1 Introduction
This is the first in the series of articles concerning random walks on metric spaces with contracting elements. This series is a reformulation of the preprint [Reference ChoiCho22a] announced by the author, aiming for a more concise and systematic exposition.
Let
$G$
be a countable group of isometries of a metric space
$(X, d)$
with basepoint
$o \in X$
. We consider the random walk generated by a probability measure
$\mu$
on
$G$
, which entails the product
$Z_{n} = g_{1} \cdots g_{n}$
of independent random isometries
$g_{i}$
’s chosen with law
$\mu$
. We are interested in the asymptotic behavior of a random path
$(Z_{n})_{n \gt 0}$
seen by
$X$
, or in other words, the behavior of a random orbit path
$(Z_{n} o)_{ n \gt 0}$
on
$X$
. For instance, we can ask the following questions:
-
– Does the random variable
$\frac {1}{n} d(o, Z_{n} o)$ converge to a constant almost surely?
-
– Does the random variable
$\frac {1}{\sqrt {n}} d(o, Z_{n} o)$ converge in law to a Gaussian law?
-
– How fast does
$\mathbb{P}\left (an \leqslant d(o, Z_{n} o) \leqslant bn\right )$ decay for
$0 \leqslant a \leqslant b$ ?
These questions are associated with the so-called moment conditions. For each
$p\gt 0$
we define the
$p$
-th moment of
$\mu$
by

and the exponential moment (with a parameter
$K\gt 0$
) of
$\mu$
by

In the classical setting of
$X = \mathbb{R}$
, the previous three questions are answered when
$\mu$
has finite first moment, finite second moment and finite exponential moment, respectively.
A particularly interesting examples come from isometric actions on non-positively curved spaces. This setting includes Gromov hyperbolic groups ([Reference Benoist and QuintBQ16], [Reference Bounlanger, Mathieu, Sert and SistoBMSS22], [Reference GouëzelGou22]); relatively hyperbolic groups ([Reference SistoSis17], [Reference Qing, Rafi and TiozzoQRT20]); groups with nontrivial Floyd boundary ([Reference Gekhtman, Gerasimov, Potyagailo and YangGGPY21]); the mapping class group of a finite-type hyperbolic surface acting on Teichmüller space ([Reference Kaimanovich and MasurKM96], [Reference HorbezHor18], [Reference Dahmani and HorbezDH18], [Reference Baik, Choi and KimBCK21]) or the curve complex ([Reference MaherMah10a], [Reference MaherMah10b], [Reference MaherMah11]); the outer automorphism group of a finite-rank free group acting on the Culler-Vogtmann Outer space ([Reference HorbezHor18], [Reference Dahmani and HorbezDH18]) and the free factor complex; groups acting on CAT(0) spaces ([Reference Karlsson and MargulisKM99], [Reference Karlsson and LedrappierKL06], [Reference FernósFer18], [Reference BarsLB22b], [Reference BarsLB22a]).
In this paper, we propose a unified theory for random walks on the aforementioned spaces. We first study the case where
$X$
possesses strongly contracting isometries (see Convention 2.11), and
$\mu$
is non-elementary (see Subsection 2.4). This condition is mild enough to cover all the aforementioned spaces (except for Outer space, which will be studied carefully in [Reference ChoiCho22b] due to the asymmetry issue). At the same time, this is just the right amount of restriction that leads to limit laws under optimal moment conditions.
We also present a parallel theory for metric spaces with weakly contracting isometries (see Convention 7.2). As a result, we obtain limit laws on hierarchically hyperbolic groups (HHGs) with optimal moment conditions. We describe the case of mapping class group for concreteness.
Theorem A.
Let
$G$
be the mapping class group of a finite-type surface, let
$d$
be a word metric on
$G$
, let
$(Z_{n})_{n\geqslant 0}$
be the random walk generated by a non-elementary probability measure
$\mu$
on
$G$
, and let

be the drift of
$\mu$
on
$G$
. Then for each
$0 \lt L \lt \lambda$
, the probability
$\mathbb{P} (d(id, Z_{n}) \leqslant Ln)$
decays exponentially as
$n$
tends to infinity.
This is an analogue of the result of Gouëzel [[Reference GouëzelGou22],Theorem 1.3], who established the exponential bound for Gromov hyperbolic spaces. Note that, for every admissible probability measure
$\mu$
on the mapping class group
$G$
, the spectral radius of
$\mu$
is strictly smaller than
$1$
due to the non-amenability of
$G$
[Reference KestenKes59]. Combining this with the exponential growth of
$G$
, one can obtain
$L\gt 0$
for which
$\mathbb{P}( d(id, Z_{n}) \leqslant Ln)$
decays exponentially. Hence, the nontrivial part of Theorem A is that
$L$
can be as close to
$\lambda$
as we want.
We also obtain the deviation inequalities with optimal moment conditions (see Proposition 4.13). Combining this with Mathieu-Sisto’s theory [Reference Mathieu and SistoMS20], we establish the central limit theorem (CLT) and law of the iterated logarithms (LIL) on mapping class groups.
Theorem B.
Let
$G$
be the mapping class group of a finite-type hyperbolic surface, let
$d$
be a word metric on
$G$
, and let
$(Z_{n})_{n\geqslant 0}$
be the random walk generated by a non-elementary probability measure
$\mu$
on
$G$
with finite second moment. Then there exists
$\sigma (\mu )\geqslant 0$
such that
$\frac {1}{\sqrt {n}} (d(id, Z_{n}) - n \lambda (\mu ))$
converges in law to the Gaussian law
$\mathcal{N}(0, \sigma (\mu ))$
of variance
$\sigma (\mu )^{2}$
, and moreover,

In acylindrically hyperbolic groups, Mathieu and Sisto established CLT for random walks with finite exponential moment ([[Reference Mathieu and SistoMS20],Theorem 13.4]). We strengthen their result by weakening the moment condition.
Lastly, we address the geodesic tracking of random paths.
Theorem C.
Let
$G$
be the mapping class group of a finite-type surface, let
$d$
be a word metric on
$G$
, and let
$(Z_{n})_{n\geqslant 0}$
be the random walk generated by a non-elementary measure
$\mu$
on
$G$
.
-
(i) Let
$p\gt 0$ and suppose that
$\mu$ has finite
$p$ -th moment. Then for almost every sample path
$(Z_{n})_{n\geqslant 0}$ , there exists a geodesic
$\gamma$ on
$G$ such that
\begin{align*} \lim _{n \rightarrow \infty } \frac {1}{n^{1/p}} d(Z_{n}, \gamma ) = 0. \end{align*}
-
(ii) If
$\mu$ has finite exponential moment, then there exists
$K \lt \infty$ such that the following holds. For almost every sample path
$(Z_{n})_{n\geqslant 0}$ , there exists a geodesic
$\gamma$ on
$G$ such that
\begin{align*} \limsup _{n \rightarrow \infty } \frac {1}{\log n} d(Z_{n}, \gamma ) \lt K. \end{align*}
For finitely supported random walks, Sisto established the deviation rate
$d(Z_{n}, \gamma ) = O(\sqrt {n \log n})$
([[Reference SistoSis17],Theorem 1.2]). Later, Qing, Rafi and Tiozzo obtained the rate
$d(Z_{n}, \gamma ) = O(\log ^{3g - 3 + b}(t))$
, where
$g$
and
$b$
denote the genus and the number of punctures of the surface ([[Reference Qing, Rafi and TiozzoQRT20],Theorem C]). We refine these results by suggesting the deviation rate
$O(\log (t))$
for random walks with finite exponential moment.
In full generality, the main results hold in the setting of Convention 2.11 and Convention 7.2. In particular, Theorem A, B and C apply to random walks on rank-1 CAT(0) spaces. This extends the author’s previous work [Reference ChoiCho23] that deals with Gromov hyperbolic spaces and Teichmüller space, and recovers several results by Le Bars [Reference BarsLB22b], [Reference BarsLB22a].
To obtain the main theorems, we blend the pioneering theories due to Gouëzel [Reference GouëzelGou22] and due to Mathieu and Sisto [Reference Mathieu and SistoMS20]. Gouëzel’s method effectively captures the alignment of the orbit path on
$X$
(see Subsection 4.1), while Mathieu-Sisto’s technique provides the desired limit theorems when appropriate deviation inequalities are given. Both of these theories rely on the Gromov hyperbolicity of the ambient space. Our contribution is to replace the Gromov hyperbolicity with weaker notion of hyperbolicity. In particular, we obtain large deviation principle, CLT and geodesic tracking on (possibly non-proper) CAT(0) spaces. Moreover, we generalize Mathieu-Sisto’s theory by lifting the moment condition, leading to the exponential bounds for the escape to infinity and CLT for random walks without finite exponential moment.
1.1 Context
Random walks on groups have often been studied via their actions on Gromov hyperbolic spaces. For instance, random walks on Teichmüller space and Outer space have been understood by coupling them with the curve complex and the free factor complex, respectively ([Reference HorbezHor18], [Reference Dahmani and HorbezDH18]). A similar strategy was recently pursued for proper CAT(0) spaces by Le Bars ([Reference BarsLB22b], [Reference BarsLB22a]), building upon a new hyperbolic model for CAT(0) spaces ([Reference Petyt, Spriano and ZalloumPSZ24]).
These strategies eventually depend on the following ingredients:
-
– the non-atomness of the stationary measure on the Gromov boundary ([[Reference Maher and TiozzoMT18],Proposition 5.1]);
-
– CLT for martingales arising from Busemann cocycles ([[Reference Benoist and QuintBQ16],Theorem 4.7]);
-
– linear progress with exponential decay ([[Reference MaherMah12],Theorem 1.2]), or
-
– linear progress using the acylindricity of the action ([[Reference Mathieu and SistoMS20],Theorem 9.1, Proposition 9.4]).
The first two items require a nice (e.g., compact) boundary structure of
$X$
. These boundary structures are also available in some class of non-Gromov-hyperbolic spaces (such as Teichmüller space, Outer space and finite-dimensional CAT(0) cube complices – see [Reference FernósFer18], [Reference FernósFer18], [Reference Fernós, Lécureux and MathéusFLM24]) but are hard to come by in the general case.
To establish the third item, Maher considered a stopping time that arises when a random path penetrates nested shadows, which relies on moment conditions: see [Reference MaherMah12] and [Reference SunderlandSun20]. For the last item, Mathieu and Sisto assumed finite exponential moment condition to couple the random paths on
$G$
with the corresponding paths on
$G$
in probability.
It is not straightforward to apply the aforementioned strategies to, say, random walks on (non-proper) CAT(0) spaces. Even in well-known settings such as Gromov hyperbolic groups, moment conditions are often necessary. Our goal is to lift these restrictions: we want a structure for random walks on a wide class of spaces
$X$
that:
-
– does not assume global Gromov hyperbolicity of
$X$ ;
-
– does not rely on any boundary structure of
$X$ ;
-
– does not assume any moment condition a priori, and
-
– effectively captures the ‘alignment’ of a sample path on
$X$ .
The first goal was studied by Sisto in [Reference SistoSis18]. Not assuming global Gromov hyperbolicity of
$X$
, Sisto presented a random walk theory using strongly contracting isometries, which are found in both Gromov hyperbolic spaces and CAT(0) spaces. Note that the existence of strongly contracting isometries also has implications on the growth problem and counting problem ([Reference Arzhantseva, Cashen and TaoACT15], [Reference YangYan14], [Reference YangYan19], [Reference YangYan20], [Reference LegaspiLeg22] and [Reference CoulonCou22]).
The second goal was pursued by Mathieu and Sisto for acylindrically hyperbolic groups in [Reference Mathieu and SistoMS20], establishing deviation inequalities without referring to the boundary of
$X$
.
The first and the second goals were also pursued by Boulanger, Mathieu, Sert and Sisto in [Reference Bounlanger, Mathieu, Sert and SistoBMSS22]. They discuss Corollary 6.5 for Gromov hyperbolic spaces and pointed out the versatility of Schottky sets in other spaces. For more detail, see Section 6.
All the goals except the first one were achieved in Gouëzel’s recent paper [Reference GouëzelGou22]. In [Reference GouëzelGou22], Gouëzel establishes Theorem A for Gromov hyperbolic spaces by recording the Schottky directions aligned along a random path. Such a recording, called the set of pivotal times, grows linearly with exponential decay. More importantly, this growth is uniform and is independent of the intermediate non-Schottky steps.
Our theory achieves the 4 goals in the setting of Convention 2.11. For this purpose, we combine Gouëzel’s pivotal time construction with Sisto’s theory of random walks involving strongly contracting isometries. This was indirectly pursued for Teichmüller space in [Reference ChoiCho23]. Our usage of strongly contracting isometries is also hugely influenced by Yang’s series of papers ([Reference YangYan14], [Reference YangYan19], [Reference YangYan20]) in the context of counting problems.
Although strongly contracting isometries are found in various groups, it is not known whether the Cayley graph of a mapping class group possesses strongly contracting isometries. A related issue arises when one considers a group
$G$
that is quasi-isometric to another group
$H$
. Having a strongly contracting isometry is not passed through quasi-isometries: it is even not preserved under the change of finite generating set of a group [[Reference Arzhantseva, Cashen, Gruber and HumeACGH19],Theorem 4.19].
This is why we provide a parallel theory in the language of weakly contracting isometries. We note that having a weakly contracting infinite quasigeodesic is stable under quasi-isometry. Strictly speaking, our setting is not stable under quasi-isometry: we consider two coarsely equivariant
$G$
-actions, one involving weak contraction and the other one involving strong contraction. Nevertheless, the present theory is an attempt towards QI-invariant random walk theory. We record recent breakthrough in this direction by Goldborough and Sisto [Reference Goldsborough and SistoGS21], showing that certain QI-invariant group-theoretic property (that involves an action on a hyperbolic space) guarantees a CLT for simple random walks.
1.2 Strategy
Morally, contracting directions constitute a tree-like structure. As a toy model, consider

acting on its Cayley graph
$X$
. A geodesic
$\gamma = abaaba$
in
$X$
is composed of edges
$e_{1} = [id, a]$
,
$e_{2} = [a, ab]$
,
$e_{3} = [ab, aba]$
and so on. The geodesicity of
$\gamma$
forces the local alignment among
$e_{i}$
’s:
$e_{i}$
projects onto
$e_{i+1}$
at the beginning point of
$e_{i+1}$
and
$e_{i+1}$
projects onto
$e_{i}$
at the ending point of
$e_{i}$
. Conversely, this local alignment implies that
$\gamma$
is geodesic. (This is false when
$e_{i}$
’s are directions in a flat, e.g.,
$e_{1} = [id, c]$
,
$e_{2} = [c, cd]$
and
$e_{3} = [cd, cdc^{-1}]$
.) The same conclusion holds even if we insert edges in the flats in between
$e_{i}$
’s. For example, consider

Observe that
$(id, e_{1})$
,
$(e_{1}, e_{2})$
and
$(e_{2}, g)$
satisfy the local alignment conditions. This forces that
$e_{1}$
and
$e_{2}$
are subsegments of any geodesic between
$id$
and
$g$
even if such a geodesic is not unique due to flat parts. We will formulate this more precisely in the alignment lemma in Section 3.
We will then construct many independent “tree-like” directions. In our example, the set

consists of
$2^{M}$
directions in the free factor. We have the following property:
-
(i) For any
$x \in X$ ,
$d(id, [x, s^{\pm 1}])\lt M$ for all but at most 1 element
$s \in S_{M, m}$ .
-
(ii) For all
$s \in S_{M, m}$ , the geodesic
$[s^{-1}, s]$ passes through
$id$ .
This property will be captured by the notion of Schottky sets (Definition 3.15). Note that one can increase the cardinality of
$S_{M, m}$
by taking larger
$M$
.
Let us now consider the random walk
$Z$
generated by a probability measure
$\mu$
with
$\mu (a), \mu (b) \gt 0$
. Then for any
$M, m\gt 0$
, each element of the Schottky set
$S_{M, m}$
is admitted by
$\mu ^{\ast Mm}$
. By decomposing
$\mu ^{\ast Mm}$
into a uniform measure on
$S_{M, m}$
and the remainder, a random path
$(Z_{n})_{n}$
can be modelled by the concatenation of some non-Schottky isometries
$w_{i}$
’s and Schottky isometries
$s_{i}$
’s, where the timing for Schottky progresses are given by a renewal process. That means, for a large
$K$
, a random word
$Z_{n} = g_{1} \cdots g_{n}$
is of the form

where
$s_{i}$
’s are drawn from
$S_{M, m}$
. Now Gouëzel’s construction of pivotal times provides a large
$K'$
such that the following holds: among
$\{1, \ldots , n/K\}$
, we can pick indices
$i(1) \lt \ldots \lt i(n/KK')$
at which the Schottky segment is aligned along the entire progress, i.e.,
$w_{0}s_{1} \cdots w_{i(k)} [id, s_{i(k)}]$
’s are subsegments of
$[id, Z_{n}]$
$(\ast$
). Now pick
$x \in X$
. We have plenty of Schottky isometries available for the slot
$s_{i(k)}$
’s. By choosing the right choice among them (i.e., by pivoting), we can also assure that
$(x, w_{0}s_{1} \cdots w_{i(k)} [id, s_{i(k)}])$
is aligned. Combined with
$(\ast$
), this means that we have a bound of
$d(id, [x, Z_{n}])$
in terms of an initial subsegment
$w_{0}s_{1} \cdots w_{i(k)}$
of the random path. All these phenomena are exponentially generic (see Lemma 4.10). We subsequently obtain deviation inequalities (Proposition 4.13), central limit theorem and geodesic tracking. A more involved combinatorial model for random paths leads to the large deviation principle.
In this example, the contracting property of a tree-like edge
$e$
is as strong as possible: any geodesic
$\gamma$
connecting the left and the right of
$e$
’s passes through
$e$
. We study two variants of such a contracting property. If we require that
$\gamma$
passes through a bounded neighborhood of
$e$
, we say that
$e$
is strongly contracting. If we require that
$\gamma$
passes through a
$\log (\operatorname {diam}(e))$
-neighborhood of
$e$
, than we say that
$e$
is weakly contracting. The argument so far also works for strongly contracting directions, up to a finite error. A more delicate argument is required for weakly contracting isometries. We will deal with these notions in Part I and Part II, respectively.
2 Preliminaries
Before entering Part I, we review basic notions and lemmata. We fix a metric space
$(X, d)$
and a basepoint
$o \in X$
. For
$x, y , z \in X$
, we define the Gromov product of
$x$
and
$z$
with respect to
$y$
by

2.1 Paths
Let
$A$
and
$B$
be subsets of
$X$
.
$A$
is
$K$
-coarsely contained in
$B$
if
$A$
is contained in the
$K$
-neighborhood of
$B$
.
$A$
and
$B$
are
$K$
-coarsely equivalent if
$A$
is
$K$
-coarsely contained in
$B$
and vice versa.
$A$
is
$K$
-coarsely connected if for every
$x, y \in A$
there exists a chain
$x = a_{0}, a_{1}, \ldots , a_{n} = y$
of points in
$A$
such that
$d(a_{i}, a_{i+1}) \leqslant K$
for each
$i$
.
A path on
$X$
is a map
$\gamma : I \rightarrow X$
from a
$1$
-coarsely connected subset
$I$
of
$\mathbb{R}$
, called a domain, to
$X$
. A subdomain
$J$
of
$I$
is of the form
$I \cap [a, b]$
for some
$a, b \in \mathbb{R}$
. The restriction of
$\gamma$
on
$J$
is called a subpath of
$\gamma$
. We denote this subpath by
$\gamma |_{[a, b]}$
.
For paths
$\gamma : I \rightarrow X$
and
$\gamma ' : I' \rightarrow X$
, we say that
$\gamma '$
is a reparametrization of
$\gamma$
when there exists a non-decreasing map
$\rho : I' \rightarrow I$
such that
$\gamma ' = \gamma \circ \rho$
. We say that two paths
$\kappa : I \rightarrow X$
and
$\eta : J \rightarrow X$
are
$K$
-fellow traveling if there exists a reparametrization
$\kappa ' : J \rightarrow X$
of
$\kappa$
such that
$d(\kappa '(t), \eta (t)) \leqslant K$
for every
$t \in J$
. In this case, note that the images of
$\kappa$
and
$\eta$
are within Hausdorff distance
$K$
and the endpoints of
$\kappa$
and
$\eta$
are pairwise
$K$
-near. By abuse of notation, for a path
$\kappa : I \rightarrow X$
,
$\kappa$
will often refer to the set-theoretical image
$\kappa (I)$
of
$\kappa$
. For instance, when we say that a path
$\kappa : I \rightarrow X$
is
$K$
-close to a point
$x$
, it means
$d(\kappa (t), x) \lt K$
for some
$t \in I$
.
We say that
$X$
is geodesic if for each pair of points
$x, y \in X$
there exists a geodesic connecting
$x$
to
$y$
. Given two points
$x, y \in X$
, we denote by
$[x, y]$
an arbitrary geodesic connecting
$x$
to
$y$
.
Let
$[x, y]$
be a geodesic on
$X$
and
$A_{1}, \ldots , A_{N}$
be subsets of
$[x, y]$
. We say that
$A_{1}, \ldots , A_{N}$
are in order from left to right if
$d(x, x_{1}) \leqslant d(x, x_{2}) \leqslant \ldots \leqslant d(x, x_{N})$
for any choices of
$x_{i} \in A_{i}$
.
We will construct a path for a sequence of isometries as follows. Given a sequence
$\alpha = (\phi _{1}, \ldots , \phi _{k})$
of isometries of
$X$
, we denote the product of its entries
$\phi _{1} \cdots \phi _{k}$
by
$\Pi (\alpha )$
. Now let

for each
$m \in \mathbb{Z}$
and
$i = 0, \ldots , k-1$
; see Figure 1. We let
$\Gamma ^{m}(\alpha ) := (x_{0}, x_{1}, \ldots , x_{mk})$
when
$m \geqslant 0$
and
$\Gamma ^{m}(\alpha ) := (x_{0}, x_{-1}, \ldots , x_{mk})$
when
$m \lt 0$
. For
$m = \pm 1$
, we also use a simpler notation

In other words, we write:

Given a path
$\gamma = (y_{1}, \ldots , y_{N})$
, we denote by
$\bar {\gamma }$
its reversal, defined by

