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

Random walks and contracting elements III: Outer space and outer automorphism group

Inhyeok Choi Cornell University
310 Malott Hall, Ithaca, NY, USA, 14850
June E Huh Center for Mathematical Challenges, KIAS
85 Hoegiro Dongdaemun-gu, Seoul 02455, Republic of Korea
[email protected]
Abstract.

Continuing from [Cho22b], we study random walks on (possibly asymmetric) metric spaces using the bounded geodesic image property (BGIP) of certain isometries. As an application, we show that a generic outer automorphism of the free group of rank at least 3 has different forward and backward expansion factors. This answers a question of Handel and Mosher in [HM07b]. Together with this, we also revisit limit laws on Outer space including SLLN, CLT, LDP and the genericity of a triangular fully irreducible outer automorphisms.

Keywords. Random walk, Outer space, Outer automorphism group, Laws of large numbers, Central limit theorem, Expansion factor, Fully irreducible automorphisms

MSC classes: 20F67, 30F60, 57M60, 60G50

1. Introduction

This is the third in a series of articles concerning random walks on metric spaces with contracting elements. This series is a reformulation of the previous preprint [Cho22a] announced by the author. In this article, we adopt the following convention:

Let XX be a space with possibly an asymmetric metric. We say that a subset AA of XX exhibits bounded geodesic image property (BGIP in short) if the closest point projection of a geodesic γ\gamma onto AA is uniformly bounded whenever γ\gamma is far away from AA. An isometry gg of XX has BGIP if the orbit {gno}n\{g^{n}o\}_{n\in\operatorname{\mathbb{Z}}} is a bi-quasigeodesic with BGIP (see Definition 2.4). Two BGIP isometries gg and hh of XX are independent if there orbits have unbounded Hausdorff distance.

Convention 1.1.

Throughout, we assume that:

  • (X,d)(X,d) is a geodesic metric space, possibly with an asymmetric metric;

  • GG is a countable group of isometries of XX, and

  • GG contains two independent BGIP isometries.

We also fix a basepoint oXo\in X.

The main purpose of this article is to generalize the random walk theory in [Cho22b] to asymmetric metric spaces. Our main result concerns the mismatch of the forward translation length τ(g)\tau(g) and the backward translation length τ(g1)\tau(g^{-1}) of a generic isometry gg of XX.

Theorem A (Asymmetry of a generic translation length).

Let (X,G)(X,G) be as in Convention 1.1. Let (Zn)n0(Z_{n})_{n\geq 0} be the random walk generated by a non-elementary, asymptotically asymmetric measure μ\mu on GG. Then for any K>0K>0, we have

limn(|τ(Zn)τ(Zn1)|<K)=0.\lim_{n\rightarrow\infty}\operatorname{\mathbb{P}}\Big{(}\big{|}\tau(Z_{n})-\tau(Z_{n}^{-1})\big{|}<K\Big{)}=0.

As an application, we obtain the following corollary:

Corollary 1.2.

Let XX be Culler-Vogtmann Outer space of rank N3N\geq 3 and GG be the corresponding outer automorphism group of the free group FNF_{N}. Let (Zn)n0(Z_{n})_{n\geq 0} be an admissible random walk on GG. Then for any K>0K>0, we have

limn(ω:|τ(Zn)τ(Zn1)|<K)=0.\lim_{n\rightarrow\infty}\operatorname{\mathbb{P}}\Big{(}\operatorname{\omega}:\big{|}\tau(Z_{n})-\tau(Z_{n}^{-1})\big{|}<K\Big{)}=0.

Corollary 1.2 asserts that a generic outer automorphism has different expansion factor than its inverse. This was suggested by Handel and Mosher in [HM07b]. There, they proved the asymmetry for a large class of automorphisms, namely, the class of parageometric fully irreducibles ([HM07b, Theorem 1]). Nonetheless, Corollary 1.2 does not follow from Handel and Mosher’s result, as a generic outer automorphism is ageometric and not parageometric ([KMPT22, Corollary C]). Meanwhile, a companion result in [HM07a, Theorem 1.1] provides a constant CC depending on the rank NN of the free group such that, for any fully irreducible element ϕOut(FN)\phi\in\operatorname{Out}(F_{N}), the expansion factor λ\lambda of ϕ\phi and λ\lambda^{\prime} of ϕ1\phi^{-1} satisfy 1/Clog(λ/λ)C1/C\leq\log(\lambda/\lambda^{\prime})\leq C. In other words, the translation lengths of ϕ\phi and ϕ1\phi^{-1} are within bounded ratio (see also [AKB12, Theorem 23]).

Together with this, we also recover Horbez’s SLLN [Hor16a, Theorem 5.7] and CLT [Hor18, Theorem 0.2] for displacement, and Dahmani-Horbez’s SLLN for translation length [DH18, Theorem 1.4] on Outer space, the latter with a weaker and optimal moment condition. We also present a CLT for translation length and its converse, which seems new for Outer space. Moreover, we obtain optimal deviation inequalities on Outer space; see [Hor18] for previously known deviation inequalities. Using them, we also establish the geodesic tracking of random walks. Finally, we also discuss the exponential genericity of (atoroidal) fully irreducible automorphisms, which is a recurring theme in [MT18], [TT16] and [KMPT22]; note that we do not require moment conditions here.

In order to apply our general theory to Outer space, we crucially utilize the BGIP of fully irreducible outer automorphisms. Namely, we modify Kapovich-Maher-Pfaff-Taylor’s observation ([KMPT22, Theorem 7.8]) into the following form:

Theorem 1.3.

Let N3N\geq 3. Then every fully irreducible outer automorphism in Out(FN)\operatorname{Out}(F_{N}) has BGIP with respect to the closest point projection.

1.1. Structure of the article

In Section 2, we recall the notion of bounded geodesic image property (BGIP) and prove relevant lemmata. These lemmata were used in [Cho22b] to establish the alignment of sample paths of a random walk, which we rephrase in the language of BGIP. In Section 3, we summarize and generalize the limit laws discussed in [Cho22b] and [Cho22c] while pointing out a subtle difference. In Section 4, we review facts about the outer automorphism group and Outer space. The main result of this section is Theorem 1.3, the BGIP of fully irreducible outer automorphisms. Combining the results in Section 3 and Section 4, we obtain limit laws for Out(FN)\operatorname{Out}(F_{N}) in Section 5.

Acknowledgments

The author thanks Hyungryul Baik, Talia Fernós, Ilya Gekhtman, Thomas Haettel, Joseph Maher, Hidetoshi Masai, Catherine Pfaff, Yulan Qing, Kasra Rafi, Samuel Taylor and Giulio Tiozzo for helpful discussions. In particular, the author is grateful to Samuel Taylor for providing the author with a simpler proof of Theorem 1.3. The author is also grateful to the American Institute of Mathematics and the organizers and the participants of the workshop “Random walks beyond hyperbolic groups” in April 2022 for helpful and inspiring discussions.

The author was partially supported by Samsung Science & Technology Foundation grant No. SSTF-BA1702-01. This work constitutes a part of the author’s PhD thesis.

2. Preliminaries

In this section, we gather necessary notions and facts about geodesics, paths, and bounded geodesic image property (BGIP).

2.1. Asymmetric metric spaces

Definition 2.1 (Metric space).

An (asymmetric) metric space (X,d)(X,d) is a set XX equipped with a function d:X×X0d:X\times X\rightarrow\mathbb{R}_{\geq 0} that satisfies the following:

  • (non-degeneracy) for each x,yXx,y\in X, d(x,y)=0d(x,y)=0 if and only if x=yx=y;

  • (triangle inequality) for each x,y,zXx,y,z\in X, d(x,z)d(x,y)+d(y,z)d(x,z)\leq d(x,y)+d(y,z);

  • (local symmetry) for each xXx\in X, there exist ϵ,K>0\epsilon,K>0 such that d(y,z)Kd(z,y)d(y,z)\leq Kd(z,y) holds for y,z{aX:min(d(x,a),d(a,x))<ϵ}y,z\in\left\{a\in X:\min\left(d(x,a),d(a,x)\right)<\epsilon\right\}.

In this situation, we say that dd is a metric on XX. dd is said to be symmetric if d(x,y)=d(y,x)d(x,y)=d(y,x) holds for all x,yXx,y\in X. We define a symmetric metric called the symmetrization of dd by

dsym(x,y):=d(x,y)+d(y,x).d^{sym}(x,y):=d(x,y)+d(y,x).

We endow (X,d)(X,d) with the topology induced by dsymd^{sym}.

We define the Gromov product between xx and yy based at zz by

(x,y)z:=12[d(x,z)+d(z,y)d(x,y)].(x,y)_{z}:=\frac{1}{2}[d(x,z)+d(z,y)-d(x,y)].

From now on, we fix an (asymmetric) metric space (X,d)(X,d). The diameter of a set AXA\subseteq X is defined by

diam(A):=sup{d(x,y):x,yA},\operatorname{\operatorname{diam}}(A):=\sup\{d(x,y):x,y\in A\},

and the (directed) distances between sets A,BXA,B\subseteq X are defined by

d(A,B)\displaystyle d(A,B) :=inf{d(x,y):xA,yB},\displaystyle:=\inf\{d(x,y):x\in A,y\in B\},
dsym(A,B)\displaystyle d^{sym}(A,B) :=inf{dsym(x,y):xA,yB}.\displaystyle:=\inf\{d^{sym}(x,y):x\in A,y\in B\}.

For R>0R>0, the RR-neighborhood of a set AXA\subseteq X is defined by

𝒩R(A):={x:dsym(x,A)<R}.\mathcal{N}_{R}(A):=\{x:d^{sym}(x,A)<R\}.

The Hausdorff distance between A,BXA,B\subseteq X is defined by

dH(A,B):=inf{R>0:A𝒩R(B)andB𝒩R(A)}.d_{H}(A,B):=\inf\{R>0:A\subseteq\mathcal{N}_{R}(B)\,\,\textrm{and}\,\,B\subseteq\mathcal{N}_{R}(A)\}.

Given A,BXA,B\subseteq X, we say that AA is KK-coarsely contained by BB if A𝒩K(B)A\subseteq\mathcal{N}_{K}(B), and that AA and BB are KK-coarsely equivalent if dH(A,B)<Kd_{H}(A,B)<K. We say that AA is KK-coarsely connected if for each x,yAx,y\in A there exists a a chain x=a0,a1,an=yx=a_{0},a_{1}\ldots,a_{n}=y of points in AA such that dsym(ai,ai+1)<Kd^{sym}(a_{i},a_{i+1})<K.

A path is a map from an interval II or a set II of consecutive integers into XX. Given a path γ:IX\gamma:I\rightarrow X and a<ba<b in \mathbb{R}, we denote its restriction to I[a,b]I\cap[a,b] by γ|[a,b]\gamma|_{[a,b]}. Also, we define the reversal γ¯\bar{\gamma} of γ\gamma by reversing the orientation, i.e., by precomposing the map ttt\mapsto-t from I-I to II.

We say that two paths γ:IX\gamma:I\rightarrow X, η:JX\eta:J\rightarrow X are KK-fellow traveling if there exist a non-decreasing surjective map φ:IJ\varphi:I\rightarrow J such that dsym(γ(t),(ηφ)(t))<Kd^{sym}(\gamma(t),(\eta\circ\varphi)(t))<K for each tIt\in I.

Definition 2.2 (Quasigeodesics).

A path γ:IX\gamma:I\rightarrow X from an interval or a set of consecutive integers II is called a KK-quasigeodesic if

(2.1) 1K|ts|Kd(γ(s),γ(t))K|ts|+K\frac{1}{K}|t-s|-K\leq d(\gamma(s),\gamma(t))\leq K|t-s|+K

holds for all s,tIs,t\in I such that s<ts<t. If Inequality 2.1 holds for all s,tIs,t\in I, we say that γ\gamma is a KK-bi-quasigeodesic. An 11-quasigeodesic from an interval to XX is called a geodesic.

A metric space XX is said to be geodesic if every ordered pair of points can be connected by a geodesic, i.e., for every x,yXx,y\in X there exists a geodesic γ:[a,b]X\gamma:[a,b]\rightarrow X such that γ(a)=x\gamma(a)=x and γ(b)=y\gamma(b)=y. We denote γ\gamma by [x,y][x,y].

Remark 2.3.

In many asymmetric metric spaces (including Culler-Vogtmann Outer space), the reversal of a geodesic may not be a geodesic. Meanwhile, geodesics on asymmetric metric spaces are continuous thanks to the local symmetry of the metric.

We will frequently use Inequality 2.1 in the following form. For any points p,qp,q on a KK-bi-quasigeodesic γ\gamma, we have

(2.2) diam(γ1(p)γ1(q))Kd(p,q)+K2,\operatorname{\operatorname{diam}}\left(\gamma^{-1}(p)\cup\gamma^{-1}(q)\right)\leq Kd(p,q)+K^{2},
(2.3) d(q,p)Kdiam(γ1(p)γ1(q))+KK2d(p,q)+K3+K.d(q,p)\leq K\operatorname{\operatorname{diam}}\left(\gamma^{-1}(p)\cup\gamma^{-1}(q)\right)+K\leq K^{2}d(p,q)+K^{3}+K.

Fixing a basepoint oXo\in X, we define the translation length of an isometry gg of XX by

τ(g):=limn1nd(o,gno).\tau(g):=\lim_{n\rightarrow\infty}\frac{1}{n}d(o,g^{n}o).

The triangle inequality tells us that τ(g)\tau(g) does not depend on the choice of oo. Meanwhile, because the metric dd can be asymmetric, τ(g)\tau(g) and τ1(g)\tau^{-1}(g) are not equal in general.

2.2. Bounded geodesic image property (BGIP)

In this subsection, we fix a (possibly asymmetric) metric space (X,d)(X,d). Given a subset AA of XX, we define the closest point projection πA:X2A\pi_{A}:X\rightarrow 2^{A} onto AA by

πA(x):={aA:d(x,a)=d(x,A)}.\pi_{A}(x):=\{a\in A:d(x,a)=d(x,A)\}.
Definition 2.4 (Bounded geodesic image property).

A subset AXA\subseteq X of a geodesic metric space XX is said to satisfy the KK-bounded geodesic image property, or KK-BGIP in short, if the following hold:

  1. (1)

    for each zXz\in X, πA(z)\pi_{A}(z) is nonempty;

  2. (2)

    for each geodesic η\eta such that dsym(η,A)>Kd^{sym}(\eta,A)>K, we have diam(πA(η))K\operatorname{\operatorname{diam}}(\pi_{A}(\eta))\leq K.

A KK-bi-quasigeodesic that satisfies KK-BGIP is called a KK-BGIP axis.

Observation 2.5.

The reversal of a KK-BGIP axis is again a KK-BGIP axis.

Note that strongly contracting property (as in [Cho22b, Definition 2.1]) and BGIP are not equivalent on asymmetric metric spaces. Hence, we need additional arguments to establish the analogues of results in [Cho22b]. Roughly speaking, we need to replace d(x,y)d(x,y), the distance between two points xx and yy, with its symmetrization dsym(x,y)d^{sym}(x,y) in the arguments in [Cho22b] and [Cho22c]. For example, we now define that:

Definition 2.6 ([BF09, Definition 5.8]).

Bi-infinite paths κ=(xi)i\kappa=(x_{i})_{i\in\operatorname{\mathbb{Z}}}, η=(yi)i\eta=(y_{i})_{i\in\operatorname{\mathbb{Z}}} are said to be independent if the map (n,m)dsym(xn,ym)(n,m)\mapsto d^{sym}(x_{n},y_{m}) is proper, i.e., for any M>0M>0, {(n,m):dsym(xn,ym)<M}\{(n,m):d^{sym}(x_{n},y_{m})<M\} is bounded.

Isometries g,hg,h of XX are said to be independent if their orbits are independent.

Definition 2.7.

A subgroup of Isom(X)\operatorname{Isom}(X) is said to be non-elementary if it contains two independent BGIP isometries.

From now on, we always discuss under Convention 1.1: (X,d)(X,d) is a geodesic metric space, with possibly asymmetric metric, GG is a countable group of isometries of XX, and GG contains two independent BGIP isometries.

2.3. Random walks

We recall the notations in [Cho22b, Subsection 2.4].

Let μ\mu be a probability measure GG, which comes with its reflected version μˇ\check{\mu} defined by μˇ(g):=μ(g1)\check{\mu}(g):=\mu(g^{-1}). The random walk generated by μ\mu is the Markov chain on GG with the transition probability p(g,h):=μ(g1h)p(g,h):=\mu(g^{-1}h); this can be defined on a probability space on which the step elements {gn}n\{g_{n}\}_{n\in\operatorname{\mathbb{Z}}} are measurable, where gig_{i}’s are i.i.d.s distributed according to μ\mu. Such a probability space is called the probability space for μ\mu. Equivalently, (Ω,)(\Omega,\operatorname{\mathbb{P}}) is a probability space for μ\mu if there is a measure preserving map from (Ω,)(\Omega,\operatorname{\mathbb{P}}) to (G,μ)(G^{\operatorname{\mathbb{Z}}},\mu^{\operatorname{\mathbb{Z}}}).

Given a step path (gn)n(G,μ)(g_{n})_{n\in\operatorname{\mathbb{Z}}}\in(G^{\operatorname{\mathbb{Z}}},\mu^{\operatorname{\mathbb{Z}}}), we define the (bi-infinite) sample path (Zn)n(Z_{n})_{n\in\operatorname{\mathbb{Z}}} by

Zn={g1gnn>0idn=0g01gn+11n<0.Z_{n}=\left\{\begin{array}[]{cc}g_{1}\cdots g_{n}&n>0\\ id&n=0\\ g_{0}^{-1}\cdots g_{n+1}^{-1}&n<0.\end{array}\right.

We also introduce the notation gˇn=gn+11\check{g}_{n}=g_{-n+1}^{-1} and Zˇn=Zn\check{Z}_{n}=Z_{-n}. Note that we have an isomorphism (G,μ)(G>0,μˇ>0)×(G>0,μ>0)(G^{\operatorname{\mathbb{Z}}},\mu^{\operatorname{\mathbb{Z}}})\rightarrow(G^{\operatorname{\mathbb{Z}}_{>0}},\check{\mu}^{\operatorname{\mathbb{Z}}_{>0}})\times(G^{\operatorname{\mathbb{Z}}_{>0}},\mu^{\operatorname{\mathbb{Z}}_{>0}}) by (gn)n((gˇn)n>0,(gn)n>0)(g_{n})_{n\in\operatorname{\mathbb{Z}}}\mapsto((\check{g}_{n})_{n>0},(g_{n})_{n>0}). In view of this, we sometimes write the bi-infinite sample path as ((Zˇn)n0,(Zn)n0)((\check{Z}_{n})_{n\geq 0},(Z_{n})_{n\geq 0}).

When a constant M0M_{0} is understood, we will denote the M0M_{0}-long subpath of the sample path ending at ZioZ_{i}o by 𝐘i\mathbf{Y}_{i} In other words, we denote the sequence (ZiM0(ω)o,ZiM0+1(ω)o,,Zi1(ω)o,Zi(ω)o)\big{(}Z_{i-M_{0}}(\operatorname{\omega})o,Z_{i-M_{0}+1}(\operatorname{\omega})o,\ldots,Z_{i-1}(\operatorname{\omega})o,Z_{i}(\operatorname{\omega})o\big{)} by 𝐘i(ω)\mathbf{Y}_{i}(\operatorname{\omega}).

We define the support of μ\mu, denoted by suppμ\operatorname{supp}\mu, as the set of elements of GG that are assigned nonzero value by μ\mu. We denote by μN\mu^{N} the product measure of NN copies of μ\mu, and by μN\mu^{\ast N} the NN-th convolution of μ\mu.

A probability measure μ\mu on GG is said to be non-elementary if the semigroup suppμ\operatorname{\langle\langle}\operatorname{supp}\mu\operatorname{\rangle\rangle} generated by the support of μ\mu contains two independent BGIP isometries g,hg,h of XX. μ\mu is said to be admissible if suppμ\operatorname{\langle\langle}\operatorname{supp}\mu\operatorname{\rangle\rangle} equals the entire group GG. μ\mu is said to be non-arithmetic if there exist N>0N>0 and g,hsuppμNg,h\in\operatorname{supp}\mu^{\ast N} such that τ(g)τ(h)\tau(g)\neq\tau(h). Finally, we say that μ\mu is asymptotically asymmetric if there exists N>0N>0 and g,hsuppμNg,h\in\operatorname{supp}\mu^{\ast N} such that

τ(g)τ(g1)τ(h)τ(h1).\tau(g)-\tau(g^{-1})\neq\tau(h)-\tau(h^{-1}).

The random walk (Zn)n0(Z_{n})_{n\geq 0} generated by μ\mu is said to be admissible (non-elementary, non-arithmetic or asymptotically asymmetric, resp.) if μ\mu is admissible (non-elementary, non-arithmetic or asymptotically asymmetric, resp.).

For a given p>0p>0, we define the pp-th moment of μ\mu by

𝔼μ[d(o,go)p]:=gGd(o,go)pμ(g).\operatorname{\mathbb{E}}_{\mu}[d(o,go)^{p}]:=\sum_{g\in G}d(o,go)^{p}\,\mu(g).

Note that 𝔼μ[d(o,go)p]\operatorname{\mathbb{E}}_{\mu}[d(o,go)^{p}] and

𝔼μˇ[d(o,go)p]:=gGd(o,go)pμˇ(g)=gGd(go,o)pμ(g)\operatorname{\mathbb{E}}_{\check{\mu}}[d(o,go)^{p}]:=\sum_{g\in G}d(o,go)^{p}\,\check{\mu}(g)=\sum_{g\in G}d(go,o)^{p}\,\mu(g)

are distinct in general, and the finitude of the former does not imply that of the latter. This technicality leads to a subtle difference between limit laws for symmetric and asymmetric metric spaces. However, many asymmetric metric spaces (including Outer space) satisfy the following coarse symmetry: there exists a global constant K>0K>0 such that d(x,y)Kd(x,y)d(x,y)\leq Kd(x,y) for x,yGox,y\in Go. Under such a coarse symmetry, a measure μ\mu has finite pp-th moment if and only if its reflected version μˇ():=μ(1)\check{\mu}(\cdot):=\mu(\cdot^{-1}) does so. Hence, this subtlety will not matter for Outer space and many other spaces.

2.4. Properties of BGIP axes

In [Cho22b], we summarized some useful properties of contracting axes that were established earlier by many authors in [ACT15], [Sis18], [Yan19]. Here, we record the analogous properties for BGIP axes.

Lemma 2.8 (cf. [Cho22b, Lemma 3.1]).

Let K>1K>1, let γ\gamma be a KK-BGIP axis and let η:IX\eta:I\rightarrow X be a geodesic such that diam(πγ(η))>K\operatorname{\operatorname{diam}}(\pi_{\gamma}(\eta))>K. Then there exist t<tt<t^{\prime} in II such that πγ(η)\pi_{\gamma}(\eta) and η|[t,t]\eta|_{[t,t^{\prime}]} are 20K320K^{3}-coarsely equivalent, and moreover,

diam(πγ(η|(,t])η(t))<7K3,diam(πγ(η|[t,+))η(t))<7K3.\operatorname{\operatorname{diam}}\big{(}\pi_{\gamma}(\eta|_{(-\infty,t]})\cup\eta(t)\big{)}<7K^{3},\quad\operatorname{\operatorname{diam}}\big{(}\pi_{\gamma}(\eta|_{[t^{\prime},+\infty)})\cup\eta(t^{\prime})\big{)}<7K^{3}.

This was proved for symmetric metric spaces in [ACT15, Lemma 2.14], [Sis18, Lemma 2.4] and [Yan19, Lemma 2.4(4)]. We give a proof for asymmetric metric spaces, which is an adaptation of [CCT23, Lemma 2.2], for completeness.

Proof.

Let N=NK(γ)N=N_{K}(\gamma) be the KK-dsymd^{sym}-neighborhood of γ\gamma, N¯\bar{N} be its closure and S=ηN¯S=\eta\cap\bar{N}. Then SS is closed. Moreover, since diam(πγ(η))>K\operatorname{\operatorname{diam}}(\pi_{\gamma}(\eta))>K, SS is nonempty by the KK-BGIP of γ\gamma. We now claim:

Observation 2.9.

Let xSx\in S, let yγy\in\gamma and suppose that dsym(x,y)Kd^{sym}(x,y)\leq K. Then d(y,πγ(x))Kd(y,\pi_{\gamma}(x))\leq K and d(πγ(x),y)2K3+Kd(\pi_{\gamma}(x),y)\leq 2K^{3}+K.

Proof of Observation 2.9.

We have

d(y,πγ(x))d(y,x)+d(x,πγ(x))d(y,x)+d(x,y)K.d\big{(}y,\pi_{\gamma}(x)\big{)}\leq d\big{(}y,x\big{)}+d\big{(}x,\pi_{\gamma}(x)\big{)}\leq d\big{(}y,x\big{)}+d\big{(}x,y\big{)}\leq K.

Since yy and πγ(x)\pi_{\gamma}(x) both lie in a KK-bi-quasigeodesic γ\gamma, Inequality 2.3 implies d(πγ(x),y)2K3+Kd\big{(}\pi_{\gamma}(x),y\big{)}\leq 2K^{3}+K. ∎

Let tt and tt^{\prime} be the infimum and the supremum of η1(S)\eta^{-1}(S). Then η|(,t]\eta|_{(-\infty,t]} is a geodesic disjoint from NN so diam(πγ(η|(,t]))<K\operatorname{\operatorname{diam}}\big{(}\pi_{\gamma}(\eta|_{(-\infty,t]})\big{)}<K holds. Furthermore, η(t)\eta(t) belongs to ηNS\eta\cap\partial N\subseteq S, so there exists yγy\in\gamma such that dsym(η(t),y)=Kd^{sym}(\eta(t),y)=K. Then Observation 2.9 implies

d(η(t),πγ(η(t)))\displaystyle d\big{(}\eta(t),\pi_{\gamma}(\eta(t))\big{)} d(η(t),y)+d(y,πγ(η(t)))2K,\displaystyle\leq d(\eta(t),y)+d\big{(}y,\pi_{\gamma}(\eta(t))\big{)}\leq 2K,
d(πγ(η(t)),η(t))\displaystyle d\big{(}\pi_{\gamma}(\eta(t)),\eta(t)\big{)} d(πγ(η(t)),y)+d(y,η(t))2K3+2K.\displaystyle\leq d\big{(}\pi_{\gamma}(\eta(t)),y\big{)}+d(y,\eta(t))\leq 2K^{3}+2K.

