,e_1,\dots ) \mapsto t_n(\omega ):=(e_1,\dots ,e_n)$ .

The next definition will be useful for the rest of the section.

Definition 6.4. A combinatorial path $\gamma $ in $\mathcal {X}$ is a good representative of $\omega \in \Sigma ^\times $ if there exists a path $\omega _0\in \Sigma ^\times $ starting at an initial state and ending at the initial vertex of $\omega $ and satisfying the following. If $w_0, w_0w\in L_\phi $ are the words corresponding to $\omega _0$ and $\omega _0 \omega $ respectively, then $\gamma $ is the portion of the path $\gamma _{w_0w}$ representing $w_0w$ from $\gamma ^-=\tau _{\mathcal {X}}(w_0)$ to $\gamma ^+=\tau _{\mathcal {X}}(w_0w)$ .

Note that good representatives are geodesic. Also, by Theorem 5.11 (5), any two good representatives of the same path in $\Sigma ^\times $ differ by a translation by an element in $\overline {\Gamma }$ . In consequence, if $\omega \in P(\Sigma ^\times )$ then there exists a well-defined conjugacy class $\beta (\omega )\in \mathbf {conj}(\overline {\Gamma })$ represented by any $g\in \overline {\Gamma }$ such that $\gamma ^+=g\gamma ^-$ for $\gamma $ a good representative of $\omega $ . Clearly $\omega \in P_n(\Sigma ^\times )$ implies $\ell _{\mathcal {X}}[\beta (\omega )]=n$ .

We also consider lifts of paths in $\Sigma ^\times $ to $\mathcal {Z}$ . First, we extend the equation (5.1) to define a map $\alpha :\Sigma ^\times \rightarrow (S_{\overline {\mathcal {Z}}})^\ast $ . If $\gamma $ is a good representative of $\omega \in \Sigma ^\times $ defined using the path $\omega _0$ as above, then $\widetilde \gamma $ is the portion of $\widetilde \gamma _{\alpha (\omega _0\omega )}$ starting at $\widetilde \gamma _{\alpha (\omega _0)}^+$ and ending at $\widetilde \gamma _{\alpha (\omega _0\omega )}^+$ , where $\widetilde \gamma _{\alpha (\omega _0)}$ and $\widetilde \gamma _{\alpha (\omega _0\omega )}$ are given by Theorem 5.11 (4). In this way the path $\widetilde \gamma $ represents the word $\alpha (\omega )$ . Different choices of $\omega _0,\omega _0'$ may give different lifts $\widetilde \gamma , \widetilde \gamma '$ even if $\tau _{\mathcal {X}}(w_0)=\tau _{\mathcal {X}}(w_0')$ , but under this assumption we have $\phi (\widetilde \gamma )=\phi (\widetilde \gamma ') =\gamma $ .

A key feature of the automaton $\mathcal {A}_{\overline {\Gamma },\phi }$ is that it keeps track of translation lengths for the actions of $\Gamma $ on the cuboid complexes $\mathcal {X}^{\mathfrak {w}}$ and $\mathcal {X}_\ast ^{\mathfrak {w}_\ast }$ , via the potential on $(\Sigma ,\sigma )$ defined below.

Definition 6.5. Let $r=r^{\mathfrak {w}}_{\mathcal {X}},\psi =\psi ^{\mathfrak {w}_\ast }_{\mathcal {X}_\ast }: E_\phi \rightarrow {\mathbb {R}}$ be the functions such that

(6.2) $$ \begin{align} r(e)=\overline{\mathfrak{w}}(\pi_\phi(e)) \ \ \text{ and } \ \ \psi(e)=\psi( v_0\xrightarrow{e} v_1)= \overline{\mathfrak{w}}_\ast(\Psi(v_0))+\overline{\mathfrak{w}}_\ast(\pi_\phi(e)), \end{align} $$

where $\Psi $ is the function from Theorem 5.11 (4).

Remark 6.6. In the definition of $\psi $ above, we note that $\overline {\mathfrak {w}}_\ast (\pi _\phi (e))$ is not necessarily zero since $S_{\overline {\mathcal {X}}}$ and $S_{\overline {\mathcal {X}}_\ast }$ are not necessarily disjoint. Also, our assumption that $\mathbb {W} \cup \mathbb {W}_\ast =\mathbb {H}(\mathcal {Z})$ implies that $S_{\overline {\mathcal {X}}} \cup S_{{\overline {\mathcal {X}}}_\ast }=S_{\overline {\mathcal {Z}}}$ , so that $\Psi (v_0)$ always belongs to $(S_{{\overline {\mathcal {X}}}_\ast })^\ast $ .

By abuse of notation we extend these functions to potentials $r,\psi :\Sigma \rightarrow {\mathbb {R}}$ via

$$\begin{align*}r(e_1,e_2,\dots)=r(e_1) \ \text{ and } \ \psi(e_1,e_2,\dots)=\psi(e_1).\end{align*}$$

Clearly r and $\psi $ are constant on 2-cylinders, and r is positive. The next lemma can be seen as a weak analog of [Reference Calegari and Fujiwara14, Lemma 3.8] for pairs of word metrics on hyperbolic groups.

Lemma 6.7. Let $r,\psi :\Sigma \rightarrow {\mathbb {R}}$ be the potentials defined above. If $n\geq 1$ and $\omega \in \operatorname {\mathrm {Fix}}_n(\Sigma )$ has truncation $t_n(\omega )\in P_n(\Sigma ^\times )$ , then $\ell _{\mathcal {X}}[\beta (t_n(\omega ))]=n$ and the nth Birkhoff sums at $\omega $ satisfy

(6.3) $$ \begin{align} r^n(\omega)=r(\omega)+r(\sigma(\omega))+\cdots +r(\sigma^{n-1}(\omega))=\ell^{\mathfrak{w}}_{\mathcal{X}}[\beta(t_n(\omega))] \end{align} $$

and

(6.4) $$ \begin{align} \psi^n(\omega)=\psi(\omega)+\psi(\sigma(\omega))+\cdots +\psi(\sigma^{n-1}(\omega))=\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[\beta(t_n(\omega))]. \end{align} $$

Proof of Lemma 6.7

Let $\omega \in \operatorname {\mathrm {Fix}}_n(\Sigma )$ be as in the statement of the lemma and let $\omega _n=t_n(\omega )=(v_0\xrightarrow {e_1} \cdots \xrightarrow {e_n}v_n)\in P_n(\Sigma ^\times )$ . As we noted previously we have $\ell _{\mathcal {X}}[\beta (\omega _n)]=n$ , so the main content of the lemma are the identities (6.3) and (6.4), which we now prove.

For any $k{\kern-1pt}\geq{\kern-1pt} 1$ , let $\omega _n^{(k)}{\kern-1.3pt}\in{\kern-1pt} P(\Sigma ^\times )$ be the concatenation of k copies of $\omega _n$ and let ${(\gamma ^{(k)})_k{\kern-1pt}={\kern-1pt}(\gamma _{({\omega _n}^{(k)})})_k{\kern-1pt}\subset{\kern-1pt} \mathcal {X}}$ be a sequence of good representatives of $\omega _n^{(k)}$ with a common starting vertex $\gamma ^-=(\gamma ^{(k)})^-$ . Let $q\in \Gamma $ satisfy $(\gamma ^{(1)})^+=q\gamma ^-$ , so that $[q]=\beta (\omega _n)$ and $(\gamma ^{(k)})^+=q^k\gamma ^-$ for each k. Also, since $\omega $ is a closed path, we have that $\gamma ^{(1)}$ is a fundamental domain for a q-invariant geodesic in $\mathcal {X}^{\mathfrak {w}}$ , and hence

$$ \begin{align*} r^n(\omega) &=\overline{\mathfrak{w}}(\pi_\phi(e_1))+\cdots +\overline{\mathfrak{w}}(\pi_\phi(e_n))\\ & =d_{\mathcal{X}}^{\mathfrak{w}}((\gamma^{(1)})^-,(\gamma^{(1)})^+)=d_{\mathcal{X}}^{\mathfrak{w}}(\gamma^-,q\gamma^-)=\ell_{\mathcal{X}}^{\mathfrak{w}}[q]=\ell_{\mathcal{X}}^{\mathfrak{w}}[\beta(t_n(\omega))]. \end{align*} $$

This proves (6.3).

To prove (6.4), let L be such that $\phi _\ast :\mathcal {Z} \rightarrow \mathcal {X}_\ast ^{\mathfrak {w}_\ast }$ is L-Lipschitz and consider the lifts $\widetilde \gamma ^{(k)}\subset \mathcal {Z}$ of the geodesic $\gamma ^{(k)}$ with a common starting vertex $\widetilde {\gamma }^-=(\widetilde {\gamma }^{(k)})^-$ . We project these paths to $\mathcal {X}_\ast $ by defining $\gamma ^-_\ast :=\phi _\ast (\widetilde \gamma ^-)$ and $(\gamma _\ast ^{(k)})^+:=\phi _\ast ((\widetilde \gamma ^{(k)})^+)$ , and for all k we get

(6.5) $$ \begin{align} |d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(\gamma_\ast^-,(\gamma_\ast^{(k)})^+)-d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(\gamma^-_\ast,q^k\gamma_\ast^-)|\leq d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}((\gamma^{(k)}_\ast)^+,q^k \gamma^-_\ast)\leq Ld_{\mathcal{Z}}((\widetilde \gamma^{(k)})^+,q^k\widetilde \gamma^-). \end{align} $$

The last term in the inequality above is bounded by a number independent of k. Indeed, since $\phi ((\widetilde \gamma ^{(k)})^+)=\phi (q^k\widetilde \gamma ^-)=q^k \gamma ^-$ , we have that both $(\widetilde \gamma ^{(k)})^+$ and $q^k\widetilde \gamma ^-$ belong to the preimage of $q^k \gamma ^-$ under $\phi $ , so their distance is bounded by a number independent of k because $\phi :\mathcal {Z} \rightarrow \mathcal {X}$ is a quasi-isometry.

Also, we note that $d_{\mathcal {X}_\ast }^{\mathfrak {w}_\ast }(\gamma _\ast ^-,(\gamma _\ast ^{(k)})^+)=\overline {\mathfrak {w}}_\ast (\alpha (\omega _n^{(k)}))$ , which equals $k\cdot \overline {\mathfrak {w}}_\ast (\alpha (\omega _n))$ since $\alpha (\omega _n^{(k)})$ is the concatenation of k copies of $\alpha (\omega _n)$ . By Theorem 5.11 (4), it follows that

$$ \begin{align*} d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(\gamma_\ast^-,(\gamma_\ast^{(k)})^+) & =k\cdot \overline{\mathfrak{w}}_\ast(\alpha(\omega_n)) \\ & =k \cdot (\overline{\mathfrak{w}}_\ast(\Psi(v_0)\pi_\phi(e_1))+\overline{\mathfrak{w}}_\ast(\Psi(v_1)\pi_\phi(e_2))+\cdots +\overline{\mathfrak{w}}_\ast(\Psi(v_{n-1})\pi_\phi(e_n))) \\ & =k \cdot (\psi(v_0\xrightarrow{e_1} v_1)+\psi(v_1\xrightarrow{e_2} v_2)+\cdots +\psi(v_{n-1}\xrightarrow{e_n} v_n)) \\ & = k \cdot (\psi(\omega)+\psi(\sigma(\omega))+\cdots +\psi(\sigma^{n-1}(\omega)))=k\psi^n(\omega). \end{align*} $$

Therefore, combining this with (6.5) and after dividing by k and letting k tend to infinity we obtain $\ell _{\mathcal {X}_\ast }^{\mathfrak {w}_\ast }[\beta (t_n(\omega ))]=\ell ^{\mathfrak {w}_\ast }_{\mathcal {X}_\ast }[q]=\psi ^n(\omega )$ , as desired.

To apply the results from Section 3 we require a mixing (or at least transitive) dynamical system. To obtain such a system we consider a maximal recurrent component $\mathcal {C}$ of the graph $\mathcal {G}_\phi $ . As before we let $\Sigma _{\mathcal {C}}^\times \subset \Sigma ^\times $ and $\Sigma _{\mathcal {C}}\subset \Sigma $ be the subsets corresponding to paths in $\mathcal {C}$ , and note that $\Sigma _{\mathcal {C}}$ is $\sigma $ -invariant. Similarly we define $P(\Sigma ^\times _{\mathcal {C}})$ , $P_n(\Sigma ^\times _{\mathcal {C}})$ and $P_{\leq n}(\Sigma _{\mathcal {C}}^\times )$ , and we identify $P_n(\Sigma _{\mathcal {C}}^\times )$ with $\operatorname {\mathrm {Fix}}_n(\Sigma _{\mathcal {C}})$ .

Let $r_{\mathcal {X}}^{\mathfrak {w}} : \Sigma _{\mathcal {C}} \to {\mathbb {R}}_{>0}$ be the (constant on $2$ -cylinders) restriction to $\Sigma _{\mathcal {C}}$ of the potential introduced in Definition 6.5 and consider the suspension flow:

$$\begin{align*}\Sigma_{\mathcal{C}}^{r_{\mathcal{X}}^{\mathfrak{w}}} := \{(\omega,t) \in \Sigma_{\mathcal{C}} \times {\mathbb{R}} : 0 \le t \le r_{\mathcal{X}}^{\mathfrak{w}}(\omega) \}/\sim, \end{align*}$$

where each $(\omega ,r_{\mathcal {X}}^{\mathfrak {w}}(\omega ))$ is identified with $(\sigma (\omega ), 0)$ and the flow $\sigma ^{r_{\mathcal {X}}^{\mathfrak {w}}}=(\sigma ^{r_{\mathcal {X}}^{\mathfrak {w}}}_s)_{s\in {\mathbb {R}}_{>0}}$ acts as $\sigma ^{r_{\mathcal {X}}^{\mathfrak {w}}}_s(\omega ,t) = (\omega ,t+s)$ .