For example, the reversal of
$\Gamma ^{-}(\alpha )$
is denoted by
$\bar {\Gamma }^{-}(\alpha )$
, which is


Figure 1. Axes associated with a sequence of isometries
$s = (\phi _{1}, \phi _{2}, \phi _{3}, \phi _{4})$
. Points inside the darker shadow constitute
$\Gamma ^{+}(s)$
, and those inside the lighter shadow constitute
$\Gamma ^{2}(s)$
. Points in the dashed region constitute
$\Gamma ^{-}(s)$
.
2.2 Strong contraction
Given a subset
$A$
of
$X$
, we define the closest point projection
$\pi _{A} : X \rightarrow 2^{A}$
onto
$A$
by

Note that
$\pi _{A}(x)$
is nonempty for each
$x \in X$
when
$A$
is a closed and locally compact set.
Definition 2.1. Let
$K\gt 0$
. A subset
$A$
of
$X$
is
$K$
-strongly contracting if the following holds for the closest point projection
$\pi _{A}$
:

for all
$x, y \in X$
that satisfy
$d_{X}(x, y) \leqslant d_{X}(x, A)$
.
A
$K$
-strongly contracting
$K$
-quasigeodesic is called a
$K$
-contracting axis. A lemma follows:
Lemma 2.2.
Let
$A$
be a
$K$
-strongly contracting subset of
$X$
. Then the closest point projection
$\pi _{A} : X \rightarrow A$
is
$(1, 4K)$
-coarsely Lipschitz, i.e., for each
$x, y \in X$
we have

This lemma is well-known in various forms ([[Reference Arzhantseva, Cashen and TaoACT15], Lemma 2.11], [[Reference SistoSis18],Lemma 2.4] and [[Reference YangYan19],Proposition 2.4(4)]). The explicit constant
$4K$
is given as a consequence of Lemma 3.1.
Lemma 2.3 ([[Reference YangYan20], Proposition 2.2 (3)]). For each
$K\gt 1$
there exists a constant
$K' = K'(K)$
such that any subpath of a
$K$
-contracting axis is a
$K'$
-contracting axis.
Lemma 2.4 ([[Reference Arzhantseva, Cashen and TaoACT15], Lemma 2.15], [[Reference YangYan20], Proposition 2.2(2)]). Let
$A$
and
$A'$
be coarsely equivalent subsets of
$X$
. Then
$A$
is strongly contracting if and only if
$A'$
is strongly contracting.
Definition 2.5. An isometry
$g$
of
$X$
is strongly contracting if its orbit
$\{g^{i} o\}_{i \in \mathbb{Z}}$
is a strongly contracting quasigeodesic.
Definition 2.6. We say that isometries
$g$
and
$h$
of
$X$
are independent if for any
$x \in X$
the map

is proper, i.e.,
$\{ (m, n) : d(g^{m} o, h^{n} o) \lt M\}$
is bounded for each
$M \gt 0$
.
The following lemma will be proved in Subsection 3.1.
Lemma 2.7.
Two strongly contracting isometries
$g$
and
$h$
of
$X$
are independent if and only if
$\pi _{\{ g^{i} o : i \in \mathbb{Z}\}}(\{h^{i} o : i \in \mathbb{Z}\})$
and
$\pi _{\{ h^{i} o : i \in \mathbb{Z}\}}(\{g^{i} o : i \in \mathbb{Z}\})$
have finite diameters.
2.3 Weak contraction
This subsection only matters in Part II; readers interested in Part I may skip this subsection.
Definition 2.8. Let
$K\gt 0$
and
$A \subset X$
. A
$K$
-projection onto
$A$
is a
$K$
-coarsely Lipschitz map
$\pi : X \rightarrow A$
such that
$d(a, \pi (a)) \leqslant K$
for each
$a \in A$
. Note that for each
$x \in X$
we have

A set
$A$
is
$K$
-weakly contracting if there exists a
$K$
-projection
$\pi _{A}$
such that

holds for all
$x, y \in X$
that satisfy
$d(x, y) \leqslant d(x, A)/K$
.
Lemma 2.9.
For each
$K, M\gt 1$
there exists
$K' \gt K$
such that the following holds.
Let
$x, y \in X$
. Let
$A$
be a
$K$
-weakly contracting set such that
$d(x, A) \geqslant K'$
and such that
$\operatorname {diam}\big (\pi _{A}(x)\cup \pi _{A}(y)\big ) \geqslant K'$
. Then there exists
$p \in [x, y]$
such that
$\operatorname {diam}\big (\pi _{A}(x)\cup \pi _{A}(p)\big ) \leqslant 2K'$
and such that either:

Proof.
We set
$K' = K^{2}\big ( M(M+7)(K+1) + 1\big )$
.
Let
$\eta : [0, L] \rightarrow X$
be a geodesic connecting
$x$
to
$y$
. Note that for

The
$K$
-coarse Lipschitzness of
$\pi _{A}$
and Inequality 2 imply

Hence, by replacing
$y$
with
$\eta (\tau )$
, we may assume
$\operatorname {diam}\big (\pi _{A}(x) \cup \pi _{A}(\eta (t))\big ) \leqslant K' + K$
for
$t \in [0, L]$
. If
$d(\eta (t), A) \lt \frac {1}{M} d(x, A)$
for some
$t \in [0, L]$
, then we are done; suppose not. We inductively take

The process halts at step
$N$
when
$t_{N}$
reaches
$L$
. We then have

for each
$i$
. Using Inequality 2, we deduce

Since the LHS is at least
$K'$
, we have
$N \geqslant K' /K \geqslant 2KM(M+7)(K+1) + 1$
.
Meanwhile,
$t_{i} - t_{i-1} = d(x, A)/MK$
holds for
$i \leqslant N-1$
. This implies

and considering the assumption
$d(x, A) \geqslant K'\geqslant K$
we deduce

Now using Inequality 1 twice, we get

Lemma 2.10.
For each
$K\gt 1$
there exists
$K'\gt 0$
satisfying the following.
Let
$A$
be a
$K$
-weakly contracting set, let
$x, y \in X$
, let
$p$
be a point on
$[x, y]$
and define

Then we have

Proof.
Let
$M:= 2K+4$
, let
$K_{1} := K'(K, M)$
be as in Lemma 2.9, and let
$K':= 9 MK_{1}$
.
Suppose to the contrary that Inequality 3 does not hold. Our goal is to find a triple
$x', y', z'$
on
$[x, y]$
, in order from left to right, such that

If we find such triple, then we have

This will then lead to the contradiction

Let
$\eta : [0, L] \rightarrow X$
be the geodesic connecting
$p$
to
$x$
and let
$t_{0} = 0$
. Given
$t_{i-1} \in [0, L)$
, we pick
$t_{i} \in [t_{i-1}, L]$
such that

If such
$t_{N}$
does not exist at step
$N$
, we let
$t_{N} = L$
and stop.
Recall that we are assuming

Hence,
$d(\eta (t_{i}), A) \geqslant M^{i} K' \geqslant K'$
for
$i = 0, \ldots , N-1$
. (
$\ast$
) Since
$\eta$
is bounded, the process must stop at some
$N$
. We always have
$t_{N} = L$
and
$\eta (t_{N}) = x$
. We discuss possible scenarios:
-
(i)
$d\big (\pi _{A}(\eta (t_{N-1}))\cup \pi _{A}(\eta (t_{N}))\big ) \gt 2K_{1}$ . Recall Lemma 2.9: there exists
$\tau \in [t_{N-1}, t_{N}]$ such that
$\operatorname {diam} \big (\pi _{A}(\eta (t_{N-1})\cup \eta (\tau )\big ) \leqslant 2K_{1}$ and either
$d(\eta (\tau ), A) \geqslant Md(\eta (t_{N-1}), A)$ or
$d(\eta (\tau ), A) \leqslant \frac {1}{M} d(\eta (t_{N-1}), A)$ . Since the first possibility is excluded, we conclude that
$d(\eta (\tau ), A) \leqslant{} \frac {1}{M}d(\eta (t_{N-1}), A)$ . There are two subcases.
-
(a)
$N \geqslant 2$ : in this case,
$d(\eta (t_{N-1}), A) \geqslant M d(\eta (t_{N-2}), A)$ holds by our choice in Display 4. By (
$\ast$ ), we also know that
$d(\eta (t_{N-1}), A) \geqslant K'$ . Lastly,
$\pi _{A}\{\eta (t_{N-2}), \eta (t_{N-1}), \eta (\tau )\}$ has diameter at most
$4K_{1}$ . Hence, we can take
$x' = \eta (\tau )$ ,
$y' = \eta (t_{N-1})$ and
$z' = \eta (t_{N-2})$ .
-
(b)
$N = 1$ : in this case, we have
$d(\eta (\tau ), A) \leqslant \frac {1}{M} d(p, A)$ . We first pick
$x' = \eta (\tau )$ and will pick
$y'$ and
$z'$ later.
-
-
(ii)
$\operatorname {diam} \big (\pi _{A}(\eta (t_{N-1})), \pi _{A}(\eta (t_{N})\big ) \leqslant 2K_{1}$ . Then we have
\begin{align*} D_{1} = \operatorname {diam} \big (\pi _{A}(\eta (0)), \pi _{A}(\eta (t_{N})\big ) \leqslant \sum _{i=1}^{N} \operatorname {diam} \big (\pi _{A}(\eta (t_{i-1})), \pi _{A}(\eta (t_{i}))\big ) \leqslant 2K_{1} N. \end{align*}
$K'\geqslant 4K_{1}$ ,
$K' \geqslant e^{2}$ and
$e\lt 4\lt 2K+4=M$ , we deduce
\begin{align*} d(x, A) \leqslant \frac {1}{K'} e^{D_{1} / K'} d(p, A) \leqslant \frac {1}{K'} e^{N} d(p, A) \leqslant (2K+4)^{N-2} d(\eta (t_{0}) , A) \leqslant \frac {1}{M} d(\eta (t_{N-1}), A). \end{align*}
$N\geqslant 2$ , we can pick
$x' = x=\eta (t_{N})$ ,
$y'= \eta (t_{N-1})$ and
$z' = \eta (t_{N-2})$ and deduce a similar contradiction. When
$N = 1$ , we set
$x' = x$ .
So far, we have obtained either the desired triple
$(x', y', z')$
, or a point
$x' \in [x, p]$
such that

A similar discussion on
$[p, y]$
also gives either the desired triple, or a point
$z' \in [p, y]$
such that
$d(\pi _{A}(z'), \pi _{A}(p)) \leqslant 2K'$
and
$d(z', A) \leqslant \frac {1}{M}d(p, A)$
. If we fall into the latter cases in both discussions, we let
$y' = p$
and deduce the contradiction.
2.4 Random walks
Let
$\mu$
be a probability measure on a discrete group
$G$
acting on a metric space
$(X, d)$
. We denote by
$\check {\mu }$
the reflected version of
$\mu$
, which by definition satisfies
$\check {\mu }(g) := \mu (g^{-1})$
. The random walk generated by
$\mu$
is the Markov chain on
$G$
with the transition probability
$p(g, h) := \mu (g^{-1} h)$
.
Consider the step space
$(G^{\mathbb{Z}}, \mu ^{\mathbb{Z}})$
, the product space of
$G$
equipped with the product measure of
$\mu$
. Each element
$(g_{n})_{n \in \mathbb{Z}}$
of the step space is called a step path, and there is a corresponding (bi-infinite) sample path
$(Z_{n})_{n \in \mathbb{Z}}$
under the correspondence

We also introduce the notation
$\check {g}_{n} = g_{-n+1}^{-1}$
and
$\check {Z}_{n} = Z_{-n}$
. Note that we have an isomorphism
$(G^{\mathbb{Z}}, \mu ^{\mathbb{Z}}) \rightarrow (G^{\mathbb{Z}_{\gt 0}}, \mu ^{\mathbb{Z}_{\gt 0}}) \times (G^{\mathbb{Z}_{\gt 0}}, \check {\mu }^{\mathbb{Z}_{\gt 0}})$
by
$(g_{n})_{n \in \mathbb{Z}} \mapsto ((g_{n})_{n \gt 0} , (\check {g}_{n})_{n \gt 0})$
. In view of this, we sometimes write the bi-infinite sample path as
$((Z_{n})_{n\geqslant 0}, (\check {Z}_{n})_{n\geqslant 0})$
, where the distribution of
$(Z_{n})_{n}$
and
$(\check {Z}_{n})_{n}$
are independent.
In certain circumstances, it is beneficial to consider a probability space
$(\Omega , \mathbb{P})$
where the step distributions for the random walk is defined, together with some other RVs. For this purpose, we say that
$(\Omega , \mathbb{P})$
is a probability space for
$\mu$
if there is a measure-preserving map from
$(\Omega , \mathbb{P})$
to
$(G^{\mathbb{Z}_{\gt 0}}, \mu ^{\mathbb{Z}_{\gt 0}})$
, or equivalently, if independent step RVs
$\{g_{n}(\omega )\}_{n\gt 0}$
are defined and distributed according to
$\mu$
. We similarly define a probability space
$(\check {\Omega }, \mathbb{P})$
for
$\check {\mu }$
, together with RVs
$\{g_{n}(\check {\omega })\}_{n \gt 0}$
. Then the product space
$(\Omega \times \check {\Omega }, \mathbb{P})$
models the (bi-infinite) random walk generated by
$\mu$
. We often omit
$\omega$
while writing e.g.
$g_{n} = g_{n}(\omega )$
and
$Z_{n} = Z_{n}(\omega )$
. To make a distinction, we mark RVs on
$\check {\Omega }$
with the ‘check’ sign, e.g.,
$\check {g}_{n} := g_{n}(\check {\omega })$
,
$\check {Z}_{n} := Z_{n}(\check {\omega })$
.
We define the support of
$\mu$
, denoted by
$\textrm {supp} \mu$
, as the set of elements in
$G$
that are assigned nonzero values of
$\mu$
. We denote by
$\mu ^{N}$
the product measure of
$N$
copies of
$\mu$
, and by
$\mu ^{\ast N}$
the
$N$
-th convolution measure of
$\mu$
. We say that
$\mu$
is non-elementary if the subsemigroup generated by the support of
$\mu$
contains two independent strongly contracting isometries
$g, h$
of
$X$
. By taking suitable powers, we may assume that
$g$
and
$h$
belong to the same
$\textrm {supp} \mu ^{\ast N}$
for some
$N \gt 0$
.
When a constant
$M_{0}$
(to be fixed later) is understood, we use the notation

Similarly, we denote
$(Z_{i-M_{0}}(\check {\omega })o, \ldots , Z_{i}(\check {\omega }) o)$
by
$\mathbf{Y}_{i}(\check {\omega })$
.
Part I: random walks with strongly contracting isometries
In Part I, we develop a theory of random walks that involve strongly contracting isometries. The following convention is employed throughout Part I.
Convention 2.11. We assume that:
-
–
$(X, d)$ is a geodesic metric space;
-
–
$G$ is a countable group of isometries of
$X$ , and
-
–
$G$ contains two independent strongly contracting isometries.
We also fix a basepoint
$o \in X$
.
We emphasize that no further requirements (properness, WPD-ness, etc.) are imposed on
$X$
or
$G$
. Convention 2.11 includes the following situations:
-
(i)
$(X, d)$ is a geodesic Gromov hyperbolic space and
$G$ contains independent loxodromics, e.g.
-
(a)
$(X, d)$ is the curve complex of a finite-type hyperbolic surface and
$G$ is the mapping class group, or
-
(b)
$(X, d)$ is the complex of free factors of the free group of rank
$N \geqslant 3$ and
$G$ is the outer automorphism group
$\textrm {Out}(F_{N})$ ;
-
-
(ii)
$X$ is Teichmüller space of finite type,
$G$ is the corresponding mapping class group, and
$d$ is either the Teichmüller metric
$d_{\mathcal{T}}$ [Reference MinskyMin96] or the Weil-Petersson metric
$d_{WP}$ [Reference Bestvina and FujiwaraBF02];
-
(iii)
$(X, d)$ is the Cayley graph of a braid group modulo its center
$B_{n}/Z(B_{n})$ with respect to its Garside generating set, and
$G$ is the braid group
$B_{n}$ [Reference Calvez and WiestCW21];
-
(iv)
$(X, d)$ is the Cayley graph of a group
$G$ with nontrivial Floyd boundary [Reference KarlssonKar03], [Reference Gerasimov and PotyagailoGP13];
-
(v)
$(X, d)$ is the Cayley graph of a
$Gr'(1/6)$ -labeled graphical small cancellation group
$G$ [Reference Arzhantseva, Cashen, Gruber and HumeACGH19];
-
(vi)
$(X, d)$ is a (not necessarily proper nor finite-dimensional) CAT(0) space and
$G$ contains independent rank-1 isometries; e.g.,
$G$ is an irreducible right-angled Artin group and
$(X, d)$ is the universal cover of its Salvetti complex.
3 Alignment I: strongly contracting axes
In this section, we will formulate and prove the following claim. Let
$(\kappa _{i})_{i=1}^{n}$
be a sequence of long enough contracting axes. Suppose that each pair of consecutive axes is aligned:
$\kappa _{i}$
(
$\kappa _{i+1}$
, resp.) projects onto
$\kappa _{i+1}$
(
$\kappa _{i}$
, resp.) near the beginning point of
$\kappa _{i+1}$
(the ending point of
$\kappa _{i}$
, resp.). Then the axes are globally aligned:
$\kappa _{i}$
projects onto
$\kappa _{j}$
near the beginning point (ending point, resp.) of
$\kappa _{j}$
when
$i \lt j$
(
$i\gt j$
, resp.).
3.1 Contracting geodesics
The goal of this subsection is to establish Corollary 3.5. We begin by recalling a lemma that appeared as [[Reference Arzhantseva, Cashen and TaoACT15], Lemma 2.14], [[Reference SistoSis18], Lemma 2.4] and [[Reference YangYan19], Lemma 2.4(4)]. For a version with explicit constant, see [[Reference Chawla, Choi and TiozzoCCT23], Lemma 2.2].
Lemma 3.1.
Let
$A$
be a
$K$
-strongly contracting set and let
$\eta : I \rightarrow X$
be a geodesic such that
$\operatorname {diam}(\pi _{A}(\eta )) \gt K$
. Then there exist
$t \lt t'$
in
$I$
such that
$\pi _{A}(\eta )$
and
$\eta |_{[t, t']}$
are
$4K$
-coarsely equivalent, and moreover, such that

Lemma 3.2.
For each
$K\gt 1$
there exists
$K'=K'(K) \gt K$
that satisfies the following.
Let
$\eta : J \rightarrow X$
be a
$K$
-quasigeodesic whose endpoints are
$x$
and
$y$
, let
$A$
be a subset of
$\eta$
such that
$d(x, A) \lt K$
,
$d(y, A) \lt K$
, and let
$\gamma : J' \rightarrow X$
be a geodesic that is
$K$
-coarsely equivalent to
$A$
. Then
$\eta$
and
$\gamma$
are also
$K'$
-coarsely equivalent, and moreover, there exists a
$K'$
-quasi-isometry
$\varphi : J \rightarrow J'$
such that
$d(\eta (t), (\gamma \circ \varphi )(t)) \lt K'$
for each
$t \in J$
.
Proof.
Without loss of generality, let
$J = [a, b]$
,
$J' = [c, d]$
and
$\eta (a) = x$
,
$\eta (b) = y$
. For each
$s \in J'$
, we can pick
$t_{s} \in J$
such that
$d(\gamma (s), \eta (t_{s})) \lt K$
as
$\gamma$
is coarsely contained in
$A$
. Note that

Similarly,
$|t_{s_{1}} - t_{s_{2}}| \geqslant \frac {1}{K} |s_{1} - s_{2}| - 1 - 2K$
holds. Hence,
$s \mapsto t_{s}$
is a
$3K^{2}$
-quasi-isometric embedding.
It remains to show that
$\{t_{s} : s \in J'\}$
is coarsely equivalent to
$J$
. Note that
$A$
is
$3K$
-coarsely connected, as it is
$K$
-coarsely contained in an
$1$
-connected set
$\gamma$
. It follows that
$\eta ^{-1}(A)$
is
$4K^{2}$
-coarsely connected subset of
$[a, b]$
. Moreover, since
$x$
and
$y$
are
$K$
-close to
$A$
, we have
$d(a, \eta ^{-1}(A)), d(b, \eta ^{-1}(A)) \lt 2K^{2}$
. Combined together,
$[a, b]$
is
$4K^{2}$
-coarsely contained in
$\eta ^{-1}(A)$
.
Next, for each
$p \in A$
there exists
$s \in J'$
such that
$d(\gamma (s), p) \lt K$
. This implies
$d(\eta (t_{s}), p) \lt 2K$
and
$\operatorname {diam}(t_{s}, \eta ^{-1}(p)) \lt 3K^{2}$
. Hence,
$\eta ^{-1}(A)$
is
$3K^{2}$
-coarsely contained in
$\{t_{s} : s \in J'\}$
.
$K$
-quasi-isometries between intervals are
$K'$
-coarsely equivalent to a monotone map for some
$K'=K'(K)$
. (for an explicit
$K'$
, see the proof of [[Reference SankaranSan06], Theorem 1.2]). Hence, we have:
Corollary 3.3.
For each
$K\gt 1$
there exists
$K'=K'(K) \gt K$
that satisfies the following.
Let
$\eta : J \rightarrow X$
be a
$K$
-quasigeodesic connecting
$x$
to
$y$
, let
$A$
be a subset of
$\eta$
such that
$d(x, A) \lt K$
and
$d(y, A) \lt K$
, and let
$\gamma : J' \rightarrow X$
be a geodesic that is
$K$
-coarsely equivalent to
$A$
. Then
$\eta$
and
$\gamma$
are
$K'$
-fellow traveling.
Combining Lemma 3.1 and Lemma 3.2, we observe an instance of the Morseness of contracting axes ([[Reference Arzhantseva, Cashen, Gruber and HumeACGH17], Theorem 1.3], [[Reference SistoSis18], Lemma 2.8.(2)], [[Reference YangYan14], Lemma 2.2]).
Corollary 3.4.
For each
$K\gt 1$
there exists a constant
$K' \gt K$
that satisfies the following. Let
$\eta : J \rightarrow X$
be a
$K$
-contracting axis and
$\gamma : J' \rightarrow X$
be a geodesic that share the endpoints. Then
$\eta$
and
$\gamma$
are
$K'$
-fellow traveling.
Corollary 3.5.
For each
$K\gt 1$
there exists a constant
$K' = K'(K)$
that satisfies the following.
Let
$\kappa : I \rightarrow X$
and
$\eta : J \rightarrow X$
be
$K$
-contracting axes. Suppose that
$\operatorname {diam}(\pi _{\kappa }(\eta )) \gt K'$
. Then there exist
$t \lt t'$
in
$I$
and
$s \lt s'$
in
$J$
such that the following sets are all
$K'$
-coarsely equivalent:

Moreover, we have

Proof.
For simplicity, we focus on the case where
$\kappa$
,
$\eta$
have endpoints.
Let
$\gamma : J' \rightarrow X$
be a geodesic that connects the endpoints of
$\eta$
. Then
$\gamma$
and
$\eta$
are coarsely equivalent by Corollary 3.4. Lemma 2.2 tells us that
$\pi _{\kappa }(\gamma )$
is coarsely equivalent to
$\pi _{\kappa }(\eta )$
and hence large. By Lemma 3.1, there exist
$u \lt u'$
in
$J'$
such that
$\pi _{\kappa }(\gamma )$
and
$\gamma |_{[u, u']}$
are coarsely equivalent and such that
$\gamma |_{(-\infty , u]}$
and
$\gamma |_{[u', +\infty )}$
project onto
$\kappa$
near
$\gamma (u)$
and
$\gamma (u')$
, respectively.
Note again that
$\eta$
and
$\gamma$
are fellow traveling by Corollary 3.4 and
$\pi _{\kappa }$
is coarsely Lipschitz. This enables us to replace
$\gamma$
with
$\eta$
: there exist
$s \lt s'$
in
$J$
such that
$\pi _{\kappa }(\eta )$
and
$\eta |_{[s, s']}$
are coarsely equivalent and such that
$\eta |_{(-\infty , s]}$
and
$\eta |_{[s', +\infty )}$
project onto
$\kappa$
near
$\eta (s)$
and
$\eta (s')$
, respectively.
Since
$\pi _{\kappa }(\eta ) \subseteq \kappa$
and
$\eta |_{[s, s']}$
are nearby, each point
$\eta (t)$
in
$\eta |_{[s, s']}$
is near a point
$\kappa (s_{t})$
of
$\kappa$
. This
$\kappa (s_{t})$
projects onto
$\eta$
near
$\eta (t)$
. It follows that
$\pi _{\eta }(\kappa )$
coarsely contains
$\eta |_{[s, s']}$
and hence
$\pi _{\kappa }(\eta )$
.
This implies that
$\pi _{\eta }(\kappa )$
is also large, and we have another round: there exist
$t \lt t'$
in
$I$
such that
$\pi _{\eta }(\kappa )$
and
$\kappa |_{[t, t']}$
are coarsely equivalent. Moreover,
$\pi _{\kappa }(\eta )$
coarsely contains
$\pi _{\eta }(\kappa )$
. Hence, the two projections are coarsely equivalent, and

are all coarsely equivalent.
We now digress to the proof of Lemma 2.7.
Proof of Lemma 2.7. Let
$\eta$
and
$\kappa$
denote the axes of
$g$
and
$h$
, i.e.,
$\eta : i \mapsto g^{i} o$
and
$\kappa : j \mapsto h^{j} o$
. Let
$\eta$
and
$\kappa$
be
$K$
-contracting axes for some
$K\gt 0$
.
Suppose that
$\pi _{\kappa }(\eta )$
has finite diameter, i.e., there exists
$M$
such that

Then for each
$i \in \mathbb{Z}$
and
$|j|\gt M+2K^{2}$
, the diameter of
$\pi _{\kappa }(\eta (i) ) \cup \kappa (j)$
is greater than
$K$
and
$[\eta (i), \kappa (j)]$
is
$2K$
-close to
$\pi _{\kappa }(\eta (i))$
. This forces that

Similarly, if
$\pi _{\eta }(\kappa )$
has finite diameter, then there exists
$M'$
such that
$d(\eta (i), \kappa (j)) \geqslant \frac {1}{K} |i| - M'$
holds for all
$j$
and
$|i| \gt M'$
. Hence
$d(g^{i} o, h^{j} o)$
is a proper function, and
$g$
and
$h$
are independent.
Now suppose that
$\pi _{\kappa }(\eta )$
has infinite diameter. By Corollary 3.5,
$\eta$
and
$\kappa$
have subpaths
$\eta '$
and
$\kappa '$
, respectively, that are coarsely equivalent to
$\pi _{\kappa }(\eta )$
, of infinite diameter. This means that
$\eta$
and
$\kappa$
are not independent.
3.2 Alignment
Let us now define the notion of alignment.
Definition 3.6. For
$i=1, \ldots , n$
, let
$\kappa _{i}$
be a path on
$X$
whose beginning and ending points are
$x_{i}$
and
$y_{i}$
, respectively. We say that
$(\kappa _{1}, \ldots , \kappa _{n})$
is
$C$
-aligned if

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

Figure 2. Schematics for an aligned sequence of paths.
Note that if
$(\kappa _{i}, \ldots , \kappa _{j})$
and
$(\kappa _{j}, \ldots , \kappa _{k})$
are
$C$
-aligned, then
$(\kappa _{i}, \ldots , \kappa _{j}, \ldots , \kappa _{k})$
is also
$C$
-aligned. We allow degenerate paths, e.g., the case where
$\kappa _{1}$
or
$\kappa _{n}$
is a point.
Combining Lemma 3.1 and Corollary 3.3, we obtain the following consequence of alignment.
Corollary 3.7.
For each
$C, K\gt 1$
, there exists
$K' = K'(K, C) \gt \max (K, C)$
such that the following holds.
Let
$x, y \in X$
and let
$\kappa$
be a
$K$
-contracting axis such that
$\operatorname {diam}(\kappa ) \gt K+2C$
and such that
$(x, \kappa , y)$
is
$C$
-aligned. Then
$[x, y]$
contains a subsegment
$\eta$
that is
$4K$
-coarsely contained in
$\kappa$
and is
$K'$
-fellow traveling with
$\kappa$
.
Our first lemma states that the alignment of two strongly contracting axes is governed by the projections of their endpoints to the other axis.
Lemma 3.8.
For each
$C, K\gt 1$
, there exists
$D= D(K, C)\gt \max (K, C)$
such that the following holds.
Let
$\kappa , \eta$
be
$K$
-contracting axes. If
$\big (\kappa , (\textrm {beginning point of}\eta )\big )$
and
$\big ((\textrm {beginning point of}\kappa ), \eta \big )$
are each
$C$
-aligned, then
$(\kappa , \eta )$
is
$D$
-aligned.
Proof.
For simplicity, let us assume that the domains of
$\kappa$
and
$\eta$
are closed intervals, say,
$I = [t_{0}, t_{1}]$
and
$J = [s_{0}, s_{1}]$
, respectively.
It suffices to show that
$\pi _{\kappa }(\eta )$
and
$\pi _{\eta }(\kappa )$
are both small. Suppose not. Then Corollary 3.5 provides
$t \lt t'$
in
$I$
and
$s \lt s'$
in
$J$
such that

are all coarsely equivalent and large. Moreover,
$\pi _{\eta }(\kappa (t_{0}))$
is near
$\kappa (t)$
and
$\pi _{\eta }(\kappa (t_{1}))$
is near
$\kappa (t')$
. Similarly,
$\pi _{\kappa }(\eta (s_{0}))$
is near
$\eta (s)$
and
$\pi _{\kappa }(\eta (s_{1}))$
is near
$\eta (s')$
.
Since
$\pi _{\kappa }(\eta )$
and
$\pi _{\eta }(\kappa )$
are large, both
$t' - t$
and
$s' - s$
are large. Since
$[\kappa (t), \kappa (t')]$
and
$[\eta (s), \eta (s')]$
are coarsely equivalent, one of the following is true:
-
–
$\kappa (t)$ is near
$\eta (s)$ and
$\kappa (t')$ is near
$\eta (s')$ ; or,
-
–
$\kappa (t)$ is near
$\eta (s')$ and
$\kappa (t')$ is near
$\eta (s)$ .
This leads to the following contradictions:
-
– If
$\kappa (t)$ is near
$\eta (s)$ , then
$\eta (s_{0})$ projects onto
$\kappa$ near
$\kappa (t)$ . Since
$t_{1} - t \geqslant t' - t$ is large, this projection cannot be near
$\kappa (t_{1}) = y$ .
-
– If
$\kappa (t)$ is near
$\eta (s')$ , then
$\kappa (t_{0})$ projects onto
$\eta$ near
$\eta (s')$ . Since
$s' - s_{0} \geqslant s' - s$ is large, this projection cannot be near
$\eta (s_{0}) = x'$ .
Hence,
$\pi _{\kappa }(\eta )$
and
$\pi _{\eta }(\kappa )$
cannot be large and the conclusion follows.
The following lemma was inspired by Behrstock’s inequality for subsurface projections and curve complexes [[Reference BehrstockBeh06], Theorem 4.3].
Lemma 3.9 ([[Reference SistoSis18], Lemma 2.5]). For each
$D, K\gt 1$
, there exists
$E = E(K, D)\gt \max ( K, D)$
that satisfies the following.
Let
$\kappa$
,
$\eta$
be
$K$
-contracting axes in
$X$
. Suppose that
$(\kappa , \eta )$
is
$D$
-aligned. Then for any
$p \in X$
, either
$(p, \eta )$
is
$E$
-aligned or
$(\kappa , p)$
is
$E$
-aligned.
We are now ready to prove the main result of this section.
Proposition 3.10.
For each
$D, K\gt 1$
, there exist
$E = E(K, D)\gt \max (K, D)$
and
$L = L(K,D)\gt \max (K, D)$
that satisfy the following.
Let
$x, y \in X$
and let
$\kappa _{1}, \ldots , \kappa _{n}$
be
$K$
-contracting axes whose domains are longer than
$L$
. Suppose that
$(x, \kappa _{1}, \ldots , \kappa _{n}, y)$
is
$D$
-aligned. Then
$(x, \kappa _{i}, y)$
is
$E$
-aligned for each
$i$
.
Proof.
Let
$E = E(K, D)$
be as in Lemma 3.9 and let
$L = 3KE + K^{2}$
. Our claim is that
$(x, \kappa _{i})$
and
$(\kappa _{i}, y)$
are
$E$
-aligned for each
$i$
. By symmetry, it suffices to prove the alignment of
$(x, \kappa _{i})$
.
Let
$\kappa$
be a
$K$
-contracting axis whose domain is longer than
$L$
. Then the endpoints of
$\kappa$
are at least
$3E$
-apart. Consequently, no point
$p$
in
$X$
satisfy the following at the same time:

From this observation, we inductively deduce

for
$i=1, \ldots , n$
, where the latter implication follows from Lemma 3.9.
The above proposition can be strengthened as follows. First, we record an immediate consequence of the definition of fellow-traveling.
Lemma 3.11.
Let
$E\gt 0$
and
$x, y \in X$
. Let
$\kappa$
be a path that
$E$
-fellow travels with a subsegment of
$[x, y]$
. Then
$(x, \kappa , y)$
is
$4E$
-aligned.
Proposition 3.12.
For each
$D, K\gt 1$
, there exist
$E = E(K, D)\gt \max (K, D)$
and
$L = L(K,D)\gt \max (K, D)$
that satisfy the following.
Let
$x, y \in X$
and let
$\kappa _{1}, \ldots , \kappa _{n}$
be
$K$
-contracting axes whose domains are longer than
$L$
and such that
$(x, \kappa _{1}, \ldots , \kappa _{n}, y)$
is
$D$
-aligned. Then the geodesic
$[x, y]$
has subsegments
$\eta _{1}, \ldots , \eta _{n}$
, in order from left to right, that are longer than
$100E$
and such that
$\eta _{i}$
and
$\kappa _{i}$
are
$0.1E$
-fellow traveling for each
$i$
. In particular,
$(x, \kappa _{i}, y)$
are
$E$
-aligned for each
$i$
.
Proof.
Let
$E_{1} = E(K, D)$
and
$L_{1} = L(K, D)$
be as in Proposition 3.10. Let
$K_{1} = E_{1} + 8K$
and let
$E = 10 K'(K, K_{1})$
, where
$K'(K, K_{1})$
is as in Corollary 3.7. Let also
$L = L_{1} + 101K(K+E) + 2K$
.
We will inductively prove a variant of the given statement, namely:
If
$(x, \kappa _{1})$
is
$K_{1}$
-aligned and
$(\kappa _{1}, \ldots , \kappa _{n}, y)$
is
$D$
-aligned, then the conclusion holds.
First, we know that
$(\kappa _{1}, y)$
is
$E_{1}$
-aligned by Proposition 3.10. Since
$(x, \kappa _{1}, y)$
is
$K_{1}$
-aligned and
$\kappa _{1}$
is long enough, Corollary 3.7 provides a subsegment
$\eta _{1} = [x_{1}', y_{1}']$
of
$[x, y]$
that is
$4K$
-coarsely contained in
$\kappa _{1}$
and is
$0.1E$
-fellow traveling with
$\kappa _{1}$
. We then have

If
$n=1$
, this finishes the proof. If not, note that
$y_{1}'$
is
$4K$
-close to
$\kappa _{1}$
. Lemma 2.2 implies that
$(y_{1}', \kappa _{2})$
is
$(D + 8K)$
-aligned, and hence
$K_{1}$
-aligned. Now the induction hypothesis implies that
$[y_{1}', y]$
has subsegments
$\eta _{2}, \ldots , \eta _{n}$
, in order from left to right, that are longer than
$100E$
and such that
$\eta _{i}$
and
$\kappa _{i}$
$0.1E$
-fellow travel for
$i\geqslant 2$
. Then
$\eta _{1}, \ldots , \eta _{n}$
become the desired subsegments.
Using Proposition 3.12, we can recover the following results by Yang.
Lemma 3.13 ([[Reference YangYan14], Lemma 4.4], [[Reference YangYan19], Proposition 2.9]). For each
$D, M\gt 0$
and
$K\gt 1$
, there exist
$E = E(K, D, M)\gt D$
and
$L = L(K,D)\gt D$
that satisfies the following.
Let
$\kappa _{1}, \ldots , \kappa _{n}$
be
$K$
-contracting axes whose domains are longer than
$L$
. Suppose that
$(\kappa _{1}, \ldots , \kappa _{n})$
is
$D$
-aligned and
$d(\kappa _{i}, \kappa _{i+1}) \lt M$
for each
$i$
. Then the concatenation
$\kappa _{1} \cup \ldots \cup \kappa _{n}$
of
$\kappa _{1}, \ldots , \kappa _{n}$
is an
$E$
-contracting axis.
Lemma 3.14 ([[Reference YangYan14], Corollary 3.2]). For each
$D\gt 0$
and
$K\gt 1$
, there exist
$E = E(K, D)\gt D$
and
$L = L(K,D)\gt D$
that satisfies the following.
For each
$i \in \mathbb{Z}$
, let
$\kappa _{i}$
be a
$K$
-contracting axis whose beginning and ending points are
$x_{i}$
and
$y_{i}$
, respectively, and whose domain is longer than
$L$
. Suppose that
$(\ldots , \kappa _{i}, \kappa _{i+1}, \ldots )$
is
$D$
-aligned. Then the concatenation of
$(\ldots , [x_{i-1}, y_{i-1}], [y_{i-1}, x_{i}], [x_{i}, y_{i}], [y_{i}, x_{i+1}], \ldots )$
is an
$E$
-quasigeodesic.
3.3 Schottky sets
Using the previous concatenation lemmata, we will construct arbitrarily many independent contracting isometries. Recall again the notation introduced in Subsection 2.1.
Definition 3.15 (cf. [[Reference GouëzelGou22], Definition 3.11]). Let
$K \gt 0$
and let
$S \subseteq G^{n}$
be a set of sequences of isometries. We say that
$S$
is
$K$
-Schottky if:
-
(i)
$\Gamma ^{+}(s)$ and
$\Gamma ^{-}(s)$ are
$K$ -contracting axes for all
$s \in S$ ;
-
(ii) for each
$x \in X$ we have
\begin{align*} \#\Big \{ s \in S : \big (x, \Gamma ^{+}(s)\big ) \textrm {and} \big (x, \Gamma ^{-}(s)\big ) \textrm {are} K-\textrm {aligned}\Big \} \geqslant \# S - 1; \end{align*}
-
(iii) for each
$s \in S$ ,
$\big (\bar {\Gamma }^{-}(s), \Gamma ^{+}(s)\big )$ is
$K$ -aligned.
Once a Schottky set
$S$
is understood, its element
$s$
is called a Schottky sequence and the translates of
$\Gamma ^{\pm }(s)$
are called Schottky axes. We say that
$S$
is large enough if its cardinality is at least 400.
Let
$\mu$
be a probability measure on
$G$
. If each element
$s$
of
$S$
is attained by the product measure of
$\mu$
, i.e.,
$S \subseteq (\textrm {supp} \mu )^{n}$
, then we say that
$S$
is a Schottky set for
$\mu$
.
An intuitive example was given in the introduction. Consider
$S_{M}:= \big \{ s_{1}s_{2} \cdots s_{M} : s_{i} \in \{a, b\} \big \}$
in
$F_{2}= \langle a, b \rangle$
. For any infinite ray on
$F_{2}$
, at most 1 element
$s \in S_{m}$
heads into the direction:

for each infinite ray
$\xi$
. Moreover,
$s$
and
$s^{-1}$
diverge early for any
$s \in S_{M}$
:

These properties are also satisfied by the set of
$N$
-th powers of elements of
$S_{M}$

Definition 3.16. Given a constant
$K_{0}\gt 0$
, we define:
-
–
$D_{0}= D(K_{0}, K_{0})$ be as in Lemma 3.8,
-
–
$E_{0}= E(K_{0}, D_{0})$ ,
$L_{0} = L(K_{0}, D_{0})$ be as in Proposition 3.12.
A
$K_{0}$
-Schottky set
$S$
whose elements have domains longer than
$L_{0}$
is called a long enough
$K_{0}$
-Schottky set. In other words, when
$S \subseteq G^{n}$
is
$K_{0}$
-Schottky and
$n \gt L_{0}$
,
$S$
is called a long enough
$K_{0}$
-Schottky set. In this case, note that the endpoints of
$\Gamma ^{+}(s)$
are
$100E_{0}$
-apart for each
$s \in S$
.
This definition is motivated by the alignment lemmata. Note that the
$D_{0}$
-alignment of a sequence of Schottky axes
$(\gamma _{1}, \ldots , \gamma _{N})$
is a local condition, between consecutive pairs of axes. Proposition 3.12 then promotes this into the global alignment, i.e., the
$E_{0}$
-alignment of
$(\gamma _{i}, \gamma _{j})$
for any
$i \lt j$
, given that the involved Schottky set is long enough. The following definition is designed to capture this local-to-global phenomenon.
Definition 3.17. Let
$S$
be a Schottky set, let
$x, y\in X$
and let
$\kappa _{1}, \ldots , \kappa _{N}$
be Schottky axes. We say that
$(x, \kappa _{1}, \ldots , \kappa _{N}, y)$
is
$C$
-semi-aligned if it is a subsequence of a
$C$
-aligned sequence of
$x$
,
$y$
and Schottky axes, i.e., if there exist Schottky axes
$\eta _{1}, \ldots , \eta _{N'}$
and
$1 \leqslant i(1) \lt \ldots \lt i(N) \leqslant N'$
such that:
-
(i)
$(x, \eta _{1}, \ldots , \eta _{N'}, y)$ is
$C$ -aligned,
-
(ii)
$\kappa _{k} = \eta _{i(k)}$ for
$k = 1, \ldots , N$ .
Here, we also say that
$(x, \kappa _{1}, \ldots , \kappa _{N})$
and
$(\kappa _{1}, \ldots , \kappa _{N}, y)$
are
$C$
-semi-aligned.
Lemma 3.18.
Let
$S$
be a long-enough
$K_{0}$
-Schottky set. Let
$x, y \in X$
, and for each
$i=1, \ldots , N$
, let
$\kappa _{i}$
be a Schottky axis whose beginning and ending points are
$x_{i}$
and
$y_{i}$
, respectively.
-
(i) If
$(x_{1}, \kappa _{2})$ and
$(\kappa _{1}, x_{2})$ are
$K_{0}$ -aligned, then
$(\kappa _{1}, \kappa _{2})$ is
$D_{0}$ -aligned.
-
(ii) If
$(x, \kappa _{1}, \ldots , \kappa _{N}, y)$ is
$D_{0}$ -semi-aligned, then
$(\kappa _{i}, \kappa _{j})$ is
$E_{0}$ -aligned for each
$i\lt j$ . Moreover,
$\kappa _{i}$ is
$0.1E_{0}$ -coarsely contained in
$[x, y]$ and
$(x, \kappa _{i}, y)$ is
$E_{0}$ -aligned for each
$i$ . We also have
\begin{align*} \begin{aligned} d(x, x_{1}) + \sum _{i=1}^{N} d(x_{i}, y_{i}) + \sum _{i=1}^{N-1} d(y_{i}, x_{i+1}) + d(y_{N}, y) &\leqslant d(x, y) + E_{0} N, \\ d(x, x_{1}) + \sum _{i=1}^{N-1} d(y_{i}, x_{i+1}) + d(y_{N}, y) &\leqslant d(x, y) - 50E_{0} N. \end{aligned} \end{align*}
Proof.
(1) By Lemma 3.8. (2) Proposition 3.12 explains the first two claims in Item (ii). More explicitly,
$[x, y]$
contains subsegments
$[x_{1}', y_{1}']$
,
$\ldots$
,
$[x_{N}', y_{N}']$
, in order from left to right, such that:
-
(a)
$[x_{i}', y_{i}']$ and
$\kappa _{i}$ are
$0.1E_{0}$ -coarsely equivalent;
-
(b)
$d(x_{i}', x_{i}) \lt 0.1E_{0}$ and
$d(y_{i}', y_{i}) \lt 0.1E_{0}$ ;
-
(c)
$d(x_{i}', y_{i}') \gt 100E_{0}$
for each
$i$
. This implies that