We deduce that

diam(η(t)πγ(η|(,t]))\displaystyle\operatorname{\operatorname{diam}}\big{(}\eta(t)\cup\pi_{\gamma}(\eta|_{(-\infty,t]})\big{)} dsym(η(t),πγ(η(t)))+diam(πγ(η|(,t]))\displaystyle\leq d^{sym}\big{(}\eta(t),\pi_{\gamma}(\eta(t))\big{)}+\operatorname{\operatorname{diam}}\big{(}\pi_{\gamma}(\eta|_{(-\infty,t]})\big{)}
<2K3+4K+K7K3.\displaystyle<2K^{3}+4K+K\leq 7K^{3}.

For a similar reason, we have diam(η(t)πγ(η|[t,+)))<7K3\operatorname{\operatorname{diam}}\big{(}\eta(t^{\prime})\cup\pi_{\gamma}(\eta|_{[t^{\prime},+\infty)})\big{)}<7K^{3}.

We next claim that dsym(η(s),πγ(η))20K3d^{sym}(\eta(s),\pi_{\gamma}(\eta))\leq 20K^{3} for each s[t,t]s\in[t,t^{\prime}]. If η(s)S\eta(s)\in S, then there exists yγy\in\gamma such that dsym(η(s),y)=Kd^{sym}(\eta(s),y)=K. Then Observation 2.9 implies that

(2.4) dsym(η(s),πγ(η(s)))\displaystyle d^{sym}\big{(}\eta(s),\pi_{\gamma}(\eta(s))\big{)} dsym(η(s),y)+dsym(y,πγ(η(s)))\displaystyle\leq d^{sym}(\eta(s),y)+d^{sym}\big{(}y,\pi_{\gamma}(\eta(s))\big{)}
K+(K+2K3+K)7K3.\displaystyle\leq K+(K+2K^{3}+K)\leq 7K^{3}.

If η(s)S\eta(s)\notin S, then we take the connected component (s1,s2)[t,t]η1(S)(s_{1},s_{2})\in[t,t^{\prime}]\setminus\eta^{-1}(S) of ss. Then the geodesic η|[s1,s2]\eta|_{[s_{1},s_{2}]} is disjoint from the KK-neighborhood of γ\gamma so the diameter of πγ(η|[s1,s2])\pi_{\gamma}(\eta|_{[s_{1},s_{2}]}) is at most KK. Moreover, η(s1)\eta(s_{1}) and η(s2)\eta(s_{2}) belong to ηN\eta\cap\partial N: let p,qγp,q\in\gamma be points such that dsym(η(s1),p),dsym(η(s2),q)=Kd^{sym}(\eta(s_{1}),p),d^{sym}(\eta(s_{2}),q)=K. By Observation 2.9, we have

d(πγ(η(s1)),η(s))\displaystyle d\big{(}\pi_{\gamma}(\eta(s_{1})),\eta(s)\big{)} d(πγ(η(s1)),η(s1))+d(η(s1),η(s))\displaystyle\leq d\big{(}\pi_{\gamma}(\eta(s_{1})),\eta(s_{1})\big{)}+d\big{(}\eta(s_{1}),\eta(s)\big{)}
d(πγ(η(s1)),η(s1))+d(η(s1),η(s2))\displaystyle\leq d\big{(}\pi_{\gamma}(\eta(s_{1})),\eta(s_{1})\big{)}+d\big{(}\eta(s_{1}),\eta(s_{2})\big{)}
d(πγ(η(s1)),η(s1))+d(η(s1),πγ(η(s1)))\displaystyle\leq d\big{(}\pi_{\gamma}(\eta(s_{1})),\eta(s_{1})\big{)}+d\big{(}\eta(s_{1}),\pi_{\gamma}(\eta(s_{1}))\big{)}
+diam(πγ(η|[s1,s2]))+d(πγ(η(s2)),η(s2))\displaystyle\quad+\operatorname{\operatorname{diam}}\big{(}\pi_{\gamma}(\eta|_{[s_{1},s_{2}]})\big{)}+d\big{(}\pi_{\gamma}(\eta(s_{2})),\eta(s_{2})\big{)}
d(πγ(η(s1)),p)+d(p,η(s1))+d(η(s1),γ)\displaystyle\leq d\big{(}\pi_{\gamma}(\eta(s_{1})),p\big{)}+d(p,\eta(s_{1}))+d(\eta(s_{1}),\gamma)
+diam(πγ(η|[s1,s2]))+d(πγ(η(s2)),q)+d(q,η(s2))\displaystyle\quad+\operatorname{\operatorname{diam}}\big{(}\pi_{\gamma}(\eta|_{[s_{1},s_{2}]})\big{)}+d\big{(}\pi_{\gamma}(\eta(s_{2})),q\big{)}+d\big{(}q,\eta(s_{2})\big{)}
(2K3+K)+K+K+(2K3+K)+K.\displaystyle\leq(2K^{3}+K)+K+K+(2K^{3}+K)+K.

Also, we have

d(η(s),πγ(η(s1)))\displaystyle d\big{(}\eta(s),\pi_{\gamma}(\eta(s_{1}))\big{)} d(η(s),η(s2))+d(η(s2),πγ(η(s2)))+d(πγ(η(s2)),πγ(η(s1)))\displaystyle\leq d(\eta(s),\eta(s_{2}))+d\big{(}\eta(s_{2}),\pi_{\gamma}(\eta(s_{2}))\big{)}+d\big{(}\pi_{\gamma}(\eta(s_{2})),\pi_{\gamma}(\eta(s_{1}))\big{)}
d(η(s1),η(s2))+d(η(s2),γ)+diam(πγ(η|[s1,s2]))\displaystyle\leq d(\eta(s_{1}),\eta(s_{2}))+d(\eta(s_{2}),\gamma\big{)}+\operatorname{\operatorname{diam}}\big{(}\pi_{\gamma}(\eta|_{[s_{1},s_{2}]})\big{)}
(d(η(s1),γ)+diam(πγ(η|[s1,s2]))+d(πγ(η(s2)),q)+d(q,η(s2)))\displaystyle\leq\Big{(}d\big{(}\eta(s_{1}),\gamma\big{)}+\operatorname{\operatorname{diam}}\big{(}\pi_{\gamma}(\eta|_{[s_{1},s_{2}]})\big{)}+d\big{(}\pi_{\gamma}(\eta(s_{2})),q\big{)}+d\big{(}q,\eta(s_{2})\big{)}\Big{)}
+d(η(s2),γ)+diam(πγ(η|[s1,s2]))\displaystyle+d\big{(}\eta(s_{2}),\gamma\big{)}+\operatorname{\operatorname{diam}}\big{(}\pi_{\gamma}(\eta|_{[s_{1},s_{2}]})\big{)}
K+K+(2K3+K)+K+K.\displaystyle\leq K+K+(2K^{3}+K)+K+K.

In summary, we have dsym(p,η(s))20K3d^{sym}(p,\eta(s))\leq 20K^{3} as desired.

Next, we claim that dsym(πγ(η(s)),η|[t,t])<20K3d^{sym}\big{(}\pi_{\gamma}(\eta(s)),\eta|_{[t,t^{\prime}]}\big{)}<20K^{3} for each sIs\in I. If η(s)S\eta(s)\in S, then dsym(πγ(η(s)),η(s))7K3d^{sym}\big{(}\pi_{\gamma}(\eta(s)),\eta(s)\big{)}\leq 7K^{3} by Inequality 2.4.

If η(s)S\eta(s)\notin S, then we take the connected component (s1,s2)Iη1(S)(s_{1},s_{2})\in I\setminus\eta^{-1}(S) of ss, with either η(s1)\eta(s_{1}) or η(s2)\eta(s_{2}) being in N\partial N. Without loss of generality, suppose that η(s2)N\eta(s_{2})\in\partial N and let yγy\in\gamma be such that dsym(η(s2),y)=Kd^{sym}(\eta(s_{2}),y)=K. Observation 2.9 and the KK-BGIP of γ\gamma implies that

dsym(πγ(η(s)),η(s2))\displaystyle d^{sym}\big{(}\pi_{\gamma}(\eta(s)),\eta(s_{2})\big{)} 2diam(πγ(η|[s,s2]))+dsym(πγ(η(s2)),y)+dsym(y,η(s2))\displaystyle\leq 2\operatorname{\operatorname{diam}}\big{(}\pi_{\gamma}(\eta|_{[s,s_{2}]})\big{)}+d^{sym}\big{(}\pi_{\gamma}(\eta(s_{2})),y)+d^{sym}(y,\eta(s_{2})\big{)}
2K+(2K3+2K)+K10K3.\displaystyle\leq 2K+(2K^{3}+2K)+K\leq 10K^{3}.\qed
Corollary 2.10.

Let K>1K>1 and let γ\gamma be a KK-BGIP axis. Then the closest point projection πγ():Xγ\pi_{\gamma}(\cdot):X\rightarrow\gamma is (1,14K3)(1,14K^{3})-coarsely Lipschitz, i.e.,

d(πγ(x),πγ(y))d(x,y)+14K3(x,yX).d(\pi_{\gamma}(x),\pi_{\gamma}(y)\big{)}\leq d(x,y)+14K^{3}\quad(\forall\,x,y\in X).
Proof.

Let η:IX\eta:I\rightarrow X represent the geodesic from xx to yy. If πγ(η)\pi_{\gamma}(\eta) has diameter smaller than KK, the conclusion follows. If not, then by Lemma 2.8, there exist t<tt<t^{\prime} in II such that d(πγ(x),η(t))<7K3d(\pi_{\gamma}(x),\eta(t))<7K^{3} and d(η(t),πγ(y))<7K3d(\eta(t^{\prime}),\pi_{\gamma}(y))<7K^{3}. We then have

d(πγ(x),πγ(y))\displaystyle d(\pi_{\gamma}(x),\pi_{\gamma}(y)) d(πγ(x),η(t))+d(η(t),η(t))+d(η(t),πγ(y))\displaystyle\leq d(\pi_{\gamma}(x),\eta(t))+d(\eta(t),\eta(t^{\prime}))+d(\eta(t^{\prime}),\pi_{\gamma}(y))
<7K3+diam(I)+7K3d(x,y)+14K3.\displaystyle<7K^{3}+\operatorname{\operatorname{diam}}(I)+7K^{3}\leq d(x,y)+14K^{3}.\qed

The following lemma was proved in [Cho22b, Section 3.1] for symmetric metrics; the same proof works here as well, after replacing quasi-geodesics with bi-quasigeodesics.

Lemma 2.11 ([Cho22b, Lemma 3.2]).

For each K>1K>1 there exists a constant K>KK^{\prime}>K such that the following holds.

Let η:JX\eta:J\rightarrow X be a KK-bi-quasigeodesic whose endpoints are xx and yy, let AηA\subseteq\eta be a subset of η\eta such that d(x,A)<Kd(x,A)<K, d(y,A)<Kd(y,A)<K, and let γ:JX\gamma:J^{\prime}\rightarrow X be a geodesic that is KK-coarsely equivalent to AA. Then η\eta and γ\gamma are also KK^{\prime}-coarsely equivalent, and moreover, there exists a KK^{\prime}-quasi-isometry φ:JJ\varphi:J\rightarrow J^{\prime} such that dsym(η(t),(γφ)(t))<Kd^{sym}(\eta(t),(\gamma\circ\varphi)(t))<K^{\prime} for each tJt\in J.

Corollary 2.12 ([Cho22b, Corollary 3.3]).

For each K>1K>1 there exists a constant K>KK^{\prime}>K such that the following holds.

Let η:JX\eta:J\rightarrow X be a KK-BGIP axis and let γ:JX\gamma:J^{\prime}\rightarrow X be a geodesic connecting the endpoints of η\eta. Then there exists a KK^{\prime}-quasi-isometry φ:JJ\varphi:J\rightarrow J^{\prime} such that dsym(η(t),(γφ)(t))<Kd^{sym}(\eta(t),(\gamma\circ\varphi)(t))<K^{\prime} for each tJt\in J; in particular, η\eta and γ\gamma are KK^{\prime}-coarsely equivalent.

By using Corollary 2.12 and Corollary 2.10, we deduced the following in [Cho22b, Section 3.1]; the proof there is phrased in terms of symmetric metrics, but the same proof works for asymmetric metrics as well.

Corollary 2.13 ([Cho22b, Corollary 3.4]).

For each K>1K>1 there exists a constant K=K(K)K^{\prime}=K^{\prime}(K) that satisfies the following.

Let κ:IX\kappa:I\rightarrow X and η:JX\eta:J\rightarrow X be KK-BGIP axes. Suppose that diam(πκ(η))>K\operatorname{\operatorname{diam}}(\pi_{\kappa}(\eta))>K^{\prime}. Then there exist t<tt<t^{\prime} in II and s<ss<s^{\prime} in JJ such that the following sets are all KK^{\prime}-coarsely equivalent:

κ|[t,t],η|[s,s],πκ(η),πη(κ).\kappa|_{[t,t^{\prime}]},\,\,\eta|_{[s,s^{\prime}]},\,\,\pi_{\kappa}(\eta),\,\,\pi_{\eta}(\kappa).

Moreover, we have

diam(πκ(η|(,s])η(s))<K,diam(πκ(η|[s,+))η(s))<K.\operatorname{\operatorname{diam}}\big{(}\pi_{\kappa}(\eta|_{(-\infty,s]})\cup\eta(s)\big{)}<K^{\prime},\quad\operatorname{\operatorname{diam}}\big{(}\pi_{\kappa}(\eta|_{[s^{\prime},+\infty)})\cup\eta(s^{\prime})\big{)}<K^{\prime}.

The following fact is proven in [Yan20, Proposition 2.2.(3)]. For completeness, we sketch the proof in the setting of asymmetric metric spaces.

Lemma 2.14.

For each K>1K>1, there exists K=K(K)>1K^{\prime}=K^{\prime}(K)>1 such that every subpath of a KK-BGIP axis is again a KK^{\prime}-BGIP axis.

Proof.

Let γ:IX\gamma:I\rightarrow X be a KK-BGIP axis and let κ=γ|[a,b]:[a,b]X\kappa=\gamma|_{[a,b]}:[a,b]\rightarrow X be the restriction for some a<ba<b in II. Our goal is to find a constant KK^{\prime}, depending only on KK, such that diam(πκ(η))>K\operatorname{\operatorname{diam}}(\pi_{\kappa}(\eta))>K^{\prime} implies ηNK(κ)\eta\cap N_{K^{\prime}}(\kappa)\neq\emptyset for each geodesic η=[x,y]\eta=[x,y]. Without loss of generality, we can assume that κ\kappa is longer than a large enough constant.

Before the proof, we make a simple observation: for xXx\in X, if πγ(x)\pi_{\gamma}(x) intersects κ\kappa, then πκ(x)\pi_{\kappa}(x) equals πγ(x)κ\pi_{\gamma}(x)\cap\kappa, which is nonempty.

Let x,yXx,y\in X. First consider the case that πγ(x)\pi_{\gamma}(x) is far from γκ\gamma\setminus\kappa and deep inside κ\kappa. In this case, Lemma 2.8 tells us that [x,y][x,y] passes through a bounded neighborhood of πγ(x)\pi_{\gamma}(x) unless πγ(y)\pi_{\gamma}(y) is KK-close to πγ(x)\pi_{\gamma}(x), in which case πγ(x)\pi_{\gamma}(x) and πγ(y)\pi_{\gamma}(y) are both contained in κ\kappa and diam(πκ(x)πκ(y))=diam(πγ(x)πγ(y))K\operatorname{\operatorname{diam}}(\pi_{\kappa}(x)\cup\pi_{\kappa}(y))=\operatorname{\operatorname{diam}}(\pi_{\gamma}(x)\cup\pi_{\gamma}(y))\leq K. Similarly, when πγ(y)\pi_{\gamma}(y) lies deep in κ\kappa, then diam(πκ(x)πκ(y))>K\operatorname{\operatorname{diam}}(\pi_{\kappa}(x)\cup\pi_{\kappa}(y))>K implies that [x,y][x,y] passes nearby κ\kappa.

Now consider the case that πγ(x)\pi_{\gamma}(x) and πγ(y)\pi_{\gamma}(y) are both near γκ\gamma\setminus\kappa. Let γL\gamma_{L} and γR\gamma_{R} be the left and the right components of γκ\gamma\setminus\kappa, respectively. If πγ(x)\pi_{\gamma}(x) (πγ(y)(\pi_{\gamma}(y), resp.) intersects with a bounded neighborhood of γL\gamma_{L} (γR\gamma_{R}, resp.), then they are far away (because κ\kappa is assumed to be long enough). Lemma 2.8 says that some subgeodesic η\eta^{\prime} of [x,y][x,y] is coarsely equivalent to a subset of γ\gamma that coarsely connects γL\gamma_{L} and γR\gamma_{R}. Lemma 2.11 then tells us that η\eta^{\prime} coarsely contains the entire subsegement of γ\gamma in between γL\gamma_{L} and γR\gamma_{R}, which is κ\kappa. Hence, [x,y][x,y] passes nearby κ\kappa.

By symmetry, it remains to handle the case that πγ(x)\pi_{\gamma}(x) and πγ(y)\pi_{\gamma}(y) are both contained in a bounded neighborhood γL\gamma_{L}. In this case, we claim that πκ(x)\pi_{\kappa}(x) is close to γ(a)\gamma(a). To see this, pick any s[a,b]s\in[a,b] with large sas-a. Also, pick t(,a]t\in(-\infty,a] such that γ(t)πγ(x)\gamma(t)\in\pi_{\gamma}(x). We now apply Lemma 2.8 to the geodesic [x,γ(s)][x,\gamma(s)] and obtain its subsegment η\eta^{\prime} which is coarsely equivalent to a subset of γ\gamma containing γ(t)\gamma(t) and γ(s)\gamma(s). Lemma 2.11 then tells us that η\eta^{\prime} is close to each γ(u)\gamma(u) for tust\leq u\leq s, including u=au=a. Hence, [x,γ(s)][x,\gamma(s)] passes nearby γ(a)\gamma(a), with γ(s)\gamma(s) and γ(a)\gamma(a) being far away from each other (since sas-a is large). This implies that γ(a)\gamma(a) is closer than γ(s)\gamma(s) to xx, and γ(s)πκ(x)\gamma(s)\notin\pi_{\kappa}(x). This implies that πκ(x)\pi_{\kappa}(x) consists of points close to γ(a)\gamma(a) as desired.

For the same reason, πκ(y)\pi_{\kappa}(y) is also close to γ(a)\gamma(a) and πκ(x)πκ(y)\pi_{\kappa}(x)\cup\pi_{\kappa}(y) has small diameter. This ends the proof. ∎

2.5. BGIP axes and alignment

We are now ready to discuss the alignment of BGIP axes.

Definition 2.15 ([Cho22b, Definition 3.5]).

Given paths κi\kappa_{i} from xix_{i} to xix^{\prime}_{i} for each i=1,,ni=1,\ldots,n, we say that (κ1,,κn)(\kappa_{1},\ldots,\kappa_{n}) is CC-aligned if

diam(xiπκi(κi+1))<C,diam(xi+1πκi+1(κi))<C.\operatorname{\operatorname{diam}}\left(x_{i}^{\prime}\cup\pi_{\kappa_{i}}(\kappa_{i+1})\right)<C,\quad\operatorname{\operatorname{diam}}\left(x_{i+1}\cup\pi_{\kappa_{i+1}}(\kappa_{i})\right)<C.

hold for i=1,,n1i=1,\ldots,n-1.

Here, we regard points as degenerate paths. For example, for a point xx and a path γ\gamma, we say that (x,γ)(x,\gamma) is CC-aligned if

diam(beginning point of γπγ(x))<C.\operatorname{\operatorname{diam}}\big{(}\textrm{beginning point of $\gamma$}\cup\pi_{\gamma}(x)\big{)}<C.
Observation 2.16 ([Cho22c, Observation 2.4]).

Let gg be an isometry of XX, let 1kn1\leq k\leq n, let K>0K>0 and let γ1,,γn\gamma_{1},\ldots,\gamma_{n} be paths on XX.

  1. (1)

    If (γ1,,γk)(\gamma_{1},\ldots,\gamma_{k}) and (γk,,γn)(\gamma_{k},\ldots,\gamma_{n}) are KK-aligned, then there concatenation (γ1,,γn)(\gamma_{1},\ldots,\gamma_{n}) is also KK-aligned.

  2. (2)

    If (γ1,,γn)(\gamma_{1},\ldots,\gamma_{n}) is KK-aligned, then (γ¯n,,γ¯1)(\bar{\gamma}_{n},\ldots,\bar{\gamma}_{1}) is also KK-aligned.

  3. (3)

    If (γ1,,γn)(\gamma_{1},\ldots,\gamma_{n}) is KK-aligned, then (gγ1,,gγn)(g\gamma_{1},\ldots,g\gamma_{n}) is also KK-aligned.

By combining Lemma 2.8 and Lemma 2.11, we deduce:

Corollary 2.17 ([Cho22b, Corollary 3.6]).

For each C,K>1C,K>1, there exists K=K(K,C)>K,CK^{\prime}=K^{\prime}(K,C)>K,C that satisfies the following.

Let x,yXx,y\in X and let κ\kappa be a KK-BGIP axis such that diam(κ)>K\operatorname{\operatorname{diam}}(\kappa)>K^{\prime} and such that (x,κ,y)(x,\kappa,y) is CC-aligned. Then [x,y][x,y] contains a subsegment η\eta that is contained in the 20K320K^{3}-dsymd^{sym}-neighborhood of κ\kappa and is KK^{\prime}-fellow traveling with κ\kappa.

The following was proved in [Cho22b] for symmetric metrics. Using Lemma 2.8 and Corollary 2.10 this time, the same proof works for asymmetric metrics as well.

Lemma 2.18 ([Cho22b, Lemma 3.7]).

For each C>0C>0 and K>1K>1, there exists D=D(K,C)>CD=D(K,C)>C that satisfies the following property.

Let κ,η\kappa,\eta be KK-BGIP axes whose beginning points are xx and xx^{\prime}, respectively. Suppose that (κ,x)(\kappa,x^{\prime}) and (x,η)(x,\eta) are CC-aligned. Then (κ,η)(\kappa,\eta) is DD-aligned.

In [Sis18], Sisto proved the following property for constricting geodesics with respect to special paths. The same proof works for BGIP axes; we give a rephrasing of Sisto’s proof for convenience.

Lemma 2.19 ([Sis18, Lemma 2.5]).

For each D>0D>0 and K>1K>1, there exists E=E(K,D)>K,DE=E(K,D)>K,D that satisfies the following.

Let κ\kappa and η\eta be KK-BGIP axes in XX. Suppose that (κ,η)(\kappa,\eta) is DD-aligned. Then for any pXp\in X, either (p,η)(p,\eta) is EE-aligned or (κ,p)(\kappa,p) is EE-aligned.

Proof.

Let I=[a,b]I=[a,b] and J=[c,d]J=[c,d] be the domains of κ\kappa and η\eta, respectively. The assumption says that πκ(η)\pi_{\kappa}(\eta) is close to κ(b)\kappa(b) and πη(κ)\pi_{\eta}(\kappa) is close to η(c)\eta(c). Now, consider the geodesic γ=[p,κ(b)]\gamma=[p,\kappa(b)] from pp to κ(b)\kappa(b), and pick the earliest point qγq\in\gamma that belongs to the KK-closed neighborhood of κη\kappa\cup\eta. (Such point exists because κ(b)NK(κη)\kappa(b)\in N_{K}(\kappa\cup\eta).)

  1. (1)

    If q𝒩K(κ)¯q\in\overline{\mathcal{N}_{K}(\kappa)}, then dsym(q,κ)=Kd^{sym}(q,\kappa)=K and [x,q][x,q] does not intersect the KK-neighborhood of η\eta. Then πη(x)\pi_{\eta}(x) is contained in the diam(πη([x,q]))\operatorname{\operatorname{diam}}(\pi_{\eta}([x,q]))-neighborhood of πη(q)\pi_{\eta}(q), and πη([x,q])\pi_{\eta}([x,q]) has diameter at most KK by the KK-BGIP of κ\kappa. Now, πη(q)\pi_{\eta}(q) is (d(q,κ)+14K3)\big{(}d(q,\kappa)+14K^{3}\big{)}-close to πη(κ)\pi_{\eta}(\kappa) by Corollary 2.10, which is close to η(c)\eta(c). It follows that πη(x)\pi_{\eta}(x) is contained in a uniform neighborhood of η(c)\eta(c), and (x,η)(x,\eta) is aligned.

  2. (2)

    If q𝒩K(η)¯q\in\overline{\mathcal{N}_{K}(\eta)}, then dsym(q,η)=Kd^{sym}(q,\eta)=K and [x,q][x,q] is does not intersect the KK-neighborhood of κ\kappa. By a symmetric argument, πκ(x)\pi_{\kappa}(x) is uniformly close to κ(b)\kappa(b) and (κ,x)(\kappa,x) is aligned.∎

Using Corollary 2.17 and Lemma 2.19 as ingredients, we proved the following in [Cho22b]:

Proposition 2.20 ([Cho22b, Proposition 3.11]).

For each D>0D>0 and K>1K>1, there exist E=E(K,D)>K,DE=E(K,D)>K,D and L=L(K,D)>K,DL=L(K,D)>K,D that satisfy the following.