Note that any closed $\sigma ^{r_{\mathcal {X}}^{\mathfrak {w}}}$ -orbit $\tau $ in $\Sigma _{\mathcal {C}}^{r_{\mathcal {X}}^{\mathfrak {w}}}$ corresponds to a closed $\sigma $ -orbit in $\Sigma _{\mathcal {C}}$ . More precisely, such an orbit $\tau $ must be of the form $\tau =\{(\omega ,t) : 0\leq t \leq (r_{\mathcal {X}}^{\mathfrak {w}})^n(\omega ) \}$ for some $\omega \in \Sigma _{\mathcal {C}}$ such that $\sigma ^n(\omega )=\omega $ . In this case the period of $\tau $ equals $l_\tau =(r_{\mathcal {X}}^{\mathfrak {w}})^n(\omega )$ .

We fix a smooth function $\Delta : [0,1] \to {\mathbb {R}}_{\ge 0}$ such that $\Delta (0) = \Delta (1) = 1$ and $\int _0^1 \Delta (t) \ dt =1$ and define $\Phi :\Sigma _{\mathcal {C}}^{r_{\mathcal {X}}^{\mathfrak {w}}}\rightarrow {\mathbb {R}}_{\ge 0}$ according to

(6.6) $$ \begin{align} \Phi(\omega,t) = \Delta\left(\frac{t}{r_{\mathcal{X}}^{\mathfrak{w}}(\omega)} \right) \frac{\psi_{\mathcal{X}_\ast}^{\mathfrak{w}_\ast}(\omega)}{r_{\mathcal{X}}^{\mathfrak{w}}(\omega)} \ \text{ for each } (\omega,t) \in \Sigma_{\mathcal{C}}^{r_{\mathcal{X}}^{\mathfrak{w}}} \text{ with }0\leq t\leq r_{\mathcal{X}}^{\mathfrak{w}}(\omega). \end{align} $$

This function has the property that, for any closed $\sigma ^{r_{\mathcal {X}}^{\mathfrak {w}}}$ -orbit $\tau $ in $\Sigma _{\mathcal {C}}^{r_{\mathcal {X}}^{\mathfrak {w}}}$ with period $l_\tau $ and corresponding periodic $\sigma $ -orbit $\omega ,\sigma (\omega ),\dots ,\sigma ^n(\omega )=\omega $ in $\Sigma _{\mathcal {C}}$ we have

$$\begin{align*}\int_\tau \Phi := \int_0^{l_\tau} \Phi(\sigma^{r_{\mathcal{X}}^{\mathfrak{w}}}_t(\omega,0)) \ dt = \ell_{\mathcal{X}_\ast}^{\mathfrak{w}_\ast}[\beta(t_n(\omega))], \end{align*}$$

where we have used Lemma 6.7. For $\tau , \omega $ , and n as above we adopt the notation

$$\begin{align*}\beta(\tau) = \beta(t_n(\omega)), \end{align*}$$

which defines a map $\beta : P(\Sigma _{\mathcal {C}}^{r_{\mathcal {X}}^{\mathfrak {w}}}) \to \mathbf {conj}(\overline {\Gamma })$ from the set $P(\Sigma _{\mathcal {C}}^{r_{\mathcal {X}}^{\mathfrak {w}}})$ of periodic orbits of $\Sigma _{\mathcal {C}}^{r_{\mathcal {X}}^{\mathfrak {w}}}$ into $\mathbf {conj}(\overline {\Gamma })$ . By Lemma 6.7, the period of any $\tau \in P(\Sigma _{\mathcal {C}}^{r_{\mathcal {X}}^{\mathfrak {w}}})$ equals $\ell _{\mathcal {X}}^{\mathfrak {w}}[\beta (\tau )]$ .

For such a suspension flow $(\Sigma ^{r_{\mathcal {X}}^{\mathfrak {w}}}_{\mathcal {C}},\sigma ^{r_{\mathcal {X}}^{\mathfrak {w}}})$ , the Manhattan curve for $(\mathcal {X}^{\mathfrak {w}},\mathcal {X}^{\mathfrak {w}_\ast }_\ast )$ can be described in terms of the pressures related to $\Phi $ , as stated in the next proposition.

Proposition 6.8. Let $\Gamma ,\mathcal {X},\mathcal {X}_\ast ,\mathcal {Z},\mathfrak {w},\mathfrak {w}_\ast $ satisfy Convention 6.3 and let $\mathcal {A}_{\overline {\Gamma },\phi }$ and $r_{\mathcal {X}}^{\mathfrak {w}},\psi _{\mathcal {X}_\ast }^{\mathfrak {w}_\ast }:\Sigma \rightarrow {\mathbb {R}}$ be given by Theorem 5.11 and Definition 6.5 respectively. If $\mathcal {C}$ is a maximal recurrent component of $\mathcal {G}_\phi $ and $\Phi :\Sigma ^{r_{\mathcal {X}}^{\mathfrak {w}}}_{\mathcal {C}} \rightarrow {\mathbb {R}}$ is given by (6.6), then for any $s\in {\mathbb {R}}$ we have

$$ \begin{align*}\theta_{\mathcal{X}^{\mathfrak{w}_\ast}_\ast/\mathcal{X}^{\mathfrak{w}}}(s) = {P}_{\mathcal{C}}(-s\Phi), \end{align*} $$

where $ {P}_{\mathcal {C}}(-s\Phi )$ is the pressure of the potential $-s\Phi $ on the suspension $(\Sigma _{\mathcal {C}}^{r_{\mathcal {X}}^{\mathfrak {w}}},\sigma ^{r_{\mathcal {X}}^{\mathfrak {w}}})$ .

In particular, this result implies that the pressure $s \mapsto {P}_{\mathcal {C}}(-s\Phi )$ is independent of the choice of maximal recurrent component $\mathcal {C}$ .

Remark 6.9. As it will be clear from the proof of Proposition 6.8, if $\mathcal {C}$ is any maximal recurrent component of $\mathcal {G}_\phi $ then for any $s\in {\mathbb {R}}$ we also have

$$\begin{align*}\theta_{\mathcal{X}_\ast/\mathcal{X}}(s)=\text{P}_{\mathcal{C}}(-s\psi),\end{align*}$$

where $\psi =\psi _{\mathcal {X}_\ast }^{\mathfrak {w}_\ast }$ for $\mathfrak {w}_\ast \equiv 1$ the constant orthotope structure and $\text {P}_{\mathcal {C}}(-s\psi )$ is the pressure of the potential $-s\psi $ on $(\Sigma _{\mathcal {C}},\sigma )$ .

The rest of this subsection is devoted to proving this proposition, and using the formalism necessary in its proof to deduce Theorem 1.4. For the sequel we fix a compact set $K\subset \mathcal {X}$ such that $\mathcal {X}=\Gamma K$ and assume that $o\in K$ . Since $\mathcal {G}_\phi $ is finite and pruned we also fix $N>0$ such that any $\omega \in \Sigma ^\times $ has a good representative $\gamma $ satisfying $d_{\mathcal {X}}(\gamma ^-,o)\leq N$ . Otherwise explicit, for any $\omega \in \Sigma ^\times $ we fix a good representative $\gamma =\gamma _\omega $ that minimizes $d_{\mathcal {X}}(\gamma ^-,o)$ and define

$$\begin{align*}\Gamma_{\mathcal{C}}:=\{g\in \Gamma \colon \text{ there exists }\omega\in \Sigma_{\mathcal{C}}^\times \text{ such that }\gamma_\omega^+\in gK\}.\end{align*}$$

We also write $B_n=\{g\in \Gamma : d_{\mathcal {X}}(go,o)\leq n\}$ for each $n\geq 0$ .

Lemma 6.10. Let C be the constant from Theorem 5.11 (1). Then

$$\begin{align*}\sup_{x\in \mathcal{X}^0}{\#\{\omega\in \Sigma^\times \colon \gamma^+_\omega=x\}}\leq C(N+1).\end{align*}$$

Proof. Let $x\in \mathcal {X}^0$ and $\omega \in \Sigma ^\times $ be such that $\gamma _\omega ^+=x$ , so that $\gamma _\omega $ is constructed from a path $\omega _0\in \Sigma ^\times $ such that $\omega '=\omega _0\omega $ also belongs to $\Sigma ^\times $ . Then $\omega _0$ is a prefix of $\omega '$ of length at most N and $\omega '$ represents the unique word $w\in L_\phi $ satisfying $\tau _{\mathcal {X}}(w)=x$ . From this we deduce

$$ \begin{align*} \#\{\omega\in \Sigma^\times \colon \gamma^+_\omega=x\} & \leq (N+1)\cdot \# \{\omega' \in \Sigma^\times \colon \omega' \text{ starts at an initial state and }\gamma^+_{\omega'}=x\} \\ & \leq (N+1)\cdot \#\{\omega' \in \Sigma^\times \colon \omega' \text{ represents }w \text{ and starts at an initial state}\}, \end{align*} $$

and the lemma follows from Theorem 5.11 (1).

Lemma 6.11. The set $\Gamma _{\mathcal {C}}$ has positive lower density for the action on $\mathcal {X}$ . That is

$$\begin{align*}\liminf_{n\to \infty}{\frac{\#(\Gamma_{\mathcal{C}} \cap B_n)}{\# B_n}}>0.\end{align*}$$

Proof. For each n we let $(\Sigma _{\mathcal {C}}^\times )_{\leq n}$ denote the set of paths in $\mathcal {C}$ of length at most n. First we claim that there exist $B>1$ and $0\leq \lambda < e^{v_{\mathcal {X}}}$ (depending only on the adjacency matrix of $\mathcal {C}$ ) such that

(6.7) $$ \begin{align} \# (\Sigma_{\mathcal{C}}^\times)_{\leq n} \geq B^{-1}e^{n v_{\mathcal{X}}}-B\lambda^{n } \end{align} $$

for all n large enough. To show this, let A be the adjacency matrix of $\mathcal {C}$ , which is irreducible since $\mathcal {C}$ is recurrent. Moreover, $\mathcal {C}$ being maximal implies that the spectral radius of A equals $e^{v_{\mathcal {X}}}$ . Suppose that A has $p\geq 1$ eigenvalues of absolute value $e^{v_{\mathcal {X}}}$ and let $0\leq \lambda <e^{v_{\mathcal {X}}}$ be any number greater than the absolute value of all the other eigenvalues of A. By [Reference Dahmani, Futer and Wise26, Theorem 3.1], for each $k\geq 1$ the matrix $A^{kp}$ has $e^{kpv_{\mathcal {X}}}$ as eigenvalue of multiplicity p and all its other eigenvalues have absolute value less than $\lambda ^{kp}$ . In particular, its trace satisfies

$$\begin{align*}\operatorname{\mathrm{tr}} (A^{kp}) \geq pe^{kpv_{\mathcal{X}}}-(\dim A -p)\lambda^{kp}.\end{align*}$$

But $\operatorname {\mathrm {tr}} (A^{kp})$ equals the number of closed paths of length $kp$ in $\mathcal {C}$ , so if $n=kp+r$ with $0\leq r<p$ an integer and $k\geq 1$ , then

$$ \begin{align*} \#(\Sigma_{\mathcal{C}}^\times)_{\leq n}\geq \#(\Sigma_{\mathcal{C}}^\times)_{\leq kp} & \geq \operatorname{\mathrm{tr}}(A^{kp})\geq pe^{kpv_{\mathcal{X}}}-(\dim A -p)\lambda^{kp}\\ & \geq (pe^{-pv_{\mathcal{X}}})e^{nv_{\mathcal{X}}}-(\dim A -p)\lambda^{n}. \end{align*} $$

This concludes the proof of the claim. Now consider n large enough and $\omega \in (\Sigma _{\mathcal {C}}^\times )_{\leq n}$ . Since the $\Gamma $ -translates of K cover $\mathcal {X}$ we have $\gamma _\omega ^+\in gK$ for some $g\in \Gamma $ , so that $d_{\mathcal {X}}(\gamma ^+_\omega ,go)\leq D$ with D being the diameter of K. In addition, by definition we have $d_{\mathcal {X}}(\gamma _\omega ^-,o)\leq N$ and hence $d_{\mathcal {X}}(o,go)\leq n+D+N$ . By Lemma 6.10 this implies that

$$ \begin{align*} \#(\Sigma_{\mathcal{C}}^\times)_{\leq n} \leq \#(\Gamma_{\mathcal{C}} \cap B_{n+D+N})\cdot \sup_{g\in \Gamma}\# \{\omega \in \Sigma_{\mathcal{C}}^\times \colon \gamma^+_\omega \in gK\} \leq (N+1)C\cdot \#K \cdot \#(\Gamma_{\mathcal{C}} \cap B_{n+D+N}), \end{align*} $$

where C is the constant from Theorem 5.11 (1).

Combining this with (6.7) we get

$$ \begin{align*} \#(\Gamma_{\mathcal{C}} \cap B_n) &\geq ((N+1)C\#K)^{-1}B^{-1}e^{(n-D-N)v_{\mathcal{X}}}-((N+1)C\#K)^{-1}B\lambda^{n-D-N} \\ &= [ ((N+1)(C\#K)B)^{-1}e^{-(D+N)v_{\mathcal{X}}}]e^{nv_{\mathcal{X}}}-[((N+1)C\#K)^{-1}B\lambda^{-D-N}]\lambda^{n}. \end{align*} $$

Finally, the since the action on $\mathcal {X}$ has a contracting element, [Reference Yang88, Theorem 1.8 (2)] implies that there exists $C'>0$ such that $\#B_n \leq C'e^{nv_{\mathcal {X}} }$ for all n large enough, and the conclusion follows.

The next two results are used to give a uniform comparison of the number of conjugacy classes in $\Gamma $ with a bound on their translation lengths (with respect to $\mathcal {X}^{\mathfrak {w}}$ ) and the number of periodic orbits in the suspension $(\Sigma ^{r_{\mathcal {X}}^{\mathfrak {w}}}_{\mathcal {C}},\sigma ^{r_{\mathcal {X}}^{\mathfrak {w}}})$ with bounded period. The assumption of having a contracting element is essential in the proof of the Lemma 6.13. For $T\geq R \geq 0$ , recall that $P(\Sigma _{\mathcal {C}}^{r_{\mathcal {X}}^{\mathfrak {w}}}, R, T)$ denotes the set of periodic orbits in $(\Sigma _{\mathcal {C}}^{r_{\mathcal {X}}^{\mathfrak {w}}},\sigma ^{r_{\mathcal {X}}^{\mathfrak {w}}})$ with period on the interval $[T-R,T+R]$ .