We now associate long enough and large Schottky sets with non-elementary measures.
Proposition 3.19 (cf. [[Reference GouëzelGou22], Proposition 3.12]). Let
$\mu$
be a non-elementary probability measure on
$G$
. Then for each
$N\gt 0$
, there exists
$K = K(N) \gt 0$
such that for each
$L\gt 0$
there exists a
$K$
-Schottky set of cardinality
$N$
in
$(\textrm {supp} \mu )^{n}$
for some
$n \gt L$
.
Proof.
Since
$\mu$
is non-elementary, the semigroup generated by
$\textrm {supp} \mu$
contains independent strongly contracting isometries
$a$
and
$b$
. By taking suitable powers, we may assume that
$a = \Pi (\alpha )$
and
$b = \Pi (\beta )$
for some
$\alpha , \beta \in (\textrm {supp} \mu )^{L_{0}}$
for some
$L_{0}\gt 0$
. There exists
$K_{0} \gt 0$
such that:
-
(i)
$\Gamma ^{+}(\alpha )$ ,
$\Gamma ^{-}(\beta )$ are
$K_{0}$ -contracting axes, and
-
(ii)
$\operatorname {diam}(o \cup \pi _{\gamma }(\eta )) \lt K_{0}$ for distinct axes
$\gamma$ ,
$\eta$ among
\begin{align*} \Gamma ^{+}(\alpha ),\,\,\Gamma ^{-}(\alpha ),\,\,\Gamma ^{+}(\beta ), \,\,\Gamma ^{-}(\beta ). \end{align*}
The above statements still hold with the same
$K_{0}$
when
$\alpha$
and
$\beta$
are replaced with their self-concatenations, thanks to Lemma 2.3 and Lemma 2.7. Let:
-
–
$K_{1} = E(K_{0}, K_{0}) \gt K_{0}$ be as in Lemma 3.9;
-
–
$K_{2} = E(K_{0}, K_{1}) \gt K_{1}$ ,
$L_{2} = L(K_{0}, K_{1})$ be as in Proposition 3.12;
-
–
$K_{3} = E(K_{0}, K_{2}) \gt K_{2}$ ,
$L_{3} = L(K_{0}, K_{2})$ be as in Proposition 3.12;
-
–
$L_{3}' = 3K_{0}K_{3}$ ;
-
–
$K_{4} = E(K_{0}, K_{0}, 0)$ ,
$L_{4} = L(K_{0}, K_{0})$ be as in Lemma 3.13.
By self-concatenating
$\alpha$
and
$\beta$
if necessary, we may assume that

Since
$\Gamma ^{+}(\alpha )$
is a
$K_{0}$
-quasigeodesic whose domain
$L_{0}$
-long, the endpoints of
$\Gamma ^{+}(\alpha )$
are at least
$(L_{0}/K_{1} - K_{1})$
-apart. Since
$L_{0}$
is greater than
$3K_{0} K_{3} \geqslant 2K_{0} K_{1} + K_{0}^{2}$
, the endpoints of
$\Gamma ^{+}(\alpha )$
are
$2K_{1}$
-far. In particular, no set
$A \subseteq X$
can be simultaneously contained in the
$K_{1}$
-neighborhoods of the two endpoints of
$\Gamma ^{+}(\alpha )$
. Hence, the statements

are mutually exclusive for any
$x \in X$
. Similarly, the statements

are mutually exclusive.
Let
$S_{0}$
be the set of sequences of
$NL_{0}$
isometries that are concatenations of
$\alpha$
’s and
$\beta$
’s, i.e.,

Note that
$\#S_{0} = 2^{N}$
is greater than
$N$
. We claim that for each
$m\gt 0$
, the set

is a
$((K_{0} L_{0} + K_{0})N + K_{2} + K_{4})$
-Schottky set.
Step 1: Investigating
$\Gamma ^{m}(s)$
.
Pick
$s =(\phi _{1}, \ldots , \phi _{NL_{0}})$
and
$s' = (\phi _{1}', \ldots , \phi _{NL_{0}}')$
in
$S_{0}$
. Recall the notation

for
$n \in \mathbb{Z}$
and
$i = 0, \ldots , NL_{0}-1$
. We now define “sub-axes”

for each
$i \gt 0$
. These are translates of
$\Gamma ^{\pm }(\alpha )$
and
$\Gamma ^{\pm }(\beta )$
. Our initial choices of
$K_{0}$
and
$L_{0}$
guarantee that:
-
–
$\Gamma _{i}(s)$ is a
$K_{0}$ -contracting axis whose domain is longer than
$L_{2}, L_{3}, L_{3}'$ and
$L_{4}$ for each
$i \in \mathbb{Z}$ ;
-
–
$(\Gamma _{i}(s), \Gamma _{i+1}(s))$ and
$(\Gamma _{-i}(s'), \Gamma _{-(i+1)}(s'))$ are
$K_{0}$ -aligned for each
$i \gt 0$ . Moreover,
$(\bar {\Gamma }_{-1}(s'), \Gamma _{1}(s))$ is
$K_{0}$ -aligned.
Lemma 3.13 tells us that
$\cup _{i \gt 0} \Gamma _{i}(s)$
is a
$K_{4}$
-contracting axis. In particular,
$\Gamma ^{m}(s)$
is a
$K_{4}$
-contracting axis for each
$m\gt 0$
. Similarly,
$\Gamma ^{-m}(s')$
is a
$K_{4}$
-contracting axis for each
$m\gt 0$
.
Now note that the following sequence of sub-axes is
$K_{0}$
-aligned:

Let
$i \gt 0$
and let
$p \in \Gamma _{-i} (s')$
. Then Proposition 3.12 tells us that
$d\big (p, \Gamma _{1}(s)\big ) \lt d\big (p, \Gamma _{j}(s)\big )$
for each
$j\gt 1$
and that
$\big (p, \Gamma _{1}(s)\big )$
is
$K_{2}$
-aligned. It follows that
$\big (p, \cup _{i \gt 0} \Gamma _{i}(s)\big )$
is
$K_{2}$
-aligned. For this reason (and its symmetric counterpart),
$(\bar {\Gamma }^{-m}(s'), \Gamma ^{m}(s))$
is
$K_{2}$
-aligned for each
$m\gt 0$
.
Next, fix
$x \in X$
and consider the condition

If Condition 5 holds, then for each
$i\gt N$

is
$K_{2}$
-aligned and
$d(x, \Gamma _{N}(s)) \lt d(x, \Gamma _{i}(s))$
holds by Proposition 3.12. Hence,
$\pi _{\cup _{i \gt 0} \Gamma _{i}(s)}(x)$
is contained in
$\Gamma _{1}(s) \cup \cdots \cup \Gamma _{N}(s)$
. Meanwhile, recall that for each
$i$
,
$\Gamma _{i}(s)$
is a
$K_{0}$
-quasigeodesic whose domain is
$L_{0}$
-long. Hence, we have

Combining these ingredients, we observe that

holds for every
$m\gt 0$
. For a similar reason, the condition

implies
$\operatorname {diam}\left ( \pi _{\Gamma ^{m}(s)}(x) \cup o \right ) \leqslant (K_{0} L_{0} + K_{0})N$
for all
$m \lt 0$
. In summary,
Observation 3.20.
If
$s \in S_{0}^{(m)}$
satisfies Condition
5
and
6
, then
$\big (x, \Gamma ^{m}(s)\big )$
is
$(K_{0} L_{0} + K_{0})N$
-aligned for all
$m \in \mathbb{Z}$
.
Step 2. Comparing two distinct axes.
We now pick
$m\gt 0$
and consider an element of
$S_{0}^{(m)}$
which violates these conditions.
Observation 3.21.
If
$s = (\phi _{1}, \ldots , \phi _{mNL_{0}}) \in S_{0}^{(m)}$
violates Condition
5
, then all the other elements
$s' = (\phi _{1}', \ldots , \phi _{mNL_{0}}') \in S_{0}^{(m)}$
satisfy Condition
5
and Condition
6
.
To show this, let
$k \in \{1, \ldots , N\}$
be the first index such that
$(\phi _{L_{0}(k-1)+1}, \ldots , \phi _{L_{0}k})$
and
$(\phi _{L_{0}(k-1) + 1}', \ldots , \phi _{L_{0}k}')$
differ. Let us denote
$x_{i}(s)$
by
$x_{i}$
and
$x_{i}(s')$
by
$x_{i}'$
. Note that the path

is the concatenation of
$K_{0}$
-aligned
$K_{0}$
-contracting axes

Recall that
$s$
violates Condition 5:
$(\bar {\Gamma }_{N}(s), x) = (\eta _{1}, x)$
is not
$K_{2}$
-aligned. Since
$(\eta _{1}, \eta _{2})$
is
$K_{0}$
-aligned, Lemma 3.9 tells us that
$(x, \eta _{2})$
is
$K_{1}$
-aligned. Then
$(x, \eta _{2}, \ldots , \eta _{2(N-k+1)})$
is
$K_{1}$
-aligned and Proposition 3.12 tells us that
$(x, \eta _{2(N-k+1)}) = (x, \Gamma _{N}(s'))$
is
$K_{2}$
-aligned. Hence,
$s'$
satisfies Condition 5.
Similarly, by considering the
$K_{0}$
-aligned sequence

we can deduce that
$(x, \Gamma _{-N}(s'))$
is
$K_{2}$
-aligned as desired.
A similar argument leads to the following.
Observation 3.22.
If
$s \in S_{0}^{(m)}$
violates Condition
6
, then all the other elements in
$S_{0}^{(m)}$
satisfy Condition
5
and Condition
6
.
Step 3: Summary.
We claim that
$S_{0}^{(m)}$
is
$((K_{0} L_{0} + K_{0})N + K_{2}+K_{4})$
-Schottky. The first and the third requirements for Schottky sets were already observed before, so it remains to discuss the second requirement. Considering Observation 3.20, it suffices to show that Condition 5 and Condition 6 are satisfied by all but at most 1 element of
$S_{0}^{(m)}$
. Observation 3.21 and 3.22 imply that this is the case.
Given these observations, we can finish the proof by taking
$K = (K_{0} L + K_{0} )N + K_{2} + K_{4}$
,
$m =L$
and by taking any subset
$S \subseteq S_{0}^{(m)}$
such that
$\#S = N$
.
4 Pivoting and limit laws
In this section, we establish the notion of pivotal times and pivoting. We will then deduce CLT, LIL and geodesic tracking of random walks using probabilistic estimates about pivotal times. The proof of a key probabilistic estimate will be postponed to Section 5.
4.1 Pivotal times: statement
Let
$\mu$
be a non-elementary probability measure on
$G$
and let
$S$
be a long enough and large Schottky set for
$\mu$
. Then for sufficiently small
$\epsilon \gt 0$
, an
$n$
-step random path
$(g_{1}, \ldots , g_{n})$
in the
$\mu$
-random walk contains at least
$\epsilon n$
subsegments

The appearance of Schottky sequences in a random path does not necessarily imply something about
$Z_{n} = g_{1} \cdots g_{n}$
. For example, every Schottky sequence might be cancelled out with the next step, resulting in
$Z_{n} = id$
. We nonetheless claim that for a high probability, certain number of Schottky axes survive. More explicitly, we seek indices
$j(1) \lt \ldots \lt j(M)$
, called the pivotal times, such that the Schottky axes arising at these indices are aligned along
$[o, Z_{n} o]$
:

We will observe that for a high probability, a random path has sufficiently many pivotal times. Then, we will freeze the steps except at the pivotal slots and choose the Schottky sequences at the pivotal times from
$S$
. More explicitly, we will realize a structure where
$\mathbf{Y}_{j(k)}$
’s are i.i.d.s on the uniform measure on
$\{\Gamma (s) : s \in S\}$
: once this is guaranteed, we can control the direction
$[o, Z_{n} o]$
and establish the deviation inequality.
We now formulate the discussion above.
Definition 4.1. Let
$\mu$
be a non-elementary probability measure on
$G$
, let
$(\Omega , \mathbb{P})$
be a probability space for
$\mu$
, let
$K_{0}, M_{0}\gt 0$
and let
$S$
be a long enough
$K_{0}$
-Schottky set contained in
$(\textrm {supp} \mu )^{M_{0}}$
, i.e.,
$M_{0}$
is as large as described in Definition 3.16.
A subset
$\mathcal{E}$
of
$\Omega$
, accompanied with the choice of a subset
$\mathcal{P}(\mathcal{E}) = \{j(1) \lt j(2) \lt \ldots \}\subseteq M_{0} \mathbb{Z}_{\gt 0}$
, is called a pivotal equivalence class if:
-
(i) for each
$i \notin \{j(k) - l : k \geqslant 1, l =0, \ldots , M_{0} - 1 \}$ ,
$g_{i}(\omega )$ is fixed on
$\mathcal{E}$ ;
-
(ii) for each
$\omega \in \mathcal{E}$ and
$k \geqslant 1$ , the following is a Schottky sequence:
\begin{align*} s_{k}(\omega ) := \big (g_{j(k) - M_{0} + 1}(\omega ),\, g_{j(k) - M_{0} + 2}(\omega ), \, \ldots , \, g_{j(k)} (\omega ) \big ) \in S; \end{align*}
-
(iii) for each
$\omega \in \mathcal{E}$ ,
$(o, \, \mathbf{Y}_{j(1)} (\omega ), \, \mathbf{Y}_{j(2)} (\omega ), \, \ldots )$ is
$D_{0}$ -semi-aligned, and
-
(iv) on
$\mathcal{E}$ ,
$\{s_{1}(\omega ), s_{2}(\omega ), \ldots \}$ are i.i.d.s distributed according to the uniform measure on
$S$ .
We say that
$\mathcal{P}(\mathcal{E})$
is the set of pivotal times for
$\mathcal{E}$
.
When a pivotal equivalence class
$\mathcal{E} \subseteq \Omega$
is understood, with the set of pivotal times
$\mathcal{P}(\mathcal{E})$
, for each element
$\omega$
of
$\mathcal{E}$
we call
$\mathcal{P}(\mathcal{E})$
the set of pivotal times for
$\omega$
and write it as
$\mathcal{P}(\omega )$
.
When the probability space
$(\Omega , \mathbb{P})$
for
$\mu$
is partitioned into pivotal equivalence classes
$\{\mathcal{E}_{\alpha }\}_{\alpha }$
, then belonging to the same
$\mathcal{E}_{\alpha }$
becomes an equivalence relation. Choosing a different element from the same pivotal equivalence class is called pivoting. But note that the choice of pivotal equivalence classes is not canonical: given an
$\omega \in \Omega$
, there are several ways to define the pivotal equivalence class for
$\omega$
. Proposition 4.2 below describes a particular choice of pivotal equivalence classes that will be useful.
Let
$k$
be a positive integer. We say that a pivotal equivalence class
$\mathcal{E}$
avoids
$k$
if
$k$
is not in
$\{j - l : j \in \mathcal{P}(\mathcal{E}), \, l = 0, \ldots , M_{0} - 1\}$
; in this case,
$g_{k}$
is fixed on
$\mathcal{E}$
.
Proposition 4.2.
Let
$\mu$
be a non-elementary probability measure on
$G$
and let
$S$
be a long enough and large Schottky set for
$\mu$
. Then there exist a probability space
$(\Omega , \mathbb{P})$
for
$\mu$
and a constant
$K\gt 0$
such that, for each
$n \geqslant 0$
, we have a measurable partition
$\mathscr{P}_{n} = \{\mathcal{E}_{\alpha }\}_{\alpha }$
of
$\Omega$
into pivotal equivalence classes avoiding
$1, \ldots , \lfloor n/2 \rfloor + 1$
and
$n+1$
that satisfies

for each choice of
$g_{1}, \ldots , g_{\lfloor n/2 \rfloor + 1}, g_{n+1} \in G$
and
$k \geqslant n$
.
We postpone the proof of Proposition 4.2 to the next section and first see its consequence.
4.2 Pivoting
Let
$K_{0}, N_{0}\gt 0$
and let
$S$
be a long enough
$K_{0}$
-Schottky set with cardinality
$N_{0}$
. Given isometries
$u_{i}$
’s, let us draw a choice
$s = (s_{1}, s_{2}, \ldots , s_{n})$
from
$S^{n}$
with the uniform measure and define

Let
$\kappa _{i} := U_{i-1} \Gamma ^{+}(s_{i}) = u_{0} \Pi (s_{1}) \cdots u_{i-1} \Gamma ^{+}(s_{i})$
. We claim that:
Lemma 4.3. We have

for each
$1 \leqslant k \leqslant n$
and
$x \in X$
.
Proof. We prove the first estimate only; the second one follows similarly. Consider the statement

Thanks to the Schottky property, at most 1 choice of
$s_{1}$
from
$S$
violates this statement. Fixing that bad choice, consider the statement

Again, at most 1 choice of
$s_{2}$
from
$S$
violates this. Keeping this manner, we conclude the following: except at most 1 bad choice among
$S^{k}$
,

holds for at least one
$i \leqslant k$
. This happens for probability at least
$1- (1/N_{0})^{k}$
.
Now fix another set of isometries
$\check {u}_{i}$
’s and another
$K_{0}$
-Schottky set
$\check {S}$
with cardinality
$N_{0}$
. We draw
$\check {s} = (\check {s}_{1}, \check {s}_{2}, \ldots , \check {s}_{n})$
from
$\check {S}^{n}$
with the uniform measure, independently from
$s$
, and define

Let
$\eta _{i} := \check {U}_{i-1} \Gamma ^{+}(\check {s}_{i})$
. Recall that
$\bar {\eta }_{i}$
denotes the reversal of
$\eta _{i}$
.
Lemma 4.4. We have

for each
$1 \leqslant k \leqslant n$
.
Proof. Consider the statements

Thanks to the Schottky property, at most 1 choice of
$s_{1}$
from
$S$
violates the first statement. Similarly, given
$s_{1}$
, at most 1 choice of
$\check {s}_{1}$
from
$\check {S}$
violates the second statement. In short, the two statements hold for all but at most
$2N_{0}$
choices of
$(s_{1}, \check {s}_{1}) \in S\times \check {S}$
.
Fixing a bad choice
$(s_{1}, \check {s}_{1})$
, consider the statements

Again, at most
$2N_{0}$
choices of
$(s_{2}, \check {s}_{2}) \in S \times \check {S}$
violates the statements. Keeping this manner, we conclude the following: for probability at least
$1 - (2/N_{0})^{k}$
, there exists
$i \leqslant k$
such that

In other words,
$(\textrm {ending point of}\,\,\bar {\eta }_{i}, \kappa _{i})$
and
$(\bar {\eta }_{i}, \textrm {ending point of}\,\, \kappa _{i})$
are
$K_{0}$
-aligned. Lemma 3.8 then tells us that
$(\bar {\eta }_{i}, \kappa _{i})$
is
$D_{0}$
-aligned.
Applying Lemma 4.3 and 4.4 to pivotal equivalence classes, we obtain the following corollaries.
Corollary 4.5.
Let
$\mu$
be a non-elementary probability measure on
$G$
, let
$K_{0}, N_{0}\gt 0$
and let
$S$
be a long enough
$K_{0}$
-Schottky set for
$\mu$
with cardinality
$N_{0}$
. Let
$\mathcal{E}$
be a pivotal equivalence class for
$\mu$
with
$\mathcal{P}(\mathcal{E}) = \{j(1) \lt j(2) \lt \ldots \}$
and let
$x \in X$
. Then for each
$k \geqslant 1$
we have

Moreover, for any
$m\geqslant 1$
,
$n \geqslant j(m)$
and
$k = 1, \ldots , m$
, we have

Corollary 4.6.
Let
$\mu$
be a non-elementary probability measure on
$G$
and let
$\check {\mu }$
be its reflected version, let
$K_{0}, N_{0}\gt 0$
and let
$S$
and
$\check {S}$
be long enough
$K_{0}$
-Schottky sets for
$\mu$
and
$\check {\mu }$
, respectively, with cardinality
$N_{0}$
. Let
$\mathcal{E}$
be a pivotal equivalence class for
$\mu$
with
$\mathcal{P}(\mathcal{E}) = \{j(1) \lt j(2) \lt \ldots \}$
, and let
$\check {\mathcal{E}}$
be a pivotal equivalence class for
$\check {\mu }$
with
$\mathcal{P}(\check {\mathcal{E}}) = \{\check {j}(1) \lt \check {j}(2) \lt \ldots \}$
. Then we have

We now record a small consequence of pivoting.
Corollary 4.7.
Let
$(Z_{n})_{n\gt 0}$
be the random walk generated by a non-elementary probability measure
$\mu$
on
$G$
with finite first moment. Then there exists a strictly positive quantity
$\lambda (\mu ) \in (0, +\infty ]$
, called the drift of
$\mu$
, such that

Remark 4.8. The statement in Corollary 4.7 holds true even without the moment condition. This will be the consequence of Theorem 6.4 in Section 6 .
Proof.
By Kingman’s subadditive ergodic theorem,
$\lambda (\mu ) = \lim _{n} \frac {1}{n} d(o, Z_{n} o)$
exists and is constant almost surely. It remains to show that
$\lambda (\mu ) \gt 0$
.
Since
$\mu$
is non-elementary, Proposition 3.19 provides a long enough and large Schottky set
$S$
for
$\mu$
. Given this, Proposition 4.2 provides a constant
$K\gt 0$
and a measurable partition
$\mathscr{P} = \{\mathcal{E}_{\alpha }\}_{\alpha }$
into pivotal equivalence classes such that

for each
$k$
. Now let
$n \gt 0$
and let
$\mathcal{E}$
be a pivotal equivalence class with
$\mathcal{P}(\mathcal{E}) = \{j(1) \lt j(2) \lt \ldots \}$
such that
$\# (\mathcal{P}(\mathcal{E}) \cap \{1, \ldots , n\}) \geqslant n/K$
, i.e.,
$j(\lfloor n/ K \rfloor ) \leqslant n$
. Corollary 4.5 tells us that

By Lemma 3.18, we then have

We sum up these conditional probabilities on
$\{\omega : \#(\mathcal{P}(\mathcal{E} \cap \{1, \ldots , n\}) \geqslant n/K\}$
to conclude

The Borel-Cantelli lemma then implies
$d(o, Z_{n} o) \geqslant 50 E_{0} n/2K$
eventually almost surely.
4.3 Deviation inequality
Let
$\mu$
be a non-elementary probability measure on
$G$
and let
$S$
be a long enough and large
$K_{0}$
-Schottky set contained in
$(\textrm {supp} \mu )^{M_{0}}$
for some
$K_{0}, M_{0} \gt 0$
. Consider a bi-infinite path
$\big ((Z_{n}(\omega ))_{n \gt 0}, (Z_{n}(\check {\omega }))_{n \gt 0}\big )$
arising from the random walk generated by
$\mu$
. Recall:

For each
$k\geqslant M_{0}$
, we investigate whether there exists
$M_{0} \leqslant i \leqslant k$
such that:
-
(i)
$(g_{i-M_{0} + 1}, \ldots , g_{i})$ is a Schottky sequence;
-
(ii)
$(\check {Z}_{m} o, \mathbf{Y}_{i}(\omega ), Z_{n} o)$ is
$D_{0}$ -semi-aligned for all
$n \geqslant k$ and
$m \geqslant 0$ .
We define
$\upsilon = \upsilon (\check {\omega }, \omega )$
as the minimal index
$k$
with the auxiliary index
$i \leqslant k$
as described above.