Let xx and yy be points in XX and let κ1,,κn\kappa_{1},\ldots,\kappa_{n} be KK-BGIP axes whose domains are longer than LL and such that (x,κ1,,κn,y)(x,\kappa_{1},\ldots,\kappa_{n},y) is DD-aligned. Then the geodesic [x,y][x,y] has subsegments η1,,ηn\eta_{1},\ldots,\eta_{n}, in order from left to right, that are longer than 100E100E and such that ηi\eta_{i} and κi\kappa_{i} are 0.1E0.1E-fellow traveling for each ii. In particular, (x,κi,y)(x,\kappa_{i},y) are EE-aligned for each ii.

Often, an isometry gIsom(X)g\in\operatorname{Isom}(X) gives rise to a periodic aligned sequence of BGIP axes (,gi1γn,giγ1,,giγn,gi+1γ1,)(\ldots,g^{i-1}\gamma_{n},g^{i}\gamma_{1},\ldots,g^{i}\gamma_{n},g^{i+1}\gamma_{1},\ldots). The following lemma enables to conclude that gg is an BGIP isometry in this situation.

Lemma 2.21 ([Yan19, Proposition 2.9]).

For each D,M>0D,M>0 and K>1K>1, there exist E=E(K,D,M)>DE=E(K,D,M)>D and L=L(K,D)>DL=L(K,D)>D that satisfies the following.

Let κ1,,κn\kappa_{1},\ldots,\kappa_{n} be KK-BGIP axes whose domains are longer than LL. Suppose that (κ1,,κn)(\kappa_{1},\ldots,\kappa_{n}) is DD-aligned and d(κi,κi+1)<Md(\kappa_{i},\kappa_{i+1})<M for each ii. Then the concatenation κ1κn\kappa_{1}\cup\ldots\cup\kappa_{n} of κ1,,κn\kappa_{1},\ldots,\kappa_{n} is an EE-BGIP axis.

Lemma 2.22 ([Yan19, Proposition 2.9]).

For each D,M>0D,M>0 and K>1K>1, there exist E=E(K,D,M)>DE=E(K,D,M)>D and L=L(K,D)>DL=L(K,D)>D that satisfies the following.

Let κ1,,κn\kappa_{1},\ldots,\kappa_{n} be KK-BGIP axes whose domains are longer than LL, and suppose that κi\kappa_{i} is connecting yi1y_{i-1} to xix_{i}. Suppose that (x0,κ1,,κn)(x_{0},\kappa_{1},\ldots,\kappa_{n}) is DD-aligned. Then the concatenation [x0,y0]κ1[x1,y1]κ2[x_{0},y_{0}]\cup\kappa_{1}\cup[x_{1},y_{1}]\cup\kappa_{2}\cup\ldots is an EE-quasigeodesic.

Here, note that the resulting path in Lemma 2.22 may not be bi-quasigeodesic.

We now recall the concept of Schottky sets. Given a sequence α=(a1,,an)Gn\alpha=(a_{1},\ldots,a_{n})\in G^{n}, we employ the following notations:

Π(α)\displaystyle\Pi(\alpha) :=a1a2an,\displaystyle:=a_{1}a_{2}\cdots a_{n},
Γ(α)\displaystyle\Gamma(\alpha) :=(o,a1o,a1a2o,,Π(α)o).\displaystyle:=\big{(}o,a_{1}o,a_{1}a_{2}o,\ldots,\Pi(\alpha)o\big{)}.
Definition 2.23 ([Cho22b, Definition 3.14]).

Let K>0K>0 and SGMS\subseteq G^{M} be a set of sequences of MM isometries. We say that SS is KK-Schottky if the following hold:

  1. (1)

    Γ(α)\Gamma(\alpha) is a KK-BGIP axis for all αS\alpha\in S;

  2. (2)

    for each xXx\in X, we have

    #{αS:(x,Γ(α)) and (Γ(α),Π(α)x) are K-aligned}#S1;\#\Big{\{}\alpha\in S:\textrm{$\left(x,\Gamma(\alpha)\right)$ and $\left(\Gamma(\alpha),\Pi(\alpha)x\right)$ are $K$-aligned}\Big{\}}\geq\#S-1;
  3. (3)

    for each αS\alpha\in S, (Γ(α),Π(α)Γ(α))\left(\Gamma(\alpha),\Pi(\alpha)\Gamma(\alpha)\right) is KK-aligned.

We say that SS is large enough if its cardinality is at least 400.

Definition 2.24 ([Cho22c, Definition 2.8]).

Given a constant K0>0K_{0}>0, we define:

  • D0=D(K0,K0)D_{0}=D(K_{0},K_{0}) be as in Lemma 2.18,

  • D1=E(K0,D0)D_{1}=E(K_{0},D_{0}), L=L(K0,D0)L=L(K_{0},D_{0}) be as in Proposition 2.20,

  • E0=E(K0,D1)E_{0}=E(K_{0},D_{1}), L=L(K0,D1)L^{\prime}=L(K_{0},D_{1}) be as in Proposition 2.20,

  • L=L(K0,D1)L^{\prime}=L(K_{0},D_{1}) be as in Lemma 2.21.

A K0K_{0}-Schottky set SGnS\subseteq G^{n} is called a fairly long K0K_{0}-Schottky set if:

  1. (1)

    n>max{L,L,L′′}n>\max\{L,L^{\prime},L^{\prime\prime}\}, and

  2. (2)

    d(o,Π(α)o)10E0d(o,\operatorname{\Pi}(\alpha)o)\geq 10E_{0} for all αS\alpha\in S.

When the Schottky parameter K0K_{0} is understood, the constants D0,D1,E0D_{0},D_{1},E_{0} always denote the ones defined above. Once a fairly long Schottky set SS is understood, its element ss is called a Schottky sequence and the translates of Γ±(s)\Gamma^{\pm}(s) are called Schottky axes. When a probability measure μ\mu on GG is given in addition such that S(suppμ)nS\subseteq(\operatorname{supp}\mu)^{n}, we say that SS is a fairly long Schottky set for μ\mu.

Definition 2.25 ([Cho22b, Definition 3.16]).

Let SS be a fairly long Schottky set and let K>0K>0. We say that a sequence of Schottky axes is KK-semi-aligned if it is a subsequence of a KK-aligned sequence of Schottky axes.

More precisely, for Schottky axes γ1,,γn\gamma_{1},\ldots,\gamma_{n}, we say that (γ1,,γn)(\gamma_{1},\ldots,\gamma_{n}) is KK-semi-aligned if there exist Schottky axes η1,,ηm\eta_{1},\ldots,\eta_{m} such that (η1,,ηm)(\eta_{1},\ldots,\eta_{m}) is KK-aligned and if there exists a subsequence {i(1)<<i(n)}{1,,m}\{i(1)<\ldots<i(n)\}\subseteq\{1,\ldots,m\} such that γl=ηi(l)\gamma_{l}=\eta_{i(l)} for l=1,,nl=1,\ldots,n.

Similarly, for points x,yXx,y\in X and Schottky axes γ1,,γn\gamma_{1},\ldots,\gamma_{n}, we say that (x,γ1,,γn,y)(x,\gamma_{1},\ldots,\gamma_{n},y) is KK-semi-aligned if it is a subsequence of a KK-aligned sequence (x,η1,,ηm,y)(x,\eta_{1},\ldots,\eta_{m},y) for some Schottky axes η1,,ηm\eta_{1},\ldots,\eta_{m}.

Proposition 2.20 implies the following corollary.

Corollary 2.26.

Let SS be a fairly long K0K_{0}-Schottky set, with constants D0,D1,E0D_{0},D_{1},E_{0} as in Definition 2.23. Let x,yXx,y\in X and γ1,,γN\gamma_{1},\ldots,\gamma_{N} be Schottky axes. Then (x,γ1,,γN,y)(x,\gamma_{1},\ldots,\gamma_{N},y) is D1D_{1}-aligned (E0(E_{0}, resp.) whenever it is D0D_{0}-semi-aligned (D1D_{1}-semi-aligned, resp.), and moreover, [x,y][x,y] contains subsegments η1,,ηN\eta_{1},\ldots,\eta_{N}, each longer than 100D1100D_{1} (100E0100E_{0}, resp.) and in order from closest to farthest from xx, such that ηi\eta_{i} and γi\gamma_{i} are 0.1D10.1D_{1}-fellow traveling (0.1E00.1E_{0}-fellow traveling, resp.) for each ii.

Using Lemma 2.14, Lemma 2.18 and Proposition 2.20, we proved the following in [Cho22b]:

Proposition 2.27 ([Cho22b, Proposition 3.12]).

Let (X,G,o)(X,G,o) be as in Convention 1.1 and let μ\mu be a non-elementary probability measure on GG. Then for each N,L>0N,L>0, there exists K=K(N)>0K=K(N)>0 and n>Ln>L such that there exists a KK-Schottky set of cardinality NN in (suppμ)n(\operatorname{supp}\mu)^{n}.

3. Limit laws

Recall again that we are fixing an asymmetric metric space XX and its countable isometry group GG involving two independent BGIP isometries. We will now describe the results established in Section 4, 5 and 6 of [Cho22b].

3.1. Results from [Cho22b]

In [Cho22b, Section 5], the author constructed pivotal times, first in a discrete model and next in random walks, following the idea of Gouëzel [Gou22, Section 4A]. The arguments in [Cho22b, Section 5] relied the following ingredients:

  1. (1)

    Existence of large and fairly long Schottky set for non-elementary probability measure μ\mu (Proposition 2.27),

  2. (2)

    Alignments of points and Schottky axes (the properties of Schottky set described in Definition 2.23), and

  3. (3)

    Alignment lemma that guarantees the alignment of two Schottky axes given the alignments of one Schottky axis and an endpoint of another Schottky axis (Lemma 2.18).

Since we have the versions of alignment and Schottky set for BGIP isometries, the entire proof works verbatim. As a consequence, we obtain the following proposition. Recall again the notation

𝐘i(ω):=(ZiM0(ω)o,ZiM0+1(ω)o,,Zi1(ω)o,Zi(ω)o).\mathbf{Y}_{i}(\operatorname{\omega}):=\big{(}Z_{i-M_{0}}(\operatorname{\omega})o,Z_{i-M_{0}+1}(\operatorname{\omega})o,\ldots,Z_{i-1}(\operatorname{\omega})o,Z_{i}(\operatorname{\omega})o\big{)}.
Definition 3.1 ([Cho22b, Definition 4.1]).

Let (X,G,o)(X,G,o) be as in Convention 1.1, let μ\mu be a non-elementary probability measure on GG, let (Ω,)(\Omega,\operatorname{\mathbb{P}}) be a probability space for μ\mu, let K0,M0>0K_{0},M_{0}>0 and let SS be a fairly long K0K_{0}-Schottky set contained in (suppμ)M0(\operatorname{supp}\mu)^{M_{0}}.

A subset \mathcal{E} of Ω\Omega is called a pivotal equivalence class if there exists a set 𝒫()={j(1)<j(2)<}M0>0\mathcal{P}(\mathcal{E})=\{j(1)<j(2)<\ldots\}\subseteq M_{0}\operatorname{\mathbb{Z}}_{>0}, called the set of pivotal times, such that the following hold:

  1. (1)

    for each i{j(k)l:k1,l=0,,M01}i\notin\{j(k)-l:k\geq 1,l=0,\ldots,M_{0}-1\}, gi(ω)g_{i}(\operatorname{\omega}) is fixed on \mathcal{E};

  2. (2)

    for each ω\operatorname{\omega}\in\mathcal{E} and k1k\geq 1, sk(ω):=(gj(k)M0+1(ω),gj(k)M0+2(ω),,gj(k)(ω))s_{k}(\operatorname{\omega}):=\big{(}g_{j(k)-M_{0}+1}(\operatorname{\omega}),g_{j(k)-M_{0}+2}(\operatorname{\omega}),\ldots,g_{j(k)}(\operatorname{\omega})\big{)} is a Schottky sequence;

  3. (3)

    for each ω\operatorname{\omega}\in\mathcal{E}, the sequence (o,𝐘j(1)(ω),𝐘j(2)(ω),)\big{(}o,\mathbf{Y}_{j(1)}(\operatorname{\omega}),\mathbf{Y}_{j(2)}(\operatorname{\omega}),\ldots\big{)} is D0D_{0}-semi-aligned;

  4. (4)

    on \mathcal{E}, {s1(ω),s2(ω),}\{s_{1}(\operatorname{\omega}),s_{2}(\operatorname{\omega}),\ldots\} are i.i.d.s distributed according to the uniform measure on SS.

For ω\operatorname{\omega}\in\mathcal{E}, we call 𝒫()\mathcal{P}(\mathcal{E}) the set of pivotal times for ω\operatorname{\omega} and write it as 𝒫(ω)\mathcal{P}(\operatorname{\omega}).

We say that a pivotal equivalence class \mathcal{E} avoids an integer kk if kk is not in {jl:j𝒫(),l=0,,M01}\{j-l:j\in\mathcal{P}(\mathcal{E}),l=0,\ldots,M_{0}-1\}.

Proposition 3.2 ([Cho22b, Proposition 4.2]).

Let (X,G,o)(X,G,o) be as in Convention 1.1, let μ\mu be a non-elementary probabiltiy measure on GG and let SS be a large and fairly long Schottky set for μ\mu. Then there exist a probability space (Ω,)(\Omega,\operatorname{\mathbb{P}}) for μ\mu and a constant K>0K>0 such that, for each n0n\geq 0, we have a measurable partition 𝒫n={α}α\mathscr{P}_{n}=\{\mathcal{E}_{\alpha}\}_{\alpha} of Ω\Omega by pivotal equivalence classes avoiding 11, \ldots, n/2+1\lfloor n/2\rfloor+1 and n+1n+1, that satisfies

(ω:#(𝒫(ω){1,,k})k/K|g1,,gn/2+1,gn+1)Kek/K\operatorname{\mathbb{P}}\left(\operatorname{\omega}:\#(\mathcal{P}(\operatorname{\omega})\cap\{1,\ldots,k\})\leq k/K\,|\,g_{1},\ldots,g_{\lfloor n/2\rfloor+1},g_{n+1}\right)\leq Ke^{-k/K}

for each choice of g1,,gn/2+1,gn+1Gg_{1},\ldots,g_{\lfloor n/2\rfloor+1},g_{n+1}\in G and for each knk\geq n.

Furthermore, again by using the property of Schottky set (Definition 2.23) and the alignment lemma (Lemma 2.18), we obtained the following:

Corollary 3.3 ([Cho22b, Corollary 4.5]).

Let (X,G,o)(X,G,o) be as in Convention 1.1, let μ\mu be a non-elementary probability measure on GG, let K0,N0>0K_{0},N_{0}>0 and let SS be a long enough K0K_{0}-Schottky set for μ\mu with cardinality N0N_{0}. Let \mathcal{E} be a pivotal equivalence class for μ\mu with 𝒫()={j(1)<j(2)<}\operatorname{\mathcal{P}}(\mathcal{E})=\{j(1)<j(2)<\ldots\} and let xx be a point in XX. Then for each k1k\geq 1 we have

((x,𝐘j(k)(ω),𝐘j(k+1)(ω),)is D0-semi-aligned|)1(1/N0)k.\operatorname{\mathbb{P}}\left(\big{(}x,\,\operatorname{\mathbf{Y}}_{j(k)}(\operatorname{\omega}),\,\operatorname{\mathbf{Y}}_{j(k+1)}(\operatorname{\omega}),\ldots\big{)}\,\,\textrm{is $D_{0}$-semi-aligned}\,\Big{|}\,\mathcal{E}\right)\geq 1-(1/N_{0})^{k}.

Moreover, for any m1m\geq 1, nj(m)n\geq j(m) and k=1,,mk=1,\ldots,m, we have

((𝐘j(1)(ω),,𝐘j(mk+1)(ω),Zn(ω)o)is D0-semi-aligned|)1(1/N0)k.\operatorname{\mathbb{P}}\left(\big{(}\operatorname{\mathbf{Y}}_{j(1)}(\operatorname{\omega}),\,\ldots,\,\operatorname{\mathbf{Y}}_{j(m-k+1)}(\operatorname{\omega}),\,Z_{n}(\operatorname{\omega})o\big{)}\,\,\textrm{is $D_{0}$-semi-aligned}\,\Big{|}\,\mathcal{E}\right)\geq 1-(1/N_{0})^{k}.
Corollary 3.4 ([Cho22b, Corollary 4.6]).

Let (X,G,o)(X,G,o) be as in Convention 1.1, let μ\mu be a non-elementary probability measure on GG and μˇ\check{\mu} be its reflected version, let K0,N0>0K_{0},N_{0}>0 and let SS and Sˇ\check{S} be long enough K0K_{0}-Schottky sets for μ\mu and μˇ\check{\mu}, respectively, with cardinality N0N_{0}. Let \mathcal{E} be a pivotal equivalence class for μ\mu with 𝒫()={j(1)<j(2)<}\operatorname{\mathcal{P}}(\mathcal{E})=\{j(1)<j(2)<\ldots\}, and let ˇ\check{\mathcal{E}} be a pivotal equivalence class for μˇ\check{\mu} with 𝒫(ˇ)={jˇ(1)<jˇ(2)<}\operatorname{\mathcal{P}}(\check{\mathcal{E}})=\{\check{j}(1)<\check{j}(2)<\ldots\}. Then for each k1k\geq 1 we have

((𝐘¯j(k)(ωˇ),𝐘jˇ(k)(ω))is D0-semi-aligned|)1(2/N0)k.\operatorname{\mathbb{P}}\left(\big{(}\bar{\operatorname{\mathbf{Y}}}_{j(k)}(\check{\operatorname{\omega}}),\,\operatorname{\mathbf{Y}}_{\check{j}(k)}(\operatorname{\omega})\big{)}\,\,\textrm{is $D_{0}$-semi-aligned}\,\Big{|}\,\mathcal{E}\right)\geq 1-(2/N_{0})^{k}.

Combining these, we proved in the proof of [Cho22b, Corollary 4.7] that:

Proposition 3.5 ([Cho22b, Corollary 4.7]).

Let (X,G,o)(X,G,o) be as in Convention 1.1, let μ\mu be a non-elementary probability measure on GG, let (Ω,)(\Omega,\operatorname{\mathbb{P}}) be a probability space for μ\mu, and let SS be a large and fairly long Schottky set for μ\mu. Then there exists K>0K>0 such that

(ω:(o,𝐘j(1),,𝐘j(n/2K),Zn(ω)o) is D0-semi-aligned for some j(1),,j(n/2K))1Ken/K.\operatorname{\mathbb{P}}\left(\operatorname{\omega}:\begin{array}[]{c}\textrm{$\big{(}o,\mathbf{Y}_{j(1)},\ldots,\mathbf{Y}_{j(\lfloor n/2K\rfloor)},Z_{n}(\operatorname{\omega})o\big{)}$ is $D_{0}$-semi-aligned}\\ \textrm{ for some $j(1),\ldots,j(\lfloor n/2K\rfloor)\in\operatorname{\mathbb{Z}}$}\end{array}\right)\geq 1-Ke^{-n/K}.

Now, Corollary 2.26 and our choices of the constants D0,D1D_{0},D_{1} tells us that, if (x,κ1,,κm,y)(x,\kappa_{1},\ldots,\kappa_{m},y) is a D0D_{0}-semi-aligned sequence, then d(x,y)d(x,y) is larger than 50D1m50D_{1}m. Combining this with Proposition 3.5, we deduce the strict positivity of the escape rate:

Theorem 3.6 ([Cho22b, Corollary 4.7]).

Let (X,G,o)(X,G,o) be as in Convention 1.1 and let (Zn)n(Z_{n})_{n} be the random walk generated by a non-elementary probability measure μ\mu on GG. Then there exists a strictly positive quantity λ(μ)(0,+]\lambda(\mu)\in(0,+\infty], called the drift of μ\mu, such that

λ(μ):=limn1nd(o,Zno)almost surely.\lambda(\mu):=\lim_{n\rightarrow\infty}\frac{1}{n}d(o,Z_{n}o)\quad\textrm{almost surely.}

We next turn to deviation inequalities. We use the parametrization (ωˇ,wˇ)(G>0,μˇ>0)×(G>0,μ>0)(\check{\operatorname{\omega}},\check{w})\in(G^{\operatorname{\mathbb{Z}}_{>0}},\check{\mu}^{\operatorname{\mathbb{Z}}_{>0}})\times(G^{\operatorname{\mathbb{Z}}_{>0}},\mu^{\operatorname{\mathbb{Z}}_{>0}}). In [Cho22b], we defined:

Definition 3.7 ([Cho22b, Section 4.3]).

Let (X,G,o)(X,G,o) be as in Convention 1.1, let μ\mu be a non-elementary probability measure on GG and let SS be a large and fairly long K0K_{0}-Schottky set in (suppμ)M0(\operatorname{supp}\mu)^{M_{0}} for some K0,M0>0K_{0},M_{0}>0.

For (ωˇ,ω)(G>0,μˇ>0)×(G>0,μ>0)(\check{\operatorname{\omega}},\operatorname{\omega})\in(G^{\operatorname{\mathbb{Z}}_{>0}},\check{\mu}^{\operatorname{\mathbb{Z}}_{>0}})\times(G^{\operatorname{\mathbb{Z}}_{>0}},\mu^{\operatorname{\mathbb{Z}}_{>0}}), we define υ=υ(ωˇ,ω)\operatorname{\upsilon}=\operatorname{\upsilon}(\check{\operatorname{\omega}},\operatorname{\omega}) to be the minimal index such that there exists M0iυM_{0}\leq i\leq\operatorname{\upsilon} satisfying:

  1. (1)

    𝐘i(ω):=(ZiM0o,,Zi1o,Zio)\mathbf{Y}_{i}(\operatorname{\omega}):=(Z_{i-M_{0}}o,\ldots,Z_{i-1}o,Z_{i}o) is a Schottky axis;

  2. (2)

    (Zˇmo,𝐘i(ω),Zno)(\check{Z}_{m}o,\mathbf{Y}_{i}(\operatorname{\omega}),Z_{n}o) is D0D_{0}-semi-aligned for all nυn\geq\operatorname{\upsilon} and m0m\geq 0.

We then define υˇ=υˇ(ωˇ,ω)\check{\operatorname{\upsilon}}=\check{\operatorname{\upsilon}}(\check{\operatorname{\omega}},\operatorname{\omega}) to be the minimal index such that there exists M0iυˇM_{0}\leq i\leq\check{\operatorname{\upsilon}} satisfying:

  1. (1)

    𝐘¯i(ωˇ):=(Zˇio,Zˇi1o,,ZˇiM0o)\bar{\mathbf{Y}}_{i}(\check{\operatorname{\omega}}):=(\check{Z}_{i}o,\check{Z}_{i-1}o,\ldots,\check{Z}_{i-M_{0}}o) is a Schottky axis;

  2. (2)

    (Zˇno,𝐘¯i(ωˇ),Zmo)(\check{Z}_{n}o,\bar{\mathbf{Y}}_{i}(\check{\operatorname{\omega}}),Z_{m}o) is D0D_{0}-semi-aligned for all nυˇn\geq\check{\operatorname{\upsilon}} and m0m\geq 0.

Using Corollary 3.3 and 3.4, we deduced the following:

Lemma 3.8 ([Cho22b, Lemma 4.9]).

Let (X,G,o)(X,G,o) be as in Convention 1.1, let μ\mu be a non-elementary probability measure on GG and let SS be a large and long enough Schottky set for μ\mu. Then there exists K>0K^{\prime}>0 such that

(3.1) (υ(ωˇ,ω)k|gk+1,gˇ1,,gˇk+1)Kek/K\operatorname{\mathbb{P}}\left(\operatorname{\upsilon}(\check{\operatorname{\omega}},\operatorname{\omega})\geq k\,\Big{|}\,g_{k+1},\check{g}_{1},\ldots,\check{g}_{k+1}\right)\leq K^{\prime}e^{-k/K^{\prime}}

holds for every k0k\geq 0 and every choice of gk+1,gˇ1,,gˇk+1Gg_{k+1},\check{g}_{1},\ldots,\check{g}_{k+1}\in G, and

(3.2) (υˇ(ωˇ,ω)k|gˇk+1,g1,,gk+1)Kek/K\operatorname{\mathbb{P}}\left(\check{\operatorname{\upsilon}}(\check{\operatorname{\omega}},\operatorname{\omega})\geq k\,\Big{|}\,\check{g}_{k+1},g_{1},\ldots,g_{k+1}\right)\leq K^{\prime}e^{-k/K^{\prime}}

holds for every k0k\geq 0 and every choice of gˇk+1,g1,,gk+1G\check{g}_{k+1},g_{1},\ldots,g_{k+1}\in G.

This martingale-like estimates enable us to compute the moments of d(o,Zυo)d(o,Z_{\operatorname{\upsilon}}o) and d(o,Zˇυˇo)d(o,\check{Z}_{\check{\operatorname{\upsilon}}}o) based on the moments of μ\mu and μˇ\check{\mu}, respectively. In [Cho22b, Lemma 4.8], these quantities were used to estimate the Gromov product between the backward and the forward sample paths. This lemma needs a modification in our setting due to the asymmetry of the metric:

Lemma 3.9 (cf. [Cho22b, Lemma 4.8]).