Lemma 6.12. For any $R>0$ there exists a polynomial Q (depending on R) that is nondecreasing on ${\mathbb {R}}_{>0}$ and such that for any $[g]\in \mathbf {conj}(\overline {\Gamma })$ we have

(6.8) $$ \begin{align} \# \{\tau\in P(\Sigma_{\mathcal{C}}^{r_{\mathcal{X}}^{\mathfrak{w}}}, R, T) : \beta(\tau) = [g] \}\leq Q(T). \end{align} $$

Proof. To solve the lemma it is enough to show that there exists a polynomial $\widetilde {Q}$ that is increasing on ${\mathbb {R}}_{>0}$ and such that for any $[g]\in \mathbf {conj}(\overline {\Gamma })$ we have

(6.9) $$ \begin{align} \# \{\omega\in P(\Sigma^\times) \colon \beta(\omega)=[g]\}\leq \widetilde{Q}(\ell_{\mathcal{X}}[g]). \end{align} $$

Indeed, since $\mathfrak {w}$ is nonvanishing, the identity map $ {Id}:\mathcal {X} \rightarrow \mathcal {X}^{\mathfrak {w}}$ is a quasi-isometry and $\ell _{\mathcal {X}}[g]\leq L\ell _{\mathcal {X}}^{\mathfrak {w}}[g]$ for any $[g]\in \mathbf {conj}(\Gamma )$ , for $L=(\min \{\mathfrak {w}(\mathfrak {h}):\mathfrak {h} \in \mathbb {H}(\mathcal {X})\})^{-1}$ . Also, if $\tau \in P(\Sigma _{\mathcal {C}}^{r_{\mathcal {X}}^{\mathfrak {w}}}, R, T)$ satisfies $\beta (\tau )=[g]$ , then $l_\tau =\ell _{\mathcal {X}}^{\mathfrak {w}}[g]\in [T-R,T+R]$ . Since any periodic orbit in $P(\Sigma _{\mathcal {C}}^{r_{\mathcal {X}}^{\mathfrak {w}}})$ corresponds to an (orbit determined by an) element in $P(\Sigma _{\mathcal {C}}^\times )\subset P(\Sigma ^\times )$ , for any $[g]$ such that the left-hand side in (6.8) is nonzero we have

$$ \begin{align*} \# \{\tau\in P(\Sigma_{\mathcal{C}}^{r_{\mathcal{X}}^{\mathfrak{w}}}, R, T) : \beta(\tau) = [g] \} & \leq \# \{\omega\in P(\Sigma^\times) : \beta(\omega) = [g] \} \\ & \leq \widetilde{Q}(\ell_{\mathcal{X}}[g]) \leq \widetilde{Q}(LT+LR)=:Q(T). \end{align*} $$

To prove (6.9) we claim that for any group $\overline {\Gamma }$ acting properly, cocompactly and co-specially on a $\mathrm {CAT}(0)$ cube complex $\mathcal {X}$ and for any $x\in \mathcal {X}^{0}$ and $R\geq 0$ there exists a polynomial $\hat {Q}$ that is increasing on ${\mathbb {R}}_{>0}$ and such that

(6.10) $$ \begin{align} \#\{g\in [g] \colon d_{\mathcal{X}}(gx,x)\leq \ell_{\mathcal{X}}[g]+R\}\leq \hat{Q}(\ell_{\mathcal{X}}[g]) \end{align} $$

for any $[g]\in \mathbf {conj}(\overline {\Gamma })$ .

To see how this claim proves the lemma, fix $[g]\in \mathbf {conj}(\overline {\Gamma })$ and let $\omega \in P(\Sigma ^\times )$ be such that $\beta (\omega )=[g]$ . If $\gamma =\gamma _{\omega }$ is a good representative such that $d_{\mathcal {X}}(\gamma ^-,o)\leq N$ , we let $q\in [g]$ be such that $\gamma ^+=q\gamma ^-$ . Then $d_{\mathcal {X}}(qo,\gamma ^+)=d_{\mathcal {X}}(o,\gamma ^-)\leq N$ and we have

$$\begin{align*}|d_{\mathcal{X}}(o,qo)-\ell_{\mathcal{X}}[g]|=|d_{\mathcal{X}}(o,qo)-d_{\mathcal{X}}(\gamma^-,\gamma^+)|\leq 2N.\end{align*}$$

In addition, there exists a constant $\hat {C}>0$ such that for any $[g]\in \mathbf {conj}(\overline {\Gamma })$ and any $q\in [g]$ satisfying $|d_{\mathcal {X}}(o,qo)-\ell _{\mathcal {X}}[g]|\leq 2N$ , the set

$$\begin{align*}\{\omega \in P(\Sigma^\times) \colon \beta(\omega)=[g] \text{ and } d_{\mathcal{X}}(\gamma_\omega^+,qo)\leq N\}\end{align*}$$

has cardinality at most $\hat {C}$ . Indeed, if B is the set of vertices at distance at most N from o, then this cardinality is bounded above by

$$ \begin{align*} \sum_{x\in B}{\# \{ \omega \in P(\Sigma^\times) \colon \gamma^+_\omega=qx\} } & \leq \# B \cdot \sup_{x\in \mathcal{X}^0}{\# \{ \omega \in P(\Sigma^\times) \colon \gamma^+_\omega = x\}} \leq \# B \cdot C(N+1) =: \hat{C}, \end{align*} $$

where for the last inequality we used Lemma 6.10.

Applying this to our case of interest, we deduce

$$ \begin{align*} \# \{\omega\in P(\Sigma^\times) \colon \beta(\omega)=[g]\}\leq \hat{C} \cdot \# \{q\in [g] \colon |d_{\mathcal{X}}(qo,o)-\ell_{\mathcal{X}}[g]|\leq 2N\}\leq \hat{C} \cdot \hat{Q}(\ell_{\mathcal{X}}[g]), \end{align*} $$

where $\hat {Q}$ is the polynomial given by the claim for $x=o$ and $R=2N$ .

To prove the claim (6.10), by equivariantly embedding $\mathcal {X}$ as a convex subcomplex of the universal cover of a Salvetti complex we can assume that $\overline {\Gamma }$ is a right-angled Artin group with standard (symmetric) generating set S and $\mathcal {X}$ is the universal cover of its Salvetti complex. Then $\mathcal {X}^{1}$ is the Cayley graph for $\overline {\Gamma }$ with respect to S, and since the expected conclusion is independent of the base point we can assume that $x=o$ is the identity element of $\overline {\Gamma }$ , so that $d_{\mathcal {X}}(gx,x)=|g|_S$ is the word length of g for any $g\in \overline {\Gamma }$ .

Now we fix $[g]\in \mathbf {conj}(\overline {\Gamma })$ , set $\ell =\ell _{\mathcal {X}}[g]=\ell _S[g]$ and consider the sets $E_n[g]=\{g\in [g]\colon |g|_S\leq \ell +n\}.$ Note that $\# E_0[g]\leq \ell $ since any two conjugate elements that minimize the word length are actually cyclically conjugated with respect to some minimal word representations. Also, if $g\in E_n[g]$ , and $n>0$ , then indeed $n\geq 2$ and g is represented by a word of the form $x_1a^{\pm }x_2a^{\mp }x_3$ , where $a\in S$ is a standard generator and all the letters in the words $x_1$ and $x_3$ commute with a. Then the element $g'$ represented by the word $x_1x_2x_3$ belongs to $E_{n-2}[g]$ , and there are at most $\#{S}\cdot (\ell +n)(\ell +n-1)/2$ ways to reconstruct g from $g'$ . Therefore, we have

$$ \begin{align*}\#E_n[g]\leq \#{S}\cdot(\ell+n)(\ell+n-1)/2\cdot \#E_{n-2}[g]\end{align*} $$

for each n, and hence $ \#\{g\in [g] \colon |g|_S\leq \ell +R\}\leq \#E_{2R}[g]\leq \hat {Q}(\ell )$ for

$$ \begin{align*}\hat{Q}(t)=(\#S)^R\cdot(t+2R)(t+2R-1)\cdots (t+1)t/2^R.\end{align*} $$

This concludes the proof of the claim and the lemma.

Lemma 6.13. There exists $C'>0$ such that for any nontorsion conjugacy class $[g]\in \mathbf {conj}(\Gamma )$ we can find a representative $\hat {g}\in [g]$ and a closed path $\omega _{[g]}\in P(\Sigma _{\mathcal {C}}^\times )$ satisfying

$$\begin{align*}\min\{d_{\mathcal{X}}(\hat{g} o,\gamma^+_{\omega_{[g]}}),d_{\mathcal{X}}(\hat{g}^{-1}o,\gamma^+_{\omega_{[g]}} )\} \leq C',\end{align*}$$

and additionally

$$\begin{align*}\max\{|\ell^{\mathfrak{w}}_{\mathcal{X}}[g]-\ell^{\mathfrak{w}}_{\mathcal{X}}[\beta(\omega_{[g]})]|, |\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]-\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[\beta(\omega_{[g]})]|\}\leq C'.\end{align*}$$

Remark 6.14. As we will see in the proof, the constant $C'$ above depends on the initial data from Convention 6.3 and the maximal component $\mathcal {C}$ . The constant also depends on the existence of a contracting element in $\Gamma $ , as we rely on the work of Yang [Reference Yang88, Theorem C].

We will need the next lemma, which follows immediately from Remarks 2.1 and 2.2.

Lemma 6.15. Let $\gamma $ be a g-invariant geodesic in $\mathcal {Z}$ for some $g\in \Gamma $ . Then the image of $\gamma $ under $\phi $ (resp. $\phi _\ast $ ) in $\mathcal {X}$ (resp. $\mathcal {X}_\ast $ ) is a (possibly nonparametrized) g-invariant geodesic. For $\mathcal {X}_\ast $ we allow the degenerate case that $\phi _\ast (\gamma )$ is a point (which happens if and only if $\ell _{\mathcal {X}_\ast }[g]=0$ ).

Proof of Lemma 6.13

Before starting with the proof we provide a brief sketch. Given the conjugacy class $[g]$ , we first find an appropriate representative $g\in [g]$ such that $d_{\mathcal {Z}}(\widetilde o,g \widetilde o)$ is uniformly comparable to $\ell _{\mathcal {Z}}[g]$ , with similar versions for $\ell _{\mathcal {X}}^{\mathfrak {w}}$ and $\ell _{\mathcal {X}_\ast }^{\mathfrak {w}_\ast }$ . Then, we use Lemma 6.11 and the existence of a contracting element to find an element $s\in \Gamma $ and a path $\omega '\in \Sigma _{\mathcal {C}}^\ast $ such that both $so$ and $sgo$ are within uniformly bounded distance from the good representative path $\gamma _{\omega '}$ . Using the recurrence of $\mathcal {C}$ , we construct the close path $\omega _{[g]}$ as the concatenation of an appropriate subpath of $\omega '$ and a path in $\Sigma _{\mathcal {C}}^\ast $ of uniformly bounded length. For the path representative $\gamma =\gamma _{\omega _{[g]}}$ we then find $s'\in \overline {\Gamma }$ such that $d_{\mathcal {X}}(s'so,\gamma ^-)$ and $d_{\mathcal {X}}(s'sgo,\gamma ^+)$ are uniformly bounded. The rest of the proof consists of verifying that $\hat {g}=s'sg(s's)^{-1}$ satisfies all the desired inequalities for the appropriate constant $C'$ .

Now we start with the proof of the lemma, for which we consider the following constants. Let $M_1$ be such that any two vertices in $\mathcal {C}$ can be joined by a path in $\mathcal {C}$ of length at most $M_1$ (in both directions). This number exists since $\mathcal {C}$ is recurrent. Also, the projection $\phi :\mathcal {Z} \rightarrow \mathcal {X}$ is a quasi-isometry since it is $\Gamma $ -equivariant and the action of $\Gamma $ on both $\mathcal {Z}$ and $\mathcal {X}$ is proper and cocompact, so let $M_2>0$ be such that

$$\begin{align*}d_{\mathcal{Z}}(x,y)\leq M_2d_{\mathcal{X}}(\phi(x),\phi(y))+M_2\end{align*}$$

for all $x,y\in \mathcal {Z}$ . In addition, let $M_3$ be the diameter of $\phi ^{-1}(K)\subset \mathcal {Z}$ and fix a constant $M_4$ larger than N and the diameter of K. Finally, let L be the maximum of all the weights $\mathfrak {w}(\mathfrak {h})$ or $\mathfrak {w}_\ast (\mathfrak {h}_\ast )$ among hyperplanes $\mathfrak {h} \in \mathbb {H}(\mathcal {X})$ and $\mathfrak {h}_\ast \in \mathbb {H}(\mathcal {X}_\ast )$ , so that the four functions

$$\begin{align*}\phi: \mathcal{Z} \rightarrow \mathcal{X}^{\mathfrak{w}}, \ \ \phi_\ast: \mathcal{Z} \rightarrow \mathcal{X}_\ast^{\mathfrak{w}_\ast}, \ \ {Id}:\mathcal{X} \rightarrow \mathcal{X}^{\mathfrak{w}}, \ \ {Id}:\mathcal{X}_\ast \rightarrow \mathcal{X}_\ast^{\mathfrak{w}_\ast}\end{align*}$$

are L-Lipschitz.