Figure 3. Persistent progress and
$\upsilon$
. Here, all of the backward loci
$(\check {Z}_{n} o)_{n \geqslant 0}$
are on the left of the persistent progress
$Z_{i}\Gamma ^{+}(\alpha )$
, while the forward loci after
$Z_{\varsigma } o$
are all on the right.
A motivating observation for the definition of
$\upsilon (\check {\omega }, \omega )$
is as follows.
Lemma 4.9.
Let
$\Omega = G^{\mathbb{Z}_{\gt 0}} \times G^{\mathbb{Z}_{\gt 0}}$
be the space of (bi-directional) step paths in
$G$
, let
$K_{0}\gt 0$
and let
$S$
be a long enough
$K_{0}$
-Schottky set. Then for each
$(\check {\omega }, \omega ) \in \Omega$
, we have

for all
$m\geqslant 0$
and
$n, k \geqslant \upsilon (\check {\omega }, \omega )$
.
Proof.
Let
$i \leqslant \upsilon (\check {\omega }, \omega )$
be the index such that
$(\check {Z}_{m'} o, \mathbf{Y}_{i}(\omega ), Z_{n'} o)$
is
$D_{0}$
-semi-aligned for all
$n' \geqslant \upsilon (\check {\omega }, \omega )$
and
$m' \geqslant 0$
. Lemma 3.18 tells us that

Let us now pick
$n, k \geqslant \upsilon (\check {\omega }, \omega )$
and
$m \geqslant 0$
. Then we have

Hence,

We now provide a probabilistic estimate for
$\upsilon (\check {\omega }, \omega )$
.
Lemma 4.10.
Let
$\mu$
be a non-elementary probability measure on
$G$
, let
$K_{0}\gt 0$
and let
$S$
be a long enough and large
$K_{0}$
-Schottky set for
$\mu$
. Then there exists
$K'\gt 0$
such that

holds for all
$k \geqslant 0$
and all choices of
$g_{k+1}, \check {g}_{1}, \ldots , \check {g}_{k+1} \in G$
.
Proof.
Let
$S$
be a long enough and large
$K_{0}$
-Schottky set in
$(\textrm {supp} \mu )^{M_{0}}$
for some
$M_{0}\gt 0$
. Let
$\check {S}$
be the reflected version of
$S$
, that means,

Then
$\check {S}$
is a long enough and large
$K_{0}$
-Schottky set for
$\check {\mu }$
. Let
$K\gt 0$
be the constant determined for
$S$
and
$\check {S}$
in Proposition 4.2. We now fix
$k$
and
$g_{k+1}, \check {g}_{1}, \ldots , \check {g}_{k+1} \in G$
.
Let
$\mathscr{P}_{k} = \{\mathcal{E}_{\alpha }\}_{\alpha }$
be the partition of
$\Omega$
into pivotal equivalence classes avoiding
$1, \ldots , \lfloor k/2 \rfloor +1$
and
$k+1$
, given by Proposition 4.2. Let also
$\check {\mathscr{P}}_{2k} = \{\check {\mathcal{E}}_{\alpha }\}_{\alpha }$
be the partition of
$\check {\Omega }$
into pivotal equivalence classes avoiding
$1, \ldots , k+1$
and
$2k+1$
, given by Proposition 4.2. We have

Let us enumerate
$\mathcal{P}(\omega )$
by
$\{j(1) \lt j(2) \lt \ldots \}$
, and
$\mathcal{P}(\check {\omega })$
by
$\{\check {j}(1) \lt \check {j}(2) \lt \ldots \}$
. Let
$\mathcal{E} \in \mathscr{P}_{k}$
and
$\check {\mathcal{E}} \in \check {\mathscr{P}}_{2k}$
be pivotal equivalence classes in
$A$
and
$\check {A}$
, respectively. In
$\check {\mathcal{E}}\times \mathcal{E}$
, let
$B$
be the set of
$(\check {\omega }, \omega )$
that satisfies the following:
-
(i) for
$x \in \{o, \check {Z}_{1} o, \ldots , \check {Z}_{2k} o\}$ , the following sequence is
$D_{0}$ -semi-aligned:
\begin{align*} \big (x, \,\mathbf{Y}_{j(\lceil k/3K\rceil )}(\omega ), \,\mathbf{Y}_{j(\lceil k/3K\rceil + 1)}(\omega ), \,\ldots \big ); \end{align*}
-
(ii) for each
$n \geqslant k$ and
$m \geqslant 2k$ , the following are
$D_{0}$ -semi-aligned:
\begin{align*} \begin{aligned} &\big (o, \,\mathbf{Y}_{j(1)}(\omega ), \, \mathbf{Y}_{j(2)}(\omega ), \, \ldots , \, \mathbf{Y}_{j(\lceil 2n/3K\rceil )}(\omega ), \,Z_{n} o\big ),\\ &\big (o, \,\mathbf{Y}_{\check {j}(1)}(\check {\omega }), \, \mathbf{Y}_{\check {j}(2)}(\check {\omega }), \, \ldots , \, \mathbf{Y}_{\check {j}(\lceil 2m/3K\rceil )}(\check {\omega }), \,\check {Z}_{m} o\big ); \end{aligned} \end{align*}
-
(iii)
$\big (\bar {\mathbf{Y}}_{\check {j}(i)}(\check {\omega }),\, \mathbf{Y}_{j(i)}(\omega ) \big )$ is
$D_{0}$ -aligned for some
$i \leqslant k/3K$ .
The first item is handled by Lemma 4.3: it holds for probability at least
$1-2k \cdot (1/400)^{k/3K}$
.
Next, recall that for each
$n \geqslant k$
, there are at least
$n/K$
pivotal times for
$\mathcal{E}$
before
$n$
. Also, for each
$m \geqslant 2k$
, there are at least
$m/K$
pivotal times for
$\check {\mathcal{E}}$
before
$m$
. Hence, we can apply Lemma 4.3 and deduce that the following are
$D_{0}$
-semi-aligned:

for probability at least
$1-(1/400)^{n/3K-1}$
and
$1-(1/400)^{m/3K - 1}$
, respectively. Taking intersection for
$n \geqslant k$
and
$m \geqslant 2k$
, we observe that Item (ii) holds for probability at least
$1-3 \cdot (1/400)^{k/3K - 1}$
.
Finally, Item (iii) is handled by Lemma 4.4: it holds for probability at least
$1- (1/200)^{k/3K - 1}$
. Combining these, we deduce

It remains to prove that
$\upsilon (\check {\omega }, \omega ) \leqslant k$
for
$(\check {\omega }, \omega ) \in B$
. First, by definition of
$A$
,
$j(\lceil k/3K\rceil )$
is smaller than
$k$
and

is Schottky. Next, for each
$n \geqslant k$
,
$(o, \mathbf{Y}_{j(1)}(\omega ), \mathbf{Y}_{j(2)}(\omega ), \ldots , \mathbf{Y}_{j(\lceil 2n/3K \rceil )}(\omega ), Z_{n} o)$
is
$D_{0}$
-semi-aligned. Hence,
$(o, \mathbf{Y}_{j(\lceil k/3K\rceil )}(\omega ), Z_{n} o)$
is also
$D_{0}$
-semi-aligned.
We now investigate the alignment of
$(\check {Z}_{m} o, \mathbf{Y}_{j(\lceil k/3K\rceil )}(\omega ))$
. For
$m \leqslant 2k$
, this is guaranteed by item (1). When
$m \geqslant 2k$
, we appeal to item (2) and (3). Namely, the sequence

is
$D_{0}$
-semi-aligned. In particular,
$\big (\check {Z}_{m} o, \bar {\mathbf{Y}}_{j(\lceil k/3K \rceil )} (\omega ) \big )$
is
$D_{0}$
-semi-aligned.
Here is a corollary of Lemma 4.10 that we will use in Section 6.
Corollary 4.11 ([[Reference GouëzelGou22], Lemma 4.14])
Let
$\mu$
be a non-elementary probability measure on
$G$
and let
$(Z_{n})_{n}$
be the random walk generated by
$\mu$
. Then for each
$\epsilon \gt 0$
, there exists
$C\gt 0$
such that

Proof.
Let us pick
$K_{0} \gt 0$
and a long enough and large
$K_{0}$
-Schottky set
$S$
for
$\mu$
. Let
$K'$
be the constant as in Lemma 4.10. Given
$\epsilon \gt 0$
, we take
$N\gt 1$
large enough so that
$K' e^{-N / K'} \leqslant \epsilon /4$
. Then, the definition of the RV
$\upsilon (\check {\omega }, \omega )$
and Lemma 4.10 tells us that

When
$(g^{-1} o, \mathbf{Y}_{i}, Z_{n} o)$
is
$D_{0}$
-semi-aligned, the second inequality in Lemma 3.18(ii) implies

This bound also holds for
$n \leqslant N$
:

Given these, the proof ends by taking large enough
$C\gt 0$
such that

Corollary 4.12.
Let
$\mu$
be a non-elementary probability measure on
$G$
whose expectation is infinite. Then
$\mu ^{\ast m}$
has infinite expectation for each
$m\gt 0$
. In particular, the drift
$\lambda (\mu ) := \lim _{m \rightarrow \infty } \frac {1}{m} \mathbb{E}_{\mu ^{\ast m}} [d(o, go)]$
is infinity.
Proof.
Let
$\epsilon =0.2$
and let
$C = C(\mu , \epsilon )$
be as in Corollary 4.11. Let
$(g_{1}, \ldots , g_{m})$
be distributed according to
$\mu ^{m}$
. Then by Corollary 4.11, we have

Now integrating over
$g_{1} \in \textrm {supp} \mu$
with law
$\mu$
, we get

Similarly, fixing the Schottky set
$S$
for
$\mu$
, we similarly define
$\check {\upsilon } = \check {\upsilon }(\check {\omega }, \omega )$
as the minimal index
$k$
that is associated with another index
$i \leqslant k$
such that:
-
(i)
$(\check {g}_{i}^{-1}, \ldots , \check {g}_{i-M_{0} + 1}^{-1})$ is a Schottky sequence;
-
(ii)
$(\check {Z}_{m}o, \bar {\mathbf{Y}}_{i}(\check {\omega }), o)$ is
$D_{0}$ -semi-aligned for all
$m \geqslant k$ , and
-
(iii)
$(\bar {\mathbf{Y}}_{i}(\check {\omega }), Z_{n} o)$ is
$D_{0}$ -semi-aligned for all
$n\geqslant 0$ .
Then we similarly have

Thanks to these exponential bounds, we can establish the deviation inequality.
Proposition 4.13.
Let
$p\gt 0$
and let
$((\check {Z}_{n})_{n}, (Z_{n})_{n})$
be the (bi-directional) random walk generated by a non-elementary probability measure
$\mu$
on
$G$
with finite
$p$
-th moment. Then the random variable
$\sup _{n, m \geqslant 0} (\check {Z}_{m} o, Z_{n} o)_{o}$
has finite
$2p$
-th moment.
Note the difference between this proposition and [[Reference ChoiCho23], Proposition 5.6, 5.8]; we are taking the global suprema, not the limit suprema.
Proof.
Let
$K'$
be the constant for
$\mu$
as in Lemma 4.10 and let

By triangle inequality,
$d(o, Z_{k} o) \lt D_{l}$
and
$d(o, \check {Z}_{k} o) \leqslant \check {D}_{l}$
for all
$k \leqslant l$
. We begin by claiming

Since
$\mathbb{P}(\max \{\upsilon , \check {\upsilon }\} \geqslant k)$
is summable by Inequality 9 and 10, Borel-Cantelli implies that

Note that the RHS of Inequality 11 is at least
$ \check {D}_{l}^{p} D_{l}^{p}$
.
Now at
$i = l$
, we have either
$\check {D}_{l} \geqslant D_{l}$
or
$\check {D}_{l} \leqslant D_{l}$
. In the first case
$l \geqslant \upsilon$
must hold. Then for
$m \geqslant 0$
and
$n \geqslant l$
, we have

by Lemma 4.9. Moreover, for
$m \geqslant 0$
and
$n \leqslant l$
, we have

In the second case
$l \geqslant \check {\upsilon }$
must hold, and for a similar reason
$(\check {Z}_{m} o, Z_{n} o)_{o}^{2p}$
is dominated by
$\check {D}_{l}^{p} D_{l}^{p}$
. Inequality 11 now follows.
We now need a small observation:
Fact 4.14.
For
$s_{1}, s_{2}, t_{1}, t_{2} \geqslant 0$
, the following holds:

Proof of Fact 4.14. The fact follows from the following inequality in [[Reference Benoist and QuintBQ16], Section 5.4]:

We give its proof for completeness. Assume
$t \geqslant s$
without loss of generality. When
$p \leqslant 1$
, the concavity of
$f(x) = x^{p}$
implies the inequality. When
$p \gt 1$
, we divide the cases. If
$s\lt t/2$
, then

If
$s \geqslant t/2$
, then we have

By Fact 4.14, the expectations of
$|\check {D}_{i+1}^{p} D_{i+1}^{p} - \check {D}_{i}^{p} D_{i}^{p}| \left (1_{\check {D}_{i} \geqslant D_{i}} 1_{i \lt \upsilon } + 1_{\check {D}_{i} \leqslant D_{i}} 1_{i \lt \check {\upsilon }}\right )$
for
$i \geqslant 0$
are summable as soon as there exists
$K''\gt 0$
such that

for each
$0 \leqslant n_{1}, n_{2} \leqslant p$
with
$n_{1} + n_{2} \geqslant \min (p, 1)$
. We discuss the case
$n_{2} \gt 0$
; the other case
$n_{1} \gt 0$
can be handled in the same way.
We will take advantage of the fact that
$\mathbb{E}[\check {D}_{i}^{p} D_{i}^{p}]$
is bounded. Namely, the expectation of
$\check {D}_{i}^{p-n_{1}} D_{i}^{p-n_{2}}$
on the set
$\{D_{i} \gt c\}$
is small for large
$c$
. Next, on the set
$\{D_{i} \leqslant c\}$
, we will bound the expectation of
$\check {D}_{i}^{p-n_{1}} D_{i}^{p-n_{2}} 1_{D_{i} \lt c}1_{i \lt \upsilon }$
by using the exponential bound on
$\mathbb{P}( i \lt \upsilon )$
(that suppresses
$D_{i}^{p-n_{2}} \lt c^{p-n_{2}}$
) independent of the distribution of
$\check {D}_{i}$
.
We first discuss the term
$\mathbb{E}\left [ d(o, \check {g}_{i+1})^{n_{1}} d(o, g_{i+1})^{n_{2}} \check {D}_{i}^{p-n_{1}} D_{i}^{p-n_{2}}\cdot 1_{\check {D}_{i} \geqslant D_{i} }1_{i\lt \upsilon } \right ]$
. Let us fix
$\check {g}_{i+1}$
and
$g_{i+1}$
for the moment, and let
$c:= e^{i/2pK'}$
. We then have a decomposition

The first term is controlled as follows:

In the final step, we used the following fact for each
$r\gt 0$
and
$i \gt 0$
:

Next, we apply Lemma 4.10 to the second term of the RHS of Equation 13 and observe:

Here,
$c^{p-n_{2}}$
is dominated by
$c^{p} = e^{i/2K'}$
. Overall, we have

We now multiply
$ d(o, \check {g}_{i+1})^{n_{1}} d(o, g_{i+1})^{n_{2}}$
and integrate. As a result, we observe

for some constant
$C(\mu ) \lt +\infty$
determined by the distribution of
$\mu$
, independent of
$i$
. Note that
$\mu$
has finite
$q$
-th moment for every
$0 \leqslant q \leqslant p$
thanks to Jensen’s inequality.
We similarly deal with the term
$\mathbb{E}\left [ d(o, \check {g}_{i+1})^{n_{1}} d(o, g_{i+1})^{n_{2}} \check {D}_{i}^{p-n_{1}} D_{i}^{p-n_{2}}\cdot 1_{\check {D}_{i} \leqslant D_{i} }1_{i\lt \check {\upsilon }} \right ]$
. Fixing
$g_{i+1}$
and
$\check {g}_{i+1}$
first, we split the expectation based on the dichotomy for
$\check {D}_{i}$
:

Here, a crucial observation is that
$\check {D}_{i}^{p-n_{1}} D_{i}^{p-n_{2}} 1_{\check {D}_{i} \gt c} 1_{\check {D}_{i} \leqslant D_{i}} 1_{i\lt \check {\upsilon }}$
is dominated by
$\check {D}_{i}^{p-n_{1}} D_{i}^{p-n_{2}} 1_{D_{i} \gt c}$
. The remaining step is analogous to the previous computations:

We then multiply
$d(o, \check {g}_{i+1})^{n_{1}} d(o, g_{i+1})^{n_{2}}$
and integrate over
$\check {g}_{i+1}$
and
$g_{i+1}$
to obtain a summable bound. This concludes the Inequality 12.
The previous proof also yields the following corollary.
Corollary 4.15.
Let
$p\gt 0$
and let
$\big ((\check {Z}_{n})_{n\gt 0}, (Z_{n})_{n\gt 0}\big )$
be the (bi-directional) random walk generated by a non-elementary probability measure
$\mu$
on
$G$
with finite
$p$
-th moment. Then there exists
$K\gt 0$
such that

Proof. In view of the previous proof, it suffices to check

The RHS is at least
$\check {D}_{l}^{p} D_{l}^{p}$
for
$l = \min \{ i : 1_{\check {D}_{i} \geqslant D_{i}} 1_{i \lt \upsilon } + 1_{\check {D}_{i} \leqslant D_{i}} 1_{i \lt \check {\upsilon }}=0\}$
. Note that either
$\check {D}_{l} \geqslant D_{l}$
or
$\check {D}_{l} \leqslant D_{l}$
holds. In the first case, we are forced to have
$l \geqslant \upsilon$
; then

In the second case, we are forced to have
$l \geqslant \check {\upsilon }$
; then

We now discuss random walks with finite exponential moment.
Corollary 4.16.
Let
$\big ((\check {Z}_{n})_{n\gt 0}, (Z_{n})_{n\gt 0}\big )$
be the (bi-directional) random walk generated by a non-elementary probability measure
$\mu$
on
$G$
with finite exponential moment. Then there exists
$K\gt 0$
such that

Proof.
Let
$K'$
be as in Lemma 4.10 and
$D_{i} = \sum _{k=1}^{i} d(o, g_{k} o)$
. Then
$e^{d(o, Z_{\upsilon } o)/K}$
is dominated by
$\sum _{i \leqslant \upsilon } e^{D_{i}/K}$
. Hence, we need to show that
$\mathbb{E}[e^{D_{i}/K} 1_{i \lt \upsilon }]$
is summable. Let
$K, c \gt 0$
and observe

By taking
$K$
large enough, we can make
$\mathbb{E}_{\mu }[ \operatorname {exp}(2d(o, go)/K)] \leqslant e^{1/4K'}$
. Then we take
$c = iK/2K'$
and conclude
$\mathbb{E}[e^{D_{i}/K} 1_{i\lt \upsilon }] \lt (K' + 1) e^{-i/4K'}$
.
4.4 Limit theorems
The second-moment deviation inequality implies the following CLT:
Theorem 4.17.
Let
$(X, G, o)$
be as in Convention
2.11
and let
$(Z_{n})_{n\gt 0}$
be the random walk generated by a non-elementary probability measure
$\mu$
on
$G$
with finite second moment. Then the following limit (called the asymptotic variance of
$\mu$
) exists:

and the random variable
$\frac {1}{\sqrt {n}} [d(o, Z_{n}o) - \lambda (\mu ) n]$
converges in law to the Gaussian law
$\mathcal{N}(0, \sigma (\mu ))$
with zero mean and variance
$\sigma ^{2}(\mu )$
.
Proof.
Since
$\mu$
has finite second moment, Proposition 4.13 implies that
$\sup _{n, m \geqslant 0} (\check {Z}_{m} o, Z_{n} o)_{o}$
has finite 4-th moment, and hence finite second moment. Now Theorem 4.1 and 4.2 of [Reference Mathieu and SistoMS20] lead to the conclusion.
Remark 4.18. In fact, the following non-degeneracy statement holds:
Fact 4.19.
Let
$(X, G, o)$
be as in Convention
2.11
and let
$(Z_{n})_{n}$
be the random walk generated by a non-elementary probability measure
$\mu$
on
$G$
. Then the asymptotic variance
$\sigma ^{2}(\mu ) := \lim _{n}\frac {1}{n} Var[d(o, Z_{n} o)]$
is nonzero if and only if
$\mu$
is non-arithmetic, i.e., there exists
$N\gt 0$
and two elements
$g, h \in (\textrm {supp} \mu ^{\ast N})$
of
$\textrm {supp} \mu ^{\ast N}$
with distinct translation lengths.
The strict positivity of
$\sigma ^{2}(\mu )$
for non-arithmetic random walks on Gromov hyperbolic spaces and Teichmüller space was discussed in [Reference ChoiCho23]; see Theorem B and Claim 6.2 of [Reference ChoiCho23]. Since the argument in [Reference ChoiCho23] also applies to the general case, we omit the proof here.
We next discuss the law of the iterated logarithms.
Theorem 4.20.
Let
$(X, G, o)$
be as in Convention
2.11
and let
$(Z_{n})_{n\gt 0}$
be the random walk generated by a non-elementary probability measure
$\mu$
on
$G$
with finite second moment. Then for almost every sample path
$(Z_{n})_{n}$
we have

where
$\lambda (\mu )$
is the drift of
$\mu$
and
$\sigma ^{2}(\mu )$
is the asymptotic variance of
$\mu$
.
We proved the LIL based on the uniform
$4$
th order deviation inequality in [Reference ChoiCho23]. We give another argument because we will only have second-order deviation inequality in Part II.
Proof. In the proof of the LIL in [Reference ChoiCho23] (see [[Reference ChoiCho23], Claim 7.1]), the author proved:
Lemma 4.21.
Let
$K\gt 0$
and let
$\{U_{k, i}\}_{i, k \in \mathbb{Z}_{\gt 0}}$
be RVs such that for each
$k$
,
$\{U_{k, i}\}_{i}$
are i.i.d.s with zero mean and variance at most
$K$
. Then for each
$\epsilon \gt 0$
, there exists
$M\gt 0$
such that