Let (X,G,o)(X,G,o) be as in Convention 1.1, let μ\mu be a non-elementary probability measure on GG and let SS be a fairly long Schottky set for μ\mu. Then for each (ωˇ,ω)(G>0,μˇ>0)×(G>0,μ>0)(\check{\operatorname{\omega}},\operatorname{\omega})\in(G^{\operatorname{\mathbb{Z}}_{>0}},\check{\mu}^{\operatorname{\mathbb{Z}}_{>0}})\times(G^{\operatorname{\mathbb{Z}}_{>0}},\mu^{\operatorname{\mathbb{Z}}_{>0}}), we have

(3.3) (Zˇmo,Zno)o12[d(Zko,o)+d(o,Zko)]=12dsym(o,Zko)(\check{Z}_{m}o,Z_{n}o)_{o}\leq\frac{1}{2}\left[d(Z_{k}o,o)+d(o,Z_{k}o)\right]=\frac{1}{2}d^{sym}(o,Z_{k}o)

for all m0m\geq 0 and n,kυ(ωˇ,ω)n,k\geq\operatorname{\upsilon}(\check{\operatorname{\omega}},\operatorname{\omega}). Moreover, we have

(3.4) (Zˇmo,Zno)o12[d(Zˇmin(m,υˇ)o,o)+d(o,Zmin(n,υ)o)](\check{Z}_{m}o,Z_{n}o)_{o}\leq\frac{1}{2}\left[d(\check{Z}_{\min(m,\check{\operatorname{\upsilon}})}o,o)+d(o,Z_{\min(n,\operatorname{\upsilon})}o)\right]

for all m,n0m,n\geq 0.

Proof.

For the first assertion, let iυ(ωˇ,ω)i\leq\operatorname{\upsilon}(\check{\operatorname{\omega}},\operatorname{\omega}) be the index such that (Zˇmo,𝐘i(ω),Zno)(\check{Z}_{m}o,\mathbf{Y}_{i}(\operatorname{\omega}),Z_{n^{\prime}}o) is D0D_{0}-semi-aligned for all nυ(ωˇ,ω)n^{\prime}\geq\operatorname{\upsilon}(\check{\operatorname{\omega}},\operatorname{\omega}) and m0m\geq 0. Since (Zˇmo,𝐘i(ω),Zno)(\check{Z}_{m}o,\mathbf{Y}_{i}(\operatorname{\omega}),Z_{n}o) is D0D_{0}-semi-aligned, Corollary 2.26 tells us the geodesic [Zˇmo,Zno][\check{Z}_{m}o,Z_{n}o] contains a point qq in the 0.1E00.1E_{0}-dsymd^{sym}-neighborhood of ZiM0oZ_{i-M_{0}}o, the beginning point of 𝐘i(ω)\mathbf{Y}_{i}(\operatorname{\omega}). We then have

d(Zˇmo,Zno)\displaystyle d(\check{Z}_{m}o,Z_{n}o) =d(Zˇmo,q)+d(q,Zno)\displaystyle=d(\check{Z}_{m}o,q)+d(q,Z_{n}o)
(d(Zˇmo,ZiM0o)d(q,ZiM0o))+(d(ZiM0o,Zno)d(ZiM0o,q))\displaystyle\geq\big{(}d(\check{Z}_{m}o,Z_{i-M_{0}}o)-d(q,Z_{i-M_{0}}o)\big{)}+\big{(}d(Z_{i-M_{0}}o,Z_{n}o)-d(Z_{i-M_{0}}o,q)\big{)}
d(Zˇmo,o)d(ZiM0o,o)+d(o,Zno)d(o,ZiM0o)0.1E0.\displaystyle\geq d(\check{Z}_{m}o,o)-d(Z_{i-M_{0}}o,o)+d(o,Z_{n}o)-d(o,Z_{i-M_{0}}o)-0.1E_{0}.

This implies that

(3.5) (Zˇmo,Zno)o12(d(ZiM0o,o)+d(o,ZiM0o))+0.05E0.(\check{Z}_{m}o,Z_{n}o)_{o}\leq\frac{1}{2}\left(d(Z_{i-M_{0}}o,o)+d(o,Z_{i-M_{0}}o)\right)+0.05E_{0}.

Meanwhile, (o,𝐘i(ω),Zko)(o,\mathbf{Y}_{i}(\operatorname{\omega}),Z_{k}o) and (Zko,𝐘¯i(ω),o)(Z_{k}o,\bar{\mathbf{Y}}_{i}(\operatorname{\omega}),o) are both D0D_{0}-semi-aligned as well. By Corollary 2.26, [o,Zko][o,Z_{k}o] contains a subsegment longer than 100E0100E_{0} that is 0.1E00.1E_{0}-fellow traveling with 𝐘i(ω)\mathbf{Y}_{i}(\operatorname{\omega}). It follows that d(o,ZiM0o)d(o,Zko)99E0d(o,Z_{i-M_{0}}o)\leq d(o,Z_{k}o)-99E_{0}. ((\ast) Similarly, by applying Corollary 2.26 to [Zko,o][Z_{k}o,o], we deduce that d(ZiM0o,o)d(Zko,o)99E0d(Z_{i-M_{0}}o,o)\leq d(Z_{k}o,o)-99E_{0}. (\ast\ast) By combining these with Inequality 3.5, we conclude the first claim.

Let us now see the second assertion. When mυˇ(ωˇ,w)m\leq\check{\operatorname{\upsilon}}(\check{\operatorname{\omega}},w) and nυ(ωˇ,ω)n\leq\operatorname{\upsilon}(\check{\operatorname{\omega}},\operatorname{\omega}), the conclusion follows from the definition of the Gromov product.

Next, consider the case that mυˇm\geq\check{\operatorname{\upsilon}} and nυn\geq\operatorname{\upsilon}. Together with the choice of iυ(ωˇ,ω)i\leq\operatorname{\upsilon}(\check{\operatorname{\omega}},\operatorname{\omega}), let jυˇ(ωˇ,ω)j\leq\check{\operatorname{\upsilon}}(\check{\operatorname{\omega}},\operatorname{\omega}) be the index such that (Zˇn′′o,𝐘¯j(ωˇ),Zmo)(\check{Z}_{n^{\prime\prime}}o,\bar{\mathbf{Y}}_{j}(\check{\operatorname{\omega}}),Z_{m}o) is D0D_{0}-semi-aligned for all n′′υˇ(ωˇ,ω)n^{\prime\prime}\geq\check{\operatorname{\upsilon}}(\check{\operatorname{\omega}},\operatorname{\omega}) and m0m\geq 0. Then (Zˇn′′o,𝐘¯j(ωˇ),Zmo)(\check{Z}_{n^{\prime\prime}}o,\bar{\mathbf{Y}}_{j}(\check{\operatorname{\omega}}),Z_{m}o) is D0D_{0}-semi-aligned, hence D1D_{1}-aligned by Corollary 2.26, for n′′υˇn^{\prime\prime}\geq\check{\operatorname{\upsilon}} and m=iM0,,im=i-M_{0},\ldots,i. Similarly, (Zˇmo,𝐘i(ω),Zno)(\check{Z}_{m}o,\mathbf{Y}_{i}(\operatorname{\omega}),Z_{n^{\prime}}o) is D1D_{1}-aligned for nυn^{\prime}\geq\operatorname{\upsilon} and m=jM0,,jm=j-M_{0},\ldots,j. We conclude that

(Zˇmo,𝐘¯j(ωˇ),𝐘i(ω),Zno)is D1-aligned.\big{(}\check{Z}_{m}o,\bar{\mathbf{Y}}_{j}(\check{\operatorname{\omega}}),\mathbf{Y}_{i}(\operatorname{\omega}),Z_{n}o\big{)}\,\,\textrm{is $D_{1}$-aligned}.

By Proposition 2.20, there exist p,q[Zˇmo,Zno]p,q\in[\check{Z}_{m}o,Z_{n}o], with pp coming first, such that p𝒩0.1E0(ZˇjM0o)p\in\mathcal{N}_{0.1E_{0}}(\check{Z}_{j-M_{0}}o) and q𝒩0.1E0(ZiM0o)q\in\mathcal{N}_{0.1E_{0}}(Z_{i-M_{0}}o). We then have

d(Zˇmo,Zno)\displaystyle d(\check{Z}_{m}o,Z_{n}o) =d(Zˇmo,p)+d(p,q)+d(q,Zno)\displaystyle=d(\check{Z}_{m}o,p)+d(p,q)+d(q,Z_{n}o)
d(Zˇmo,p)+d(q,Zno)\displaystyle\geq d(\check{Z}_{m}o,p)+d(q,Z_{n}o)
d(Zˇmo,ZˇjM0o)+d(ZiM0o,Zno)0.2E0\displaystyle\geq d(\check{Z}_{m}o,\check{Z}_{j-M_{0}}o)+d(Z_{i-M_{0}}o,Z_{n}o)-0.2E_{0}
d(Zˇmo,o)+d(o,Zno)d(ZˇjM0o,o)d(o,ZiM0o)0.2E0.\displaystyle\geq d(\check{Z}_{m}o,o)+d(o,Z_{n}o)-d(\check{Z}_{j-M_{0}}o,o)-d(o,Z_{i-M_{0}}o)-0.2E_{0}.

Meanwhile, by ((\ast) and ((\ast\ast), we have d(ZˇjM0o,o)d(Zˇlo,o)+99E0d(\check{Z}_{j-M_{0}}o,o)\leq d(\check{Z}_{l}o,o)+99E_{0} and d(o,ZiM0o)d(o,Zko)+99E0d(o,Z_{i-M_{0}}o)\leq d(o,Z_{k}o)+99E_{0}. Combining these yields the conclusion.

Finally, consider the case that mυˇm\geq\check{\operatorname{\upsilon}} and nυn\leq\operatorname{\upsilon}. Then (Zˇmo,𝐘¯j(wˇ),Zno)(\check{Z}_{m}o,\bar{\mathbf{Y}}_{j}(\check{w}),Z_{n}o) is D1D_{1}-aligned. By Proposition 2.20, there exist p[Zˇmo,Zno]p\in[\check{Z}_{m}o,Z_{n}o] such that p𝒩0.1E0(ZˇjM0o)p\in\mathcal{N}_{0.1E_{0}}(\check{Z}_{j-M_{0}}o). We then have

d(Zˇmo,Zno)\displaystyle d(\check{Z}_{m}o,Z_{n}o) d(Zˇmo,p)d(Zˇmo,o)d(ZˇjM0o,o)0.05E0.\displaystyle\geq d(\check{Z}_{m}o,p)\geq d(\check{Z}_{m}o,o)-d(\check{Z}_{j-M_{0}}o,o)-0.05E_{0}.

This implies that (Zˇmo,Zno)o12[d(ZˇjM0o,o)+d(o,Zno)](\check{Z}_{m}o,Z_{n}o)_{o}\leq\frac{1}{2}\left[d(\check{Z}_{j-M_{0}}o,o)+d(o,Z_{n}o)\right]. Meanwhile, by ()(\ast\ast), we have d(ZˇjM0o,o)d(Zˇlo,o)+99E0d(\check{Z}_{j-M_{0}}o,o)\leq d(\check{Z}_{l}o,o)+99E_{0}. Combining these results yields the conclusion. The case mυˇm\leq\check{\operatorname{\upsilon}} and nυn\geq\operatorname{\upsilon} can be handled with a similar argument. ∎

Because Inequality 3.3 involves dsym(o,Zko)d^{sym}(o,Z_{k}o) instead of d(o,Zko)d(o,Z_{k}o), [Cho22c, Proposition 4.10] now requires the finite pp-th moment of μ\mu with respect to dsymd^{sym}, or equivalently, the finite pp-th moments of μ\mu and μˇ\check{\mu}:

Proposition 3.10 ([Cho22b, Proposition 4.10]).

Let (X,G,o)(X,G,o) be as in Convention 1.1, let p>0p>0 and let ((Zˇn)n>0,(Zn)n>0)((\check{Z}_{n})_{n>0},(Z_{n})_{n>0}) be the (bi-directional) random walk generated by a non-elementary probability measure μ\mu on GG. Suppose that both μ\mu and μˇ\check{\mu} has finite pp-th moment. Then the random variable supn,m0(Zˇmo,Zno)o\sup_{n,m\geq 0}(\check{Z}_{m}o,Z_{n}o)_{o} has finite 2p2p-th moment.

Proof.

Note that the assumption implies that 𝔼μ[dsym(o,go)p]<+\operatorname{\mathbb{E}}_{\mu}[d^{sym}(o,go)^{p}]<+\infty. The proof is done almost verbatim to the proof of [Cho22b, Proposition 4.10], after replacing d(,)d(\cdot,\cdot) with 12dsym(,)\frac{1}{2}d^{sym}(\cdot,\cdot) throughout. ∎

By using Inequality 3.4 instead, we obtain the following weaker result:

Proposition 3.11 ([Cho22b, Proposition 4.10]).

Let (X,G,o)(X,G,o) be as in Convention 1.1, let p>0p>0 and let ((Zˇn)n>0,(Zn)n>0)((\check{Z}_{n})_{n>0},(Z_{n})_{n>0}) be the (bi-directional) random walk generated by a non-elementary probability measure μ\mu on GG with finite pp-th moment. Then the random variable supn,m0(Zˇmo,Zno)o\sup_{n,m\geq 0}(\check{Z}_{m}o,Z_{n}o)_{o} has finite pp-th moment.

Proof.

Let KK^{\prime} be as in Lemma 3.8. We define Dk:=i=1kd(o,gio)D_{k}:=\sum_{i=1}^{k}d(o,g_{i}o) and Dˇk:=i=1kd(gˇio,o)=i=1kd(o,g1io)\check{D}_{k}:=\sum_{i=1}^{k}d(\check{g}_{i}o,o)=\sum_{i=1}^{k}d(o,g_{1-i}o). Note the following consequence of the triangle inequality:

d(o,Z1o),,d(o,Zko)Dk,d(Zˇ1o,o),,d(Zˇko,o)Dˇk.d(o,Z_{1}o),\ldots,d(o,Z_{k}o)\leq D_{k},\quad d(\check{Z}_{1}o,o),\ldots,d(\check{Z}_{k}o,o)\leq\check{D}_{k}.

Also note the following computation: if non-negative constants A,BA,B and CC satisfy A12(B+C)A\leq\frac{1}{2}(B+C), then

Ap(12(B+C))p(max(B,C))pBp+Cp.A^{p}\leq\left(\frac{1}{2}(B+C)\right)^{p}\leq\left(\max(B,C)\right)^{p}\leq B^{p}+C^{p}.

Noting this inequality, by Lemma 3.9, we have

supm,n0(Zˇmo,Zno)op\displaystyle\sup_{m,n\geq 0}(\check{Z}_{m}o,Z_{n}o)_{o}^{p} supm,n0(d(Zˇmin(m,υˇ)o,o)p+d(o,Zmin(n,υ)o)p)\displaystyle\leq\sup_{m,n\geq 0}\left(d(\check{Z}_{\min(m,\check{\operatorname{\upsilon}})}o,o)^{p}+d(o,Z_{\min(n,\operatorname{\upsilon})}o)^{p}\right)
Dυp+Dˇυˇp.\displaystyle\leq D_{\operatorname{\upsilon}}^{p}+\check{D}_{\check{\operatorname{\upsilon}}}^{p}.

Hence, it suffices to control 𝔼[Dυp]\operatorname{\mathbb{E}}[D_{\operatorname{\upsilon}}^{p}] and 𝔼[Dˇυˇp]\operatorname{\mathbb{E}}[\check{D}_{\check{\operatorname{\upsilon}}}^{p}] on (G>0,μˇ>0)×(G>0,μ>0)(G^{\operatorname{\mathbb{Z}}_{>0}},\check{\mu}^{\operatorname{\mathbb{Z}}_{>0}})\times(G^{\operatorname{\mathbb{Z}}_{>0}},\mu^{\operatorname{\mathbb{Z}}_{>0}}). The computations are analogous so we will only show that 𝔼[Dυp]\operatorname{\mathbb{E}}[D_{\operatorname{\upsilon}}^{p}] is finite. First, when p1p\leq 1, the concavity of f(t)=tpf(t)=t^{p} tells us that Dk+1pDkp(Dk+1Dk)p=d(o,gk+1o)pD_{k+1}^{p}-D_{k}^{p}\leq(D_{k+1}-D_{k})^{p}=d(o,g_{k+1}o)^{p} for each kk. Hence, we compute

(3.6) 𝔼[Dυp]\displaystyle\operatorname{\mathbb{E}}[D_{\operatorname{\upsilon}}^{p}] =𝔼[i=0(Di+1pDip)1i<υ]i=0𝔼[d(o,gi+1o)p1i<υ]\displaystyle=\operatorname{\mathbb{E}}\left[\sum_{i=0}^{\infty}\left(D_{i+1}^{p}-D_{i}^{p}\right)1_{i<\operatorname{\upsilon}}\right]\leq\sum_{i=0}^{\infty}\operatorname{\mathbb{E}}\left[d(o,g_{i+1}o)^{p}1_{i<\operatorname{\upsilon}}\right]
i=0𝔼[d(o,gi+1o)p[i<υ|gi+1]]i=0𝔼[d(o,gi+1o)pKei/K]\displaystyle\leq\sum_{i=0}^{\infty}\operatorname{\mathbb{E}}\left[d(o,g_{i+1}o)^{p}\cdot\operatorname{\mathbb{P}}\left[i<\operatorname{\upsilon}\,\big{|}\,g_{i+1}\right]\right]\leq\sum_{i=0}^{\infty}\operatorname{\mathbb{E}}\left[d(o,g_{i+1}o)^{p}\cdot K^{\prime}e^{-i/K^{\prime}}\right] (Lemma 3.8)\displaystyle(\because\textrm{Lemma \ref{lem:Devi}})
K1e1/K𝔼μ[d(o,go)p]<+.\displaystyle\leq\frac{K^{\prime}}{1-e^{-1/K^{\prime}}}\operatorname{\mathbb{E}}_{\mu}[d(o,go)^{p}]<+\infty.

When p1p\geq 1, we have

Dk+1pDkp2pd(o,gk+1o)p+2pDkp1d(o,gk+1o).D_{k+1}^{p}-D_{k}^{p}\leq 2^{p}d(o,g_{k+1}o)^{p}+2^{p}D_{k}^{p-1}d(o,g_{k+1}o).

Using this, we can bound 𝔼[Dυp]=𝔼[i=0(Di+1pDip)1i<υ]\operatorname{\mathbb{E}}[D_{\operatorname{\upsilon}}^{p}]=\operatorname{\mathbb{E}}\left[\sum_{i=0}^{\infty}\left(D_{i+1}^{p}-D_{i}^{p}\right)1_{i<\operatorname{\upsilon}}\right] by a linear combination of i=0𝔼[d(o,gi+1o)p1i<υ]\sum_{i=0}^{\infty}\operatorname{\mathbb{E}}\left[d(o,g_{i+1}o)^{p}1_{i<\operatorname{\upsilon}}\right] and i=0𝔼[Dip1d(o,gi+1o)1i<υ]\sum_{i=0}^{\infty}\operatorname{\mathbb{E}}\left[D_{i}^{p-1}d(o,g_{i+1}o)1_{i<\operatorname{\upsilon}}\right]. The former summation was bounded by a multiple of 𝔼μ[d(o,go)p]\operatorname{\mathbb{E}}_{\mu}[d(o,go)^{p}] in Display 3.6. We now treat the summands of the latter summation as follows:

(3.7) 𝔼[Dip1d(o,gi+1o)1i<υ]\displaystyle\operatorname{\mathbb{E}}\left[D_{i}^{p-1}d(o,g_{i+1}o)1_{i<\operatorname{\upsilon}}\right] 𝔼[ei/2Kd(o,gi+1o)1i<υ]+𝔼[Dip11Di>ei/2K(p1)d(o,gi+1o)1i<υ]\displaystyle\leq\operatorname{\mathbb{E}}\left[e^{i/2K^{\prime}}\cdot d(o,g_{i+1}o)1_{i<\operatorname{\upsilon}}\right]+\operatorname{\mathbb{E}}\left[D_{i}^{p-1}1_{D_{i}>e^{i/2K^{\prime}(p-1)}}\cdot d(o,g_{i+1}o)1_{i<\operatorname{\upsilon}}\right]
ei/2K𝔼[d(o,gi+1o)[i<υ|gi+1]]+𝔼[Dipei/2K(p1)d(o,gi+1o)]\displaystyle\leq e^{i/2K^{\prime}}\cdot\operatorname{\mathbb{E}}\left[d(o,g_{i+1}o)\operatorname{\mathbb{P}}\left[i<\operatorname{\upsilon}\,\big{|}\,g_{i+1}\right]\right]+\operatorname{\mathbb{E}}\left[D_{i}^{p}\cdot e^{-i/2K^{\prime}(p-1)}\cdot d(o,g_{i+1}o)\right]
Kei/2K𝔼[d(o,gi+1o)p]\displaystyle\leq K^{\prime}e^{-i/2K^{\prime}}\operatorname{\mathbb{E}}[d(o,g_{i+1}o)^{p}]
+ei/2K(p1)𝔼[d(o,gi+1o)]ip𝔼[d(o,g1o)p++d(o,gio)p].\displaystyle+e^{-i/2K^{\prime}(p-1)}\cdot\operatorname{\mathbb{E}}[d(o,g_{i+1}o)]\cdot i^{p}\operatorname{\mathbb{E}}[d(o,g_{1}o)^{p}+\ldots+d(o,g_{i}o)^{p}].

Here, the final term due to the fact that 𝔼[(X1++Xn)p]𝔼[(nmaxiXi)p]=np𝔼[maxiXip]np𝔼[X1p++Xnp]\operatorname{\mathbb{E}}[(X_{1}+\ldots+X_{n})^{p}]\leq\operatorname{\mathbb{E}}[(n\max_{i}X_{i})^{p}]=n^{p}\cdot\operatorname{\mathbb{E}}[\max_{i}X_{i}^{p}]\leq n^{p}\operatorname{\mathbb{E}}[X_{1}^{p}+\ldots+X_{n}^{p}] for nonnegative RVs XiX_{i}’s. Since this summand is exponentially decaying, its summation is finite as desired. ∎

Using Inequality 3.4, we also obtain the exponential deviation inequality as below.

Proposition 3.12 ([Cho22b, Corollary 4.12]).

Let (X,G,o)(X,G,o) be as in Convention 1.1 and let ((Zˇn)n>0,(Zn)n>0)((\check{Z}_{n})_{n>0},(Z_{n})_{n>0}) be the (bi-directional) random walk generated by a non-elementary random walk μ\mu on GG with finite exponential moment. Then there exists K>0K>0 such that

𝔼[exp(supn,m0(Zˇmo,Zno)o/K)]<K.\operatorname{\mathbb{E}}\left[\operatorname{exp}\left(\sup_{n,m\geq 0}(\check{Z}_{m}o,Z_{n}o)_{o}/K\right)\right]<K.
Proof.

As in the proof of Proposition 3.11, we let Dk:=i=1kd(o,gio)D_{k}:=\sum_{i=1}^{k}d(o,g_{i}o) and Dˇk:=i=1kd(gˇio,o)=i=1kd(o,g1io)\check{D}_{k}:=\sum_{i=1}^{k}d(\check{g}_{i}o,o)=\sum_{i=1}^{k}d(o,g_{1-i}o). Then by Lemma 3.9, we have

exp(supm,n0(Zˇmo,Zno)o/K)\displaystyle\operatorname{exp}\left(\sup_{m,n\geq 0}(\check{Z}_{m}o,Z_{n}o)_{o}/K\right) supm,n0(exp(d(Zˇmin(m,υˇ)o,o)/K)+exp(d(o,Zmin(n,υ)o)/K))\displaystyle\leq\sup_{m,n\geq 0}\left(\operatorname{exp}\big{(}d(\check{Z}_{\min(m,\check{\operatorname{\upsilon}})}o,o)/K\big{)}+\operatorname{exp}\big{(}d(o,Z_{\min(n,\operatorname{\upsilon})}o)/K\big{)}\right)
exp(Dυ/K)+exp(Dˇυˇ/K).\displaystyle\leq\operatorname{exp}(D_{\operatorname{\upsilon}}/K)+\operatorname{exp}(\check{D}_{\check{\operatorname{\upsilon}}}/K).

Also note that

𝔼[exp(Dυ/K)]i=0𝔼[exp(Di+1/K)1i<υ],𝔼[exp(Dˇυˇ/K)]i=0𝔼[exp(Dˇi+1/K)1i<υˇ].\operatorname{\mathbb{E}}[\operatorname{exp}(D_{\operatorname{\upsilon}}/K)]\leq\sum_{i=0}^{\infty}\operatorname{\mathbb{E}}\left[\operatorname{exp}(D_{i+1}/K)1_{i<\operatorname{\upsilon}}\right],\quad\operatorname{\mathbb{E}}[\operatorname{exp}(\check{D}_{\check{\operatorname{\upsilon}}}/K)]\leq\sum_{i=0}^{\infty}\operatorname{\mathbb{E}}\left[\operatorname{exp}(\check{D}_{i+1}/K)1_{i<\check{\operatorname{\upsilon}}}\right].

Hence, it suffices to prove that 𝔼[exp(Di+1/K)1i<υ]\operatorname{\mathbb{E}}\left[\operatorname{exp}(D_{i+1}/K)1_{i<\operatorname{\upsilon}}\right] (and its symmetric counterpart) are exponentially summable for large enough KK. This follows from the proof of [Cho22b, Corollary 4.12]. ∎

From these deviation inequalities, we deduced several limit theorems in [Cho22b, Section 4.4]. These include:

Theorem 3.13 ([Cho22b, Theorem 4.13]).

Let (X,G,o)(X,G,o) be as in Convention 1.1 and let (Zn)n(Z_{n})_{n} be the random walk generated by a non-elementary probability measure μ\mu on GG with finite second moment. Then the following limit (called the asymptotic variance of μ\mu) exists:

σ2(μ):=limn1nVar[d(o,Zno)],\sigma^{2}(\mu):=\lim_{n\rightarrow\infty}\frac{1}{n}Var[d(o,Z_{n}o)],

and the random variable 1n[d(o,Zno)λ(μ)n]\frac{1}{\sqrt{n}}[d(o,Z_{n}o)-\lambda(\mu)n] converges in law to the Gaussian law 𝒩(0,σ(μ))\mathcal{N}(0,\sigma(\mu)) with zero mean and variance σ2(μ)\sigma^{2}(\mu).