We now start the proof, so we let $g\in \Gamma $ represent the nontorsion conjugacy class $[g]\in \mathbf {conj}(\Gamma )$ . Then g fixes a bi-infinite combinatorial axis $\widetilde {\lambda }$ in the cubical barycentric subdivision $\dot {\mathcal {Z}}$ [Reference Haglund46, Theorem 1.4]. After conjugating by an element of $\Gamma $ we can assume that $d_{\mathcal {Z}}(\widetilde o,\widetilde {\lambda })\leq M_3$ , so in particular we have $|d_{\mathcal {Z}}(\widetilde o,g \widetilde o)-\ell _{\mathcal {Z}}[g]|\leq 2M_3.$

By Lemma 6.15 and the fact that $\phi ,\phi _\ast $ are Lipschitz, the images $\lambda =\phi (\widetilde \lambda )$ and $\lambda _\ast =\phi _\ast (\widetilde \lambda )$ are also g-invariant (unparametrized) geodesics satisfying $d^{\mathfrak {w}}_{\mathcal {X}}( o,\lambda )\leq LM_3$ and $d^{\mathfrak {w}_\ast }_{\mathcal {X}_\ast }(o_\ast ,\lambda _\ast )\leq LM_3$ , which gives us

(6.11) $$ \begin{align} |d^{\mathfrak{w}}_{\mathcal{X}}(o,g o)-\ell^{\mathfrak{w}}_{\mathcal{X}}[g]|\leq 2LM_3 \ \ \text{ and } \ \ |d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(o_\ast,g o_\ast)-\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]|\leq 2LM_3. \end{align} $$

The action of $\Gamma $ on $\mathcal {X}$ is proper, cocompact and has a contracting element, and hence by [Reference Yang88, Theorem C] there exists a constant $\epsilon>0$ satisfying the following for any $h\in \Gamma $ . Let $\mathcal {V}_h$ denote the set of all the group elements $k\in \Gamma $ such that if $\gamma \subset \mathcal {X}$ is a combinatorial geodesic path with endpoints $\gamma ^{\pm }$ verifying $d_{\mathcal {X}}(\gamma ^-,o)\leq M_4$ and $d_{\mathcal {X}}(\gamma ^+,ko)\leq M_4$ , then there exists no $s\in \Gamma $ such that $d_{\mathcal {X}}(so,\gamma )\leq \epsilon $ and $d_{\mathcal {X}}(sho,\gamma )\leq \epsilon $ . Then

$$\begin{align*}\lim_{n\to \infty}{\frac{\#(\mathcal{V}_h\cap B_n)}{\#B_n}}=0\end{align*}$$

(the freedom in our choice for $M_4$ comes from the Remark after Theorem C in [Reference Yang88]). In virtue of Lemma 6.11 we conclude that the set $\Gamma _{\mathcal {C}} \backslash \mathcal {V}_h$ is nonempty for every $h\in \Gamma $ . Applying this to $h=g$ we deduce the existence of a path $\omega '\in \Sigma _{\mathcal {C}}^\times $ and $s\in \Gamma $ such that $d_{\mathcal {X}}(so,\gamma _{\omega '})\leq \epsilon $ and $d_{\mathcal {X}}(sgo,\gamma _{\omega '})\leq \epsilon $ .

Let $u,v\in \gamma _{\omega '}$ be such that $d_{\mathcal {X}}(so,u)\leq \epsilon $ and $d_{\mathcal {X}}(sgo,v)\leq \epsilon $ , and without loss of generality assume that u belongs to the portion of $\gamma _{\omega '}$ from $\gamma ^-_{\omega '}$ to v. Let $\omega =\omega _{[g]}\in \Sigma _{\mathcal {C}}^\times $ be a closed path composed by the concatenation of the subpath $\overline {\omega }'$ of $\omega '$ that determines the portion of $\gamma _{\omega '}$ from u to v and a path in $\Sigma _{\mathcal {C}}^\times $ of length at most $M_1$ from the final vertex of $\overline {\omega }'$ to its initial vertex. Let $\gamma =\gamma _{\omega }\subset \mathcal {X}$ be the good representative of $\omega $ with

(6.12) $$ \begin{align} L^{-1}d^{\mathfrak{w}}_{\mathcal{X}}(\gamma^-,o)\leq d_{\mathcal{X}}(\gamma^-,o) \leq N, \end{align} $$

and let $s'\in \overline {\Gamma }$ be such that $\gamma ^-=s'u$ . This implies

(6.13) $$ \begin{align} L^{-1}d^{\mathfrak{w}}_{\mathcal{X}}(s'so,\gamma^-)\leq d_{\mathcal{X}}(s'so,\gamma^-)\leq \epsilon \ \text{ and }\ L^{-1}d^{\mathfrak{w}}_{\mathcal{X}}(s'sgo,\gamma^+)\leq d_{\mathcal{X}}(s'sgo,\gamma^+)\leq \epsilon+M_1.\end{align} $$

Since $\omega $ is a loop we have $\gamma ^+=q\gamma ^-$ for $[q]=\beta (\omega )\in \mathbf {conj}(\overline {\Gamma })$ , and in particular from (6.11) we get

$$\begin{align*}|\ell^{\mathfrak{w}}_{\mathcal{X}}[g]-\ell^{\mathfrak{w}}_{\mathcal{X}}[\beta(\omega)]|\leq L(2\epsilon+M_1+2M_3).\end{align*}$$

Also, for $k\geq 1$ let $\omega ^{(k)}\in \Sigma _{\mathcal {C}}^\times $ be the concatenation of k copies of $\omega $ , and let $\gamma ^{(k)}\subset \mathcal {X}$ be a good representative of $\omega ^{(k)}$ so that $(\gamma ^{(k)})^{-}=\gamma ^{-}$ and $(\gamma ^{(k)})^{+}=q^k\gamma ^{-}$ . Note that $\gamma ^{(k)}$ is always a subpath of $\gamma ^{(k+1)}$ .

Now we lift each $\gamma ^{(k)}$ to $\mathcal {Z}$ to get a sequence $\widetilde \gamma ^{(k)}$ of geodesic paths in $\mathcal {Z}$ . Then $\phi (\widetilde \gamma ^{(k)})=\gamma ^{(k)}$ (up to parametrization), so that $(\gamma ^{(k)})^\pm =\phi ((\widetilde \gamma ^{(k)})^\pm )$ for all k. We denote $\widetilde \gamma ^\pm =(\widetilde \gamma ^{(1)})^\pm $ and we assume $(\widetilde \gamma ^{(k)})^-= \widetilde \gamma ^{-}$ for all k. By $\Gamma $ -equivariance of $\phi $ we have $\phi ((\widetilde \gamma ^{(k)})^+)=(\gamma ^{(k)})^+=q^k\gamma ^-=\phi (q^k\widetilde {\gamma }^{-} )$ , and hence

(6.14) $$ \begin{align} d_{\mathcal{Z}}((\widetilde \gamma^{(k)})^+,q^k\widetilde{\gamma}^{-})\leq M_3 \end{align} $$

for all k. Also, the inequalities (6.13) imply

(6.15) $$ \begin{align} d_{\mathcal{Z}}(s's\widetilde o,\widetilde{\gamma}^-)\leq M_2\epsilon+M_2 \ \ \text{ and }\ \ d_{\mathcal{Z}}(s'sg\widetilde o,\widetilde{\gamma}^+)\leq M_2(\epsilon+M_1)+M_2. \end{align} $$

We project the geodesics $\widetilde \gamma ^{(k)}$ to $\mathcal {X}_\ast ^{\mathfrak {w}_\ast }$ via $\phi _\ast $ , so we consider $\gamma _\ast ^{(k)}:=\phi _\ast (\widetilde \gamma ^{(k)})$ , which are (unparametrized) geodesics by Lemma 6.15, and as before we denote $ \gamma _\ast ^\pm =\phi _\ast (\gamma _\ast ^\pm )$ .

The length of $\gamma ^{(k)}_\ast $ in $\mathcal {X}_\ast ^{\mathfrak {w}_\ast }$ equals $\overline {\mathfrak {w}}_\ast (\alpha (\omega ^{(k)}))$ , and since the word $\alpha (\omega ^{(k)})$ is the concatenation of k copies of $\alpha (\omega )$ we have

(6.16) $$ \begin{align} d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(\gamma_\ast^{-},(\gamma_\ast^{(k)})^{+})=kd^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(\gamma_\ast^{-},\gamma_\ast^{+}) \end{align} $$

for all k. In addition, by (6.14) and (6.15) we obtain

(6.17) $$ \begin{align} d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}((\gamma_\ast^{(k)})^+,q^k\gamma^{-}_\ast)\leq L M_3, \end{align} $$

and

(6.18) $$ \begin{align} d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(s's o_\ast,\gamma_\ast^-)\leq LM_2(\epsilon+1) \ \ \text{ and } \ \ d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(s'sg o_\ast,\gamma^+_\ast)\leq LM_2(\epsilon+M_1+1). \end{align} $$

From these inequalities and (6.11) we get

$$ \begin{align*} \ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[q] \leq d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(\gamma_\ast^-,q\gamma_\ast^-) & \leq d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(\gamma_\ast^-,\gamma_\ast^+)+LM_3 \\ & \leq d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(o_\ast, g o_\ast)+L(M_3+M_2(2\epsilon+M_1+2))\\ & \leq \ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]+L(3M_3+M_2(2\epsilon+M_1+2)). \end{align*} $$

On the other hand, (6.16) and (6.17) imply

$$ \begin{align*} kd^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(\gamma_\ast^-,\gamma_\ast^+)=d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(\gamma_\ast^-,(\gamma_\ast^{(k)})^+) \leq d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(\gamma_\ast^-,q^k\gamma_\ast^-)+LM_3, \end{align*} $$

and after dividing by k and letting k tend to infinity we get

$$ \begin{align*} d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(\gamma_\ast^-,\gamma_\ast^+)\leq \ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[q]. \end{align*} $$

Combining this inequality with (6.11) and (6.18) gives us

$$ \begin{align*} \ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g] & \leq d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(s's o_\ast, s's g o_\ast)+2LM_3 \\ & \leq d^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}(\gamma_\ast^-,\gamma_\ast^+)+L(2M_3+M_2(2\epsilon+M_1+2)) \\ & \leq \ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[q]+L(2M_3+M_2(2\epsilon+M_1+2)), \end{align*} $$

and we deduce

$$ \begin{align*} |\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]-\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[\beta(\omega)]|=|\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]-\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[q]|\leq L(3M_3+M_2(2\epsilon+M_1+2)). \end{align*} $$

Finally, if we define $\hat {g}=s'sg(s's)^{-1}\in [g]$ , then by (6.12) and (6.13) we get

$$ \begin{align*}d_{\mathcal{X}}(\hat{g} o, \gamma^+)\leq d_{\mathcal{X}}(\hat{g} o, s'sgo)+\epsilon+M_1 \leq d_{\mathcal{X}}((s's)^{-1}o,o)+\epsilon+M_1\leq 2\epsilon+M_1+N.\end{align*} $$

In conclusion, the lemma follows with $C'=(L+1)(N+3M_3+M_2(2\epsilon +M_1+2))$ .

Now we prove Theorem 1.4.

Proof of Theorem 1.4

Given $\Gamma \in \mathfrak {G}$ and cubulations $\mathcal {X}$ and $\mathcal {X}_\ast $ of $\Gamma $ , we know that $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )\in \mathfrak {X}$ . Applying Proposition 5.7 we obtain a finite index subgroup $\overline {\Gamma }<\Gamma $ and a cubulation $\mathcal {Z}$ of $\Gamma $ , so that the $\Gamma $ -essential cores $\hat {\mathcal {X}}, \hat {\mathcal {X}}_\ast $ of $\mathcal {X},\mathcal {X}_\ast $ , respectively, are restriction quotients of $\mathcal {Z}$ . Taking a further finite index subgroup if necessary, and by applying Lemma 5.3, we can assume that $\overline {\mathcal {X}}=\overline {\Gamma } \backslash \mathcal {X}$ is special, and we choose this group $\overline {\Gamma }$ as data for part i).

The graph $\mathcal {G}$ for part ii) is the underlying graph for the automaton $\mathcal {A}_{\overline {\Gamma },\phi }$ from Theorem 5.11, applied to the triplet $\Gamma ,\hat {\mathcal {X}},\mathcal {Z}$ which by construction satisfies Convention 5.10. We note that the labeling $\pi =\pi _\phi $ still has image in $S_{\overline {\mathcal {X}}}$ since $\overline {\Gamma } \backslash \hat {\mathcal {X}} \rightarrow \overline {\mathcal {X}}$ is a convex isometric embedding (so the set of oriented hyperplanes in $\overline {\Gamma } \backslash \hat {\mathcal {X}}$ injects into $S_{{\overline {\mathcal {X}}}}$ ).

The triplet $(\Gamma ,\hat {\mathcal {X}},\mathcal {Z})$ also satisfies Convention 6.3 with $\hat {\mathcal {X}}_\ast =\mathcal {Z}(\mathbb {W}_\ast )$ (here $\mathfrak {w} \equiv \mathfrak {w}_\ast \equiv 1$ are the trivial orthotope structures), and the function $\psi =\psi _{\hat {\mathcal {X}}_\ast }^{\mathfrak {w}_\ast }$ is given as in Definition 6.5. This completes the data in part iii), and for a path $\omega $ in $\mathcal {G}$ , the associated loop $\overline {\gamma }_\omega $ in $\overline {\mathcal {X}}$ is the image under $\mathcal {X} \rightarrow \overline {\Gamma } \backslash \hat {\mathcal {X}} \subset \overline {\mathcal {X}}$ of any good representative $\gamma =\gamma _\omega $ as in Definition 6.4.

We are left to prove the claims (1)-(3) from the statement. Item (1) follows from Theorem 5.11 (2), and Item (2) follows from Lemma 6.7. Finally, Item (3) follows from Lemma 6.12 and Lemma 6.11.

To end the subsection we prove Proposition 6.8.

Proof of Proposition 6.8

For each $s\in {\mathbb {R}}$ and $R>0$ we consider the sums

$$\begin{align*}\mathcal{P}(R, T,s) = \sum_{|\ell^{\mathfrak{w}}_{\mathcal{X}}[g] - T| \le R} e^{-s\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]} \ \text{ and } \ \mathcal{P}_{\mathcal{C}}(R, T,s) = \sum_{\tau \in P(\Sigma_{\mathcal{C}}^{r_{\mathcal{X}}^{\mathfrak{w}}},R,T)} e^{-s \int_\tau{\Phi}}. \end{align*}$$

Since $l_\tau =\ell _{\mathcal {X}}^{\mathfrak {w}}[\beta (\tau )]$ for any closed orbit $\tau $ , by Lemmas 6.7 and 6.12 there exists a polynomial Q depending only on R such that $\mathcal {P}_{\mathcal {C}}(R, T,s) \le Q(T)\mathcal {P}(R, T,s)$ for each $s\in {\mathbb {R}}$ and $T>0$ .