We now set

Equivalently, we have
$Y_{k+1, i} = Y_{k, 2i-1} + Y_{k, 2i} - 2b_{k, i}$
. Note that
$\{b_{k, i}\}_{k, i}$
have uniformly bounded variance by Lemma 4.10 and
$\{b_{k, i} - \mathbb{E}[b_{k, i}]\}_{i}$
are i.i.d.s with zero mean for each
$k$
. We also set

We then observe the decomposition

Indeed, the RHS is unchanged when
$M$
increases by 1 and is equal to
$d(o, Z_{n}o)$
at
$M \gt \lfloor \log _{2} n \rfloor$
.
Now, fixing an
$\epsilon \gt 0$
, we take
$M\gt 0$
for
$\{b_{k, i} - \mathbb{E}[b_{k, i}]\}_{i, k}$
using Lemma 4.21. We balance each term in Display 15 by subtracting its expectation, normalize with the denominator
$\sqrt {2n \log \log n}$
and then examine the almost sure limit supremum. The classical LIL tells us that

Regarding the second term, note that
$d(Z_{2^{M} \lfloor n/2^{M} \rfloor } , Z_{n} o)$
is dominated by the sum of at most
$2^{M}$
independent steps distributed according to
$\mu$
. This implies that

and RHS is summable in
$n$
because
$\mu$
has finite second moment. By Borel-Cantelli lemma,

Next, Lemma 4.21 implies that the term

eventually falls into the interval
$[-\epsilon , +\epsilon ]$
outside a set of probability
$\epsilon$
.
It remains to deal with
$\frac {1}{\sqrt {2n \log \log n}}\sum _{k=M}^{\lfloor \log _{2}n \rfloor } (b_{k; n} - \mathbb{E}[b_{k; n}])$
. Let

Then for each
$k$
and
$n$
,
$0 \leqslant b_{k; n} \leqslant b_{2^{k} (2 \lfloor n/ 2^{k+1} \rfloor + 1)}$
holds. Moreover,
$b_{j}$
’s are identically distributed with finite variance (and hence finite expectation). This implies that

tends to 0 as
$n$
goes to infinity.
We now estimate the summation

To estimate this, for each
$y\gt 0$
let us count the number of pairs
$(i, k) \in \mathbb{Z}_{\geqslant 0}^{2}$
such that
$ \epsilon ^{2} \sqrt {2}^{k} (2i-2) \lt y$
. For each
$k \in \mathbb{Z}_{\geqslant 0}$
, there exist at most
$y / (\sqrt {2}^{k} \epsilon ^{2})$
candidates for
$i$
. Summing them up, there are at most
$C_{\epsilon } y$
such pairs
$(i, k)$
, where
$C_{\epsilon }\gt 0$
is a constant. This implies that

By Borell-Cantelli, for almost every sample path
$b_{2^{k}(2i-1)} \lt \epsilon \cdot 2^{k/4} \sqrt {2i-2}$
holds for all but finitely many
$(i,k)$
’s. In particular, for sufficiently large
$n$
, we have

for each
$k = 1, \ldots , \lfloor \log _{2} n \rfloor$
. Hence, we have

Combining these estimates with Equation 15, we observe that for probability at least
$1-\epsilon$
,

By decreasing
$\epsilon$
while increasing
$M$
, we arrive at the desired conclusion.
We finally prove the geodesic tracking by random walks.
Theorem 4.22.
Let
$(X, G, o)$
be as in Convention
2.11
, let
$p\gt 0$
and let
$(Z_{n})_{n}$
be the random walk generated by a non-elementary probability measure
$\mu$
on
$G$
with finite
$p$
-th moment. Then there exists
$K\gt 0$
such that, for almost every sample path
$(Z_{n})_{n \geqslant 0}$
, there exists a
$K$
-quasigeodesic
$\gamma$
on
$X$
satisfying

Proof.
Recall Definition 3.16: Given
$K_{0}\gt 0$
, we have defined:
-
–
$D_{0}= D(K_{0}, K_{0}) \gt K_{0}$ be as in Lemma 3.8,
-
–
$E_{0}= E(K_{0}, D_{0}) \gt D_{0}$ ,
$L_{0} = L(K_{0}, D_{0})$ be as in Proposition 3.12.
In addition to these, we define:
-
–
$E_{1}= E(K_{0}, D_{0})$ ,
$L_{1} = L(K_{0}, D_{0})$ be as in Lemma 3.14;
-
–
$E_{2}= E(K_{0}, E_{0} + 5K_{0})$ ,
$L_{2} = L(K_{0}, E_{0}+5K_{0})$ be as in Proposition 3.12.
Since
$\mu$
is non-elementary, Proposition 3.19 guarantees that there exist
$K_{0}\gt 0$
,
$M_{0}\gt L_{0} + L_{1}+L_{2}$
and a large enough
$K_{0}$
-Schottky set
$S \subseteq (\textrm {supp} \mu )^{M_{0}}$
. We fix this
$S$
from now on.
By Proposition 4.2, there exists a probability space
$(\Omega , \mathbb{P})$
with RV
$\mathcal{P}(\omega ) = \{j(1) \lt j(2) \lt \ldots \} \subseteq M_{0} \mathbb{Z}_{\gt 0}$
, the set of pivotal times, such that
$(o, \mathbf{Y}_{j(1)}(\omega ), \mathbf{Y}_{j(2)}(\omega ), \ldots )$
is
$D_{0}$
-semi-aligned.
We now define
$\Gamma (\omega )$
as the concatenation of
$[o, Z_{j(1) - M_{0}} o]$
,
$[Z_{j(1) - M_{0}} o, Z_{j(1)} o]$
,
$[Z_{j(1)} o, Z_{j(2) - M_{0}} o]$
,
$[Z_{j(2) - M_{0}} o, Z_{j(2)} o]$
,
$\ldots$
. By Lemma 3.14,
$\Gamma (\omega )$
is an
$E_{1}$
-quasigeodesic for almost every
$\omega \in \Omega$
. It remains to prove
$\lim _{n} d(Z_{n}(\omega ) o, \Gamma (\omega )) / n^{1/2p}=0$
almost everywhere.
By Corollary 4.15,
$\min [ d(o, Z_{\upsilon } o), d(o, \check {Z}_{\check {\upsilon }} o)]^{2p}$
is dominated by an integrable RV. This implies

for some function
$g$
such that
$\lim _{k} g(k)/k^{1/2p} = 0$
. Also, Lemma 4.10 tells us that

Now, for each
$k \in \mathbb{Z}_{\gt 0}$
, we consider the following sets:


Then the definition of the RV
$\upsilon (\check {\omega }, \omega )$
and
$\check {\upsilon }(\check {\omega }, w)$
, together with Inequality 8, tells us that

Thanks to Display 16 and 17, we observe that
$\mathbb{P}(A_{k}^{c} \cap B_{k}^{c})$
is also summable.
Finally, consider

Then
$C_{k}$
contains
$T^{k}(A_{k} \cup B_{k})$
, where
$T$
is the Bernoulli shift operator on the bi-infinite sample paths, which is measure preserving. Hence,
$\mathbb{P}(C_{k}^{c}) \leqslant \mathbb{P}(A_{k}^{c} \cap B_{k}^{c})$
is summable. The Borel-Cantelli lemma implies that, for almost every sample path, for each sufficiently large
$k$
there exists
$M_{0} \leqslant j'(k) \leqslant 2k$
such that
$\operatorname {diam}(Z_{k} o \cup \mathbf{Y}_{j'(k)} o) \leqslant d(Z_{k} o, Z_{j'(k)} o) + \operatorname {diam} (\mathbf{Y}_{j'(k)} o) \leqslant g(k) + K_{0}M_{0} + K_{0}$
and such that
$(o, \mathbf{Y}_{j'(k)}, Z_{n} o)$
is
$D_{0}$
-semi-aligned for
$n \geqslant 2k$
(
$\ast$
),
Let us now pick a sample path satisfying
$(\ast$
), pick a sufficiently large
$k$
, and let
$N$
be an index such that
$j(N) \geqslant 2k$
. Recall that
$(o, \mathbf{Y}_{j(1)}(\omega ), \mathbf{Y}_{j(2)}(\omega ), \ldots )$
is
$D_{0}$
-semi-aligned. By Proposition 3.12,
$[o, Z_{j(N)} o]$
have subsegments
$[x_{1}, y_{1}], \ldots , [x_{N}, y_{N}]$
, in order from left to right, such that
$[x_{i}, y_{i}]$
and
$\mathbf{Y}_{j(i)}$
are
$0.1E_{0}$
-fellow traveling for
$i = 1, \ldots , N$
. Moreover, by Corollary 3.4,
$\mathbf{Y}_{j(i)}$
and
$[Z_{j(i) - M_{0}} o, Z_{j(i)} o]$
are
$0.1E_{0}$
-fellow traveling for
$i=1, \ldots , N$
. Finally, since
$(o, \mathbf{Y}_{j'(k)}, Z_{j(N)} o)$
is
$D_{0}$
-semi-aligned, Proposition 3.12 tells us that
$[o, Z_{j(N)} o]$
also contains a subsegment
$[q_{1}, q_{2}]$
that
$0.1E_{0}$
-fellow travels with
$\mathbf{Y}_{j'(k)}$
. For convenience we let
$y_{0} = o$
and
$j(0) = 0$
.
If
$[q_{1}, q_{2}]$
overlaps with some
$[x_{i}, y_{i}]$
, this implies
$d(\mathbf{Y}_{j'(k)}, [Z_{j(i) - M_{0}} o, Z_{j(i)} o]) \leqslant E_{0}$
and hence
$d(Z_{k} o, \Gamma (\omega ) )\leqslant g(k) + E_{0} + K_{0} M_{0} + K_{0}$
. If not, then
$[q_{1}, q_{2}]$
is a subsegment of
$[y_{i-1}, x_{i}]$
for some
$i$
. Lemma 3.11 then tells us that
$(y_{i-1}, \mathbf{Y}_{j'(k)}, x_{i})$
is
$0.4E_{0}$
-aligned. Since
$d(y_{i-1}, Z_{j(i-1)} o) \leqslant 0.1E_{0}$
and
$d(x_{i}, Z_{j(i) - M_{0}} o) \leqslant 0.1E_{0}$
, Lemma 2.2 implies that
$(Z_{j(i-1)} o, \mathbf{Y}_{j'(k)}, Z_{j(i)- M_{0}} o)$
is
$(0.5E_{0} + 4K_{0})$
-aligned. By Proposition 3.12,
$[Z_{j(i-1)} o, Z_{j(i) - M_{0}} o]$
passes through the
$E_{2}$
-neighborhood of
$\mathbf{Y}_{j'(k)}$
, and
$d(Z_{k} o, \Gamma (\omega )) \leqslant g(k) + E_{2}$
.
In summary, almost every sample path
$(\check {\omega }, \omega )$
satisfies
$(\ast )$
, which leads to
$d(Z_{k} o, \Gamma (\omega )) \leqslant g(k)+E_{0}+E_{2} + K_{0}M_{0} + K_{0}= o(k^{1/2p})$
eventually. This ends the proof.
Recall Corollary 4.16: if
$\mu$
has finite exponential moment, then
$\mathbb{E}[\operatorname {exp}(d(o, Z_{\upsilon } o) / K)]$
is finite, i.e.,
$\mathbb{P}( d(o, Z_{\upsilon } o) \gt K \log k)$
is summable for some
$K\gt 0$
. By replacing
$g(k)$
in the previous proof with
$K \log k$
, we obtain:
Theorem 4.23.
Let
$(X, G, o)$
be as in Convention
2.11
and let
$(Z_{n})_{n}$
be the random walk generated by a non-elementary probability measure
$\mu$
on
$G$
with finite exponential moment. Then there exists
$K\gt 0$
such that, for almost every sample path
$(Z_{n})_{n \geqslant 0}$
, there exists a
$K$
-quasigeodesic
$\gamma$
satisfying

5. Pivotal time construction
In this section we prove Proposition 4.2 by generalizing Gouëzel’s theory in [[Reference GouëzelGou22], Section 4A] to the setting of Convention 2.11. We first construct and study pivotal times in a discrete model and then realize them on random walks. This strategy is also employed for LDP in Section 6.
5.1 Pivotal times: discrete model
Throughout the subsection, we fix a long enough
$K_{0}$
-Schottky set
$S$
with cardinality
$N_{0}$
. Given sequences of isometries
$\mathbf{w} =(w_{i})_{i=0}^{\infty }$
and
$\mathbf{v} =(v_{i})_{i=1}^{\infty }$
in
$G$
, we draw a sequence of Schottky sequences

with respect to the uniform measure on
$S^{4n}$
. We define isometries

and study the word

With the base case
$w_{0, 2}^{+} := id$
, we define its subwords for
$i\gt 0$
:

Let us also employ the notations

We define the set of pivotal times
$P_{n} = P_{n}\left (\mathbf{s}; \mathbf{w}, \mathbf{v}\right )$
and an auxiliary moving point
$z_{n} = z_{n}\left (\mathbf{s}; \mathbf{w}, \mathbf{v}\right )$
inductively. Let
$P_{0} = \emptyset$
and
$z_{0} = o$
as the base case. Given
$P_{n-1} \subseteq \{1, \ldots , n-1\}$
and
$z_{n-1} \in X$
, the data
$P_{n}$
and
$z_{n}$
at step
$n$
are determined by the following criteria.
-
(A) When
$\left (z_{n-1}, \Upsilon (\alpha _{n})\right )$ ,
$\left (\Upsilon (\beta _{n}), w_{n, 1}^{+} o\right )$ ,
$\left (w_{n, 0}^{-}o, \Upsilon (\gamma _{n})\right )$ and
$\left (\Upsilon (\delta _{n}), w_{n+1, 2}^{-}o\right )$ are
$K_{0}$ -aligned, we set
$P_{n} = P_{n-1} \cup \{n\}$ and
$z_{n} = w_{n, 1}^{+}o$ (see Figure 4).
-
(B) Otherwise, we seek
$i \in P_{n-1}$ and an integer
$j \in \{i+1, \ldots , n-1\}$ such that
$\big (\Upsilon (\delta _{i}), \Upsilon (\beta _{j})\big )$ is
$D_{0}$ -semi-aligned and such that
$\big (\Upsilon (\beta _{j}), w_{n+1, 2}^{-} o\big )$ is
$K_{0}$ -aligned. If such a pair
$(i, j)$ exists, we pick the lexicographically maximal one and let
$P_{n} := P_{n-1} \cap \{1, \ldots , i\}$ and
$z_{n} = w_{j, 1}^{-}o$ . If such a pair does not exist, then we let
$P_{n} := \emptyset$ and
$z_{n} := o$ .
We note that the set the set
$P_{n}$
depends solely on
$(w_{i})_{i=0}^{n}$
,
$(v_{i})_{i=1}^{n}$
and
$(\alpha _{i}, \beta _{i}, \gamma _{i}, \delta _{i})_{i=1}^{n}$
; it is independent from
$\{w_{i}, v_{i}, \alpha _{i}, \beta _{i}, \gamma _{i}, \delta _{i} : i \gt n\}$
.
$P_{n}$
records the Schottky axes aligned along
$[o, w_{n+1, 2}^{-} o]$
. More precisely:
Proposition 5.1.
Let
$P_{n}= \{i(1) \lt \ldots \lt i(m)\}$
. Then

is
$D_{0}$
-semi-aligned.
On Gromov hyperbolic spaces, this corresponds to [[Reference GouëzelGou22], Lemma 5.3]. Before proving the entire statement, let us prove two small parts of it.
Lemma 5.2.
For any
$\mathbf{s} \in S^{4n}$
and
$1 \leqslant i \leqslant n$
,
$\left (\Upsilon (\alpha _{i}), \Upsilon (\beta _{i})\right )$
and
$\left (\Upsilon (\gamma _{i}), \Upsilon (\delta _{i})\right )$
are
$D_{0}$
-aligned.
Proof.
Let us prove that
$\left (\Upsilon (\alpha _{i}), \Upsilon (\beta _{i})\right ) = \left ( w_{i, 1}^{-} \bar {\Gamma }^{-} (\alpha _{i}), w_{i, 1}^{-} \Gamma ^{+}(\beta _{i}) \right )$
is
$D_{0}$
-aligned, or equivalently, that
$\left (\bar {\Gamma }^{-}(\alpha _{i}), \Gamma ^{+}(\beta _{i})\right )$
is
$D_{0}$
-aligned. When
$\alpha _{i} = \beta _{i}$
, this is guaranteed by the definition of
$K_{0}$
-Schottky sets.
Now suppose
$\alpha _{i} \neq \beta _{i}$
. First,
$(\bar {\Gamma }^{-}(\alpha _{i}), o)$
is
$0$
-aligned. Second,
$\left (a_{i}^{-1} o, \Gamma ^{-}(\alpha _{i})\right )$
is not
$K_{0}$
-aligned as
$d(o, a_{i}^{-1} o) \geqslant 100E_{0} \geqslant K_{0}$
. Then by the Schottky property of
$S$
,
$\left (a_{i}^{-1} o, \Gamma ^{+}(\beta _{i})\right )$
is
$K_{0}$
-aligned. Now Lemma 3.8 tells us that
$\left (\bar {\Gamma }^{-}(\alpha _{i}), \Gamma ^{+}(\beta _{i})\right )$
is
$D_{0}$
-aligned.
The alignment of
$\left (\Upsilon (\gamma _{i}), \Upsilon (\delta _{i})\right )$
holds for the same reason.
Lemma 5.3 ([[Reference ChoiCho24], Lemma 3.2]). Let
$k \in \mathbb{Z}_{\gt 0}$
. Let
$l\lt m$
be consecutive elements in
$P_{k}$
, i.e.,
$m \in P_{k}$
and
$l = \max (P_{k} \cap \{1, \ldots , m-1\})$
. Then
$\left (\Upsilon (\delta _{l}), \Upsilon (\alpha _{m}) \right )$
is
$D_{0}$
-semi-aligned.
Proof.
$l, m \in P_{k}$
implies that
$l \in P_{l}$
and
$l, m \in P_{m}$
. In particular,
$l$
and
$m$
are newly chosen at step
$l$
and
$m$
, respectively, by fulfilling Criterion (A). Hence,
$(\Upsilon (\delta _{l}), w_{l+1, 2}^{-}o)$
and
$(z_{m-1}, \Upsilon (\alpha _{m}))$
are
$K_{0}$
-aligned (
$\ast$
), and
$z_{l} = w_{l, 1}^{+}o$
. Moreover, we have
$P_{m} = P_{m-1} \cup \{m\}$
and
$l = \max P_{m-1}$
.
If
$l = m-1$
and
$m$
was newly chosen at step
$m = l+1$
, then
$z_{m-1} = z_{l} = w_{l,1}^{+}o$
holds. Lemma 3.8 and (
$\ast$
) imply that
$\left (\Upsilon (\delta _{l}), \Upsilon (\alpha _{m})\right )$
is
$D_{0}$
-aligned.
If
$l \lt m-1$
, then
$l=\max P_{m-1}$
has survived at step
$m-1$
by fulfilling Criterion (B); there exist
$j \gt l$
such that
$\big ( \Upsilon (\delta _{l}), \Upsilon (\beta _{j})\big )$
is
$D_{0}$
-semi-aligned and
$\big (\Upsilon (\beta _{j}), w_{n+1, 2}^{-} o\big )$
is
$K_{0}$
-aligned. Furthermore,
$z_{m-1}$
equals
$w_{j, 1}^{-} o$
, the beginning point of
$\Upsilon (\beta _{j})$
.
Note that
$(z_{m-1}, \Upsilon (\alpha _{m})\big )$
is
$K_{0}$
-aligned by (
$\ast$
). Lemma 3.8 then asserts that
$\left (\Upsilon (\beta _{j}), \Upsilon (\alpha _{m})\right )$
is
$D_{0}$
-aligned. Concatenating the two
$D_{0}$
-semi-aligned sequences, we conclude that
$\left (\Upsilon (\delta _{l}), \Upsilon (\alpha _{m})\right )$
is
$D_{0}$
-semi-aligned.
Proof of Proposition 5.1. Having established Lemma 5.3, it remains to prove that:
-
–
$\left (o, \Upsilon (\alpha _{i(1)})\right )$ is
$K_{0}$ -aligned;
-
– for
$1 \leqslant t \leqslant m$ ,
$\left (\Upsilon (\alpha _{i(t)}), \Upsilon (\beta _{i(t)}), \Upsilon (\gamma _{i(t)}), \Upsilon (\delta _{i(t)})\right )$ is
$D_{0}$ -aligned;
-
–
$\left (\Upsilon (\delta _{i(m)}), w_{n+1, 2}^{-}o\right )$ is
$D_{0}$ -semi-aligned.
Note that for each
$t =1, \ldots , m$
,
$i(t)$
is newly chosen as a pivotal time at step
$i(t)$
by fulfilling Criterion (A). In particular, we have that:
-
–
$\left (\Upsilon (\alpha _{i(t)}), \Upsilon (\beta _{i(t)})\right )$ is
$D_{0}$ -aligned (Lemma 5.2);
-
–
$\left (\Upsilon (\beta _{i(t)}), \Upsilon (\gamma _{i(t)})\right )$ is
$D_{0}$ -aligned since
$\left (\Upsilon (\beta _{i(t)}), w_{n, 1}^{+}o\right )$ and
$\left (w_{i(t), 0}^{-}o, \Upsilon (\gamma _{i(t)})\right )$ are
$K_{0}$ -aligned (Lemma 3.8), and
-
–
$\left (\Upsilon (\gamma _{i(t)}), \Upsilon (\delta _{i(t)})\right )$ is
$D_{0}$ -aligned (Lemma 5.2).
This guarantees the second item.
We also note that
$P_{i(1)-1} = \emptyset$
. Indeed, any
$j$
in
$P_{i(1) - 1}$
is smaller than
$i(1)$
and would have survived in
$P_{i(1)}$
(since what happened at step
$i(1)$
was adding an element, not deleting some). Since
$i(1)$
was not deleted at any later step, such
$j$
would also not be deleted till the end and should have appeared in
$P_{n}$
. Since
$i(1)$
is the earliest pivotal time in
$P_{n}$
, no such
$j$
exists. Hence,
$z_{i(1) -1} = o$
and Criterion (A) for
$i(1)$
leads to the first item.
We now observe how
$i(m)$
survived in
$P_{n}$
. If
$i(m) = n$
, then it was newly chosen at step
$n$
by fulfilling Criterion (A). In particular,
$(\Upsilon (\delta _{n}), w_{n+1, 2}^{-}o)$
is
$K_{0}$
-aligned as desired.
If
$i(m) \neq n$
, then it has survived at step
$n$
as the last pivotal time by fulfilling Criterion (B). In particular, there exist
$j \gt i(m)$
such that
$\big (\Upsilon (\delta _{i(m)}), \Upsilon (\beta _{j}) \big )$
is
$D_{0}$
-semi-aligned and such that
$(\Upsilon (\beta _{j}), w_{n+1, 2}^{-} o)$
is
$K_{0}$
-aligned. In particular,
$\big (\Upsilon (\delta _{i(m)}), \Upsilon (\beta _{j}) , w_{n+1, 2}^{-} o\big )$
is
$D_{0}$
-semi-aligned.
Next, we study when
$P_{n} = P_{n-1} \cup \{n\}$
happens, i.e., a new pivotal time is added to the set of pivotal times. This will guide us how to pivot the direction at a pivotal time without affecting the set of pivotal times. Recall that we draw
$\alpha _{i}, \beta _{i}, \gamma _{i}, \delta _{i}$
’s from
$S$
with the uniform measure.
Lemma 5.4.
Let us fix
$\mathbf{w} = (w_{i})_{i}$
,
$\mathbf{v} = (v_{i})_{i}$
and
$\mathbf{s} \in S^{4(n-1)}$
. Then