Theorem 3.14 ([Cho22b, Theorem 4.16]).

Let (X,G,o)(X,G,o) be as in Convention 1.1 and let (Zn)n(Z_{n})_{n} be the random walk generated by a non-elementary probability measure μ\mu on GG with finite second moment. Then for almost every sample path (Zn)n(Z_{n})_{n} we have

lim supnd(o,Zno)λ(μ)n2nloglogn=σ(μ),\limsup_{n\rightarrow\infty}\frac{d(o,Z_{n}o)-\lambda(\mu)n}{\sqrt{2n\log\log n}}=\sigma(\mu),

where λ(μ)\lambda(\mu) is the drift of μ\mu and σ2(μ)\sigma^{2}(\mu) is the asymptotic variance of μ\mu.

Theorem 3.15 ([Cho22c, Theorem 4.18]).

Let (X,G,o)(X,G,o) be as in Convention 1.1 and let (Zn)n(Z_{n})_{n} be the random walk generated by a non-elementary measure μ\mu on GG.

  1. (1)

    Suppose that μ\mu and μˇ\check{\mu} have finite pp-th moment for some p>0p>0. Then for almost every sample path (Zn(ω))n0(Z_{n}(\operatorname{\omega}))_{n\geq 0}, there exists a quasigeodesic γ\gamma such that

    limn1n1/2pdsym(Zno,γ)=0.\lim_{n\rightarrow\infty}\frac{1}{n^{1/2p}}d^{sym}(Z_{n}o,\gamma)=0.
  2. (2)

    Suppose that μ\mu and μˇ\check{\mu} have finite exponential moment. Then there exists K<+K<+\infty such that for almost every sample path (Zn(ω)n0(Z_{n}(\operatorname{\omega})_{n\geq 0}, there exists a quasigeodesic γ\gamma satisfying

    limn1logndsym(Zno,γ)<K.\lim_{n\rightarrow\infty}\frac{1}{\log n}d^{sym}(Z_{n}o,\gamma)<K.

In [Cho22b, Section 6], we discussed a more complicated version of pivotal times following [Gou22, Section 5.A]. The ingredients required for the construction of pivotal times are again Definition 2.23 and Lemma 2.18. Furthermore, [Cho22b, Lemma 3.17] linked the alignment and the almost additivity of (forward) step progresses, and Proposition 2.20 now plays the same role. Hence, by following [Cho22b, Section 6], we deduce the large deviation principle:

Theorem 3.16 ([Cho22b, Theorem 6.4]).

Let (X,G,o)(X,G,o) be as in Convention 1.1 and let (Zn)n(Z_{n})_{n} be the random walk generated by a non-elementary probability measure μ\mu on GG. Let λ(μ)=limn1n𝔼[d(o,Zno)]\lambda(\mu)=\lim_{n}\frac{1}{n}\operatorname{\mathbb{E}}[d(o,Z_{n}o)] be the drift of μ\mu. Then for each 0<L<λ(μ)0<L<\lambda(\mu), the probability (d(o,Zno)Ln)\operatorname{\mathbb{P}}(d(o,Z_{n}o)\leq Ln) decays exponentially as nn tends to infinity.

Corollary 3.17 ([Cho22b, Corollary 6.5]).

Let (X,G,o)(X,G,o) be as in Convention 1.1 and let (Zn)n0(Z_{n})_{n\geq 0} be the random walk generated by a non-elementary probability measure μ\mu on GG. Then there exists a proper convex function I:[0,+]I:\mathbb{R}\rightarrow[0,+\infty], vanishing only at the drift λ(μ)\lambda(\mu), such that

infxint(E)I(x)\displaystyle-\inf_{x\in\operatorname{int}(E)}I(x) lim infn1nlog(1nd(id,Zn)E),\displaystyle\leq\liminf_{n\rightarrow\infty}\frac{1}{n}\log\operatorname{\mathbb{P}}\left(\frac{1}{n}d(id,Z_{n})\in E\right),
infxE¯I(x)\displaystyle-\inf_{x\in\bar{E}}I(x) lim supn1nlog(1nd(id,Zn)E)\displaystyle\geq\limsup_{n\rightarrow\infty}\frac{1}{n}\log\operatorname{\mathbb{P}}\left(\frac{1}{n}d(id,Z_{n})\in E\right)

holds for every measurable set EE\subseteq\mathbb{R}.

3.2. Results from [Cho22c]

We now discuss the results of [Cho22c], which consists of two parts. In the first part, we made use of the properties of the indices υ(ωˇ,ω)\operatorname{\upsilon}(\check{\operatorname{\omega}},\operatorname{\omega}) and υˇ(ωˇ,ω)\check{\operatorname{\upsilon}}(\check{\operatorname{\omega}},\operatorname{\omega}) in Definition 3.7. As we saw in Lemma 3.9, the asymmetry of the metric necessitates suitable modification.

For us, [Cho22c, Claim 3.3] during the proof of [Cho22c, Theorem B] requires a modification. This claim asserts that the discrepancy between the displacement d(o,Zno)d(o,Z_{n}o) and the translation length τ(Zn)\tau(Z_{n}) of a random isometry ZnZ_{n} is bounded by (a multiple of) the minimum of d(o,Zνo)d(o,Z_{\nu}o) and d(o,Zˇνˇo)d(o,\check{Z}_{\check{\nu}}o). This remains true if we replace d(,)d(\cdot,\cdot) with dsym(,)d^{sym}(\cdot,\cdot), and we obtain the full conclusion of [Cho22c, Theorem B.(1)] if μ\mu and μˇ\check{\mu} have finite pp-th moment.

In general, we have the following modified version of [Cho22c, Claim 3.3].

Lemma 3.18.

Let (X,G,o)(X,G,o) be as in Convention 1.1, let μ\mu be a non-elementary probability measure on GG with finite pp-th moment, and let SS be a fairly long Schottky set for μ\mu. Then for each n>0n>0, there exist RVs DnD_{n} and Dˇn\check{D}_{n} such that the following holds:

  1. (1)

    DnD_{n} and Dˇn\check{D}_{n} each have the same distribution with d(o,Zυ(ω)o)d(o,Z_{\operatorname{\upsilon}}(\operatorname{\omega})o) and d(Zυˇ(ωˇ)o)d(Z_{\check{\operatorname{\upsilon}}}(\check{\operatorname{\omega}})o) for (ωˇ,ω)(G>0,μˇ>0)×(G>0,μ>0)(\check{\operatorname{\omega}},\operatorname{\omega})\in(G^{\operatorname{\mathbb{Z}}_{>0}},\check{\mu}^{\operatorname{\mathbb{Z}}_{>0}})\times(G^{\operatorname{\mathbb{Z}}_{>0}},\mu^{\operatorname{\mathbb{Z}}_{>0}});

  2. (2)

    d(o,Zno)τ(Zn)Dˇn+Dnd(o,Z_{n}o)-\tau(Z_{n})\leq\check{D}_{n}+D_{n} holds outside a set of exponentially decaying probability (in nn).

Proof.

We follow the proof of [Cho22c, Theorem B]. Let K0,D0,D1,E0K_{0},D_{0},D_{1},E_{0} be the constants associated to the Schottky set SS.

First, we prepare two bi-infinite sequences (gi)iZ(g_{i})_{i\in Z} and (hi)i(h_{i})_{i\in\operatorname{\mathbb{Z}}} of independent RVs, all distributed according to μ\mu. Following our standard convention, ZiZ_{i} always denotes g1gig_{1}\cdots g_{i}. We now fix nn, and define

gi;0:={gii=1,,n/2hii>n/2,\displaystyle g_{i;0}:=\left\{\begin{array}[]{cc}g_{i}&i=1,\ldots,\lfloor n/2\rfloor\\ h_{i}&i>\lfloor n/2\rfloor,\end{array}\right. gˇi;0:={gni+11i=1,,nn/2hi1i>nn/2,\displaystyle\check{g}_{i;0}:=\left\{\begin{array}[]{cc}g_{n-i+1}^{-1}&i=1,\ldots,n-\lfloor n/2\rfloor\\ h_{-i}^{-1}&i>n-\lfloor n/2\rfloor,\end{array}\right.
gi;1:={gn/2+ii=1,,nn/2hii>nn/2,\displaystyle g_{i;1}:=\left\{\begin{array}[]{cc}g_{\lfloor n/2\rfloor+i}&i=1,\ldots,n-\lfloor n/2\rfloor\\ h_{i}&i>n-\lfloor n/2\rfloor,\end{array}\right. gˇi;1:={gn/2i+11i=1,,n/2hi1i>n/2.\displaystyle\check{g}_{i;1}:=\left\{\begin{array}[]{cc}g_{\lfloor n/2\rfloor-i+1}^{-1}&i=1,\ldots,\lfloor n/2\rfloor\\ h_{-i}^{-1}&i>\lfloor n/2\rfloor.\end{array}\right.

Then ((gˇi;t)i>0,(gi;t)i>0)\big{(}(\check{g}_{i;t})_{i>0},(g_{i;t})_{i>0}\big{)} is distributed according to μˇ>0×μ>0\check{\mu}^{\operatorname{\mathbb{Z}}_{>0}}\times\mu^{\operatorname{\mathbb{Z}}_{>0}} for t=0,1t=0,1. Using them, we similarly defined other RVs such as

Zi;t\displaystyle Z_{i;t} :=g1;tgi;t,\displaystyle:=g_{1;t}\cdots g_{i;t}, Zˇi;t\displaystyle\check{Z}_{i;t} :=gˇ1;tgˇi;t,\displaystyle:=\check{g}_{1;t}\cdots\check{g}_{i;t},
υt\displaystyle\operatorname{\upsilon}_{t} :=υ((gˇi;t)i>0,(gi;t)i>0),\displaystyle:=\operatorname{\upsilon}\big{(}(\check{g}_{i;t})_{i>0},(g_{i;t})_{i>0}\big{)}, υˇt\displaystyle\check{\operatorname{\upsilon}}_{t} :=υˇ((gˇi;t)i>0,(gi;t)i>0).\displaystyle:=\check{\operatorname{\upsilon}}\big{(}(\check{g}_{i;t})_{i>0},(g_{i;t})_{i>0}\big{)}.

Finally, we will set Dn:=d(o,Zυ0;0o)D_{n}:=d(o,Z_{\operatorname{\upsilon}_{0};0}o) and Dˇn:=d(Zˇυˇ0;0o,o)\check{D}_{n}:=d(\check{Z}_{\check{\operatorname{\upsilon}}_{0};0}o,o). Since ((gˇi;0)i>0,(gi;0)i>0)\big{(}(\check{g}_{i;0})_{i>0},(g_{i;0})_{i>0}\big{)} is distributed according to μˇ>0×μ>0\check{\mu}^{\operatorname{\mathbb{Z}}_{>0}}\times\mu^{\operatorname{\mathbb{Z}}_{>0}}, we can verify Item (1) of the conclusion.

We then define

An:={max{υ0,υˇ0,υ1,υˇ1}n/10}.A_{n}:=\Big{\{}\max\{\operatorname{\upsilon}_{0},\check{\operatorname{\upsilon}}_{0},\operatorname{\upsilon}_{1},\check{\operatorname{\upsilon}}_{1}\}\geq n/10\Big{\}}.

Since ((gˇi;0)i>0,(gi;0)i>0)\big{(}(\check{g}_{i;0})_{i>0},(g_{i;0})_{i>0}\big{)} and ((gˇi;1)i>0,(gi;1)i>0)\big{(}(\check{g}_{i;1})_{i>0},(g_{i;1})_{i>0}\big{)} are both distributed according to μˇ>0×μ>0\check{\mu}^{\operatorname{\mathbb{Z}}_{>0}}\times\mu^{\operatorname{\mathbb{Z}}_{>0}}, Lemma 3.8 implies that (An)\operatorname{\mathbb{P}}(A_{n}) decays exponentially in nn.

In the proof of [Cho22c, Claim 3.3], using the alignment lemma (Proposition 2.20 in the current setting), we showed the existence of 0i,jn/20\leq i,j\leq n/2 such that:

  1. (1)

    γ:=(ZiM0o,,Zio)\gamma:=(Z_{i-M_{0}}o,\ldots,Z_{i}o) and γˇ:=(Znjo,,Znj+M0o)\check{\gamma}:=(Z_{n-j}o,\ldots,Z_{n-j+M_{0}}o) are Schottky axes;

  2. (2)

    on AnA_{n}, (o,γ,Zυ0;0o)(o,\,\gamma,Z_{\operatorname{\upsilon}_{0};0}\,o) is D1D_{1}-aligned;

  3. (3)

    on AnA_{n}, (Zˇυˇ0;0o,Zn1γˇ,o)(\check{Z}_{\check{\operatorname{\upsilon}}_{0};0}o,\,Z_{n}^{-1}\check{\gamma},\,o) is D1D_{1}-aligned;

  4. (4)

    on AnA_{n}, for each k>0k>0, there exist points p0,q0,,pk1,qk1p_{0},q_{0},\ldots,p_{k-1},q_{k-1} on [o,Znko][o,Z_{n}^{k}o], from left to right, so that

    dsym(pi,ZniZiM0o)<0.1E0,dsym(qi,ZniZnj+M0o)<0.1E0.d^{sym}(p_{i},Z_{n}^{i}\cdot Z_{i-M_{0}}o)<0.1E_{0},\quad d^{sym}(q_{i},Z_{n}^{i}\cdot Z_{n-j+M_{0}}o)<0.1E_{0}.

Then we have

(3.8) d(pi,qi)\displaystyle d(p_{i},q_{i}) d(ZiM0o,Znj+M0o)0.2E0\displaystyle\geq d(Z_{i-M_{0}}o,Z_{n-j+M_{0}}o)-0.2E_{0}
(3.9) d(o,Zno)d(o,ZiM0o)d(Zn1Znj+M0o,o)0.2E0,\displaystyle\geq d(o,Z_{n}o)-d(o,Z_{i-M_{0}}o)-d(Z_{n}^{-1}Z_{n-j+M_{0}}o,o)-0.2E_{0},

and we deduce

d(o,Znko)i=1kd(pi1,qi1)k(d(o,Zno)d(o,ZiM0o)d(Zn1Znj+M0o,o)0.2E0).d(o,Z_{n}^{k}o)\geq\sum_{i=1}^{k}d(p_{i-1},q_{i-1})\geq k\cdot\left(d(o,Z_{n}o)-d(o,Z_{i-M_{0}}o)-d(Z_{n}^{-1}Z_{n-j+M_{0}}o,o)-0.2E_{0}\right).

By dividing by kk and taking the limit, we deduce that

d(o,Zno)τ(Zn)d(o,ZiM0o)+d(Zn1Znj+M0o,o)+0.2E0.d(o,Z_{n}o)-\tau(Z_{n})\leq d(o,Z_{i-M_{0}}o)+d(Z_{n}^{-1}Z_{n-j+M_{0}}o,o)+0.2E_{0}.

Meanwhile, the alignments of (o,γ,Zυ0;0o)(o,\,\gamma,Z_{\operatorname{\upsilon}_{0};0}\,o) and (Zˇυˇ0;0o,Zn1γˇ,o)(\check{Z}_{\check{\operatorname{\upsilon}}_{0};0}o,\,Z_{n}^{-1}\check{\gamma},\,o) imply that d(o,ZiM0o)+0.1E0d(o,Z_{i-M_{0}}o)+0.1E_{0} is smaller than Dn:=d(o,Zυ0;0o)D_{n}:=d(o,Z_{\operatorname{\upsilon}_{0};0}o), and that d(Zn1Znj+M0o,o)+0.1E0d(Z_{n}^{-1}Z_{n-j+M_{0}}o,o)+0.1E_{0} is smaller than Dˇn:=d(Zˇυˇ0;0o,o)\check{D}_{n}:=d(\check{Z}_{\check{\operatorname{\upsilon}}_{0};0}o,o). This concludes Item (2). ∎

Given this lemma, it follows that

(d(o,Zno)τ(Zn)ϵn1/p)(Anc)+(Dn+Dˇnϵn1/p)\operatorname{\mathbb{P}}(d(o,Z_{n}o)-\tau(Z_{n})\geq\epsilon n^{1/p})\leq\operatorname{\mathbb{P}}(A_{n}^{c})+\operatorname{\mathbb{P}}(D_{n}+\check{D}_{n}\geq\epsilon n^{1/p})

for every ϵ>0\epsilon>0. Since Dn+DˇnD_{n}+\check{D}_{n} has the same distribution with d(Zυˇ(ωˇ)o,o)+d(o,Zυ(ω)o)d(Z_{\check{\operatorname{\upsilon}}}(\check{\operatorname{\omega}})o,o)+d(o,Z_{\operatorname{\upsilon}}(\operatorname{\omega})o), which has finite pp-th moment (cf. Proposition 3.11), (Dn+Dˇnϵn1/p)\operatorname{\mathbb{P}}(D_{n}+\check{D}_{n}\geq\epsilon n^{1/p}) is summable. Using Borel-Cantelli lemma, we conclude:

Theorem 3.19.

Let (X,G,o)(X,G,o) be as in Convention 1.1, and (Zn)n>0(Z_{n})_{n>0} be the random walk generated by a non-elementary measure μ\mu on GG. If μ\mu has finite pp-th moment for some p>0p>0, then

limn1n1/p[d(o,Zno)τ(Zn)]=0a.s.\lim_{n\rightarrow\infty}\frac{1}{n^{1/p}}[d(o,Z_{n}o)-\tau(Z_{n})]=0\quad\textrm{a.s.}

We note again that if μ\mu and μˇ\check{\mu} both have finite pp-th moment, the proof of [Cho22b, Theorem B] works up to changing d(,)d(\cdot,\cdot) with dsym(,)d^{sym}(\cdot,\cdot) and give the stronger result (o(n1/2p)o(n^{1/2p})-tracking, and logarithmic tracking when p=1p=1).

Although we have weaker result than [Cho22b, Theorem B], Theorem 3.19 is sufficient to deduce the corollaries. Namely, combined with the SLLN and CLT for displacement (Theorem 3.16 and 3.13), Theorem 3.19 implies:

Corollary 3.20 ([Cho22c, Corollary 3.7]).

Let (X,G,o)(X,G,o) be as in Convention 1.1, and let (Zn)n0(Z_{n})_{n\geq 0} be the random walk generated by a non-elementary measure μ\mu on GG with finite first moment. Then

(3.10) limn1nτ(Zn)=λ\lim_{n}\frac{1}{n}\tau(Z_{n})=\lambda

holds almost surely, where λ=λ(μ)\lambda=\lambda(\mu) is the escape rate of μ\mu.

Corollary 3.21 ([Cho22c, Corollary 3.8]).

Let (X,G,o)(X,G,o) be as in Convention 1.1, and let (Zn)n0(Z_{n})_{n\geq 0} be the random walk generated by a non-elementary measure μ\mu on GG. If μ\mu has finite second moment, then there exists σ(μ)0\sigma(\mu)\geq 0 such that 1n(τ(Zn)nλ)\frac{1}{\sqrt{n}}(\tau(Z_{n})-n\lambda) and 1n(d(o,Zno)nλ)\frac{1}{\sqrt{n}}(d(o,Z_{n}o)-n\lambda) converge to the same Gaussian distribution 𝒩(0,σ(μ)2)\mathscr{N}(0,\sigma(\mu)^{2}) in law. We also have

lim supnτ(Zn)λn2nloglogn=σ(μ)almost surely.\limsup_{n\rightarrow\infty}\frac{\tau(Z_{n})-\lambda n}{\sqrt{2n\log\log n}}=\sigma(\mu)\quad\textrm{almost surely}.

Meanwhile, the exponential bound in [Cho22c, Theorem A] was established based on Inequality 3.8 and the exponential bound for displacement (Theorem 3.16). Moreover, we know that exponentially generic isometries have BGIP, this time using Lemma 2.21 instead of [Cho22c, Lemma 2.7]. Considering these, the proof of [Cho22c, Theorem A] yields the following:

Theorem 3.22.

Let (X,G,o)(X,G,o) be as in Convention 1.1 and let (Zn)n0(Z_{n})_{n\geq 0} be the random walk generated by a non-elementary measure μ\mu on GG. Let λ(μ)\lambda(\mu) be the escape rate of μ\mu, i.e., λ(μ):=limn𝔼μn[d(o,go)]/n\lambda(\mu):=\lim_{n\rightarrow\infty}\operatorname{\mathbb{E}}_{\mu^{\ast n}}[d(o,go)]/n. Then for each 0<L<λ(μ)0<L<\lambda(\mu), there exists K>0K>0 such that for each nn we have

(Zn has BGIP and τ(Zn)Ln)1Ken/K.\operatorname{\mathbb{P}}\Big{(}\textrm{$Z_{n}$ has BGIP and $\tau(Z_{n})\geq Ln$}\Big{)}\geq 1-Ke^{-n/K}.

In the second part of [Cho22c], we discussed the pivotal time construction following [Gou22, Subsection 5.A]. [Cho22c, Section 4.1] describes pivotal times in a discrete model, which applies to the current setting verbatim. We record one definition from [Cho22c, Section 4.1].

Definition 3.23 ([Cho22c, Defnition 4.2]).

Let (X,G,o)(X,G,o) be as in Convention 1.1. Given a fairly long Schottky set SS, we define

S~:={(β,γ,v)S2×G:(Γ(β),Π(β)vΠ(γ)o) and (v1o,Γ(γ)) are K0-aligned}.\tilde{S}:=\big{\{}(\beta,\gamma,v)\in S^{2}\times G:\textrm{$\big{(}\Gamma(\beta),\Pi(\beta)v\Pi(\gamma)o\big{)}$ and $\big{(}v^{-1}o,\Gamma(\gamma)\big{)}$ are $K_{0}$-aligned}\big{\}}.

In [Cho22c, Section 4.2], we made several reductions for random walks. In [Cho22c, Section 4.3], we discussed the pivoting for translation length. Thanks to the existence of Schottky set for non-elementary probability measures, (Proposition 2.27), the property of Schottky sets (Definition 2.23) and alignment lemma (Lemma 2.18, we can bring these to the current setting of Convention 1.1.

For latter use, let us record one lemma from [Cho22c, Section 4.2].

Lemma 3.24 (cf. [Cho22c, Lemma 4.6]).

Let (X,G,o)(X,G,o) be as in Convention 1.1, let μ\mu be a non-elementary probability measure on GG, let μ\mu^{\prime} be a probability measure dominated by a multiple of a convolution of μ\mu (i.e., there exists α,k>0\alpha,k>0 such that μ(g)αμk(g)\mu^{\prime}(g)\leq\alpha\mu^{\ast k}(g) for all gGg\in G) and let SGM0S\subseteq G^{M_{0}} be a fairly long K0K_{0}-Schottky set for μ\mu. Then for each nn there exist an integer m(n)m(n), a probability space Ωn\Omega_{n}, a measurable subset BnΩnB_{n}\subseteq\Omega_{n}, a measurable partition 𝒬n\mathcal{Q}_{n} of BnB_{n}, and random variables

Z\displaystyle Z G,\displaystyle\in G,
{wi,i=0,,m(n)}\displaystyle\{w_{i},i=0,\ldots,m(n)\} Gm(n)+1,\displaystyle\in G^{m(n)+1},
{vi:i=1,,m(n)}\displaystyle\{v_{i}:i=1,\ldots,m(n)\} Gm(n),\displaystyle\in G^{m(n)},
{αi,βi,γi,δi:i=1,,m(n)}\displaystyle\{\alpha_{i},\beta_{i},\gamma_{i},\delta_{i}:i=1,\ldots,m(n)\} S4m(n)\displaystyle\in S^{4m(n)}

such that the following hold:

  1. (1)

    limn+(Bn)=1\lim_{n\rightarrow+\infty}\operatorname{\mathbb{P}}(B_{n})=1 and limn+m(n)/n>0\lim_{n\rightarrow+\infty}m(n)/n>0.

  2. (2)

    On each equivalence class 𝒬n\mathcal{F}\in\mathcal{Q}_{n}, (wi)i=0m(n)(w_{i})_{i=0}^{m(n)} are constant and (αi,βi,γi,δi,vi)i=1m(n)(\alpha_{i},\beta_{i},\gamma_{i},\delta_{i},v_{i})_{i=1}^{m(n)} are i.i.d.s distributed according to (uniform measure on S4)×μ(\textrm{uniform measure on $S^{4}$})\times\mu^{\prime}.

  3. (3)

    ZZ is distributed according to μn\mu^{\ast n} on Ωn\Omega_{n} and

    Z=w0Π(α1)Π(β1)v1Π(γ1)Π(δ1)w1Π(αm(n))Π(βm(n))vm(n)Π(γm(n))Π(δm(n))wm(n)Z=w_{0}\Pi(\alpha_{1})\Pi(\beta_{1})v_{1}\Pi(\gamma_{1})\Pi(\delta_{1})w_{1}\cdots\Pi(\alpha_{m(n)})\Pi(\beta_{m(n)})v_{m(n)}\Pi(\gamma_{m(n)})\Pi(\delta_{m(n)})w_{m(n)}

    holds on AnA_{n}.

In [Cho22c, Lemma 4.6], the lemma was stated for the case μ=μ\mu^{\prime}=\mu. By employing the decomposition

μ2M0×μk×μ2M0=p(μS2×μk×μS2)+(1p)ν\mu^{2M_{0}}\times\mu^{\ast k}\times\mu^{2M_{0}}=p(\mu_{S}^{2}\times\mu^{\ast k}\times\mu_{S}^{2})+(1-p)\nu

for a suitable choice of 0<p<10<p<1 and (nonnegative) probability measure ν\nu, the proof of [Cho22c, Lemma 4.6] leads to the current general version.

In [Cho22c, Section 4.4], we made the first application of the discussion above, which is the converse of CLT. For future use, we record a slight modification of [Cho22c, Corollary 4.13].

Definition 3.25 ([Cho22c, Definition 4.7]).