For an inequality in the other direction, for any $[g]$ we use Lemma 6.13 to find a path $\omega _{[g]} \in P(\Sigma _{\mathcal {C}}^\times )$ and a representative $\hat {g}$ of $[g]$ satisfying

(6.19) $$ \begin{align} \min \{d_{\mathcal{X}}(\hat{g} o, \gamma^+_{\omega_{[g]}}),d_{\mathcal{X}}(\hat{g}^{-1} o, \gamma^+_{\omega_{[g]}})\} \leq C' \end{align} $$

and

(6.20) $$ \begin{align} \max\{|\ell^{\mathfrak{w}}_{\mathcal{X}}[g]-\ell^{\mathfrak{w}}_{\mathcal{X}}[\beta(\omega_{[g]})]|, |\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]-\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[\beta(\omega_{[g]})]|\}\leq C'. \end{align} $$

From (6.19) we get that the association $[g] \mapsto \omega _{[g]}$ is uniformly finite-to- $1$ . We extend this association to $[g]\mapsto \omega _{[g]} \mapsto \tau _{[g]}$ , where $\tau _{[g]}\in P(\Sigma _{\mathcal {C}}^{r_{\mathcal {X}}^{\mathfrak {w}}})$ is the periodic orbit corresponding to the path $\omega _{[g]}$ . Since changing the initial vertex of a closed path in $P(\Sigma _{\mathcal {C}}^\times )$ does not change the periodic orbit in $P(\Sigma _{\mathcal {C}}^{r_{\mathcal {X}}^{\mathfrak {w}}})$ , the association $\omega _{[g]} \mapsto \tau _{[g]}$ is at most (linear in $\ell _{\mathcal {X}}[g]$ )-to-1. But $\ell _{\mathcal {X}}[g]$ is comparable to $\ell ^{\mathfrak {w}}_{\mathcal {X}}[g]=\ell ^{\mathfrak {w}}_{\mathcal {X}}[\beta (\tau _{[g]})]=l_{\tau _{[g]}}$ (recall that $\mathcal {X}$ and $\mathcal {X}^{\mathfrak {w}}$ are quasi-isometric), and so from (6.20) we deduce that for each $s \in {\mathbb {R}}$ there is $C_s>0$ such that $\mathcal {P}(R,T,s) \le C^{\prime }_s F(T)\mathcal {P}_{\mathcal {C}}(R+C',T,s)$ for each $T> 0$ , where F is a degree 1 polynomial depending only on R.

It follows that for any fixed R sufficiently large and for any $s \in {\mathbb {R}}$

(6.21) $$ \begin{align} \lim_{T\to\infty} \frac{1}{T} \log \mathcal{P}(R,T,s) =\lim_{T \to \infty} \frac{1}{T} \log \left( \sum_{\tau \in P(\Sigma^{r_{\mathcal{X}}^{\mathfrak{w}}}_{\mathcal{C}},R,T)}e^{-s \int_\tau{\Phi}}\right)={P}_{\mathcal{C}}(-s\Phi), \end{align} $$

where $\text {P}_{\mathcal {C}}(-s\Phi )$ is the pressure of the potential $-s\Phi $ on the suspension $(\Sigma ^{r_{\mathcal {X}}^{\mathfrak {w}}}_{\mathcal {C}},\sigma ^{r_{\mathcal {X}}^{\mathfrak {w}}})$ .

Also, note that

$$\begin{align*}\sum_{[g] \in \mathbf{conj}(\Gamma)} e^{- t\ell^{\mathfrak{w}}_{\mathcal{X}}[g] - s\ell_{\mathcal{X}^{\mathfrak{w}_\ast}_\ast}[g]} \le e^{R|t|} \sum_{T = 1}^\infty \mathcal{P}(R,T,s) e^{-tT} \end{align*}$$

assuming the right-hand side of the above converges. Similarly we have

$$\begin{align*}\sum_{T = 1}^\infty \mathcal{P}(R,T,s) e^{-tT} \le 2R e^{R|t|}\sum_{[g] \in \mathbf{conj}(\Gamma)} e^{- t\ell^{\mathfrak{w}}_{\mathcal{X}}[g] - s\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]} \end{align*}$$

when the right-hand side converges. We deduce that for each $s \in {\mathbb {R}}$ the series

$$\begin{align*}\sum_{T = 1}^\infty \mathcal{P}(R,T,s) e^{-tT} \ \text{ and } \ \sum_{[g] \in \mathbf{conj}(\Gamma)}e^{- t\ell^{\mathfrak{w}}_{\mathcal{X}}[g] - s\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]} \end{align*}$$

have the same abscissa of convergence as t varies. Hence by (6.21) we deduce

$$\begin{align*}\theta_{\mathcal{X}^{\mathfrak{w}_\ast}_\ast/\mathcal{X}^{\mathfrak{w}}}(s) = \text{P}_{\mathcal{C}}(-s\Phi), \end{align*}$$

as desired.

6.2 Analyticity and Large deviations for pairs of cubulations

In this subsection we prove Theorems 6.1 and 6.2. For a triplet $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )\in \mathfrak {X}$ we always assume that it satisfies Convention 6.3, which is possible by Proposition 5.7. In particular, all the results and notations from this and the previous section are valid for this triplet. We first prove a large deviations principle that follows from Proposition 6.8.

Corollary 6.16. Let $(\Gamma ,\mathcal {X},\mathcal {X}_\ast )\in \mathfrak {X}$ and $\mathfrak {w},\mathfrak {w}_\ast $ be $\Gamma $ -invariant orthotope structures on $\mathcal {X},\mathcal {X}_\ast $ , and let $\mathcal {L} : [\mathrm {Dil}(\mathcal {X}^{\mathfrak {w}}, \mathcal {X}^{\mathfrak {w}_\ast }_\ast )^{-1}, \mathrm {Dil}(\mathcal {X}^{\mathfrak {w}_\ast }_\ast , \mathcal {X}^{\mathfrak {w}}) ] \to {\mathbb {R}}$ be the Legendre transform of $\theta _{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}$ . Then for any nonempty open set $U \subset {\mathbb {R}}$ and closed set $V \subset {\mathbb {R}}$ with $U \subset V$ we have that

$$ \begin{align*} -\inf_{s \in U} \mathcal{L}(s) &\le \liminf_{T\to\infty} \frac{1}{T} \log\left( \frac{1}{\#\mathfrak{C}_{\mathcal{X}^{\mathfrak{w}}}(T)} \#\left\{[g] \in \mathfrak{C}_{\mathcal{X}^{\mathfrak{w}}}(T): \frac{\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]}{ \ell^{\mathfrak{w}}_{\mathcal{X}}[g]}\in U \right\}\right)\\ &\le \limsup_{T\to\infty} \frac{1}{T} \log\left( \frac{1}{\#\mathfrak{C}_{\mathcal{X}^{\mathfrak{w}}}(T)} \#\left\{[g] \in \mathfrak{C}_{\mathcal{X}^{\mathfrak{w}}}(T): \frac{\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]}{ \ell^{\mathfrak{w}}_{\mathcal{X}}[g]}\in V \right\}\right) \le -\inf_{s \in V} \mathcal{L}(s). \end{align*} $$

In consequence, the limit

$$\begin{align*}\tau(\mathcal{X}^{\mathfrak{w}_\ast}_\ast/\mathcal{X}^{\mathfrak{w}}) := \lim_{T\to\infty} \frac{1}{\#\mathfrak{C}_{\mathcal{X}^{\mathfrak{w}}}(T)} \sum_{[g] \in \mathfrak{C}_{\mathcal{X}^{\mathfrak{w}}}(T)} \frac{\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]}{\ell^{\mathfrak{w}}_{\mathcal{X}}[g]} \end{align*}$$

exists and equals $-\theta ^{\prime }_{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}(0)$ .

Proof. Recall that for $T>R >0$ and $s \in {\mathbb {R}}$ we defined

$$\begin{align*}\mathcal{P}(R, T,s) = \sum_{|\ell^{\mathfrak{w}}_{\mathcal{X}}[g] - T| \le R} e^{-s\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]} \end{align*}$$

in the proof of Proposition 6.8 above. We saw during that proof that

$$\begin{align*}\lim_{T\to\infty} \frac{1}{T} \log \mathcal{P}(R,T,s) = \theta_{\mathcal{X}^{\mathfrak{w}_\ast}_\ast/\mathcal{X}^{\mathfrak{w}}}(s). \end{align*}$$

It follows from the Gärtner-Ellis Theorem [Reference Dembo and Zeitouni29, Theorem 2.3.6] that the large deviations principle stated in this corollary holds but with $\mathfrak {C}_{\mathcal {X}^{\mathfrak {w}}}(T)$ replaced by

$$\begin{align*}\mathfrak{C}_{\mathcal{X}^{\mathfrak{w}}}(T, R) = \{ [g]\in \mathbf{conj} \colon |\ell^{\mathfrak{w}}_{\mathcal{X}}[g] - T| < R\} \end{align*}$$

for any fixed $R>0$ sufficiently large. It is then easy to check that this large deviations principle implies the one stated in the corollary.

By this large deviations principle we know that for any $\epsilon>0$ the cardinality of the set

$$\begin{align*}E_\epsilon(T): = \left\{[g] \in \mathfrak{C}_{\mathcal{X}^{\mathfrak{w}}}(T): \left|\frac{\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]}{ \ell^{\mathfrak{w}}_{\mathcal{X}}[g]} + \theta^{\prime}_{\mathcal{X}^{\mathfrak{w}_\ast}_\ast/\mathcal{X}^{\mathfrak{w}}}(0) \right|> \epsilon \right\} \end{align*}$$

grows strictly exponentially slower than $\#\mathfrak {C}_{\mathcal {X}^{\mathfrak {w}}}(T)$ as $T\to \infty $ , that is, the quotient $\#E_\epsilon (T)/\#\mathfrak {C}_{\mathcal {X}^{\mathfrak {w}}}(T)$ decays to $0$ exponentially as $T\to \infty $ . It is then standard to deduce that

$$\begin{align*}\tau(\mathcal{X}_\ast^{\mathfrak{w}_\ast}/\mathcal{X}^{\mathfrak{w}}) := \lim_{T\to\infty} \frac{1}{\#\mathfrak{C}_{\mathcal{X}^{\mathfrak{w}}}(T)} \sum_{[g] \in \mathfrak{C}_{\mathcal{X}^{\mathfrak{w}}}(T)} \frac{\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]}{\ell^{\mathfrak{w}}_{\mathcal{X}}[g]} \end{align*}$$

exists and is equal to $ -\theta _{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}'(0)$ as required.

Proof of Theorem 6.1

We showed in Proposition 6.8 that $\theta _{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}(s)$ is equal to the pressure $ {P}_{\mathcal {C}}(-s\Phi )$ for any s. It follows that $\theta _{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}$ is analytic, convex and decreasing (see also Rmark 2.10). Also, by Corollary 6.16 we know that the limit labeled $\tau (\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}})$ in the theorem exists. Further by comparing the exponential growth rates of both sides of the inequality

$$\begin{align*}\#\left\{[g] \in \mathfrak{C}_{\mathcal{X}^{\mathfrak{w}}}(T): \left|\frac{\ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g]}{ \ell^{\mathfrak{w}}_{\mathcal{X}}[g]} - \tau(\mathcal{X}^{\mathfrak{w}_\ast}_\ast/\mathcal{X}^{\mathfrak{w}}) \right| \le \epsilon \right\} \le \#\left\{ [g] \in \mathfrak{C}_{\mathcal{X}^{\mathfrak{w}}}(T) : \ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast}[g] \le (\tau(\mathcal{X}^{\mathfrak{w}_\ast}_\ast/\mathcal{X}^{\mathfrak{w}}) + \epsilon) T\right\} \end{align*}$$

we see that

$$\begin{align*}\tau(\mathcal{X}^{\mathfrak{w}_\ast}_\ast/\mathcal{X}^{\mathfrak{w}}) \ge \frac{v_{\mathcal{X}^{\mathfrak{w}}}}{v_{\mathcal{X}^{\mathfrak{w}_\ast}_\ast}}. \end{align*}$$

Therefore to conclude the proof we need to check the equivalence of the statements $(1)$ , $(2)$ and $(3)$ when the action of $\Gamma $ on $\mathcal {X}_\ast $ (and hence on $\mathcal {X}_\ast ^{\mathfrak {w}_\ast }$ ) is proper. When this is the case we have $0<v_{\mathcal {X}^{\mathfrak {w}_\ast }_\ast } < \infty $ and the Manhattan curve $ \theta _{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}(s)$ is $0$ at $s=v_{\mathcal {X}^{\mathfrak {w}_\ast }_\ast }.$

We will prove the implications $(1) \Rightarrow (3) \Rightarrow (2) \Rightarrow (1)$ . Note that the implication $(1) \Rightarrow (3)$ follows easily from the facts that $\tau (\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}) = - \theta ^{\prime }_{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}(0)$ , $\theta _{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}(0) = v_{\mathcal {X}^{\mathfrak {w}}}, \theta _{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}(v_{\mathcal {X}^{\mathfrak {w}_\ast }_\ast }) = 0$ and $\theta _{\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}}$ is convex so has nonincreasing derivative. Also, the implication $(2) \Rightarrow (1)$ follows from the definition of the Manhattan curve. Hence we just need to prove the implication $(3) \Rightarrow (2).$