Proof.
Recall Criterion (A) for
$\#P_{n} = \# P_{n-1} + 1$
. We will investigate the four required conditions one-by-one.
First, the condition

depends only on
$\gamma _{n}$
. This holds for at least
$(\#S - 1)$
choices in
$S$
by the
$K_{0}$
-Schottky-ness of
$S$
.
Similarly, the condition

depends only on
$\delta _{n}$
, and holds for at least
$(\#S - 1)$
choices in
$S$
.
Fixing the choice of
$\gamma _{n}$
, the condition

depends only on
$\beta _{n}$
. This holds for at least
$(\#S - 1)$
choices in
$S$
.
We now additionally fix the choice of
$s = (\alpha _{1}, \beta _{1}, \gamma _{1}, \delta _{1}, \ldots , \alpha _{n-1}, \beta _{n-1}, \gamma _{n-1}, \delta _{n-1})$
; in particular,
$w_{n, 2}^{-}$
and
$z_{n-1}$
are now determined. Then the condition

depends on
$\alpha _{n}$
. This holds for at least
$(\#S - 1)$
choices of
$\alpha _{n}$
.
In summary, the probability that Criterion (A) holds is at least

We now define the set
$\tilde {\mathrm{S}}$
of triples
$(\beta , \gamma , v) \in S^{2} \times G$
that satisfy Condition 19 and 21:

We also define its section for each
$v \in G$
:

While checking Display 19 and 21, we observed that
$\#\tilde {\mathrm{S}}(v) \geqslant \#S^{2} - 2\#S$
for each
$v\in G$
. We now define pivoting.
Lemma 5.5.
Let
$\mathbf{s} = (\alpha _{1}, \beta _{1}, \gamma _{1}, \delta _{1}, \ldots , \alpha _{n}, \beta _{n}, \gamma _{n}, \delta _{n})$
be a choice drawn from
$S^{4n}$
and let
$\mathbf{w}, \mathbf{v}$
be auxiliary sequences in
$G$
.
Let
$k \in P_{n}(\mathbf{s}; \mathbf{w}, \mathbf{v})$
and let (
$\bar {\mathbf{s}}$
;
$\mathbf{w}, \bar {\mathbf{v}})$
be obtained from
$(\mathbf{s}; \mathbf{w}, \mathbf{v})$
by replacing
$(\beta _{k}, \gamma _{k}, v_{k})$
with some
$(\bar {\beta }_{k}, \bar {\gamma }_{k}, \bar {v}_{k})$
chosen from
$\tilde {S}$
.
Then
$P_{l}(\mathbf{s}; \mathbf{w}, \mathbf{v}) = P_{l}(\bar {\mathbf{s}}; \mathbf{w}, \bar {\mathbf{v}})$
for any
$1 \leqslant l \leqslant n$
.
On Gromov hyperbolic spaces, this corresponds to [[Reference GouëzelGou22], Lemma 5.7].
Proof.
Since
$\alpha _{1}, \beta _{1}, \gamma _{1}, \delta _{1}, \ldots , \alpha _{k-1}, \beta _{k-1}, \gamma _{k-1}, \delta _{k-1}$
are intact,
$P_{l}(\mathbf{s}) = P_{l}(\bar {\mathbf{s}})$
and
$\tilde {\mathrm{S}}_{l}(\mathbf{s}) = \tilde {\mathrm{S}}_{l}(\bar {\mathbf{s}})$
hold for
$l=0, \ldots , k-1$
. At step
$k$
,
$\alpha _{k}$
and
$\delta _{k}$
satisfy Condition 22 and Condition 20 since
$k \in P_{n}(\mathbf{s})$
. Furthermore,
$\bar {\beta }_{k}$
and
$\bar {\gamma }_{k}$
satisfy Condition 19 and 21 for the new choice
$\bar {v}_{k}$
:

since
$(\bar {\beta }_{k}, \bar {\gamma }_{k}, \bar {v}_{k}) \in \tilde {\mathrm{S}}$
. Hence,
$k$
is newly added in
$P_{k}(\bar {\mathbf{s}})$
and

Meanwhile,
$z_{k}$
is modified into
$\bar {z}_{k} = \bar {w}_{k, 1}^{+}o = g w_{k, 1}^{+}o=gz_{k}$
, where
$g := w_{k, 2}^{-} a_{k}\bar {b}_{k} \bar {v}_{k}\bar {c}_{k} (w_{k, 2}^{-} a_{k}b_{k}v_{k}c_{k})^{-1}$
. More generally, we have

We now claim the following for
$k \lt l \leqslant n$
:
-
(i) If
$s$ fulfills Criterion (A) at step
$l$ , then so does
$\bar {\mathbf{s}}$ .
-
(ii) If not and if
$(i, j)$ is the maximal pair of indices for
$s$ in Criterion (B) at step
$l$ , then it is also the maximal one for
$\bar {\mathbf{s}}$ at step
$l$ .
-
(iii) In both cases, we have
$P_{l}(\mathbf{s}) = P_{l}(\bar {\mathbf{s}})$ and
$\bar {z}_{l} = g z_{l}$ .
Assuming the third item for
$l-1$
:
$P_{l-1}(\mathbf{s}) = P_{l-1}(\bar {\mathbf{s}})$
and
$\bar {z}_{l-1} = g z_{l-1}$
, Equality 23 implies the first item. In this case we deduce
$P_{l}(\mathbf{s}) = P_{l-1}(\mathbf{s}) \cup \{l\} = P_{l-1}(\bar {\mathbf{s}}) \cup \{l\} = P_{l}(\bar {\mathbf{s}})$
and
$\bar {z}_{l} = \bar {w}_{l, 1}^{+}o = gw_{l, 1}^{+}o = gz_{l}$
, the third item for
$l$
.
Furthermore, Equality 23 implies that
$i$
in
$P_{l-1}(\mathbf{s}) \cap \{k, \ldots , l-1\} = P_{l-1}(\bar {\mathbf{s}})\cap \{k, \ldots , l-1\}$
and
$j\gt i$
work for
$\mathbf{s}$
in Criterion (B) if and only if they work for
$\bar {\mathbf{s}}$
. Such
$i$
can be found in
$\{k, \ldots , l-1\}$
, because
$k$
survived in
$P_{n}(\mathbf{s})$
and should not have been erased at step
$l$
. Hence, the maximal pair
$(i, j)$
for
$\mathbf{s}$
is also maximal for
$\bar {\mathbf{s}}$
. We then deduce
$P_{l}(\mathbf{s}) = P_{l-1}(\mathbf{s}) \cap \{1, \ldots , i\} = P_{l-1}(\bar {\mathbf{s}}) \cap \{1, \ldots , i\} = P_{l}(\bar {\mathbf{s}})$
and
$\bar {z}_{l} = \bar {w}_{j, 1}^{-}o = gw_{j, 1}^{-}o = gz_{l}$
(using
$j \gt i$
), the third item for
$l$
.
For
$\mathbf{s}, \mathbf{s}' \in S^{4n}$
and sequences
$\mathbf{w}, \mathbf{v}$
,
$\bar {\mathbf{v}}$
in
$G$
, we say that
$(\bar {\mathbf{s}}; \mathbf{w}, \bar {\mathbf{v}})$
is pivoted from
$(\mathbf{s}; \mathbf{w}, \mathbf{v})$
if:
-
–
$\alpha _{i} = \bar {\alpha }_{i}$ ,
$\delta _{i} = \bar {\delta }_{i}$ for all
$i \in \{1, \ldots , n\}$ ;
-
–
$(\bar {\beta }_{i}, \bar {\gamma }_{i}, \bar {v}_{i}) \in \tilde {\mathrm{S}}$ for each
$i \in P_{n}(\mathbf{s}; \mathbf{w}, \mathbf{v})$ , and
-
–
$(\beta _{i}, \gamma _{i}, v_{i})= (\bar {\beta }_{i}, \bar {\gamma }_{i}, \bar {v}_{i})$ for each
$i \in \{1, \ldots , n\} \setminus P_{n}(\mathbf{s}; \mathbf{w}, \mathbf{v})$ .
By Lemma 5.5, being pivoted from each other is an equivalence relation.
Fixing
$\mathbf{w}$
and
$\mathbf{v}$
, for each
$\mathbf{s} \in S^{4n}$
let
$\mathcal{E}_{n}(\mathbf{s})$
be the equivalence class of
$s$
:

We endow
$\mathcal{E}_{n}(\mathbf{s})$
with the conditional probability of the uniform measure on
$S^{4n}$
. We now claim that
$\#P_{n+1} - \#P_{n}$
conditioned on an equivalence class
$\mathcal{E}_{n}(\mathbf{s})$
till step
$n$
and the choice at step
$n+1$
has uniform exponential tail.
Proposition 5.6.
Fix
$\mathbf{w} = (w_{i})_{i=0}^{\infty }$
and
$\mathbf{v} = (v_{i})_{i=1}^{\infty }$
. For each
$j \geqslant 0$
and
$\mathbf{s} \in S^{4n}$
,

is less than
$(4/N_{0})^{j+1}$
.
On Gromov hyperbolic spaces, this corresponds to [[Reference GouëzelGou22], Lemma 5.8].
Proof.
An element
$\tilde {\mathbf{s}} \in \mathcal{E}_{n}(\mathbf{s})$
is determined by its coordinates
$(\tilde {\beta }_{k}, \tilde {\gamma }_{k})_{k \in P_{n}(\mathbf{s})}$
subject to the condition
$(\tilde {\beta }_{k}, \tilde {\gamma }_{k}, v_{k}) \in \tilde {\mathrm{S}}$
. We consider a finer equivalence class by additionally fixing the coordinates
$\gamma _{k}$
’s: for
$\tilde {\mathbf{s}} \in \mathcal{E}_{n}(\mathbf{s})$
, let
$\mathcal{E}_{n}'(\tilde {\mathbf{s}})$
be the set of
$\bar {\mathbf{s}} \in \mathcal{E}_{n}(\mathbf{s})$
such that
$\bar {\gamma }_{k} = \tilde {\gamma }_{k}$
for all
$k$
. Then
$\mathcal{E}_{n}(\mathbf{s})$
is partitioned into
$\{ \mathcal{E}_{n}'(\tilde {\mathbf{s}}): \tilde {\mathbf{s}} \in \mathcal{E}_{n}(\mathbf{s})\}$
, and it suffices to establish the estimates on each
$\mathcal{E}_{n}'(\tilde {\mathbf{s}})$
. Henceforth, we will prove that

is less than
$(4/N_{0})^{j+1}$
for each
$\mathbf{s} = (\alpha _{1}, \beta _{1}, \gamma _{1}, \delta _{1}, \ldots , \alpha _{n}, \beta _{n}, \gamma _{n}, \delta _{n}) \in S^{4n}$
and
$j \geqslant 0$
.
Recall that we are fixing the sequences
$\mathbf{w}$
and
$\mathbf{v}$
throughout the proof. Let us define

Then
$\mathcal{E}_{n}'(\mathbf{s})$
is parametrized by
$\prod _{i \in P_{n}(\mathbf{s})} \tilde {\mathrm{S}}_{k}$
with the uniform measure. Let

Lemma 5.4 implies that
$\mathbb{P}(\mathcal{A}) \geqslant 1-4/N_{0}$
with respect to the uniform measure on
$S^{4}$
. Note that for each element
$(\alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1})$
of
$\mathcal{A}$
, we have

Hence, for each
$\tilde {\mathbf{s}}\in \mathcal{E}_{n}'(\mathbf{s})$
,
$(\tilde {\mathbf{s}}, \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1})$
(as a choice in
$S^{4(n+1)}$
) is pivoted from
$(\mathbf{s}, \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1})$
and
$P_{n+1}(\tilde {\mathbf{s}}) = P_{n+1}(\mathbf{s}) = P_{n}(\mathbf{s}) \cup \{n+1\} = P_{n}(\tilde {\mathbf{s}}) \cup \{n+1\}$
. Thanks to this, we have

This settles the case
$j=0$
.
Now let
$j = 1$
. The event under discussion becomes void when
$\# P_{n}(\mathbf{s}) \lt 2$
. Excluding such cases, let
$l\lt m$
be the last 2 elements of
$P_{n}(\mathbf{s})$
. We now freeze the coordinates
$\beta _{k}$
’s except for
$k=m$
. Namely, for
$\tilde {\mathbf{s}} \in \mathcal{E}_{n}'(\mathbf{s})$
, let
$E^{(m)}(\tilde {\mathbf{s}})$
be the set of
$\bar {\mathbf{s}} \in \mathcal{E}_{n}'(\mathbf{s})$
such that
$\bar {\beta }_{k} = \tilde {\beta }_{k}$
for
$k \in P_{n}(\mathbf{s}) \setminus \{m\}$
. Then
$\{E^{(m)}(\tilde {\mathbf{s}}) : \tilde {\mathbf{s}} \in \mathcal{E}_{n}'(\mathbf{s}) \}$
becomes a partition of
$\mathcal{E}_{n}'(\mathbf{s})$
, and
$E^{(m)}(\tilde {\mathbf{s}})$
is parametrized by
$\bar {\beta }_{m} \in \tilde {\mathrm{S}}_{m}$
with the uniform measure. Note that
$\tilde {\mathrm{S}}_{m}$
has at least
$\#S - 1$
elements.
Fixing
$(\alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}) \in S^{4}$
, and let
$F^{(m)}(\tilde {\mathbf{s}})$
be the set of
$\bar {\mathbf{s}} \in E^{(m)}(\tilde {\mathbf{s}})$
such that
$\left (\Upsilon (\bar {\beta }_{m}), \bar {w}_{n+2, 2}^{-} o \right )$
$K_{0}$
-aligned, or more precisely,

This amounts to requiring a new Schottky condition to
$\bar {\beta }_{m}$
, in addition to the alignment of
$\big ( \bar {\Gamma }^{-1}(\bar {\beta }_{m}), v_{m} \Pi (\gamma _{m}) o\big )$
; there are at least
$\#S - 2$
choices that additionally satisfy this.
We now claim
$\# P_{n+1}(\bar {\mathbf{s}}, \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}) \geqslant \# P_{n}(\mathbf{s}) - 1$
for
$\bar {\mathbf{s}} \in F^{(m)}(\tilde {\mathbf{s}})$
. First, since
$l\lt m$
are consecutive elements in
$P_{n}(\mathbf{s}) = P_{n}(\bar {\mathbf{s}})$
, Lemma 5.3 asserts that
$\left (\Upsilon (\bar {\delta }_{l}), \Upsilon (\bar {\alpha }_{m}) \right )$
is
$D_{0}$
-semi-aligned. Moreover, Lemma 5.2 and Condition 24 imply that

are
$D_{0}$
-aligned and
$K_{0}$
-aligned, respectively. These together imply that

are
$D_{0}$
-semi-aligned and
$K_{0}$
-aligned, respectively: the pair
$(l, m)$
qualifies Criterion (B) at step
$n+1$
. Hence,
$P_{n+1}(\bar {\mathbf{s}}, \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}) \supseteq P_{n}(\bar {\mathbf{s}}) \cap \{1, \ldots , l\}$
.
As a result, for each
$\tilde {\mathbf{s}} \in \mathcal{E}_{n}(\mathbf{s})$
and
$(\alpha _{n+1}, \ldots , \delta _{n+1}) \in S^{4}$
we have

Since
$\{ E^{(m)}(\tilde {\mathbf{s}}): \tilde {\mathbf{s}} \in \mathcal{E}_{n}(\mathbf{s})\}$
partitions
$\mathcal{E}_{n}(\mathbf{s})$
, we deduce

for each
$(\alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}) \in S^{4}$
. Moreover, this probability vanishes when
$(\alpha _{n+1}, \ldots , \delta _{n+1}) \in \mathcal{A}$
. Since
$\mathbb{P}\big ((\alpha _{n+1}, \ldots , \delta _{n+1}) \in \mathcal{A} \,\big |\, (\alpha _{n+1}, \ldots , \delta _{n+1}) \in S^{4}\big ) \geqslant 1-4/N_{0}$
, we deduce that

Now let
$j =2$
. Excluding the void case, we assume that
$\#P_{n}(\mathbf{s}) \geqslant 3$
; let
$l' \lt l\lt m$
be the last 3 elements. To ease the notation, for
$\beta \in S$
we define

In other words,
$\mathbf{s}'(\beta )$
is obtained from
$\mathbf{s}$
by replacing
$\beta _{m}$
with
$\beta$
. Now let

Equivalently, we are requiring

(This equivalence relies on the fact
$P_{n}(\mathbf{s}') = P_{n}(\mathbf{s})$
due to Lemma 5.5.)
Observation 5.7. For each

$(\tilde {\beta }_{m}, \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}) \in \mathcal{A}_{1}$
if and only if
$\#P_{n+1}(\tilde {\mathbf{s}}, \alpha _{n+1}, \ldots , \delta _{n+1}) \geqslant \# P_{n}(\mathbf{s}) - 1$
.
To see this, suppose first that
$(\tilde {\beta }_{m}, \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}) \in \mathcal{A}_{1}$
. Then
$(\tilde {\mathbf{s}}, \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1})$
is pivoted from
$\left (\mathbf{s}'( \tilde {\beta }_{m}), \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}\right )$
, as they differ only at entries
$\beta _{k}$
’s for
$k\in P_{n}(\mathbf{s}) \cap \{1, \ldots , l\} \subseteq P_{n+1}\left (\mathbf{s}', \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}\right )$
. Lemma 5.5 then implies that

and
$\#P_{n+1}(\tilde {\mathbf{s}}, \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}) \geqslant \# P_{n}(\mathbf{s}) - 1$
.
Conversely, suppose
$\#P_{n+1}(\tilde {\mathbf{s}}, \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}) \geqslant \# P_{n}(\mathbf{s}) - 1$
. (
$\ast$
) Recall that
$P_{n}(\tilde {\mathbf{s}}) = P_{n}(\mathbf{s})$
, and recall that
$P_{n+1}(\tilde {\mathbf{s}}, \alpha _{n+1}, \ldots , \delta _{n+1})$
is either an initial section of
$P_{n}(\tilde {s})$
or contains
$P_{n}(\tilde {s})$
. Considering these, the assumption (
$\ast )$
implies

Then
$\left (\mathbf{s}'( \tilde {\beta }_{m}), \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}\right )$
is pivoted from
$(\tilde {\mathbf{s}}, \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1})$
, as the former choice differs from the latter choice only at entries
$(\tilde {\alpha }_{k}, \tilde {\beta }_{k}, \tilde {\gamma }_{k}$
)’s for
$k \in P_{n}(\mathbf{s}) \cap \{1, \ldots , l\} \subseteq P_{n+1}\left (\tilde {\mathbf{s}}, \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}\right )$
. Lemma 5.5 then implies that

and
$(\tilde {\beta }_{m}, \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}) \in \mathcal{A}_{1}$
.
Combining Observation 5.7 and Inequality 25, we deduce

This time, we freeze the coordinates
$\beta _{k}$
’s except for
$k=l$
: for
$\tilde {\mathbf{s}} \in \mathcal{E}_{n}'(\mathbf{s})$
, let
$E^{(l)}(\tilde {\mathbf{s}})$
be the set of
$\bar {\mathbf{s}} \in \mathcal{E}_{n}'(\mathbf{s})$
such that
$\bar {\beta }_{k} = \tilde {\beta }_{k}$
for
$k \in P_{n}(\mathbf{s}) \setminus \{l\}$
. Then
$\{E^{(l)}(\tilde {\mathbf{s}}) : \tilde {\mathbf{s}} \in \mathcal{E}_{n}'(\mathbf{s})\}$
partitions
$\mathcal{E}_{n}'(\mathbf{s})$
and
$E^{(l)}(\tilde {\mathbf{s}})$
is parametrized by
$\bar {\beta }_{l} \in \tilde {\mathrm{S}}_{l}$
with the uniform measure; note that
$\# \tilde {\mathrm{S}}_{l} \geqslant \#S - 1$
.
Fixing
$\tilde {\mathbf{s}}$
, now pick
$(\alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}) \in S^{4}$
and let
$F^{(l)}(\tilde {\mathbf{s}})$
be the set of
$\bar {\mathbf{s}} \in E^{(m)}(\tilde {\mathbf{s}})$
such that
$\left (\Upsilon (\bar {\beta }_{l}), \bar {w}_{n+2, 2}^{-} o \right )$
$K_{0}$
-aligned, i.e.,