Let (X,G,o)(X,G,o) be as in Convention 1.1, let n>0n>0 and K>0K>0, and let SS be a fairly long Schottky set. (This determines S~\tilde{S} following Definition 3.23.) We say that a sequence (wi)i=0n(w_{i})_{i=0}^{n} in GG is KK-pre-aligned if, for the isometries W0:=w0W_{0}:=w_{0} and

Vk:=WkΓ(βk+1)vk+1,Wk+1:=VkΓ(γk+1)wk+1(k=0,,n1),V_{k}:=W_{k}\Gamma(\beta_{k+1})v_{k+1},\,\,W_{k+1}:=V_{k}\Gamma(\gamma_{k+1})w_{k+1}\quad(k=0,\ldots,n-1),

the sequence

(o,W0Γ(β1),V0Γ(γ1),,Wn1Γ(βn),Vn1Γ(γn),Wno)\Big{(}o,\,W_{0}\Gamma(\beta_{1}),\,V_{0}\Gamma(\gamma_{1}),\,\ldots,\,W_{n-1}\Gamma(\beta_{n}),\,V_{n-1}\Gamma(\gamma_{n}),\,W_{n}o\Big{)}

is KK-semi-aligned for any choices of (βi,γi,vi)S~(\beta_{i},\gamma_{i},v_{i})\in\tilde{S} (i=1,,ni=1,\ldots,n).

We say that an isometry ϕG\phi\in G is KK-pre-aligned if

(Γ(γ),Π(γ)ϕΓ(β))\Big{(}\Gamma(\gamma^{\prime}),\,\,\Pi(\gamma^{\prime})\phi\cdot\Gamma(\beta)\Big{)}

is KK-semi-aligned for any choices of (β,γ,v),(β,γ,v)S~(\beta,\gamma,v),(\beta^{\prime},\gamma^{\prime},v^{\prime})\in\tilde{S}.

Corollary 3.26 (cf. [Cho22c, Corollary 4.13]).

Let (X,G,o)(X,G,o) be as in Convention 1.1, let μ\mu be a non-elementary probability measure on GG, let μ\mu^{\prime} be a probability measure dominated by a multiple of a convolution of μ\mu, and let SS be a fairly long K0K_{0}-Schottky set for μ\mu with cardinality at least 400400. Then for each N>0N>0, for each nn, there exist m(n){2s:s>0}m(n)\in\{2^{s}:s>0\}, a probability space Ωn\Omega_{n} with a measurable subset AnΩnA_{n}\subseteq\Omega_{n}, a measurable partition 𝒫n\mathcal{P}_{n} of AnA_{n} and RVs

Zn\displaystyle Z_{n} G,\displaystyle\in G,
{wi:i=0,,m(n)}\displaystyle\{w_{i}:i=0,\ldots,m(n)\} Gm(n)+1,\displaystyle\in G^{m(n)+1},
{vi:i=1,,m(n)}\displaystyle\{v_{i}:i=1,\ldots,m(n)\} Gm(n),\displaystyle\in G^{m(n)},
{βi,γi:i=1,,m(n)}\displaystyle\{\beta_{i},\gamma_{i}:i=1,\ldots,m(n)\} S2m(n)\displaystyle\in S^{2m(n)}

such that the following hold:

  1. (1)

    limn+(An)=1\lim_{n\rightarrow+\infty}\operatorname{\mathbb{P}}(A_{n})=1 and N<m(n)<2NN<m(n)<2N eventually.

  2. (2)

    On AnA_{n}, (wi)i=0m(n)(w_{i})_{i=0}^{m(n)} is a D0D_{0}-pre-aligned sequence in GG and wm(n)1w0w_{m(n)}^{-1}w_{0} is a D0D_{0}-pre-aligned isometry.

  3. (3)

    On each equivalence class 𝒫n\mathcal{E}\in\mathcal{P}_{n}, (wi)i=0m(n)(w_{i})_{i=0}^{m(n)} are constant and (βi,γi,vi)i=1m(n)(\beta_{i},\gamma_{i},v_{i})_{i=1}^{m(n)} are i.i.d.s distributed according to the measure (uniform measure on S2)×μ\left(\textrm{uniform measure on $S^{2}$}\right)\times\mu^{\prime} conditioned on S~\tilde{S}.

  4. (4)

    ZnZ_{n} is distributed according to μn\mu^{\ast n} on Ωn\Omega_{n} and

    Zn=w0Π(β1)v1Π(γ1)w1Π(βm(n))vm(n)Π(γm(n))wm(n)Z_{n}=w_{0}\Pi(\beta_{1})v_{1}\Pi(\gamma_{1})w_{1}\cdots\Pi(\beta_{m(n)})v_{m(n)}\Pi(\gamma_{m(n)})w_{m(n)}

    holds on AnA_{n}.

The difference between [Cho22c, Corollary 4.13] and Corollary 3.26 is that:

  1. (1)

    μ\mu^{\prime} is a measure dominated by some μM\mu^{\ast M^{\prime}} instead of μ=μ\mu^{\prime}=\mu;

  2. (2)

    m(n)m(n) eventually lies in [N,2N][N,2N], rather than growing linearly.

The first item was addressed in Lemma 3.24. Next, the original proof of [Cho22c, Corollary 4.13] (which depends on [Cho22c, Lemma 4.6]) can in fact handle the second item, thanks to the following fact: if we have a decomposition

μS2M0×μM×μS2M0=p(μS2×μ×μS2)+(1p)ν(ν:probability measure)\mu_{S}^{2M_{0}}\times\mu^{\ast M^{\prime}}\times\mu_{S}^{2M_{0}}=p(\mu_{S}^{2}\times\mu^{\prime}\times\mu_{S}^{2})+(1-p)\nu\quad(\nu:\textrm{probability measure})

for some p=ϵ>0p=\epsilon>0, then the decomposition of the same form is possible for all 0<p<ϵ0<p<\epsilon.

As an example, we can plug in μ=atom at g\mu^{\prime}=\textrm{atom at $g$} for gsuppμg\in\operatorname{\langle\langle}\operatorname{supp}\mu\operatorname{\rangle\rangle}. Combining this with Proposition 2.20, we obtain the following.

Proposition 3.27.

Let (X,G,o)(X,G,o) be as in Convention 1.1. Given a non-elementary probability measure μ\mu on GG, there exists K>0K>0 such that the following holds.

For each gsuppμg\in\operatorname{\langle\langle}\operatorname{supp}\mu\operatorname{\rangle\rangle}, there exists ϵ>0\epsilon>0 such that the following holds outside a set of exponentially decaying probability. For a random path (Zi)i=1(Z_{i})_{i=1}^{\infty}, there exist 0<i(1)<<i(ϵn)<n0<i(1)<\ldots<i(\epsilon n)<n such that Zi(1)o,Zi(1)gMoZ_{i(1)}o,Z_{i(1)}g^{M}o, \ldots, Zi(ϵn)o,Zi(ϵn)gMoZ_{i(\epsilon n)}o,Z_{i(\epsilon n)}g^{M}o are KK-dsymd^{sym}-close to points p1,,p2ϵnp_{1},\ldots,p_{2\epsilon n} on [o,Zno][o,Z_{n}o], respectively, such that d(o,pi)d(o,pi+1o)d(o,p_{i})\leq d(o,p_{i+1}o) for each ii.

The discussion so far was about securing the alignment. We now need to couple the alignment and (almost) additivity of the displacement along sample path. Thanks to Proposition 2.20, we can generalize such a coupling to the setting of Convention 1.1 as well.

Lemma 3.28.

Let (X,G,o)(X,G,o) be as in Convention 1.1, let SS be a large and fairly long K0K_{0}-Schottky set for μ\mu, and fix a D0D_{0}-pre-aligned sequence (wi)i=0n(w_{i})_{i=0}^{n} in GG such that wn1w0w_{n}^{-1}w_{0} is a D0D_{0}-pre-aligned isometry. Given (βi,γi,vi)S~(\beta_{i},\gamma_{i},v_{i})\in\tilde{S} for i=1,,ni=1,\ldots,n, define W0:=w0W_{0}:=w_{0} and

Vk\displaystyle V_{k} :=WkΓ(βk+1)vk+1,Wk+1:=VkΓ(γk+1)wk+1\displaystyle:=W_{k}\Gamma(\beta_{k+1})v_{k+1},\,\,W_{k+1}:=V_{k}\Gamma(\gamma_{k+1})w_{k+1} (k=0,,n1),\displaystyle(k=0,\ldots,n-1),
(x2k1,x2k)\displaystyle(x_{2k-1},x_{2k}) :=(Wk1o,Vk1Π(γk)o)\displaystyle:=\big{(}W_{k-1}o,\,\,V_{k-1}\Pi(\gamma_{k})o\big{)} (k=1,,n),\displaystyle(k=1,\ldots,n),
x0\displaystyle x_{0} :=0,x2n+1:=Wno.\displaystyle:=0,\,\,x_{2n+1}:=W_{n}o.

Let μ\mu^{\prime} be a probability measure on GG and let (βi,γi,vi)i=1n(\beta_{i},\gamma_{i},v_{i})_{i=1}^{n} be i.i.d.s distributed according to (uniform measure on S2)×μ(\textrm{uniform measure on $S^{2}$})\times\mu^{\prime}. Then the following hold:

  1. (1)

    d(x2l1,x2l)d(x_{2l-1},x_{2l}) and d(o,vlo)d(o,v_{l}o) differ by at most 2max{d(o,Π(α)o):αS}2\max\{d(o,\Pi(\alpha)o):\alpha\in S\};

  2. (2)

    {d(x2l,x2l+1):l=0,,n}\{d(x_{2l},x_{2l+1}):l=0,\ldots,n\} are constant RVs;

  3. (3)

    for each 0ijk2n+10\leq i\leq j\leq k\leq 2n+1, (xi,xk)xj(x_{i},x_{k})_{x_{j}} is bounded by E0E_{0};

  4. (4)

    for each 0ijkijk2n+10\leq i\leq j\leq k\leq i^{\prime}\leq j^{\prime}\leq k^{\prime}\leq 2n+1, (xi,xk)xj(x_{i},x_{k})_{x_{j}} and (xi,xk)xj(x_{i^{\prime}},x_{k^{\prime}})_{x_{j^{\prime}}} are independent;

  5. (5)

    the translation length τ(Wn)\tau(W_{n}) and d(x0,x2n)d(x_{0},x_{2n}) differ by at most E0E_{0}.

Combining Corollary 3.26 and Lemma 3.28, we obtained the following converse of CLT:

Proposition 3.29 ([Cho22c, Proposition 4.16]).

Let (X,G,o)(X,G,o) be as in Convention 1.1 and let (Zn)n0(Z_{n})_{n\geq 0} be the random walk on GG generated by a non-elementary measure μ\mu with infinite second moment. Then for any sequence (cn)n(c_{n})_{n} of real numbers, neither of 1n(d(o,Zno)cn)\frac{1}{\sqrt{n}}(d(o,Z_{n}o)-c_{n}) and 1n(τ(Zn)cn)\frac{1}{\sqrt{n}}(\tau(Z_{n})-c_{n}) converges in law.

The remaining part of [Cho22c] works verbatim, and we obtain the following results.

Theorem 3.30.

Let (X,G,o)(X,G,o) be as in Convention 1.1, and (Zn(1),,Zn(k))n0(Z_{n}^{(1)},\ldots,Z_{n}^{(k)})_{n\geq 0} be kk independent random walks generated by a non-elementary measure μ\mu on GG. Then there exists K>0K>0 such that

[Zn(1),,Zn(k)is q.i. embedded into a quasi-convex subset of X]1Ken/K.\operatorname{\mathbb{P}}\left[\langle Z_{n}^{(1)},\ldots,Z_{n}^{(k)}\rangle\,\,\textrm{is q.i. embedded into a quasi-convex subset of $X$}\right]\geq 1-Ke^{-n/K}.
Theorem 3.31.

Let GG be a finitely generated group acting on an asymmetric metric space XX with at least two independent BGIP elements. Then for each k>0k>0, there exists a finite generating set SS of GG such that

#{(g1,,gk)(BS(n))k:g1,,gkis q.i. embedded intoa quasi-convex subset of X}(#BS(n))k\frac{\#\left\{(g_{1},\ldots,g_{k})\in\big{(}B_{S}(n)\big{)}^{k}:\begin{array}[]{c}\langle g_{1},\ldots,g_{k}\rangle\,\,\textrm{is q.i. embedded into}\\ \textrm{a quasi-convex subset of $X$}\end{array}\right\}}{\big{(}\#B_{S}(n)\big{)}^{k}}

converges to 1 exponentially fast.

3.3. Proof of Theorem A

We first formulate a variation of Proposition 2.27.

Lemma 3.32.

Let (X,G,o)(X,G,o) be as in Convention 1.1 and let μ\mu be a non-elementary probability measure on GG. Then for each N,L>0N,L>0, there exist K=K(N)>0K=K(N)>0, n>Ln>L and a KK-Schottky set SS of cardinality NN in (suppμ)n(\operatorname{supp}\mu)^{n} satisfying the following: for every gGg\in G, there exist α,αS\alpha,\alpha^{\prime}\in S such that

(3.11) (Γ(β),Π(β)(Π(α)gΠ(α))Π(γ)o)and((Π(α)gΠ(α))1o,Γ(γ))are K-aligned(β,γS).\Big{(}\Gamma(\beta),\Pi(\beta)\cdot\big{(}\Pi(\alpha)g\Pi(\alpha^{\prime})\big{)}\cdot\Pi(\gamma)o\big{)}\,\,\textrm{and}\,\,\Big{(}\big{(}\Pi(\alpha)g\Pi(\alpha^{\prime})\big{)}^{-1}o,\Gamma(\gamma)\Big{)}\,\,\textrm{are $K$-aligned}\quad(\forall\beta,\gamma\in S).
Proof.

Let N,L>0N,L>0 be given. Without loss of generality, we can assume N>10N>10. We take K0=K(N)K_{0}=K(N) be as in Proposition 2.27, D0=D(K0,K0)D_{0}=D(K_{0},K_{0}) be as in Lemma 2.18 and K=E(D0,K0)K=E(D_{0},K_{0}), L=L(D0,K0)L^{\prime}=L(D_{0},K_{0}) be as in Proposition 2.20. Note that KK only depends on NN and not on LL.

Now, by Proposition 2.27, there exists a K0K_{0}-Schottky set SS of cardinality NN in (suppμ)n(\operatorname{supp}\mu)^{n} for some n>L+L+2K02n>L+L^{\prime}+2K_{0}^{2}. We claim that SS satisfies the desired property. First, the inequality K0<D0<KK_{0}<D_{0}<K tells us that SS is automatically a KK-Schottky set. It remains to check the condition in Display 3.11.

For this, let gGg\in G. Note that #S>10\#S>10. By the K0K_{0}-Schottky property (2) of SS, there exist αS\alpha^{\prime}\in S such that (g1o,Γ(α))(g^{-1}o,\Gamma(\alpha^{\prime})) is K0K_{0}-aligned. Fixing such an α\alpha^{\prime}, again the K0K_{0}-Schottky property (2) of SS guarantees the existence of αS\alpha\in S such that (Γ(α),Π(α)go)(\Gamma(\alpha),\Pi(\alpha)go) is K0K_{0}-aligned. Then by Lemma 2.18, (Γ(α),Π(α)gΓ(α))\big{(}\Gamma(\alpha),\Pi(\alpha)g\Gamma(\alpha^{\prime})\big{)} is D0D_{0}-aligned.

Next, note that (Π(α)o,Γ(α))(\Pi(\alpha)o,\Gamma(\alpha)) is not K0K_{0}-aligned, as d(o,πΠ(α)o(Γ(α)))=d(o,Π(α)o)n/K0K0>K0d(o,\pi_{\Pi(\alpha)o}(\Gamma(\alpha)))=d(o,\Pi(\alpha)o)\geq n/K_{0}-K_{0}>K_{0}. By Schottky property (2), we conclude that (Γ(β),Π(β)Π(α)o)(\Gamma(\beta),\Pi(\beta)\Pi(\alpha)o) is K0K_{0}-aligned for every βS{α}\beta\in S\setminus\{\alpha\}, or equivalently, (Γ(β),Π(β)Π(α)o)(\Gamma(\beta),\Pi(\beta)\Pi(\alpha)o) is K0K_{0}-aligned. In this case, since (Π(β)o,Π(β)Γ(α)o)(\Pi(\beta)o,\Pi(\beta)\Gamma(\alpha)o) is 0-aligned, we conclude that (Γ(β),Π(β)Γ(α))(\Gamma(\beta),\Pi(\beta)\Gamma(\alpha)) is D0D_{0}-aligned by Lemma 2.18. Meanwhile, Schottky property (3) guarantees that (Γ(α),Π(α)Γ(α))(\Gamma(\alpha),\Pi(\alpha)\Gamma(\alpha)) is K0K_{0}-aligned as well.

For a similar reason, (Γ(α),Π(α)Γ(γ))(\Gamma(\alpha^{\prime}),\Pi(\alpha^{\prime})\Gamma(\gamma)) is D0D_{0}-aligned for every γS\gamma\in S. In conclusion, the following sequence is D0D_{0}-aligned for every β,γS\beta,\gamma\in S:

(Γ(β),Π(β)Γ(α),Π(β)Π(α)gΓ(α),Π(β)Π(α)gΠ(α)Γ(γ)).\big{(}\Gamma(\beta),\,\Pi(\beta)\Gamma(\alpha),\,\Pi(\beta)\Pi(\alpha)g\Gamma(\alpha^{\prime}),\,\Pi(\beta)\Pi(\alpha)g\Pi(\alpha^{\prime})\Gamma(\gamma)\big{)}.

Also, the involved K0K_{0}-Schottky axes have domains longer than LL^{\prime}. We then apply Proposition 2.20 to conclude that

(Π(β)o,Π(β)Π(α)gΠ(α)Γ(γ)),(Γ(β),Π(β)Π(α)gΠ(α)Π(γ)o)\big{(}\Pi(\beta)o,\,\Pi(\beta)\Pi(\alpha)g\Pi(\alpha^{\prime})\Gamma(\gamma)\big{)},\quad\big{(}\Gamma(\beta),\,\Pi(\beta)\Pi(\alpha)g\Pi(\alpha^{\prime})\Pi(\gamma)o\big{)}

are KK-aligned as desired. ∎

We are now ready to prove Theorem A.

Proof.

Let μ\mu be a non-elementary and asymptotically asymmetric measure on GG. We take K0=K(400)K_{0}=K(400) as in Lemma 3.32. Lemma 3.32 guarantees that there exists a fairly long K0K_{0}-Schottky set S(suppμ)M0S\subseteq(\operatorname{supp}\mu)^{M_{0}} with cardinality at least 400, with the property that for every gGg\in G, there exist α,αS\alpha,\alpha^{\prime}\in S such that Display 3.11 is satisfied. This choice of SS is fixed throughout.

We now pick a large enough positive integer nn^{\prime}. Since μ\mu is asymptotically asymmetric, we can take ϕ,φsuppμM\phi,\varphi\in\operatorname{supp}\mu^{\ast M^{\prime}} for some MM^{\prime} such that

[τ(ϕ)τ(ϕ1)][τ(φ)τ(φ1)]>0.\left[\tau(\phi)-\tau(\phi^{-1})\right]-\left[\tau(\varphi)-\tau(\varphi^{-1})\right]>0.

By replacing ϕ\phi and φ\varphi with their suitable powers, we may assume that

(3.12) (d(o,ϕo)d(o,ϕ1o))(d(o,φo)d(o,φ1o))(6E0+12K0M0+2)n.\big{(}d(o,\phi o)-d(o,\phi^{-1}o)\big{)}-\big{(}d(o,\varphi o)-d(o,\varphi^{-1}o)\big{)}\geq(6E_{0}+12K_{0}M_{0}+2)n^{\prime}.

Now let α,αS\alpha,\alpha^{\prime}\in S be as in Display 3.11 for g=ϕg=\phi, and define ϕnew=Π(α)ϕΠ(α)\phi_{new}=\Pi(\alpha)\phi\Pi(\alpha^{\prime}). Then ϕnew\phi_{new} belongs to (suppμ(M+2M0))(\operatorname{supp}\mu^{\ast(M^{\prime}+2M_{0})}), and (β,γ,ϕnew)S~(\beta,\gamma,\phi_{new})\in\tilde{S} holds for all β,γS\beta,\gamma\in S. Note also that

(3.13) d(o,ϕo)d(o,ϕnewo)d(o,Π(α)o)+d(o,Π(α))o)2M0K0,\displaystyle d(o,\phi o)-d(o,\phi_{new}o)\leq d(o,\Pi(\alpha)o)+d(o,\Pi(\alpha^{\prime}))o)\leq 2M_{0}K_{0},
d(ϕo,o)d(ϕnewo,o)d(Π(α)o,o)+d(Π(α))o,o)2M0K0.\displaystyle d(\phi o,o)-d(\phi_{new}o,o)\leq d(\Pi(\alpha)o,o)+d(\Pi(\alpha^{\prime}))o,o)\leq 2M_{0}K_{0}.

and similarly d(ϕo,o)d(\phi o,o) and d(ϕnewo,o)d(\phi_{new}o,o) differ by 2M0K02M_{0}K_{0}. We can similarly take φnew(suppμ(M+2M0))\varphi_{new}\in(\operatorname{supp}\mu^{\ast(M^{\prime}+2M_{0})}) such that (β,γ,φnew)S~(\beta,\gamma,\varphi_{new})\in\tilde{S} for all β,γS\beta,\gamma\in S, and so that

(3.14) |d(o,φo)d(o,φnewo)|2M0K0,|d(φo,o)d(φnewo,o)|2M0K0.\big{|}d(o,\varphi o)-d(o,\varphi_{new}o)\big{|}\leq 2M_{0}K_{0},\quad\big{|}d(\varphi o,o)-d(\varphi_{new}o,o)\big{|}\leq 2M_{0}K_{0}.

For convenience, we introduce the notation

L1±:=d(o,ϕnew±1o),L2±:=d(o,φnew±1o)L_{1}^{\pm}:=d(o,\phi_{new}^{\pm 1}o),\quad L_{2}^{\pm}:=d(o,\varphi_{new}^{\pm 1}o)

Combining Inequality 3.12, 3.13 and 3.14, we conclude

(3.15) (d(o,ϕnewo)d(o,ϕnew1o))(d(o,φnewo)d(o,φnew1o))\displaystyle\big{(}d(o,\phi_{new}o)-d(o,\phi_{new}^{-1}o)\big{)}-\big{(}d(o,\varphi_{new}o)-d(o,\varphi_{new}^{-1}o)\big{)} (6E0+12K0M0+2)n8M0K0\displaystyle\geq(6E_{0}+12K_{0}M_{0}+2)n^{\prime}-8M_{0}K_{0}
(6E0+4K0M0+2)n.\displaystyle\geq(6E_{0}+4K_{0}M_{0}+2)n^{\prime}.

Let μ\mu^{\prime} be the measure assigning 1/21/2 to each of ϕnew\phi_{new} and φnew\varphi_{new}. Then μ\mu^{\prime} is dominated by a multiple of μ(M+2M0)\mu^{\ast(M^{\prime}+2M_{0})}. We now apply Corollary 3.26: for each nn, we obtain an integer m(n)m(n), a probability space Ωn\Omega_{n} with AnΩnA_{n}\subseteq\Omega_{n}, and a measurable partition 𝒫n\mathcal{P}_{n} of AnA_{n}, and RVs

Zn\displaystyle Z_{n} G,\displaystyle\in G,
{wi:i=0,,m(n)}\displaystyle\{w_{i}:i=0,\ldots,m(n)\} Gm(n)+1,\displaystyle\in G^{m(n)+1},
{vi:i=1,,m(n)}\displaystyle\{v_{i}:i=1,\ldots,m(n)\} Gm(n),\displaystyle\in G^{m(n)},
{βi,γi:i=1,,m(n)}\displaystyle\{\beta_{i},\gamma_{i}:i=1,\ldots,m(n)\} S2m(n)\displaystyle\in S^{2m(n)}

such that the following hold:

  1. (1)

    limn+(An)=1\lim_{n\rightarrow+\infty}\operatorname{\mathbb{P}}(A_{n})=1 and m(n)[n,2n]m(n)\in[n^{\prime},2n^{\prime}] eventually;

  2. (2)

    On AnA_{n}, (wi)i=0m(n)(w_{i})_{i=0}^{m(n)} is a D0D_{0}-pre-aligned sequence in GG and wm(n)1w0w_{m(n)}^{-1}w_{0} is a D0D_{0}-pre-aligned isometry.

  3. (3)

    On each equivalence class 𝒫n\mathcal{E}\in\mathcal{P}_{n}, (wi)i=0m(n)(w_{i})_{i=0}^{m(n)} are constant and (βi,γi,vi)i=1m(n)(\beta_{i},\gamma_{i},v_{i})_{i=1}^{m(n)} are i.i.d.s distributed according to the measure (uniform measure on S2)×μ\left(\textrm{uniform measure on $S^{2}$}\right)\times\mu^{\prime} conditioned on S~\tilde{S}.

  4. (4)

    ZnZ_{n} is distributed according to μn\mu^{\ast n} on Ωn\Omega_{n} and

    Zn=w0Π(β1)v1Π(γ1)w1Π(βm(n))vm(n)Π(γm(n))wm(n)Z_{n}=w_{0}\Pi(\beta_{1})v_{1}\Pi(\gamma_{1})w_{1}\cdots\Pi(\beta_{m(n)})v_{m(n)}\Pi(\gamma_{m(n)})w_{m(n)}

    holds on AnA_{n}.