To do so we note that

$$\begin{align*}\tau(\mathcal{X}_\ast^{\mathfrak{w}_\ast}/\mathcal{X}^{\mathfrak{w}}) = \int_{\Sigma_{\mathcal{C}}^{r_{\mathcal{X}}^{\mathfrak{w}}}} \Phi \ dm \end{align*}$$

where m is the measure of maximal entropy for $(\Sigma _{\mathcal {C}}^{r_{\mathcal {X}}^{\mathfrak {w}}},\sigma ^{r_{\mathcal {X}}^{\mathfrak {w}}})$ . However we saw in Subsection 2.5 that

$$\begin{align*}\int_{\Sigma_{\mathcal{C}}^{r_{\mathcal{X}}^{\mathfrak{w}}}} \Phi \ dm = \frac{\int_{\Sigma_{\mathcal{C}}} \psi_{\mathcal{X}_\ast}^{\mathfrak{w}_\ast}\ d\mu_1}{\int_{\Sigma_{\mathcal{C}}} r_{\mathcal{X}}^{\mathfrak{w}} \ d\mu_1} \end{align*}$$

where $\mu _1$ is the equilibrium state of $- \delta _{r_{\mathcal {X}}^{\mathfrak {w}}} r_{\mathcal {X}}^{\mathfrak {w}}$ on $\Sigma _{\mathcal {C}}$ . To simplify notation going forward we will also write $r =r_{\mathcal {X}}^{\mathfrak {w}}$ , $\psi =\psi _{\mathcal {X}_\ast }^{\mathfrak {w}_\ast } $ and $\mu _2$ for the equilibrium state of $- \delta _{\psi _{\mathcal {X}_\ast }^{\mathfrak {w}_\ast }} \psi _{\mathcal {X}_\ast }^{\mathfrak {w}_\ast }$ on $\Sigma _{\mathcal {C}}$ . We now note that by Proposition 6.8 we have that

$$\begin{align*}\text{P}_{\mathcal{C}}(-v_{\mathcal{X}^{\mathfrak{w}}} r_{\mathcal{X}}^{\mathfrak{w}}) = \text{P}_{\mathcal{C}}(-v_{{\mathcal{X}^{\mathfrak{w}_\ast}_\ast}} \psi_{\mathcal{X}_\ast}^{\mathfrak{w}_\ast}) = 0. \end{align*}$$

Here the pressures are the pressures of the potentials over the subshift (not suspension). Hence the inequality $\tau (\mathcal {X}^{\mathfrak {w}_\ast }_\ast /\mathcal {X}^{\mathfrak {w}}) \ge v_{\mathcal {X}^{\mathfrak {w}}}/v_{\mathcal {X}^{\mathfrak {w}_\ast }_\ast }$ can be rewritten as

$$\begin{align*}\frac{h_{\mu_2}(\sigma)}{\int_{\Sigma_{\mathcal{C}}} \psi \ d\mu_2} \ge \frac{h_{\mu_1}(\sigma)}{\int_{\Sigma_{\mathcal{C}}} \psi \ d\mu_1} \end{align*}$$

where $h_{\mu _1}(\sigma ), h_{\mu _2}(\sigma )$ are the entropies of $\mu _1, \mu _2$ over the component ${\mathcal {C}}$ . This inequality is true by the variational principle. Furthermore this inequality is a strict equality unless r and $\psi $ are cohomologous. This implies by Lemmas 6.7 and 6.13 that there exist $\Lambda , C>0$ such that

$$\begin{align*}|\ell^{\mathfrak{w}}_{\mathcal{X}^{\mathfrak{w}}}[g] - \Lambda \ell^{\mathfrak{w}_\ast}_{\mathcal{X}_\ast^{\mathfrak{w}_\ast}}[g]| < C \end{align*}$$

for all $[g] \in \mathbf {conj}(\Gamma )$ . This can only happen if $(2)$ holds.

Proof of Theorem 6.2

Let $\psi =\psi _{\mathcal {X}_\ast }^{\mathfrak {w}_\ast }:\Sigma \rightarrow {\mathbb {Z}}$ be the potential associated to the constant orthotope structure $\mathfrak {w}_\ast \equiv 1$ . Let $\mathcal {L}:[\mathrm {Dil}(\mathcal {X}, \mathcal {X}_\ast )^{-1}, \mathrm {Dil}(\mathcal {X}_\ast , \mathcal {X}) ]\rightarrow {\mathbb {R}}$ be the Legendre transform of $\theta _{\mathcal {X}_\ast /\mathcal {X}}$ , which by Remark 6.9 equals the Legendre transform of $s \mapsto {P}_{\mathcal {C}}(-s\psi )$ for $\mathcal {C}$ any maximal recurrent component of $\mathcal {G}_\phi $ . Hence $\mathcal {L}$ is analytic.

From our large devation principle in Corollary 6.16 we have that

$$\begin{align*}\limsup_{T\to\infty} \frac{1}{T} \log \left(\#\left\{ [g] \in \mathbf{conj}: \ell_{\mathcal{X}}[g] < T, | \ell_{\mathcal{X}_\ast}[g] - \eta \ell_{\mathcal{X}}[g] | < \frac{C}{T} \right\}\right) \le v_{\mathcal{X}} - \mathcal{L}(\eta) \end{align*}$$

for all $\eta \in (\mathrm {Dil}(\mathcal {X}, \mathcal {X}_\ast )^{-1}, \mathrm {Dil}(\mathcal {X}_\ast , \mathcal {X})).$

We now prove the lower bound. Fix a maximal component ${\mathcal {C}}$ . By Lemma 6.7 and Lemma 6.12 there exists a polynomial Q such that for any $C>0$ and $\eta \in (\mathrm {Dil}(\mathcal {X}, \mathcal {X}_\ast )^{-1}, \mathrm {Dil}(\mathcal {X}_\ast , \mathcal {X}))$

$$\begin{align*}\# \left\{\omega \in P_n(\Sigma_{\mathcal{C}}^\times): \left|\frac{\psi^n(\omega)}{n} - \eta \right| < \frac{C}{n} \right\} \le Q(n)\cdot \#\left\{ [g] \in \mathbf{conj}: \ell_{\mathcal{X}}[g] \le n, \left| \frac{\ell_{\mathcal{X}_\ast}[g]}{\ell_{\mathcal{X}}[g]} - \eta \right| < \frac{C}{n} \right\} \end{align*}$$

where $\psi $ is the potential from Definition 6.5. However, by Theorem 3.2 (and Remark 3.3 as ${\mathcal {C}}$ may only be transitive) we have that

$$\begin{align*}\limsup_{n\to\infty} \frac{1}{n} \log \left(\# \left\{\omega \in P_n(\Sigma_{\mathcal{C}}^\times): \left|\frac{\psi^n(\omega)}{n} - \eta \right| < \frac{C}{n} \right\} \right)= h - \mathcal{I}(\eta) \end{align*}$$

where $\mathcal {I}$ is the Legendre transform of the map $s \mapsto {P}_{\mathcal {C}}(-s\psi )$ and h is the topological entropy of the subshift $(\Sigma _{\mathcal {C}},\sigma )$ . However, as we saw above, $\mathcal {I}$ is precisely $\mathcal {L}$ and further by Lemma 6.11 we have $h = v_{\mathcal {X}}$ . Hence we deduce that

$$\begin{align*}\limsup_{T\to\infty} \frac{1}{T} \log \#\left\{ [g] \in \mathbf{conj}: \ell_{\mathcal{X}}[g] < T, | \ell_{\mathcal{X}_\ast}[g] - \eta \ell_{\mathcal{X}}[g] | < \frac{C}{T} \right\} \ge v_{\mathcal{X}} - \mathcal{L}(\eta) \end{align*}$$

for each $\eta \in (\mathrm {Dil}(\mathcal {X}, \mathcal {X}_\ast )^{-1}, \mathrm {Dil}(\mathcal {X}_\ast , \mathcal {X}))$ . We have shown that the limit supremum in the statement of the theorem is equal to $v_{\mathcal {X}} - \mathcal {L}$ as required.

To conclude the proof we need to explain the additional conditions mentioned in the theorem. In particular we need to show that

$$\begin{align*}0 < v_{\mathcal{X}} - \mathcal{L}(\eta) \le v_{\mathcal{X}} \text{ for all } \eta \in (\mathrm{Dil}(\mathcal{X}, \mathcal{X}_\ast)^{-1}, \mathrm{Dil}(\mathcal{X}_\ast, \mathcal{X})) \end{align*}$$

and that the upper bound inequality is an equality if and only if $\eta = \tau (\mathcal {X}_\ast /\mathcal {X}).$ All of these properties follow from the definition of $\mathcal {L}$ and the fact that $s\mapsto {P}_{\mathcal {C}}(-s\psi )$ is strictly convex.

Appendix A Convex-cocompact subgroups of cubulable relatively hyperbolic groups

In this appendix we prove Proposition 5.4. First, we recall the statement.

Proposition A.1. Let $\Gamma $ be a relatively hyperbolic group acting properly and cocompactly on the $\mathrm {CAT}(0)$ cube complex $\mathcal {X}$ . Then the following are equivalent for a subgroup $H< \Gamma $ .

  1. (1) H is convex-cocompact for the action on $\mathcal {X}$ .

  2. (2) H is relatively quasiconvex and $H\cap P$ is convex-cocompact for the action of $\Gamma $ on $\mathcal {X}$ for any maximal parabolic subgroup $P<\Gamma $ .

Proof. Under these assumptions $\Gamma $ is finitely generated, so fix $S\subset \Gamma $ a finite symmetric generating set and a $\Gamma $ -equivariant quasi-isometry $\phi : \mathcal {X} \rightarrow \mathrm {Cay}(\Gamma ,S)$ . We also fix a vertex $x_0\in \mathcal {X}$ such that $\phi (x_0)=o$ is the identity element in $\Gamma $ . Let $\mathbb {P}$ be a complete collection of representatives of conjugacy classes of maximal parabolic subgroups in $\Gamma $ , and let $\mathcal {P}=(\bigcup \mathbb {P}) \backslash \{o\}$ . We let $d_S$ denote the (graph) word metric on $\mathrm {Cay}(\Gamma ,S)$ and let $H<\Gamma $ be any subgroup.

If H is convex-cocompact, then it is undistorted, hence finitely generated and relatively quasiconvex by [Reference Hruska50, Theorem 1.5]. Also, any maximal parabolic subgroup $P<\Gamma $ is convex-cocompact by [Reference Sageev and Wise76, Theorem 1.1], and hence $H\cap P$ is also convex-cocompact by [Reference Reyes71, Lemma 2.14 & Lemma 2.15]. This proves the implication $(1) \Rightarrow (2)$ .

The implication $(2) \Rightarrow (1)$ is more involved, and for its proof we adopt the following convention. If $\gamma '$ is a parameterized curve and $x=\gamma ^{\prime }_{t^-},y=\gamma ^{\prime }_{t^+}$ belong to $\gamma '$ with $t^-\leq t^+$ , then $\gamma '|_{[x,y]}=\gamma '|_{[y,x]}$ is a set of points of form $\gamma ^{\prime }_t$ , with $t^-\leq t \leq t^+$ (if there is more than one option for $t^\pm $ , we consider any of them).

By [Reference Genevois39, Lemma 4.3] it is enough to prove the following: there exists $K>0$ such that if $\overline {\gamma } \subset \mathcal {X}$ is a (continuous) combinatorial geodesic with endpoints in $H x_0$ , then $\overline {\gamma } \subset N_K(H x_0)$ .

To find such K, consider constants $L,C$ such that the image under $\phi $ of any combinatorial geodesic $\overline {\gamma }$ in $\mathcal {X}$ is at Hausdorff distance at most L from an L-Lipschitz $(L,C)$ -quasigeodesic $\gamma :[0,\ell ]\rightarrow \mathrm {Cay}(\Gamma ,S)$ with same endpoints as $\phi (\overline {\gamma })$ (see e.g., [Reference Burago, Burago and Ivanov11, Proposition 8.3.4]).

Let $\overline {\gamma } \subset \mathcal {X}$ be a geodesic with endpoints in $Hx_0$ and let $\gamma =\gamma ([0,\ell ])\subset \mathrm {Cay}(\Gamma ,S)$ be as above, so that the endpoints of $\gamma $ belong to H. Also, let $\hat {c}$ be a geodesic in $\mathrm {Cay}(\Gamma ,S\cup \mathcal {P})$ with same endpoints as $\gamma $ and let $\hat {c}_0,\dots , \hat {c}_n$ be the (ordered) vertex set of $\hat {c}$ . We define $I=\{j_0<j_i<\cdots <j_k\}$ to be the set of all $0\leq i \leq n-1$ such that $\hat {c}_{i+1}^{-1}\hat {c}_i\in \mathcal {P}$ .

By quasiconvexity of H, there exists $\kappa $ (independent of $\gamma $ ) and $h_i\in H$ such that $d_S(\hat {c}_i,h_i)\leq \kappa $ for all $0\leq i \leq n$ , see e.g., [Reference Hruska50, Def. 6.10]. Also, by [Reference Hruska50, Lemma 8.8] there exists $A_0$ depending only on $L,C$ such that for any $0\leq i \leq n$ there exists $c_i=\gamma _{t_i}\in \gamma $ satisfying $d_S(c_i,\hat {c}_i)\leq A_0$ , for which we assume $c_0=\hat {c}_0$ and $c_n=\hat {c}_n$ . Up to increasing $A_0$ (only in terms of $L,C$ ), we can always assume that $c_i$ is a group element.

Since $\Gamma $ is finitely generated, by [Reference Hruska50, Proposition 9.4] there exists $B_0$ depending only on $A_0$ and $\kappa $ (so only on $L,C$ ) such that if $g_1,g_2\in \Gamma $ satisfy $|g_1|_S,|g_2|_S\leq A_0+\kappa $ , then

(A.1) $$ \begin{align} N_{A_0+\kappa}(g_1H) \cap N_{A_0+\kappa}(g_2P)\subset N_{B_0}(g_1Hg_1^{-1}\cap g_2Pg_2^{-1}) \end{align} $$

for any $P\in \mathbb {P}$ , where the neighborhoods are considered in $\mathrm {Cay}(\Gamma ,S)$ .

Let $\widetilde p$ be a geodesic lift of $\hat {c}$ to $\mathrm {Cay}(\Gamma ,S)$ . That is, $\widetilde p$ is obtained from $\hat {c}$ by replacing each edge corresponding to an element of $\mathcal {P}$ by a geodesic in $\mathrm {Cay}(\Gamma ,S)$ with the same endpoints. For a point $x\in \gamma $ we distinguish two cases.