Recall that (β,γ,g)S~(\beta,\gamma,g)\in\tilde{S} for any β,γS\beta,\gamma\in S and g{φnew,ϕnew}g\in\{\varphi_{new},\phi_{new}\}. Hence, the support of (uniform measure on S2)×μ(\textrm{uniform measure on $S^{2}$})\times\mu^{\prime} restricted to S~\tilde{S} is S2×{φnew,ϕnew}S^{2}\times\{\varphi_{new},\phi_{new}\}, and (βi,γi,vi)i=1m(n)(\beta_{i},\gamma_{i},v_{i})_{i=1}^{m(n)} has the law (uniform measure on S2)×μ\big{(}\textrm{uniform measure on $S^{2}$}\big{)}\times\mu^{\prime}.

Now, let us pick an equivalence class 𝒫n\mathcal{E}\in\mathcal{P}_{n}. Conditioned on \mathcal{E}, Lemma 3.28 guarantees the existence of points o=:x0,x1,,x2m(n),x2m(n)+1:=Znoo=:x_{0},x_{1},\ldots,x_{2m(n)},x_{2m(n)+1}:=Z_{n}o such that the following hold:

  1. (1)

    d(x2l1,x2l)d(x_{2l-1},x_{2l}) and d(o,vlo)d(o,v_{l}o) differ by at most 2K0M02K_{0}M_{0} for l=1,,m(n)l=1,\ldots,m(n);

  2. (2)

    {d(x2l,x2l+1):l=0,,n}\{d(x_{2l},x_{2l+1}):l=0,\ldots,n\} are constant RVs;

  3. (3)

    for each 0ijk2n+10\leq i\leq j\leq k\leq 2n+1, (xi,xk)xj(x_{i},x_{k})_{x_{j}} is bounded by E0E_{0};

  4. (4)

    the translation length τ(Wn)\tau(W_{n}) and d(x0,x2n)d(x_{0},x_{2n}) differ by at most E0E_{0}.

In view of Observation 2.5 and Observation 2.16(2), we also have that:

  1. (1)

    d(x2l,x2l1)d(x_{2l},x_{2l-1}) and d(vlo,o)d(v_{l}o,o) differ by at most 2K0M02K_{0}M_{0} for l=1,,m(n)l=1,\ldots,m(n);

  2. (2)

    {d(x2l+1,x2l):l=0,,n}\{d(x_{2l+1},x_{2l}):l=0,\ldots,n\} are constant RVs;

  3. (3)

    for each 0ijk2n+10\leq i\leq j\leq k\leq 2n+1, (xk,xi)xj(x_{k},x_{i})_{x_{j}} is bounded by E0E_{0};

  4. (4)

    the translation length τ(Wn)\tau(W_{n}) and d(x2n,x0)d(x_{2n},x_{0}) differ by at most E0E_{0}.

Now, for each ω\operatorname{\omega}\in\mathcal{E} we define

D+\displaystyle D^{+} :=l=0nd(x2l,x2l+1),\displaystyle:=\sum_{l=0}^{n}d(x_{2l},x_{2l+1}), D\displaystyle D^{-} :=l=0nd(x2l+1,x2l),\displaystyle:=\sum_{l=0}^{n}d(x_{2l+1},x_{2l}),
N(ω)\displaystyle N(\operatorname{\omega}) :=#{i=1,,m(n):vi=ϕnew}.\displaystyle:=\#\{i=1,\ldots,m(n):v_{i}=\phi_{new}\}.

Note that D+D^{+} and DD^{-} are constant on \mathcal{E}. Moreover, we have

d(x0(ω),x2m(n)(ω))\displaystyle d\big{(}x_{0}(\operatorname{\omega}),x_{2m(n)}(\operatorname{\omega})\big{)} :=l=12m(n)d(xl1,xl)+l=12m(n)1d(x0,xl+1)xl\displaystyle:=\sum_{l=1}^{2m(n)}d(x_{l-1},x_{l})+\sum_{l=1}^{2m(n)-1}d(x_{0},x_{l+1})_{x_{l}}
(D++l=1m(n)d(x2l1,x2l)2m(n)E0,D++l=1m(n)d(x2l1,x2l)+2m(n)E0).\displaystyle\in\left(D^{+}+\sum_{l=1}^{m(n)}d(x_{2l-1},x_{2l})-2m(n)\cdot E_{0},D^{+}+\sum_{l=1}^{m(n)}d(x_{2l-1},x_{2l})+2m(n)\cdot E_{0}\right).

Also, recall that d(o,vlo)d(o,v_{l}o) and d(x2l1,x2l)d(x_{2l-1},x_{2l}) differ by at most 2K0M02K_{0}M_{0}, and that

l=1m(n)d(o,vlo)=NL1++(m(n)N)L2+\displaystyle\sum_{l=1}^{m(n)}d(o,v_{l}o)=NL_{1}^{+}+(m(n)-N)L_{2}^{+}

Finally, d(x0(ω),x2m(n)(ω))d\big{(}x_{0}(\operatorname{\omega}),x_{2m(n)}(\operatorname{\omega})\big{)} and τ(Zn(ω))\tau(Z_{n}(\operatorname{\omega})) differ by at most E0E_{0}. Combining these, we conclude that

|τ(Zn(ω))(NL1++(m(n)N)L2+)|2m(n)E0+2K0M0m(n)+E0.\left|\tau(Z_{n}(\operatorname{\omega}))-\big{(}NL_{1}^{+}+(m(n)-N)L_{2}^{+}\big{)}\right|\leq 2m(n)\cdot E_{0}+2K_{0}M_{0}m(n)+E_{0}.

Similarly, using the equality

l=1m(n)d(vlo,o)=NL1+(m(n)N)L2,\sum_{l=1}^{m(n)}d(v_{l}o,o)=NL_{1}^{-}+(m(n)-N)L_{2}^{-},

we conclude that

|τ(Zn1(ω))(NL1+(m(n)N)L2)|2m(n)E0+2K0M0m(n)+E0.\left|\tau(Z_{n}^{-1}(\operatorname{\omega}))-\big{(}NL_{1}^{-}+(m(n)-N)L_{2}^{-}\big{)}\right|\leq 2m(n)\cdot E_{0}+2K_{0}M_{0}m(n)+E_{0}.

This implies that τ(Zn(ω))τ(Zn1(ω))\tau(Z_{n}(\operatorname{\omega}))-\tau(Z_{n}^{-1}(\operatorname{\omega})) falls in to the interval INI_{N} of width (6E0+4K0M0)n(6E_{0}+4K_{0}M_{0})n^{\prime}, centered at cN:=N(L1+L1)+(m(n)N)(L2+L2)c_{N}:=N(L_{1}^{+}-L_{1}^{-})+(m(n)-N)(L_{2}^{+}-L_{2}^{-}). Note that

cN+1cN=(L1+L1)(L2+L2)>(6E0+4K0M0+2)n.c_{N+1}-c_{N}=(L_{1}^{+}-L_{1}^{-})-(L_{2}^{+}-L_{2}^{-})>(6E_{0}+4K_{0}M_{0}+2)n^{\prime}.

Hence, the intervals {IN}N=0m(n)\{I_{N}\}_{N=0}^{m(n)} are disjoint and 2n2n^{\prime}-separated from each other. This implies that only at most one of them can intersect the interval [n,n][-n^{\prime},n^{\prime}].

Meanwhile, note that

maxk(N(ω)=k|ω)\displaystyle\max_{k}\operatorname{\mathbb{P}}\left(N(\operatorname{\omega})=k\,\big{|}\,\operatorname{\omega}\in\mathcal{E}\right) =maxk(l=1m(n)Bl=k)({Bi}i=1m(n) are i.i.d. Bernoulli RVs with mean 1/2)\displaystyle=\max_{k}\operatorname{\mathbb{P}}\left(\sum_{l=1}^{m(n)}B_{l}=k\right)\quad\left(\begin{array}[]{c}\textrm{$\{B_{i}\}_{i=1}^{m(n)}$ are i.i.d. }\\ \textrm{Bernoulli RVs with mean $1/2$}\end{array}\right)
=maxk2m(n)(m(n)k)=O(1/m(n))=O(1/n).\displaystyle=\max_{k}2^{-m(n)}\binom{m(n)}{k}=O\left(1/\sqrt{m(n)}\right)=O\left(1/\sqrt{n^{\prime}}\right).

From this, we obtain

(|τ(Zn(ω))τ(Zn1(ω))|n|ω)=O(1/n).\operatorname{\mathbb{P}}\left(|\tau(Z_{n}(\operatorname{\omega}))-\tau(Z_{n}^{-1}(\operatorname{\omega}))|\leq n^{\prime}\,\big{|}\,\operatorname{\omega}\in\mathcal{E}\right)=O\left(1/\sqrt{n^{\prime}}\right).

Since AnA_{n} is partitioned with such an equivalence class \mathcal{E} and takes up most of Ωn\Omega_{n} asymptotically, we conclude that

lim infn(|τ(Zn(ω))τ(Zn1(ω))|n)=O(1/n).\liminf_{n\rightarrow\infty}\operatorname{\mathbb{P}}\left(|\tau(Z_{n}(\operatorname{\omega}))-\tau(Z_{n}^{-1}(\operatorname{\omega}))|\leq n^{\prime}\right)=O\left(1/\sqrt{n^{\prime}}\right).

We now send nn^{\prime} to infinity to conclude. ∎

4. Outer space

In this section, we collect facts about the outer automorphism group and Outer space. For detailed definitions and theories, we refer the readers to the general exposition by Vogtmann [Vog15] or individual papers, e.g. [BH92], [FM11], [FM12], [AKB12], [AK11], [DH18], [DT18] and [KMPT22].

Let XX be the Culler-Vogtmann Outer space CVNCV_{N} of rank N3N\geq 3, which is the space of unit-volume marked metric graphs with fundamental group FNF_{N}. In other words, a point pCVNp\in CV_{N} corresponds to the homotopic class of a homotopy equivalence h:RNΓh:R_{N}\rightarrow\Gamma, where RNR_{N} is a fixed rose with NN petals and Γ\Gamma is a unit-volume metric graph. The corresponding space without the volume normalization is called the unprojectivized Outer space cvNcv_{N}, and there is a projectivization from cvNcv_{N} to CVNCV_{N} by dilation.

Outer space comes equipped with a canonical metric, the Lipschitz distance, which is defined as follows: for two markings h1:RNΓ1h_{1}:R_{N}\rightarrow\Gamma_{1} and h2:RNΓ2h_{2}:R_{N}\rightarrow\Gamma_{2}, the distance from Γ1\Gamma_{1} to Γ2\Gamma_{2} is defined by

dCV(Γ1,Γ2):=inf{logLip(f):ff2f11},d_{CV}(\Gamma_{1},\Gamma_{2}):=\inf\{\log\operatorname{Lip}(f):f\sim f_{2}\circ f_{1}^{-1}\},

where Lip(f)\operatorname{Lip}(f) is the (maximal) Lipschitz constant of ff. We now make a convention that differs from the traditional one. Namely, the outer automorphism group Out(FN)\operatorname{Out}(F_{N}) of rank NN acts on CVNCV_{N} by changing the basis of the marking with the inverses: given ϕOut(FN)\phi\in\operatorname{Out}(F_{N}) and h:RNΓh:R_{N}\rightarrow\Gamma representing a point of CVNCV_{N}, ϕ\phi moves hh to hϕ1:FNϕ1FNhΓh\circ\phi^{-1}:F_{N}\stackrel{{\scriptstyle\phi^{-1}}}{{\rightarrow}}F_{N}\stackrel{{\scriptstyle h}}{{\rightarrow}}\Gamma. This is a left action by isometries. We denote action by XhϕhXX\ni h\mapsto\phi\cdot h\in X.

It is known that the Lipschitz distance is asymmetric [FM11] and not uniquely geodesic. However, distances among ϵ\epsilon-thick points (i.e., those with systole at least ϵ\epsilon) have the coarse symmetry: there exists a constant C=C(ϵ)<+C=C(\epsilon)<+\infty such that for any ϵ\epsilon-thick points xx and yy, one has d(x,y)Cd(y,x)d(x,y)\leq Cd(y,x) [AKB12]. In particular, distances among the translates of the reference point oo by Out(FN)\operatorname{Out}(F_{N}) satisfy the coarse symmetry.

Just as Teichmüller space 𝒯(Σ)\operatorname{\mathcal{T}}(\Sigma) is accompanied by the curve complex 𝒞(Σ)\mathcal{C}(\Sigma) and the coarse projection π𝒞:𝒯(Σ)𝒞(Σ)\pi^{\mathcal{C}}:\operatorname{\mathcal{T}}(\Sigma)\rightarrow\mathcal{C}(\Sigma), CVNCV_{N} is accompanied by the complex of free factors N\mathcal{FF}_{N} and the coarse projection π:CVNN\pi^{\mathcal{FF}}:CV_{N}\rightarrow\mathcal{FF}_{N}. This projection is coarsely Out(FN)\operatorname{Out}(F_{N})-equivariant and coarsely Lipschitz. Moreover, geodesics in CVNCV_{N} projects to KK-unparametrized bi-quasigeodesics for some uniform K>0K>0 [BF14, Proposition 9.2].

Outer space also accomodates lots of BGIP isometries. We say that an outer automorphism ϕOut(FN)\phi\in\operatorname{Out}(F_{N}) is reducible if there exists a free product decomposition FN=C1CkCk+1F_{N}=C_{1}\ast\cdots\ast C_{k}\ast C_{k+1}, with k1k\geq 1 and Ci{e}C_{i}\neq\{e\}, such that ϕ\phi permutes the conjugacy classes of C1,,CkC_{1},\ldots,C_{k}. If not, we say that ϕ\phi is irreducible. We also say that ϕ\phi is fully ireducible (or iwip) if no power of ϕ\phi is reducible, or equivalently, no power of ϕ\phi preserves the conjugacy class of any proper free factor of FNF_{N}. We also say that ϕ\phi is atoroidal (or hyperbolic) if no power of ϕ\phi fixes any nontrivial conjugacy class in FNF_{N}. When ϕ\phi is fully irreducible, it is non-atoroidal if and only if it is geometric, i.e., induced by a pseudo-Anosov φ:ΣΣ\varphi:\Sigma\rightarrow\Sigma on a compact surface Σ\Sigma with one boundary component, via an identification of FNF_{N} with π(Σ)\pi(\Sigma). Bestvina and Feighn proved in [BF14] that ϕOut(FN)\phi\in\operatorname{Out}(F_{N}) is fully irreducible if and only if it acts on N\mathcal{FF}_{N} loxodromically.

We say that a subgroup GOut(FN)G\leq Out(F_{N}) is non-elementary if it acts on N\mathcal{FF}_{N} in a non-elementary way, or equivalently, contains two fully irreducibles with mutually distinct attracting/repelling trees. It is known that if GOut(FN)G\leq Out(F_{N}) does not fix any finite subset of NN\mathcal{FF}_{N}\cup\partial\mathcal{FF}_{N}, or equivalently, if it is not virtually cyclic nor virtually fixes the conjugacy class of a proper free factor of FNF_{N}, then GG is non-elementary [Hor16b]. Since π\pi^{\mathcal{FF}} is coarsely Lipschitz, the independence of two fully irreducibles in N\mathcal{FF}_{N} is lifted to the independence in CVNCV_{N}.

We refer the readers to [BH92], [AK10], [BF14] and [AKKP19] for the precise definition of a train-track representative of an outer automorphism. Roughly speaking, a train-track representative of ϕ\phi is a self-map f:ΓΓf:\Gamma\rightarrow\Gamma in the free homotopy class of ϕ\phi on a simplicial graph Γ\Gamma that sends vertices to vertices, restricts to an immersion on each edge of Γ\Gamma, and sends edges to immersed segments after iterations. It is due to Bestvina and Handel [BH92] that every irreducible outer automorphism admits a train-track representative, although it may not be unique.

Given such a structure, one can endow Γ\Gamma with a metric such that ff stretches each edge of Γ\Gamma by the same constant λ>1\lambda>1, which is called the expansion factor of ff. This expansion factor is uniquely determined by the choice of ϕ\phi and does not depend on the choice of ff. Moreover, in view of Skora’s interpretation of Stallings fold decompositions, one obtains a continuous path on cvNcv_{N} from Γ\Gamma to Γϕ\Gamma\circ\phi by folding a single illegal turn at each time (cf. [AKKP19]). This descends to a geodesic segment of length logλ\log\lambda (after a reparametrization) and the concatenation of its translates by powers of ϕ\phi becomes a bi-infinite, ϕ\phi-periodic geodesic. We call this a (optimal) folding axis of ϕ\phi. Algom-Kfir observed the following:

Theorem 4.1 ([AK11]).

Folding axes of fully irreducible outer automorphisms are strongly contracting.

Meanwhile, we need BGIP instead of the strongly contracting property in our setting, and the author does not know a way to promote the latter to the former. Meanwhile, I. Kapovich, Maher, Pfaff and Taylor observed the following version of BGIP in Outer space. This requires the notion of greedy folding path, whose accurate definition can be found in [FM11], [BF14] and [DH18]. In short, a greedy folding path γ:IcvN\gamma:I\rightarrow cv_{N} is obtained by folding every illegal turns at each time with speed 1, where the illegal turn structures at different forward times are identical and define a well-defined illegal turn structure. This descends to a geodesic on CVNCV_{N}, and we have the following theorem:

Theorem 4.2 ([KMPT22, Theorem 7.8]).

Let ϕOut(FN)\phi\in\operatorname{Out}(F_{N}) be a fully irreducible outer automorphism. Suppose that γ\gamma is a bi-infinite, ϕ\phi-periodic greedy folding path. Then there exist C>0C>0 such that the following holds.

Let x,yXx,y\in X be points such that dsym(πγ(x),πγ(y))Cd^{sym}(\pi_{\gamma}(x),\pi_{\gamma}(y))\geq C, and satisfy dsym(πγ(x))=γ(t1)d^{sym}(\pi_{\gamma}(x))=\gamma(t_{1}), dsym(πγ(y))=γ(t2)d^{sym}(\pi_{\gamma}(y))=\gamma(t_{2}) for some t1<t2t_{1}<t_{2}. Then any geodesic [x,y][x,y] between them contains a subsegment [z1,z2][z_{1},z_{2}] such that

dsym(z1,πγ(x))<C,dsym(z2,πγ(y))<C.d^{sym}(z_{1},\pi_{\gamma}(x))<C,\quad d^{sym}(z_{2},\pi_{\gamma}(y))<C.

This uni-directional version of BGIP is designed for outer automorphisms that have an invariant(=periodic) greedy folding line. It seems not shown that all fully irreducibles have such a line. (The author thanks Sam Taylor for pointing this out.) Nonetheless, by adapting Dowdall-Taylor’s idea and Kapovich-Maher-Pfaff-Taylor’s proof of Theorem 4.2, we can obtain the following result. This proof was kindly informed by Sam Taylor.

See 1.3

Proof.

Before we begin, we recall the following facts regarding a geodesic δ\delta-hyperbolic space YY.

  1. (1)

    (Morse property) A KK-quasigeodesic and a geodesic with the same endpoints are within Hausdorff distance K2=K2(K,δ)K_{2}=K_{2}(K,\delta).

  2. (2)

    The closest point projections of a point yYy\in Y onto a KK-quasigeodesic and a geodesic on YY with the same endpoints are within distance K3=K3(K,δ)K_{3}=K_{3}(K,\delta).

  3. (3)

    If the projections of x,yYx,y\in Y onto a KK-quasigeodesic γ\gamma contain γ(s)\gamma(s) and γ(t)\gamma(t), respectively, and if d(γ(s),γ(t))>K4=K4(K,δ)d(\gamma(s),\gamma(t))>K_{4}=K_{4}(K,\delta), then [x,y][x,y] and [x,γ(s)]γ|[s,t][γ(t),y][x,\gamma(s)]\cup\gamma|_{[s,t]}\cup[\gamma(t),y] are within Hausdoff distance K4K_{4}.

  4. (4)

    If two KK-quasigeodesics γ\gamma, γ\gamma^{\prime} are within Hausdorff distant KK and the distance between starting points is at most KK, then γ\gamma^{\prime} crosses γ\gamma up to a constant K5=K5(K,δ)K_{5}=K_{5}(K,\delta), i.e., γ\gamma and γρ\gamma^{\prime}\circ\rho are K5K_{5}-fellow traveling for some non-decreasing reparametrization ρ\rho.

Let oCVNo\in CV_{N} be an arbitrary basepoint. Let T+T^{+}, TT^{-} be the attracting and repelling trees of φ\varphi, respectively. There exist optimal greedy folding lines γ±:CVN\gamma^{\pm}:\mathbb{R}\rightarrow CV_{N} such that

(4.1) limt+γ±(t)=T±,limtγ±(t)=T\lim_{t\rightarrow+\infty}\gamma^{\pm}(t)=T^{\pm},\quad\lim_{t\rightarrow-\infty}\gamma^{\pm}(t)=T^{\mp}

([BR15], Lemma 6.7 and Lemma 7.3). Since {φio}i\{\varphi_{i}o\}_{i} is a quasigeodesic whose endpoints agree with γ+\gamma^{+}, Theorem 4.1 of [DT18] asserts that dH({φio}i,γ+)<K1d_{H}(\{\varphi^{i}o\}_{i},\gamma^{+})<K_{1} and π(γ+)\pi^{\mathcal{FF}}(\gamma^{+}) is a K1K_{1}-quasigeodesic for some K1K_{1}. Similarly, by comparing {φio}i\{\varphi_{-i}o\}_{i} and γ\gamma^{-}, we deduce that dH({φio}i,γ)<K1d_{H}(\{\varphi^{i}o\}_{i},\gamma^{-})<K_{1} and π(γ)\pi^{\mathcal{FF}}(\gamma^{-}) is a K1K_{1}-quasigeodesic. Also, γ±\gamma^{\pm} are uniformly thick.

Recall our notation: πγ±\pi_{\gamma^{\pm}} denotes the nearest point projection onto γ±\gamma^{\pm}. For now, we denote by πγ+sym(x)\pi_{\gamma^{+}}^{sym}(x) the nearest point projection of xCVNx\in CV_{N} onto γ+\gamma^{+} with respect to dsymd^{sym}. Similarly, we denote by πγsym()\pi_{\gamma^{-}}^{sym}(\cdot) the dsymd^{sym}-nearest point projection onto γ\gamma^{-}.

Let us now take xi+πγ+sym(φio)x_{i}^{+}\in\pi_{\gamma^{+}}^{sym}(\varphi^{i}o) and xiπγsym(φio)x_{i}^{-}\in\pi_{\gamma^{-}}^{sym}(\varphi^{i}o) for each ii. We recall the following result of Dahmani and Horbez ([DH18, Proposition 5.17, Corollary 5.22]; see also Section 7 of [KMPT22]): there exist B,D>0B,D>0 such that γ±\gamma^{\pm} are (B,D)(B,D)-contracting at xi±x_{i}^{\pm}’s (with a suitable crossing constant κ\kappa). In other words, a geodesic η\eta on CVNCV_{N} has the projection πη\pi^{\mathcal{FF}}\circ\eta that κ\kappa-crosses up the BB-long subsegment of πγ±\pi^{\mathcal{FF}}\gamma^{\pm} that begins from π(xi±)\pi^{\mathcal{FF}}(x_{i}^{\pm}), then η\eta has a point pp such that d(p,xi±)Dd(p,x_{i}^{\pm})\leq D. Since γ±\gamma^{\pm} are thick, d(xi±,p)d(x_{i}^{\pm},p) is (uniformly) proportional to d(p,xi±)d(p,x_{i}^{\pm}) and η\eta intersects a uniform dsymd^{sym}-neighborhood of γ±\gamma^{\pm} in such a case.

We now observe that ππγ+,ππγ\pi^{\mathcal{FF}}\pi_{\gamma^{+}},\pi^{\mathcal{FF}}\pi_{\gamma^{-}} and ππ({φio}i)π\pi_{\pi^{\mathcal{FF}}(\{\varphi^{i}o\}_{i})}\circ\pi^{\mathcal{FF}} are coarsely equivalent. First, [DT18, Lemma 4.11] asserts that πγ±\pi_{\gamma^{\pm}} and Prγ±Pr_{\gamma^{\pm}} are equivalent, where PrPr stands for the Bestvina-Feighn left projection. Then [DT18, Lemma 4.2] asserts that πPrγ±\pi^{\mathcal{FF}}Pr_{\gamma^{\pm}} and ππ(γ±)π\pi_{\pi^{\mathcal{FF}}(\gamma^{\pm})}\circ\pi^{\mathcal{FF}} are equivalent. These are then equivalent to ππ({φio}i)π\pi_{\pi^{\mathcal{FF}}(\{\varphi^{i}o\}_{i})}\circ\pi^{\mathcal{FF}}, since π(γ±)\pi^{\mathcal{FF}}(\gamma^{\pm}) and π({φio}i)\pi^{\mathcal{FF}}(\{\varphi^{i}o\}_{i}) are equivalent quasi-geodesics on the Gromov hyperbolic space \mathcal{FF}.

We now lift these projections: we claim that πγ+\pi_{\gamma^{+}}, πγ\pi_{\gamma^{-}} and π{φio}i\pi_{\{\varphi^{i}o\}_{i}} are equivalent. First, suppose that πγ+(x)\pi_{\gamma^{+}}(x) and πγ(x)\pi_{\gamma^{-}}(x) are far from each other for some xXx\in X. Since γ+\gamma^{+}, γ\gamma^{-}, {φio}i\{\varphi^{i}o\}_{i} are close to each other, we may take φio\varphi^{i}o and φjo\varphi^{j}o near πγ+(x)\pi_{\gamma^{+}}(x) and πγ(x)\pi_{\gamma^{-}}(x), respectively, and conclude that |ij||i-j| is large. This implies that π(φio)\pi^{\mathcal{FF}}(\varphi^{i}o) and π(φjo)\pi^{\mathcal{FF}}(\varphi^{j}o) are also far from each other (since φ\varphi is loxodromic on CVNCV_{N}), and consequently π(πγ+(x))\pi^{\mathcal{FF}}(\pi_{\gamma^{+}}(x)), π(πγ(x))\pi^{\mathcal{FF}}(\pi_{\gamma^{-}}(x)) are far from each other. (\ast) Since ππγ+\pi^{\mathcal{FF}}\pi_{\gamma^{+}} and ππγ\pi^{\mathcal{FF}}\pi_{\gamma^{-}} are equivalent, this cannot happen. For this reason, πγ+\pi_{\gamma^{+}} and πγ\pi_{\gamma^{-}} are equivalent.