Case 1: $x\in \gamma |_{[c_{{j_i}+1},c_{j_{i+1}}]}$ for some $j_i\in I$ (with the convention that $j_{-1}=-1$ and $j_{k+1}=n$ ). Consider geodesic paths $[c_{{j_i}+1},\hat {c}_{{j_i}+1}]$ and $[c_{j_{i+1}}, \hat {c}_{j_{i+1}}]$ in $\mathrm {Cay}(\Gamma ,S)$ , and the quasigeodesic triangle with sides

$$\begin{align*}\ell_1=[\hat{c}_{{j_i}+1},c_{{j_i}+1}] \cup \gamma|_{[c_{{j_i}+1},x]} , \ \ \ell_2= \gamma|_{[x,c_{j_{i+1}}]} \cup [c_{j_{i+1}}, \hat{c}_{j_{i+1}}], \ \ \ell_3= \widetilde p|_{[c_{j_{i+1}}, c_{{j_i}+1}]}.\end{align*}$$

We also set

$$ \begin{align*}\ell_1^-=\hat{c}_{j_i+1}, \ell_1^+=x, \ \ \ell_2^-=\hat{x}, \ell_2^+=\hat{c}_{j_{i+1}}, \ \ \text{ and } \ \ \ell_3^-=\hat{c}_{j_{i+1}}, \ell_3^+=\hat{c}_{j_i+1}.\end{align*} $$

Note that $\ell _1,\ell _2,\ell _3$ are Lipschitz quasigeodesics with constants depending only on $L,C$ and $A_0$ (hence only on $L,C$ ). Then by [Reference Druţu and Sapir30, Lemma 8.19] there exists R depending on $L,C$ such that either:

  • there exists $z\in \mathrm {Cay}(\Gamma ,S)$ with $d_S(z,\ell _i)\leq R$ for $i=1,2,3$ ; or,

  • there exist $g\in \Gamma $ and $P\in \mathbb {P}$ such that $d_S(gP,\ell _i)\leq R$ for $i=1,2,3$ .

In the first subcase, let $u_i\in \ell _i$ be such that $d_S(z,u_i)\leq R$ . Then $d_S(u_a,u_b)\leq 2R$ for all $1\leq a,b\leq 3$ , and since $\gamma $ is $(L,C)$ -quasigeodesic and $x\in \gamma |_{[u_1,u_2]}$ , we have that $d_S(x,\ell _3)\leq d_S(x,u_3)\leq d_S(x,u_1)+d_S(u_1,u_3)$ is bounded above in terms of $L,C$ and R (thus only in terms of $L,C$ ).

In the second subcase, by [Reference Druţu and Sapir30, Lemma 8.15] we can find M and $\mathfrak {d}$ depending only on $L,C$ and R (so only on $L,C$ ) and points $u_i^-,u_i^+\in \ell _i$ for $i=1,2,3$ that satisfy:

  • $d_S(u_i^\pm ,gP) \leq M$ ; and,

  • $\operatorname {\mathrm {diam}}(\ell _i|_{[\ell _i^\pm ,u_i^\pm ]}\cap \overline {N}_M(gP))\leq \mathfrak {d}$ .

Take $v_i^\pm \in gP$ such that $d_S(u_i^\pm ,v_i^\pm )\leq M$ . Then by the definition of I and after considering vertices in $\hat {c}$ that are closest to $u_3^\pm $ in $\mathrm {Cay}(\Gamma ,S)$ we get

$$\begin{align*}d_S(u_3^+,u_3^-)\leq d_{S\cup \mathcal{P}}(v_3^-,v_3^+)+2(1+M)+2\leq 5+2M. \end{align*}$$

Also, [Reference Druţu and Sapir30, Lemma 8.14] implies the existence of $D_1$ depending only on $L,C,M$ and $\mathfrak {d}$ (so only on $L,C$ ) with $d_S(u_i^+,u_{i+1}^-)\leq D_1$ (mod 3) for all i. In particular,

$$ \begin{align*}d_S(u_1^-,u_2^+)\leq d_S(u_1^-,u_3^+)+d_S(u_3^+,u_3^-)+d_S(u_3^-,u_2^+),\end{align*} $$

and as in the first subcase we conclude that x belongs to a neighborhood of $\ell _3$ depending only on L and C.

In both subcases, we deduce that $d_S(x,\ell _3)$ is bounded in terms of $L,C$ , and since $\ell _3$ is contained in a neighborhood of H depending only on $\kappa $ , we have that

(A.2) $$ \begin{align} d_S(x,H)\leq K_0 \end{align} $$

for some $K_0$ depending only on $L,C$ and $\kappa $ (hence only in terms of $L,C$ ).

Case 2: $x\in \gamma |_{[c_{j},c_{j+1}]}$ for some $j\in I$ . Suppose $\hat {c}_j^{-1}\hat {c}_{j+1}=p\in P$ for $P\in \mathbb {P}$ . Then

$$ \begin{align*} d_S(c_j^{-1}c_{j+1},(c_j^{-1}h_j)H) \leq d_S(c_j^{-1}c_{j+1},c_j^{-1}h_{j+1}) = d_S(c_{j+1},h_{j+1}) \leq A_0+\kappa, \end{align*} $$

and

$$ \begin{align*} d_S(c_j^{-1}c_{j+1}, (c_j^{-1}\hat{c}_j)P) \leq d_S(c_j^{-1}c_{j+1},c_j^{-1}\hat{c}_jp) = d_S(c_{j+1},\hat{c}_{j+1})\leq A_0. \end{align*} $$

Since $\max \{|c_j^{-1}h_j|_S,|c_j^{-1}\hat {c}_j^{-1}|_S\}\leq A_0+\kappa $ , by (A.1) we conclude

(A.3) $$ \begin{align} d_S(c_j^{-1}c_{j+1}, (c_j^{-1}h_j)H(c_j^{-1}h_j)^{-1}\cap (c_j^{-1}\hat{c}_j)P(c_j^{-1}\hat{c}_j)^{-1})\leq B_0. \end{align} $$

Note that any point $x\in \gamma $ satisfies the assumptions of one of the two cases above. Indeed, for $x=\gamma _t\in \gamma $ , let $I_-$ be the set of all the $j\in I$ such that $c_j$ is not of the form $\gamma _{t'}$ with $t'>t$ . Suppose first that $I_-$ is nonempty and let j be its maximal element. If x does not satisfy Case 2, then $c_{{j}+1}$ does not belong to $\gamma |_{[x,\gamma _\ell ]}$ . But if $j=j_i<j_k$ , then $j_{i+1}\notin I_-$ , so that $x\in \gamma |_{[c_{{j_i}+1},c_{j_{i+1}}]}$ and x satisfies Case 1.

Also, if $j=j_k$ , then $x\in \gamma |_{[c_{{j_i}+1},c_n]}$ and x also satisfies Case 1. Therefore, we can assume that $I_-$ is empty. But if I is nonempty then $x\in \gamma |_{[c_0,c_{j_1}]}$ and x satisfies Case 1, and if I is empty then $x\in \gamma |_{[c_0,c_{n}]}$ and x also satisfies Case 1.

Now, take $\overline {x}\in \overline {\gamma }$ and let $x\in \gamma $ within r from $\phi (\overline {x})$ in $\mathrm {Cay}(\Gamma ,S)$ , where r is independent of $\overline {x}$ and $\overline {\gamma }$ . If x satisfies Case 1, by (A.2) we conclude that $d_{\mathcal {X}}(\overline {x},H x_0)\leq K_1$ for $K_1$ a constant independent of $\overline {x}$ and $\overline {\gamma }$ .

If x satisfies Case 2, suppose that $x\in \gamma |_{[c_j,c_{j+1}]}$ for $j\in I$ . Then by (A.3) there exist vertices $\overline {x}^-,\overline {x}^+\subset \overline {\gamma }$ satisfying $d_{\mathcal {X}}(c_j x_0,\overline {x}^-)\leq \hat {r}$ and $d_{\mathcal {X}}(c_{j+1} x_0,\overline {x}^+)\leq \hat {r}$ , where $\hat {r}$ depends only on $\phi $ and $L,C$ .

Let F be the set of pairs $\alpha ,\beta \in \Gamma $ satisfying $|\alpha |_S,|\beta |_S\leq A_0+\kappa $ . By our assumption and [Reference Sageev and Wise76, Theorem 1.1] we can find a convex core $Z_{\alpha ,\beta }\subset \mathcal {X}$ for the group $\alpha H\alpha ^{-1}\cap \beta P \beta ^{-1}$ that contains the $\hat {r}$ -neighborhood of $x_0$ . By cocompactness, we can find $K_2>0$ such that

$$ \begin{align*} Z_{\alpha,\beta}\subset N_{K_2}((\alpha H\alpha^{-1}\cap \beta P \beta^{-1})x_0)\subset N_{K_2}((\alpha H\alpha^{-1})x_0) \end{align*} $$

for all $(\alpha ,\beta )\in F$ . Note that $K_2$ is independent of $\overline {\gamma }$ . In particular we have

$$ \begin{align*} \overline{x} \in c_j Z_{c_j^{-1}h_j,c_j^{-1}\hat{c}_j}\subset c_jN_{K_3}(c_j^{-1} H(h_j^{-1}c_j)x_0)\subset N_{K_2}(Hx_0), \end{align*} $$

where $K_3:=K_2+\max \{d_{\mathcal {X}}(\alpha x_0,x_0)\colon |\alpha |_S\leq A_0+\kappa \}$ is independent of $\overline {x}$ and $\overline {\gamma }$ . In conclusion, $\overline {\gamma }\subset N_K(H x_0)$ for $K:=\max \{K_1,K_3\}$ , and the implication $(2) \Rightarrow (1)$ follows.

Acknowledgments

Both authors are grateful to Richard Sharp for his comments and suggestions on a preliminary version of this paper. The second author would like to thank the Max Planck Institut für Mathematik for its hospitality and financial support.

Competing interest

The authors have no competing interests to declare.

References

Agol, I., ‘The virtual Haken conjecture’. With an appendix by I. Agol, D. Groves and J. F. Manning. Doc. Math. 18 (2013), 10451087.10.4171/dm/421CrossRefGoogle Scholar
Aougab, T., Clay, M. and Rieck, Y., ‘Thermodynamic metrics on outer space’, Ergodic Theory Dynam. Systems 43(3) (2023), 729793.CrossRefGoogle Scholar
Bergeron, N. and Wise, D. T., ‘A boundary criterion for cubulation’, Amer. J. Math. 134(3) (2012), 843859.CrossRefGoogle Scholar
Beyrer, J. and Fioravanti, E., ‘Cross ratios and cubulations of hyperbolic groups’, Math. Ann. 384(3–4) (2022), 15471592.CrossRefGoogle Scholar
Beyrer, J. and Fioravanti, E., ‘Cross-ratios on CAT(0) cube complexes and marked length-spectrum rigidity’, J. Lond. Math. Soc. (2) 104(5) (2021), 19732015.CrossRefGoogle Scholar
Bray, H., Canary, R. and Kao, L.-Y., ‘Pressure metrics for deformation spaces of quasifuchsian groups with parabolics’, Algebr. Geom. Topol. 23(8) (2023), 36153653.10.2140/agt.2023.23.3615CrossRefGoogle Scholar
Bray, H., Canary, R., Kao, L.-Y. and Martone, G., ‘Counting, equidistribution and entropy gaps at infinity with applications to cusped Hitchin representations’, J. Reine Angew. Math. 791 (2022), 151.10.1515/crelle-2022-0035CrossRefGoogle Scholar
Bregman, C., Charney, R. and Vogtmann, K., ‘Outer space for RAAGs’, Duke Math. J. 172(6) (2023), 10331108.10.1215/00127094-2023-0007CrossRefGoogle Scholar
Bridgeman, M., Canary, R., Labourie, F. and Sambarino, A., ‘The pressure metric for Anosov representations’, Geom. Funct. Anal. 25(4) (2015), 10891179.CrossRefGoogle Scholar
Bridson, M. and Haefliger, A., Metric spaces of non-positive curvature . Grundlehren der Mathematischen Wissenschaften, vol. 319 (Springer-Verlag, Berlin, 1999).CrossRefGoogle Scholar
Burago, D., Burago, Y. and Ivanov, S., A Course in Metric Geometry . Graduate Studies in Mathematics, vol. 33 (American Mathematical Society, Providence, RI, 2001).10.1090/gsm/033CrossRefGoogle Scholar
Burger, M., ‘Intersection, the Manhattan curve, and Patterson-Sullivan theory in rank 2’, Int. Math. Res. Not. 1993(7), 217225.CrossRefGoogle Scholar
Bowen, R. and Series, C., ‘Markov maps associated with Fuchsian groups’, Inst. Hautes Études Sci. Publ. Math. 50 (1979), 153170.CrossRefGoogle Scholar
Calegari, D. and Fujiwara, K., ‘Combable functions, quasimorphisms, and the central limit theorem’, Ergodic Theory Dynam. Systems 30(5) (2010), 13431369.CrossRefGoogle Scholar
Cannon, J. W., ‘The combinatorial structure of cocompact discrete hyperbolic groups’, Geom. Dedicata 16(2) (1984), 123148.CrossRefGoogle Scholar
Cantrell, S., ‘Mixing of the Mineyev flow, orbital counting and Poincaré series for strongly hyperbolic metrics’, Math. Ann. 392 (2025), 12531288.10.1007/s00208-025-03127-4CrossRefGoogle Scholar
Cantrell, S. and Reyes, E., Marked length spectrum rigidity from rigidity on subsets. arXiv:2304.13209, 2023.Google Scholar
S. Cantrell and, R, ‘Tanaka, Invariant measures of the topological flow and measures at infinity on hyperbolic groups’, J. Mod. Dyn. 20 (2024), 215274.CrossRefGoogle Scholar
Cantrell, S. and Tanaka, R., ‘The Manhattan curve, ergodic theory of topological flows and rigidity’, To appear in Geom. Topol., arXiv:2104.13451, 2021.Google Scholar
Caprace, P.-E. and Sageev, M., ‘Rank rigidity for CAT(0) cube complexes’, Geom. Funct. Anal. 21 (2011), 851891.Google Scholar
Charney, R. and Davis, M. W., ‘The K(π,1)-problem for hyperplane complements associated to infinite reflection groups’, J. Amer. Math. Soc. 8(3) (1995), 597627.Google Scholar
Charney, R., Stambaugh, N. and Vogtmann, K., ‘Outer space for untwisted automorphisms of right-angled Artin groups’, Geom. Topol. 21 (2017), 11311178.Google Scholar
Chepoi, V., ‘Graphs of some CAT(0) complexes’, Adv. Appl. Math. 24(2) (2000), 125179.CrossRefGoogle Scholar
Cooper, D. and Futer, D., ‘Ubiquitous quasi-Fuchsian surfaces in cusped hyperbolic 3–manifolds’, Geom. Topol. 23(1) (2019), 241298.10.2140/gt.2019.23.241CrossRefGoogle Scholar
Culler, M. and Vogtmann, K., ‘Moduli of graphs and automorphisms of free groups’, Invent. Math. 84(1) (1986), 91119.10.1007/BF01388734CrossRefGoogle Scholar
Dahmani, F., Futer, D. and Wise, D. T., ‘Growth of quasiconvex subgroups’, Math. Proc. Cambridge Philos. Soc. 167(3) (2019), 505530.CrossRefGoogle Scholar
Dai, X. and Martone, G., ‘Correlation of the renormalized Hilbert length for convex projective surfaces’, Ergodic Theory Dynam. Systems 43 (2023), 29382973.Google Scholar
Dal’bo, F., ‘Remarques sur le spectre des longueurs d’une surface et comptages’, Bol. Soc. Brasil. Mat. (N.S.) 30(2) (1999), 199221.CrossRefGoogle Scholar
Dembo, A. and Zeitouni, O., Large Deviations Techniques and Applications . Applications of Mathematics (New York), vol. 38, 2nd ed. (Springer-Verlag, New York, 1998).10.1007/978-1-4612-5320-4CrossRefGoogle Scholar
Druţu, C. and Sapir, M., ‘Tree-graded spaces and asymptotic cones of groups’. With an appendix by D. V. Osin and M. Sapir. Topology 44 (2005), 9591058.Google Scholar
Erlandsson, V., ‘A remark on the word length in surface groups’. Trans. Amer. Math. Soc. 327 (2019), 441455.10.1090/tran/7561CrossRefGoogle Scholar
Epstein, D. B. A., Cannon, J. W., Holt, D., Levy, S., Paterson, M. and Thurston, W., Word Processing in Groups (Jones and Bartlett Publishers, Boston, MA, 1992).CrossRefGoogle Scholar
Farb, B. and Margalit, D., A Primer on Mapping Class Groups . Princeton Mathematical Series, vol. 49 (Princeton University Press, 2012).Google Scholar
Fioravanti, E., ‘On automorphisms and splittings of special groups’, Compos. Math. 159(2) (2023), 232305.10.1112/S0010437X22007850CrossRefGoogle Scholar
Fioravanti, E., Levcovitz, I. and Sageev, M., ‘Coarse cubical rigidity’, J. Topol. 17(3) (2024), Paper No. e12353, 50.10.1112/topo.12353CrossRefGoogle Scholar
Gekhtman, I., Taylor, S. J. and Tiozzo, G., ‘Central limit theorems for counting measures in coarse negative curvature’, Compos. Math. 158(10) (2022), 19802013.CrossRefGoogle Scholar
Gekhtman, I., Taylor, S. J. and Tiozzo, G., ‘Counting loxodromics for hyperbolic actions’, J. Topol. 11(2) (2018), 379419.10.1112/topo.12053CrossRefGoogle Scholar
Gekhtman, I., Taylor, S. J. and Tiozzo, G., ‘Counting problems in graph products and relatively hyperbolic groups’, Israel J. Math. 237 (2020), 311371.10.1007/s11856-020-2008-xCrossRefGoogle Scholar
Genevois, A., ‘Hyperbolicities in CAT(0) cube complexes’, Enseign. Math. 65(1/2) (2019), 33100.10.4171/lem/65-1/2-2CrossRefGoogle Scholar
Godelle, E. and Paris, L., ‘K(π,1) and word problems for infinite type Artin-Tits groups, and applications to virtual braid groups’, Math. Z. 272 (2012), 13391364.10.1007/s00209-012-0989-9CrossRefGoogle Scholar
Groves, D. and Manning, J. F., ‘Specializing cubulated relatively hyperbolic groups’, J. Topol. 15 (2022), 398442.CrossRefGoogle Scholar
Guivarc’h, Y. and Hardy, J., ‘Théorèmes limites pour une classe de chaines de Markov et applications aux difféomorphismes d’Anosov’, Ann. Inst. Henri Poincaré Probab. Stat. 24(1) (1988), 7398.Google Scholar
Hagen, M. F. and Przytycki, P., ‘Cocompactly cubulated graph manifolds’, Israel J. Math. 207(1) (2015), 377394.CrossRefGoogle Scholar
Hagen, M. F. and Wise, D. T., ‘Cubulating hyperbolic free-by-cyclic groups: the general case’, Geom. Funct. Anal. 25 (2015), 134179.Google Scholar
Hagen, M. F. and Wise, D. T., ‘Cubulating hyperbolic free-by-cyclic groups: the irreducible case’, Duke Math. J. 165(9) (2016), 17531813.10.1215/00127094-3450752CrossRefGoogle Scholar
Haglund, F., ‘Isometries of CAT(0) cube complexes are semi-simple’, Ann. Math. Québec (2021). https://doi.org/10.1007/s40316-021-00186-2.Google Scholar
Haglund, F. and Wise, D. T., ‘Special cube complexes’, Geom. Funct. Anal. 17(5) (2008), 15511620.10.1007/s00039-007-0629-4CrossRefGoogle Scholar
He, Y. M., Lee, H. and Park, I., Pressure metrics in geometry and dynamics. arXiv:2407.18441, 2024.Google Scholar
Hermiller, S. and Meier, J., ‘Algorithms and geometry for graph products of groups’, J. Algebra 171(1) (1995), 230257.10.1006/jabr.1995.1010CrossRefGoogle Scholar
Hruska, G. C., ‘Relative hyperbolicity and relative quasiconvexity for countable groups’, Algebr. Geom. Topol. 10(3) (2010), 18071856.10.2140/agt.2010.10.1807CrossRefGoogle Scholar
Hurwitz, A., ‘Über die angenäherte Darstellung der Irrationalzahlen durch rationale Brüche’, Math. Ann. 39(2) (1891), 279284.10.1007/BF01206656CrossRefGoogle Scholar
Kahn, J. and Markovic, V., ‘Immersing almost geodesic surfaces in a closed hyperbolic three manifold’, Ann. Math. (2) 175(3) (2012), 11271190.CrossRefGoogle Scholar
Kao, L.-Y., ‘Entropy rigidity, pressure metric, and immersed surfaces in hyperbolic 3-Manifolds’. In Thermodynamic Formalism: CIRM Jean-Morlet Chair, Fall 2019 (Springer, International Publishing, 2021), pp. 351393.CrossRefGoogle Scholar
Kao, L.-Y., ‘Manhattan curves for hyperbolic surfaces with cusps’, Ergodic Theory Dynam. Systems 40(7) (2020), 18431874.10.1017/etds.2018.124CrossRefGoogle Scholar
Kao, L.-Y., ‘Pressure metrics and Manhattan curves for Teichmüller spaces of punctured surfaces’, Israel J. Math. 240(2) (2020), 567602.10.1007/s11856-020-2073-1CrossRefGoogle Scholar
Kifer, Y., ‘Large deviations, averaging and periodic orbits of dynamical systems’, Comm. Math. Phys. 162 (1994), 3346.10.1007/BF02105185CrossRefGoogle Scholar
Lauer, J. and Wise, D. T., ‘Cubulating one-relator groups with torsion’, Math. Proc. Cambridge Philos. Soc. 155(3) (2013), 411429.10.1017/S0305004113000285CrossRefGoogle Scholar
Li, J. and Wise, D. T., ‘No growth-gaps for special cube complexes’, Groups Geom. Dyn. 14(1) (2020), 117135.CrossRefGoogle Scholar
Marcus, B. and Tuncel, S., ‘Entropy at a weight-per-symbol and embeddings of Markov chains’, Invent. Math. 102 (1990), 235266.CrossRefGoogle Scholar
Martin, A. and Steenbock, M., ‘A combination theorem for cubulation in small cancellation theory over free products’, Ann. Inst. Fourier (Grenoble) 67 (2017), 16131670.10.5802/aif.3118CrossRefGoogle Scholar
McMullen, C., ‘Thermodynamics, dimension and the Weil–Petersson metric’, Invent. Math. 173 (2008), 365425.10.1007/s00222-008-0121-2CrossRefGoogle Scholar
Niblo, G. A. and Reeves, L. D., ‘Coxeter groups act on CAT(0) cube complexes’, J. Group Theory 6(3) (2003), 399413.10.1515/jgth.2003.028CrossRefGoogle Scholar
Odrzygóźdź, T., ‘Cubulating random groups in the square model’, Israel J. Math. 227(2) (2018), 623661.10.1007/s11856-018-1734-9CrossRefGoogle Scholar
Ollivier, Y. and Wise, D. T., ‘Cubulating random groups at density less than 1/6’, Trans. Amer. Math. Soc. 363(9) (2011), 47014733.10.1090/S0002-9947-2011-05197-4CrossRefGoogle Scholar
Parry, W., ‘Intrinsic Markov chains, Trans. Amer. Math. Soc. 112 (1964), 5566.10.1090/S0002-9947-1964-0161372-1CrossRefGoogle Scholar
Parry, W. and Pollicott, M., ‘Zeta functions and the periodic orbit structure of hyperbolic dynamics’, Astérisque 187–188 (1990), 1268.Google Scholar
Pollicott, M. and Sharp, R., ‘A Weil–Petersson metric on spaces of metric graphs’, Geom. Dedicata 172 (2014), 229244.CrossRefGoogle Scholar
Pollicott, M. and Sharp, R., ‘Large deviations, fluctuations and shrinking intervals’, Comm. Math. Phys. 290 (2009), 321334.10.1007/s00220-008-0725-9CrossRefGoogle Scholar
Przytycki, P. and Wise, D. T., ‘Graph manifolds with boundary are virtually special’, J. Topol. 7(2) (2014), 419435.10.1112/jtopol/jtt009CrossRefGoogle Scholar
Przytycki, P. and Wise, D. T., ‘Mixed 3-manifolds are virtually special’, J. Amer. Math. Soc. 31(2) (2018), 319347.Google Scholar
Reyes, E., ‘On cubulated relatively hyperbolic groups’, Geom. Topol. 27 (2023), 575640.10.2140/gt.2023.27.575CrossRefGoogle Scholar
Roller, M. A., Poc sets, median algebras and group actions. An extended study of Dunwoody’s construction and Sageev’s theorem. arXiv:1607.07747; University of Southampton preprint, 1998.Google Scholar
Rousseau-Egele, J., ‘Un théoreme de la limite locale pour une classe de transformations dilatantes et monotones par morceaux’, Ann. Probab. 11 (1983), 772788.10.1214/aop/1176993522CrossRefGoogle Scholar
Sageev, M., ‘CAT(0) cube complexes and groups’. In Geometric Group Theory, IAS/Park City Math. Ser., vol. 21 (American Mathematical Society, Providence, RI, 2014), pp. 754.CrossRefGoogle Scholar
Sageev, M., ‘Ends of group pairs and non-positively curved cube complexes’, Proc. Lond. Math. Soc. 71(3) (1995), 585617.10.1112/plms/s3-71.3.585CrossRefGoogle Scholar
Sageev, M. and Wise, D. T., ‘Cores for quasiconvex actions’, Proc. Amer. Math. Soc. 143 (2015), 27312741.10.1090/S0002-9939-2015-12297-6CrossRefGoogle Scholar
Sale, A. and Susse, T., ‘Outer automorphism groups of right-angled Coxeter groups are either large or virtually abelian’, Trans. Amer. Math. Soc. 372(11) (2019), 77857803.10.1090/tran/7897CrossRefGoogle Scholar
Schwartz, R. and Sharp, R., ‘The correlation of length spectra of two hyperbolic surfaces’, Comm. Math. Phys. 154 (1993), 423430.CrossRefGoogle Scholar
Sharp, R., ‘Comparing length functions on free groups’. In Spectrum and Dynamics, CRM Conf. Proc. Lecture Notes, vol. 52 (Amer. Math. Soc., Providence, RI, 2010), pp. 185207.10.1090/crmp/052/10CrossRefGoogle Scholar
Sharp, R., ‘Distortion and entropy for automorphisms of free groups’, Discrete Contin. Dyn. Syst. 26 (2010), 347363.10.3934/dcds.2010.26.347CrossRefGoogle Scholar
Sharp, R., ‘The Manhattan curve and the correlation of length spectra on hyperbolic surfaces’, Math. Z. 228 (1998), 745750.10.1007/PL00004643CrossRefGoogle Scholar
Sigmund, K., ‘Generic properties of invariant measures for Axiom A diffeomorphisms’, Invent. Math. 11 (1970), 99109.10.1007/BF01404606CrossRefGoogle Scholar
Stucky, B., ‘Cubulating one-relator products with torsion’, Groups Geom. Dyn. 15 (2021), 691754.10.4171/ggd/619CrossRefGoogle Scholar
Tidmore, J., ‘Cocompact cubulations of mixed 3-manifolds’, Groups Geom. Dyn. 12(4) (2018), 14291460.10.4171/ggd/474CrossRefGoogle Scholar
Wise, D. T., ‘Cubulating small cancellation groups’, Geom. Funct. Anal. 14(1) (2004), 150214.10.1007/s00039-004-0454-yCrossRefGoogle Scholar
Wise, D. T., ‘The structure of groups with a quasiconvex hierarchy’, Ann. of Math. Stud. 209 (2021), 1376.Google Scholar
Wise, D. T. and Woodhouse, D. J., ‘A cubical flat torus theorem and the bounded packing property’, Israel J. Math. 217 (2017), 263281.10.1007/s11856-017-1445-7CrossRefGoogle Scholar
Yang, W., ‘Statistically convex-cocompact actions of groups with contracting elements’, Int. Math. Res. Not. IMRN 2019(23), 72597323.Google Scholar