Now suppose that π{φio}i(x)\pi_{\{\varphi^{i}o\}_{i}}(x) and πγ±(x)\pi_{\gamma^{\pm}}(x) are far from each other for some xXx\in X. We take φjoπ{φio}i(x)\varphi^{j}o\in\pi_{\{\varphi^{i}o\}_{i}}(x) and φjo\varphi^{j^{\prime}}o near πγ±(x)\pi_{\gamma^{\pm}}(x) and conclude that |jj||j^{\prime}-j| is large. If jjj\gg j^{\prime}, then π([x,φjo])\pi^{\mathcal{FF}}([x,\varphi^{j}o]) is a quasigeodesic whose endpoints project onto π({φio}i)\pi^{\mathcal{FF}}(\{\varphi^{i}o\}_{i}) near πφjo\pi^{\mathcal{FF}}\varphi^{j^{\prime}}o and πφjo\pi^{\mathcal{FF}}\varphi^{j}o, respectively. Since jjj^{\prime}-j is large enough, this quasigeodesic crosses up long enough subsegments of π({φio}i)\pi^{\mathcal{FF}}(\{\varphi^{i}o\}_{i}) and π(γ+)\pi^{\mathcal{FF}}(\gamma^{+}) that begin at π(φjo)\pi^{\mathcal{FF}}(\varphi^{j^{\prime}}o) and π(xj)\pi^{\mathcal{FF}}(x_{j^{\prime}}), respectively. Using the (B,D)(B,D)-contraction at xj+x_{j^{\prime}}^{+} of γ+\gamma^{+}, we conclude that [x,φjo][x,\varphi^{j}o] contains a point pp nearby xj+x_{j^{\prime}}^{+}, which makes d(x,φjo)d(x,\varphi^{j^{\prime}}o) shorter than d(x,φjo)d(x,\varphi^{j}o) and leads to a contradiction. Similar contradiction occurs when jjj^{\prime}\gg j, due to the contracting property of γ\gamma^{-} at xix_{i}^{-}’s. Hence, π{φio}i(x)\pi_{\{\varphi^{i}o\}_{i}}(x) and πγ±(x)\pi_{\gamma^{\pm}}(x) are equivalent.

Now, if a geodesic η\eta on CVNCV_{N} has a large nearest point projection on {φio}i\{\varphi^{i}o\}_{i}, then it also has large projections on γ±\gamma^{\pm}. This leads to large π(πγ±(η))\pi^{\mathcal{FF}}(\pi_{\gamma^{\pm}}(\eta)), because π\pi^{\mathcal{FF}} restricted on γ±\gamma^{\pm} is a QI-embedding (again due to [DT18, Theorem 4.1]). When π(πγ±(η))\pi^{\mathcal{FF}}(\pi_{\gamma^{\pm}}(\eta)) progresses in the forward direction with respect to {φio}i\{\varphi_{i}o\}_{i}, then we employ the contracting property of γ+\gamma^{+} to conclude that dsym(η,γ+)d^{sym}(\eta,\gamma^{+}) is uniformly bounded. If it progresses in the backward direction, then we employ the contracting property of γ\gamma^{-} to conclude. ∎

We now explain more details on the classification of fully irreducible outer automorphisms. Coulbois and Hilion classified fully irreducibles into the following mutually distinct categories in [CH12]:

  1. (1)

    geometric automorphisms that have geometric attracting and repelling trees,

  2. (2)

    parageometric automorphisms that have geometric attracting tree and non-geometric attracting tree,

  3. (3)

    inverses of parageometric automorphisms that have geometric repelling tree and non-geometric attracting tree, and

  4. (4)

    pseudo-Levitt automorphisms that have non-geometric attracting and repelling trees.

Sometimes, automorphisms of category (3) and (4) are together called ageometric automorphisms.

In Theorem A we are concerned with fully irreducibles whose expansion factors differ from that of their inverses. Examples of such automorphisms are given in [HM07b]: every parageometric fully irreducible automorphism has expansion factor larger than the expansion factor of its inverse. Meanwhile, every geometric fully irreducible automorphism and its inverse have the same expansion factor, due to an analogous fact for pseudo-Anosov mapping classes. For pseudo-Levitt automorphisms, both situations can happen ([HM07b], [CH12]).

As mentioned before, fully irreducibles in Out(F2)\operatorname{Out}(F_{2}) are always geometric. In contrast, [JL08] provides an example of parageometric automorphism for each N3N\geq 3. Hence, given an admissible probability measure μ\mu on Out(FN)\operatorname{Out}(F_{N}) for N3N\geq 3, there exists n>0n>0 such that suppμn\operatorname{supp}\mu^{\ast n} contains a parageometric automorphism φ\varphi, while there exists m>0m>0 such that suppμm\operatorname{supp}\mu^{\ast m} contains identity. Then suppμmn\operatorname{supp}\mu^{\ast mn} contains φm\varphi^{m} and idid, where

τ(φm)τ(φm)>0=τ(id)τ(id).\tau(\varphi^{m})-\tau(\varphi^{-m})>0=\tau(id)-\tau(id).

For this reason, every admissible probability measure on Out(FN)\operatorname{Out}(F_{N}) is asymptotically asymmetric.

When N3N\geq 3, if a non-elementary random walk (Zn)n>0(Z_{n})_{n>0} on Out(FN)\operatorname{Out}(F_{N}) has bounded support that generates a semigroup containing a principal fully irreducible and an inverse of a principal fully irreducible, then the probability that ZnZ_{n} is pseudo-Levitt tends to 1 as nn\rightarrow\infty [KMPT22, Theorem A]. Hence, the above property of parageometric automorphisms does not imply that a generic automorphism and its inverse have different expansion factors.

5. Limit laws for Outer space

By applying the limit laws in Section 3 to the outer automorphism group (with Theorem 1.3 in hand), we obtain the following results. We first recover the escape to infinity of random walks on Out(FN)\operatorname{Out}(F_{N}) due to Horbez [Hor16a], and weaken the moment condition the SLLN for translation length in [DH18, Theorem 1.4].

Theorem 5.1.

Let (X,d)(X,d) be the Culler-Vogtmann Outer space CVNCV_{N} of rank N3N\geq 3 with the Lipschitz metric, and let (Zn)n(Z_{n})_{n} be the random walk generated by a non-elementary probability measure μ\mu on Out(FN)\operatorname{Out}(F_{N}). Then there exists a strictly positive quantity λ(μ)(0,+]\lambda(\mu)\in(0,+\infty], called the drift of μ\mu, such that

λ(μ):=limn1nd(o,Zno)almost surely.\lambda(\mu):=\lim_{n\rightarrow\infty}\frac{1}{n}d(o,Z_{n}o)\quad\textrm{almost surely.}

Moreover, for each 0<L<λ(μ)0<L<\lambda(\mu), there exists K>0K>0 such that

(Zn has BGIP and τ(Zn)Ln)1Ken/K.\operatorname{\mathbb{P}}\Big{(}\textrm{$Z_{n}$ has BGIP and $\tau(Z_{n})\geq Ln$}\Big{)}\geq 1-Ke^{-n/K}.

Next, we recover Horbez’s CLT on Out(FN)\operatorname{Out}(F_{N}) [Hor18] and establish its converse:

Theorem 5.2.

Let (X,d)(X,d) be the Culler-Vogtmann Outer space CVNCV_{N} of rank N3N\geq 3 with the Lipschitz metric, and let (Zn)n(Z_{n})_{n} be the random walk generated by a non-elementary probability measure μ\mu on Out(FN)\operatorname{Out}(F_{N}).

  1. (1)

    Suppose that μ\mu has finite second moment. Then the following limit (called the asymptotic variance of μ\mu) exists:

    σ2(μ):=limn1nVar[d(o,Zno)],\sigma^{2}(\mu):=\lim_{n\rightarrow\infty}\frac{1}{n}Var[d(o,Z_{n}o)],

    and the random variables 1n[d(o,Zno)λ(μ)n]\frac{1}{\sqrt{n}}[d(o,Z_{n}o)-\lambda(\mu)n] and 1n[τ(Zn)λ(μ)n]\frac{1}{\sqrt{n}}[\tau(Z_{n})-\lambda(\mu)n] converge in law to the Gaussian law 𝒩(0,σ(μ))\mathcal{N}(0,\sigma(\mu)) with zero mean and variance σ2(μ)\sigma^{2}(\mu). Furthermore, we have

    lim supnd(o,Zno)λ(μ)n2nloglogn=lim supnτ(Zn)λ(μ)n2nloglogn=σ(μ).\limsup_{n\rightarrow\infty}\frac{d(o,Z_{n}o)-\lambda(\mu)n}{\sqrt{2n\log\log n}}=\limsup_{n\rightarrow\infty}\frac{\tau(Z_{n})-\lambda(\mu)n}{\sqrt{2n\log\log n}}=\sigma(\mu).
  2. (2)

    Suppose that μ\mu has infinite second moment. Then neither (1nd(o,Zno)cn)n>0(\frac{1}{\sqrt{n}}d(o,Z_{n}o)-c_{n})_{n>0} nor (1nτ(Zn)cn)n>0(\frac{1}{\sqrt{n}}\tau(Z_{n})-c_{n})_{n>0} converge in law for any choice of the sequence (cn)n>0(c_{n})_{n>0}.

Next, we discuss the geodesic tracking of random walks, which is a generalization of [Kai00, Theorem 7.2], [Tio15, Theorem 5] and [MT18, Theorem 1.3] to Outer space.

Theorem 5.3.

Let (X,d)(X,d) be the Culler-Vogtmann Outer space CVNCV_{N} of rank N3N\geq 3 with the Lipschitz metric, and let (Zn)n(Z_{n})_{n} be the random walk generated by a non-elementary probability measure μ\mu on Out(FN)\operatorname{Out}(F_{N}).

  1. (1)

    Suppose that μ\mu has finite pp-th moment for some p>0p>0. Then for almost every sample path (Zn(ω))n0(Z_{n}(\operatorname{\omega}))_{n\geq 0}, there exists a quasigeodesic γ\gamma such that

    limn1n1/2pdsym(Zno,γ)=0.\lim_{n\rightarrow\infty}\frac{1}{n^{1/2p}}d^{sym}(Z_{n}o,\gamma)=0.
  2. (2)

    Suppose that μ\mu and μˇ\check{\mu} have finite exponential moment. Then there exists K<+K<+\infty such that for almost every sample path (Zn(ω)n0(Z_{n}(\operatorname{\omega})_{n\geq 0}, there exists a quasigeodesic γ\gamma satisfying

    limn1logndsym(Zno,γ)<K.\lim_{n\rightarrow\infty}\frac{1}{\log n}d^{sym}(Z_{n}o,\gamma)<K.

We also discuss the large deviation principle for the RVs {d(o,Zno)/n}n>0\{d(o,Z_{n}o)/n\}_{n>0}, by combining the strategy of [BMSS22] and [Gou22].

Theorem 5.4.

Let (X,d)(X,d) be the Culler-Vogtmann Outer space CVNCV_{N} of rank N3N\geq 3 with the Lipschitz metric, and let (Zn)n(Z_{n})_{n} be the random walk generated by a non-elementary probability measure μ\mu on Out(FN)\operatorname{Out}(F_{N}). Let λ(μ)=limn1n𝔼[d(o,Zno)]\lambda(\mu)=\lim_{n}\frac{1}{n}\operatorname{\mathbb{E}}[d(o,Z_{n}o)] be the drift of μ\mu. Then for each 0<L<λ(μ)0<L<\lambda(\mu), the probability (d(o,Zno)Ln)\operatorname{\mathbb{P}}(d(o,Z_{n}o)\leq Ln) decays exponentially as nn tends to infinity.

Theorem 5.5.

Let (X,d)(X,d) be the Culler-Vogtmann Outer space CVNCV_{N} of rank N3N\geq 3 with the Lipschitz metric, and let (Zn)n(Z_{n})_{n} be the random walk generated by a non-elementary probability measure μ\mu on Out(FN)\operatorname{Out}(F_{N}). Then there exists a proper convex function I:[0,+]I:\mathbb{R}\rightarrow[0,+\infty], vanishing only at the drift λ(μ)\lambda(\mu), such that

infxint(E)I(x)\displaystyle-\inf_{x\in\operatorname{int}(E)}I(x) lim infn1nlog(1nd(id,Zn)E),\displaystyle\leq\liminf_{n\rightarrow\infty}\frac{1}{n}\log\operatorname{\mathbb{P}}\left(\frac{1}{n}d(id,Z_{n})\in E\right),
infxE¯I(x)\displaystyle-\inf_{x\in\bar{E}}I(x) lim supn1nlog(1nd(id,Zn)E)\displaystyle\geq\limsup_{n\rightarrow\infty}\frac{1}{n}\log\operatorname{\mathbb{P}}\left(\frac{1}{n}d(id,Z_{n})\in E\right)

holds for every measurable set EE\subseteq\mathbb{R}.

Theorem 5.6.

Let (X,d)(X,d) be the Culler-Vogtmann Outer space CVNCV_{N} of rank N3N\geq 3 with the Lipschitz metric, and let (Zn)n(Z_{n})_{n} be the random walk generated by a non-elementary probability measure μ\mu on Out(FN)\operatorname{Out}(F_{N}). If μ\mu has finite pp-th moment for some p>0p>0, then

limn1n1/2p[d(o,Zno)τ(Zn)]=0a.s.\lim_{n\rightarrow\infty}\frac{1}{n^{1/2p}}[d(o,Z_{n}o)-\tau(Z_{n})]=0\quad\textrm{a.s.}

Furthermore, if μ\mu has finite first moment, then there exists K>0K>0 such that

limn1logn[d(o,Zno)τ(Zn)]<Ka.s.\lim_{n\rightarrow\infty}\frac{1}{\log n}[d(o,Z_{n}o)-\tau(Z_{n})]<K\quad\textrm{a.s.}

The following is an effective version of [TT16, Theorem 1.4] regarding genericity of q.i. embedded subgroups of Out(FN)\operatorname{Out}(F_{N}).

Theorem 5.7.

Let (X,d)(X,d) be the Culler-Vogtmann Outer space CVNCV_{N} of rank N3N\geq 3 with the Lipschitz metric and let (Zn(1),,Zn(k))n0(Z_{n}^{(1)},\ldots,Z_{n}^{(k)})_{n\geq 0} be kk independent random walks generated by a non-elementary measure μ\mu on Out(FN)\operatorname{Out}(F_{N}). Then there exists K>0K>0 such that

[Zn(1),,Zn(k)is q.i. embedded into a quasi-convex subset of X]1Ken/K.\operatorname{\mathbb{P}}\left[\langle Z_{n}^{(1)},\ldots,Z_{n}^{(k)}\rangle\,\,\textrm{is q.i. embedded into a quasi-convex subset of $X$}\right]\geq 1-Ke^{-n/K}.

By using the technique in [Cho21], [Cho22c], we obtain an analogue in the counting problem.

Theorem 5.8.

Let (X,d)(X,d) be the Culler-Vogtmann Outer space CVNCV_{N} of rank N3N\geq 3 with the Lipschitz metric. Then for each k>0k>0, there exists a finite generating set SS of Out(FN)\operatorname{Out}(F_{N}) such that

#{(g1,,gk)(BS(n))k:g1,,gkis q.i. embedded intoa quasi-convex subset of X}(#BS(n))k\frac{\#\left\{(g_{1},\ldots,g_{k})\in\big{(}B_{S}(n)\big{)}^{k}:\begin{array}[]{c}\langle g_{1},\ldots,g_{k}\rangle\,\,\textrm{is q.i. embedded into}\\ \textrm{a quasi-convex subset of $X$}\end{array}\right\}}{\big{(}\#B_{S}(n)\big{)}^{k}}

converges to 1 exponentially fast.

Since any fully irreducible has BGIP, we can apply Lemma 2.8 to have the following version of Proposition 3.27:

Proposition 5.9.

Let (X,d)(X,d) be the Culler-Vogtmann Outer space CVNCV_{N} of rank N3N\geq 3 with the Lipschitz metric and let (Zn)n>0(Z_{n})_{n>0} be the random walk generated by a non-elementary probability measure μ\mu on Out(FN)\operatorname{Out}(F_{N}). Then for each fully irreducible gg generated by the support of μ\mu, there exists K>0K>0 such that the following holds.

For each positive integer MM, there exists ϵ>0\epsilon>0 such that the following holds outside a set of exponentially decaying probability: for a random path (Zn)n>0(Z_{n})_{n>0}, there exist 0<i(1)<<i(ϵn)<n0<i(1)<\ldots<i(\epsilon n)<n and there exist disjoint subsegments η1,,ηϵn\eta_{1},\ldots,\eta_{\epsilon n} of [o,Zno][o,Z_{n}o], from closest to farthest from oo, such that ηk\eta_{k} and [Zi(k),Zi(k)gMo][Z_{i(k)},Z_{i(k)}g^{M}o] are KK-fellow traveling for k=1,,ϵnk=1,\ldots,\epsilon n.

In [KMPT22], the author observed the following stability of triangular fully irreducible automorphisms. For the definition of principal and triangular fully irreducibles, refer to [KMPT22, Section 5.3]; for us, it it sufficient to know that principal and triangular fully irreducibles are ageometric.

Lemma 5.10 ([KMPT22, Corollary 5.14]).

Let N3N\geq 3. Let φ\varphi be a fully irreducible outer automorphism in Out(FN)\operatorname{Out}(F_{N}) with the lone axis γ\gamma on CVNCV_{N} (this is satisfied when φ\varphi is a principal fully irreducible). Then for each ϵ>0\epsilon>0 there exists R1R\geq 1 such that, if ψ\psi is a fully irreducible automorphism with an axis γ\gamma^{\prime} such that some RR-long subsegments of γ\gamma and γ\gamma^{\prime} are ϵ\epsilon-fellow traveling, then ψ\psi is a triangular fully irreducible.

By combining Lemma 5.10 with Proposition 5.9, we recover version of Kapovich-Maher-Pfaff-Taylor’s result with weaker moment condition.

Theorem 5.11 (cf. [KMPT22, Theorem A]).

Let N3N\geq 3 and let μ\mu be a non-elementary probability measure on Out(FN)\operatorname{Out}(F_{N}) such that the subsemigroup suppμ\operatorname{\langle\langle}\operatorname{supp}\mu\operatorname{\rangle\rangle} generated by the support of μ\mu contains the inverse of a principal fully irreducible automorphism. Then outside a set of exponentially decaying probability, ZnZ_{n} is an ageometric triangular fully irreducible outer automorphism.

Finally, Theorem A reads as follows:

Theorem 5.12.

Let (X,d)(X,d) be the Culler-Vogtmann Outer space CVNCV_{N} of rank N3N\geq 3 with the Lipschitz metric and let (Zn)n>0(Z_{n})_{n>0} be the random walk generated by a non-elementary, asympotically asymmetric probability measure μ\mu on Out(FN)\operatorname{Out}(F_{N}) (for example, μ\mu is an admissible probability measure). Then for any K>0K>0, we have

limn(ω:|τ(Zn)τ(Zn1)|<K)=0.\lim_{n\rightarrow\infty}\operatorname{\mathbb{P}}\Big{(}\operatorname{\omega}:\big{|}\tau(Z_{n})-\tau(Z_{n}^{-1})\big{|}<K\Big{)}=0.

References

  • [ACT15] Goulnara N. Arzhantseva, Christopher H. Cashen, and Jing Tao. Growth tight actions. Pacific J. Math., 278(1):1–49, 2015.
  • [AK10] Yael Algom-Kfir. The Lipschitz metric on outer space. ProQuest LLC, Ann Arbor, MI, 2010. Thesis (Ph.D.)–The University of Utah.
  • [AK11] Yael Algom-Kfir. Strongly contracting geodesics in outer space. Geom. Topol., 15(4):2181–2233, 2011.
  • [AKB12] Yael Algom-Kfir and Mladen Bestvina. Asymmetry of outer space. Geom. Dedicata, 156:81–92, 2012.
  • [AKKP19] Yael Algom-Kfir, Ilya Kapovich, and Catherine Pfaff. Stable strata of geodesics in outer space. Int. Math. Res. Not. IMRN, (14):4549–4578, 2019.
  • [BF09] Mladen Bestvina and Koji Fujiwara. A characterization of higher rank symmetric spaces via bounded cohomology. Geom. Funct. Anal., 19(1):11–40, 2009.
  • [BF14] Mladen Bestvina and Mark Feighn. Hyperbolicity of the complex of free factors. Adv. Math., 256:104–155, 2014.
  • [BH92] Mladen Bestvina and Michael Handel. Train tracks and automorphisms of free groups. Ann. of Math. (2), 135(1):1–51, 1992.
  • [BMSS22] Adrien Bounlanger, Pierre Mathieu, Çağrı Sert, and Alessandro Sisto. Large deviations for random walks on hyperbolic spaces. Ann. Sci. Éc. Norm. Supér. (4), 2022.
  • [BR15] Mladen Bestvina and Patrick Reynolds. The boundary of the complex of free factors. Duke Math. J., 164(11):2213–2251, 2015.
  • [CCT23] Kunal Chawla, Inhyeok Choi, and Giulio Tiozzo. Genericity of contracting geodesics in groups. In preparation, 2023.
  • [CH12] Thierry Coulbois and Arnaud Hilion. Botany of irreducible automorphisms of free groups. Pacific J. Math., 256(2):291–307, 2012.
  • [Cho21] Inhyeok Choi. Pseudo-Anosovs are exponentially generic in mapping class groups. arXiv preprint arXiv:2110.06678, to appear in Geometry and Topology, 2021.
  • [Cho22a] Inhyeok Choi. Limit laws on Outer space, Teichmüller space, and CAT(0) spaces. arXiv preprint arXiv:2207.06597v1, 2022.
  • [Cho22b] Inhyeok Choi. Random walks and contracting elements I: Deviation inequality and limit laws. arXiv preprint arXiv:2207.06597v2, 2022.
  • [Cho22c] Inhyeok Choi. Random walks and contracting elements II: translation length and quasi-isometric embedding. arXiv preprint arXiv:2212.12119, 2022.
  • [DH18] François Dahmani and Camille Horbez. Spectral theorems for random walks on mapping class groups and Out(FN)(F_{N}). Int. Math. Res. Not. IMRN, (9):2693–2744, 2018.
  • [DT18] Spencer Dowdall and Samuel J. Taylor. Hyperbolic extensions of free groups. Geom. Topol., 22(1):517–570, 2018.
  • [FM11] Stefano Francaviglia and Armando Martino. Metric properties of outer space. Publ. Mat., 55(2):433–473, 2011.
  • [FM12] Stefano Francaviglia and Armando Martino. The isometry group of outer space. Adv. Math., 231(3-4):1940–1973, 2012.
  • [Gou22] Sébastien Gouëzel. Exponential bounds for random walks on hyperbolic spaces without moment conditions. Tunis. J. Math., 4(4):635–671, 2022.
  • [HM07a] Michael Handel and Lee Mosher. The expansion factors of an outer automorphism and its inverse. Trans. Amer. Math. Soc., 359(7):3185–3208, 2007.
  • [HM07b] Michael Handel and Lee Mosher. Parageometric outer automorphisms of free groups. Trans. Amer. Math. Soc., 359(7):3153–3183, 2007.
  • [Hor16a] Camille Horbez. The horoboundary of outer space, and growth under random automorphisms. Ann. Sci. Éc. Norm. Supér. (4), 49(5):1075–1123, 2016.
  • [Hor16b] Camille Horbez. A short proof of Handel and Mosher’s alternative for subgroups of Out(FN)(F_{N}). Groups Geom. Dyn., 10(2):709–721, 2016.
  • [Hor18] Camille Horbez. Central limit theorems for mapping class groups and Out(FN){\rm Out}(F_{N}). Geom. Topol., 22(1):105–156, 2018.
  • [JL08] André Jäger and Martin Lustig. Free group automorphisms with many fixed points at infinity. In The Zieschang Gedenkschrift, volume 14 of Geom. Topol. Monogr., pages 321–333. Geom. Topol. Publ., Coventry, 2008.
  • [Kai00] Vadim A. Kaimanovich. The Poisson formula for groups with hyperbolic properties. Ann. of Math. (2), 152(3):659–692, 2000.
  • [KMPT22] Ilya Kapovich, Joseph Maher, Catherine Pfaff, and Samuel J. Taylor. Random outer automorphisms of free groups: attracting trees and their singularity structures. Trans. Amer. Math. Soc., 375(1):525–557, 2022.
  • [MT18] Joseph Maher and Giulio Tiozzo. Random walks on weakly hyperbolic groups. J. Reine Angew. Math., 742:187–239, 2018.
  • [Sis18] Alessandro Sisto. Contracting elements and random walks. J. Reine Angew. Math., 742:79–114, 2018.
  • [Tio15] Giulio Tiozzo. Sublinear deviation between geodesics and sample paths. Duke Math. J., 164(3):511–539, 2015.
  • [TT16] Samuel J. Taylor and Giulio Tiozzo. Random extensions of free groups and surface groups are hyperbolic. Int. Math. Res. Not. IMRN, (1):294–310, 2016.
  • [Vog15] Karen Vogtmann. On the geometry of outer space. Bull. Amer. Math. Soc. (N.S.), 52(1):27–46, 2015.
  • [Yan19] Wen-yuan Yang. Statistically convex-cocompact actions of groups with contracting elements. Int. Math. Res. Not. IMRN, (23):7259–7323, 2019.
  • [Yan20] Wen-yuan Yang. Genericity of contracting elements in groups. Math. Ann., 376(3-4):823–861, 2020.