This amounts to requiring another Schottky condition to
$\bar {\beta }_{l}$
; there are at least
$\#S - 2$
choices that additionally satisfy this.
We now claim that
$\#P_{n+1}(\bar {\mathbf{s}}, \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}) \geqslant \# P_{n}(\mathbf{s}) - 2$
for
$\bar {\mathbf{s}} \in F^{(l)}(\tilde {\mathbf{s}})$
. First, since
$l' \lt l$
are consecutive elements in
$P_{n}(\bar {\mathbf{s}})$
, Lemma 5.3 asserts that

is
$D_{0}$
-semi-aligned. Moreover, Lemma 5.2 and Condition 26 imply that

is
$D_{0}$
-aligned and
$K_{0}$
-aligned, respectively. Combining these, we observe that the pair
$(l', l)$
qualifies Criterion (B) at step
$n+1$
. This implies
$P_{n+1}(\bar {\mathbf{s}}, \alpha _{n+1}, \beta _{n+1}, \gamma _{n+1}, \delta _{n+1}) \supseteq P_{n}(\bar {\mathbf{s}}) \cap \{1, \ldots , l'\}$
, hence the claim.
As a result, for each
$\tilde {\mathbf{s}} \in \mathcal{E}_{n}(\mathbf{s})$
and
$(\alpha _{n+1}, \ldots , \delta _{n+1}) \in S^{4}$
we have

Moreover, Observation 5.7 asserts that the above probability vanishes for those equivalence classes
$E^{(l)}(\tilde {\mathbf{s}})$
such that
$(\tilde {\beta }_{m}, \alpha _{n+1}, \ldots , \delta _{n+1}) \in \mathcal{A}_{1}$
. Since
$\mathbb{P}[\mathcal{A}_{1} | \tilde {\mathrm{S}}_{m} \times S^{4}] \leqslant (4/N_{0})^{2}$
, we conclude

We repeat this procedure for
$j\lt \# P_{n}(\mathbf{s})$
. The case
$j \geqslant \# P_{n}(\mathbf{s})$
is void.
Corollary 5.8.
Let us fix
$\mathbf{w}$
and
$\mathbf{v}$
. When
$\mathbf{s} = (\alpha _{i}, \beta _{i}, \gamma _{i}, \delta _{i})_{i=1}^{n}$
is chosen from
$S^{4n}$
with the uniform measure,
$\# P_{n}(\mathbf{s})$
is greater in distribution than the sum of
$n$
i.i.d.
$X_{i}$
, whose distribution is given by

Moreover, we have
$\mathbb{P}(\#P_{n}(\mathbf{s}) \leqslant (1-10/N_{0}) n) \leqslant e^{-n/K}$
for some
$K\gt 0$
.
Proof.
Let
$\{X_{i}\}_{i}$
be the family of i.i.d. as in Equation 28 that is also assumed to be independent from the choice
$\mathbf{s}$
. Lemma 5.4 and Proposition 5.6 together imply the following for each
$0\leqslant k \lt n$
:

Hence, there exists a nonnegative random variable
$U_{k}$
such that
$\# P_{k+1} - U_{k}$
and
$\# P_{k} + X_{k+1}$
have the same distribution.
For each
$1 \leqslant k \leqslant n$
, we claim that
$\mathbb{P}(\#P_{k} \geqslant i) \geqslant \mathbb{P} (X_{1} + \cdots + X_{k} \geqslant i)$
for each
$i$
. For
$k=1$
, we have
$\#P_{k-1}= 0$
and the claim follows from Inequality 29. Given the claim for
$k$
, we have

The second claim holds since
$X_{i}$
’s have finite exponential moment and
$\mathbb{E}[X_{i}] \geqslant 1-9/N_{0}$
.
We now describe a simpler situation when
$v_{i} = id$
for all
$i$
, i.e., we study

Before defining the pivoting, note that the conditions

are satisfied by every pair of Schottky sequences
$(\beta , \gamma ) \in S^{2}$
, as we proved in Lemma 5.2. Hence, Criterion (A) while defining the set of pivotal times is simplified as follows:
-
When
$\left (z_{n-1}, \Upsilon (\alpha _{n})\right )$ and
$\left (\Upsilon (\delta _{n}), w_{n+1, 2}^{-}o\right )$ are
$K_{0}$ -aligned, we set
$P_{n} = P_{n-1} \cup \{n\}$ and
$z_{n} = w_{n, 1}^{+}o$ (see Figure 4).
Moreover,
$\tilde {\mathrm{S}}$
contains all of
$\{(\beta , \gamma , id) : \beta , \gamma \in S\}$
. Hence, when
$\mathbf{w}$
and
$\mathbf{v}= (id)_{i=1}^{\infty }$
are fixed, the previous definition reads as follows: given a choice
$\mathbf{s} = (\alpha _{1}, \beta _{1}, \ldots , \gamma _{n}, \delta _{n})$
in
$S^{4n}$
, we say that
$\bar {\mathbf{s}} \in S^{4n}$
is pivoted from
$\mathbf{s}$
if:
-
–
$\alpha _{i} = \bar {\alpha }_{i}$ ,
$\delta _{i} = \bar {\delta }_{i}$ for all
$i \in \{1, \ldots , n\}$ ;
-
–
$(\beta _{i}, \gamma _{i})= (\bar {\beta }_{i}, \bar {\gamma }_{i})$ for each
$i \in \{1, \ldots , n\} \setminus P_{n}(\mathbf{s})$ .
Therefore, for
$\mathbf{s} \in S^{4n}$
,
$\bar {\mathbf{s}} \in \mathcal{E}_{n}(\mathbf{s})$
is parametrized by their choices
$(\bar {\beta }_{i}, \bar {\gamma }_{i})_{i \in P_{n}(\mathbf{s})}$
distributed according to the uniform measure on
$S^{2\#P_{n}(\mathbf{s})}$
.
5.2 Pivotal times in random walks
In this subsection, we define pivotal times for random walks and prove Proposition 4.2. Let
$\mu$
be a non-elementary probability measure on
$G$
and
$S \subseteq (\textrm {supp} \mu )^{M_{0}}$
be a large enough
$K_{0}$
-Schottky set with cardinality
$N_{0} \geqslant 400$
. We also fix an integer
$n\geqslant 0$
.
Let
$\mu _{S}$
be the uniform measure on
$S$
. By taking suitably small
$\alpha$
, we can decompose
$\mu ^{4M_{0}}$
as

for some probability measure
$\nu$
. We then consider Bernoulli RVs
$(\rho _{i})_{i}$
with
$\mathbb{P}(\rho _{i} = 1) = \alpha$
and
$\mathbb{P}(\rho _{i} = 0) = 1-\alpha$
,
$(\eta _{i})_{i}$
with the law
$\mu _{S}^{4}$
and
$(\nu _{i})_{i}$
with the law
$\nu$
, all independent, and define

Then
$(g_{i})_{i =1}^{\infty }$
has the law
$\mu ^{\infty }$
. Since we need to prove Proposition 4.2 by fixing the choice of
$g_{1}, \ldots , g_{\lfloor n/2\rfloor + 1}$
and
$g_{n+1}$
, we slightly modify
$\rho _{i}$
’s, namely,

Let
$\Omega$
be the ambient probability space on which the above RVs are all measurable. We denote by
$\mathscr{B}(k) := \sum _{i=0}^{k} \rho _{i}^{(n)}$
the number of the Schottky slots till
$k$
and by
$\vartheta (i) := \min \{j \geqslant 0 : \mathscr{B}(j) = i\}$
the
$i$
-th Schottky slot. We also set
$\vartheta (0) = -1$
. Note that
$\{j \geqslant 0 : \rho _{j}^{(n)} = 1\}= \{\vartheta (1) \lt \vartheta (2) \lt \ldots \}$
.
For each
$\omega \in \Omega$
and
$i \geqslant 1$
we define

In other words,
$\eta _{\vartheta (i)}$
becomes
$(\alpha _{i}, \beta _{i}, \gamma _{i}, \delta _{i})$
(with
$M_{0}$
steps each) and
$w_{i}$
is the product of intermediate steps between
$\eta _{\vartheta (i-1)}$
and
$\eta _{\vartheta (i)}$
. As in Subsection 5.1, we write
$a_{i} := \Pi (\alpha _{i})$
,
$b_{i} := \Pi (\beta _{i})$
and so on. We then have

for each
$l\gt 0$
. Following the discussion in Subsection 5.1, we define

We finally define

Recall that
$P_{k}$
is formed from
$P_{k-1}$
by adjoining a new element
$k$
or taking an initial section of
$P_{k-1}$
. Hence, any initial section
$\{i(1) \lt \ldots \lt i(N)\}$
of
$\liminf _{k} P_{k}(\omega )$
is an initial section of some
$P_{m}(\omega )$
(in fact, for all sufficiently large
$m$
). Proposition 5.1 then tells us that:
Observation 5.9.
Let
$\mathcal{P}(\omega ) = \{i(1) \lt i(2) \lt \ldots \}$
. Then

is
$D_{0}$
-semi-aligned.
Note that
$(w_{i})_{i}$
’s and
$(\alpha _{i}, \beta _{i}, \gamma _{i}, \delta _{i})_{i \gt 0}$
are independent, the latter being i.i.d. with the uniform distribution on
$S^{4}$
. By Corollary 5.8,
$P_{k}$
linearly increases:
Observation 5.10.
There exists
$K\gt 1$
such that

for every
$k\gt 0$
and every choice of
$(w_{i})_{i}$
.
Here, the growth rate is independent of
$g_{i}$
’s that are not involved in
$(\alpha _{i}, \beta _{i}, \gamma _{i}, \delta _{i})$
’s. In particular, it is independent of
$g_{1}, \ldots , g_{\lfloor n/2 \rfloor + 1}$
and
$g_{n+1}$
.
To couple the words
$w_{k, 2}^{-}$
’s and the actual random walk
$Z_{n}$
’s, we need to control
$\vartheta (i)$
’s. For each
$k \geqslant \lfloor n/4M_{0}\rfloor$
and
$L\gt 0$
, we have

By plugging in
$L = \frac {\log (1+\alpha ^{2}/2)}{3M_{0}} k$
, we obtain

for some
$K' =K'(\alpha , M_{0})\gt 1$
(independent of
$\alpha$
).
Let us now combine the ingredients and prove Proposition 4.2. Given the measure
$\mu$
, integers
$k \geqslant n$
, and the choices of
$g_{1}, \ldots , g_{\lfloor n/2 \rfloor + 1}$
and
$g_{n+1}$
, we do the above construction. Then we have

Let us fix a combination of the values of
$\{\rho _{i}^{(n)} : i \gt 0\}$
in
$A_{k}^{c}$
, which determines
$\vartheta (i)$
’s. Furthermore, we fix a combination of the values of
$\{\eta _{i} : i\gt 0\}$
. These choices determine

and consequently
$w_{i}$
’s. Note also that
$\{(\alpha _{i}, \beta _{i}, \gamma _{i}, \delta _{i}) = \eta _{\vartheta (i)} : i\gt 0\}$
are i.i.d.s distributed according to
$\mu _{S}^{4}$
. Hence, conditioned on choices of
$\{\rho _{i}^{(n)} : i \gt 0\} \in A_{k}^{c}$
and
$\{\eta _{i} : i\gt 0\}$
, we are now reduced to the combinatorial model. From Observation 5.10, we deduce that

In other words, except for probability
$\frac {K}{1-e^{-1/K}} e^{-\lfloor k/4KK'M_{0} \rfloor }$
(under the conditioning), the initial
$\lfloor k/4KK'M_{0} \rfloor$
-sections of
$P_{\lfloor k/4K'M_{0} \rfloor }(\omega )$
remains the same in
$P_{l}(\omega )$
for
$l \geqslant \lfloor k/4M_{0} \rfloor$
. Hence, it becomes an initial section of
$\liminf _{l} P_{l}(\omega )$
. This means that

Meanwhile, since
$\{\rho _{i}^{(n)} : i \gt 0\}$
is determined in
$A_{k}^{c}$
, we have
$\mathscr{B}(\lfloor k/4M_{0} \rfloor ) \gt \lfloor k/4K'M_{0}\rfloor$
and

This implies
$\{4M_{0} \vartheta (i) + 2M_{0} : i\in P_{\lfloor k/4K'M_{0} \rfloor }(\omega ) \}\subseteq \{1, \ldots , k - 2M_{0}\}$
. Combines with Display 35, this implies

Summing up the conditional probabilities, we have

Since
$\mathbb{P}(A_{k})$
decays exponentially, we conclude Inequality 7.
It remains to partition the probability space
$\Omega$
into pivotal equivalence classes that satisfy Definition 4.1, with
$\mathcal{P}(\omega )$
as the set of pivotal times. We say that
$\bar {\omega } \in \Omega$
is pivoted from
$\omega$
if they only differ in the value of
$\beta _{i}$
’s for
$i \in \liminf _{l} P_{l}(\omega )$
. Then being pivoted from each other is an equivalence relation. On an equivalence class
$\mathcal{E}$
, all random paths have the same set of pivotal times
$\mathcal{P}(\mathcal{E}) = \{j(1) \lt j(2) \lt \ldots \} \subseteq M_{0}\mathbb{Z}$
that avoids
$1, \ldots , \lfloor n/2 \rfloor$
and
$n$
. Moreover, the steps
$g_{i}$
’s are uniform across
$\mathcal{E}$
except for

which are i.i.d.s chosen from
$S$
according to
$\mu _{S}$
. Lastly, observe that

By Observation 5.9,
$(o, \mathbf{Y}_{j(1)}, \mathbf{Y}_{j(2)}, \ldots )$
is always
$D_{0}$
-semi-aligned. Proposition 4.2 is now proved.
6. Large deviation principles
In this section, we consider a more delicate pivoting that leads to the large deviation principle. Definition 6.1 and Proposition 6.2 rephrases Gouëzel’s result in [[Reference GouëzelGou22], Section 5A] in terms of strongly contracting isometries.
Definition 6.1. Let
$\mu$
and
$\nu$
be non-elementary probability measures on
$G$
and
$(\Omega , \mathbb{P})$
be a probability space for
$\mu$
. Let
$0\lt \epsilon \lt 1$
, let
$K_{0}, N \gt 0$
and let
$S \subseteq (\textrm {supp} \mu )^{M_{0}}$
be a long enough and large
$K_{0}$
-Schottky set for
$\mu$
.
A subset
$\mathcal{E}$
of
$\Omega$
is called an
$(n, N, \epsilon , \nu )$
-pivotal equivalence class for
$\mu$
, associated with the set of pivotal times

if the following hold:
-
(i) for each
$\omega \in \mathcal{E}$ and
$k \geqslant 1$ ,
\begin{align*} \begin{aligned} s_{k}(\omega ) &:= \big (g_{j(k) - M_{0} + 1}(\omega ), \,\,g_{j(k) - M_{0} + 2}(\omega ), \,\, \ldots , \,\, g_{j(k)}(\omega ) \big ),\\ s_{k}'(\omega ) &:= \big (g_{j'(k) - M_{0} + 1}(\omega ), \,\,g_{j'(k) - M_{0} + 2}(\omega ), \,\, \ldots , \,\, g_{j'(k)}(\omega ) \big ) \end{aligned} \end{align*}
-
(ii) for each
$\omega \in \mathcal{E}$ ,
$\big (o, \mathbf{Y}_{j(1)}, \mathbf{Y}_{j'(1)}, \ldots , \mathbf{Y}_{j(\#\mathcal{P}/2)}, \mathbf{Y}_{j'(\#\mathcal{P}/2)}, Z_{n} o\big )$ is
$D_{0}$ -semi-aligned;
-
(iii) for the RV defined as
\begin{align*} r_{k} := g_{j(k) + 1} g_{j(k) + 2} \cdots g_{j'(k) - M_{0}}, \end{align*}
$(s_{k}, s_{k}', r_{k})_{k \gt 0}$ on
$\mathcal{E}$ are i.i.d.s and
$r_{k}$ ’s are distributed almost according to
$\mu ^{\ast 2 M_{0} N} \ast \nu ^{\ast \frac {j'(k) - j(k)}{2M_{0}} - N-0.5}$ in the sense that the following holds for every
$g \in G$ :
\begin{align*} (1-\epsilon ) \big (\mu ^{\ast 2M_{0} N} \ast \nu ^{\ast \frac {j'(k) - j(k)}{2M_{0}} - N-0.5}\big )(g) \leqslant \mathbb{P}(r_{k} = g) \leqslant (1+\epsilon ) \big (\mu ^{\ast 2M_{0} N} \ast \nu ^{\ast \frac { j'(k) - j(k)}{2M_{0}} -N-0.5}\big )(g) \end{align*}
$g \in G$ .
Proposition 6.2.
Let
$M_{0}\gt 0$
,
$\mu$
be a non-elementary probability measure on
$G$
, let
$0\lt \epsilon \lt 1$
and let
$S \subseteq (\textrm {supp} \mu )^{M_{0}}$
be a long enough and large Schottky set for
$\mu$
with cardinality greater than
$100/\epsilon$
. Then there exists a non-elementary probability measure
$\nu$
on
$G$
such that the following holds.
For each sufficiently large integer
$N$
, there exists
$K\gt 0$
such that for each
$n$
we have a probability space
$(\Omega , \mathbb{P})$
for
$\mu$
and its measurable partition
$\mathscr{P}_{n, N, \epsilon , \nu } = \{\mathcal{E}_{\alpha }\}_{\alpha }$
into
$(n, N, \epsilon , \nu )$
-pivotal equivalence classes that satisfies

We will in fact prove a statement that is more explicit than Proposition 6.2:
Proposition 6.3.
Let
$0\lt \epsilon \lt 1$
, let
$K_{0}, M_{0} \gt 0$
and let
$S \subseteq G^{M_{0}}$
be a long enough and large
$K_{0}$
-Schottky set with
$\#S \geqslant 100/\epsilon$
. Let
$\mu$
be a probability measure on
$G$
such that
$m := \min \{ \mu ^{M_{0}}(s) : s \in S\}$
is positive. Let
$N \gt 40/m^{2} \epsilon$
and let
$\nu$
be the measure defined by

Then
$\nu$
is a non-elementary probability measure. Moreover, there exists
$K \gt 0$
depending only on
$S$
,
$m, N$
and
$\epsilon$
(but not on
$\mu$
) such that, for each
$n$
, we have a probability space
$(\Omega , \mathbb{P})$
for
$\mu$
and its measurable partition
$\mathscr{P}_{n, N, \epsilon , \nu } = \{\mathcal{E}_{\alpha }\}_{\alpha }$
into
$(n, N, \epsilon , \nu )$
-pivotal equivalence classes, associated with the set of pivotal times
$\mathcal{P}^{(n, N, \epsilon , \nu )}$
, that satisfies

Gouëzel proved Proposition 6.2 for random walks on a Gromov hyperbolic space in [[Reference GouëzelGou22], Section 5C]. We adapt his proof to our setting here.
Proof.
Let us denote the uniform measure on
$S$
by
$\mu _{S}$
. In this proof, when a probability measure
$\tau$
on
$G^{k}$
is given, we denote by
$\tau ^{\ast }$
the pushforward measure by convolution:

Let
$N_{0} = \#S$
be the cardinality of
$S$
. Note that
$10/N_{0} \leqslant \epsilon /10$
. Consider the decomposition

where
$\tau$
is a probability measure on
$G^{2M_{0}}$
with
$\tau ^{\ast } = \nu$
. Recall that
$S$
is a long enough and large
$K_{0}$
-Schottky set, so there exists
$a, b \in S$
such that
$\Pi (a)$
and
$\Pi (b)$
are independent strongly contracting isometries. Since
$\tau$
has the same support with
$\mu ^{2M_{0}}$
,
$\nu$
puts nonzero weights on
$a^{2}$
and
$b^{2}$
. Hence
$\nu$
is non-elementary.
Given the decomposition as in Equation 38, we consider Bernoulli RVs
$(\rho _{i})_{i\geqslant 0}$
with
$\mathbb{P}(\rho _{i} = 1) = 0.5m^{2}$
and
$\mathbb{P}(\rho _{i} = 0) = 1-0.5m^{2}$
,
$(\eta _{i})_{i\geqslant 0}$
with the law of
$\mu _{S}^{2}$
,
$(\tau _{i})_{i}$
with the law of
$\tau$
and
$(\xi _{i})_{i\geqslant 0}$
with the law of
$\mu ^{2M_{0}}$
, all independent. We define RVs
$\{t_{j}, t_{j}'\}_{j =1}^{\infty }$
. First,
$t_{1}$
is the smallest
$i\gt 0$
with
$\rho _{i} = 1$
, and
$t_{1}' := \min \{ i \gt t_{1} + N : \rho _{i} = 1\}$
. Inductively, we define

For convenience, we set
$t_{0}' := 0$
. We then define

Then
$(g_{i})_{i=1}^{\infty }$
is distributed according to the product measure
$\mu ^{\infty }$
[[Reference GouëzelGou22], Claim 5.11]. We let
$\mathscr{B}(k) :=\#\{j \geqslant 1 : t_{j}' \lt k\}$
. Now define

for
$i=1, \ldots , \mathscr{B}(\lfloor n/2M_{0} \rfloor )$
and define
$w_{\mathscr{B}(\lfloor n/2M_{0} \rfloor )} = g_{2M_{0}t_{\mathscr{B}(\lfloor n/2M_{0} \rfloor )}' + 1} \cdots g_{n}$
. Using these data, we define the set of pivotal times

as in Subsection 5.1.
We first determine the values of
$\rho _{j}$
’s. Observe that
$\mathscr{B}(\lfloor n/2M_{0} \rfloor )$
and
$\{t_{j}, t_{j}'\}_{j}$
depend solely on
$\{\rho _{j}\}_{j}$
and counts the renewal times in
$[0, n/2M_{0}]$
formed with a geometric distribution after a delay
$N$
. More explicitly, if we ‘omit’
$\rho _{t_{k} + i}$
’s for
$k \gt 0$
and
$i = 1, \ldots , N$
and define

then
$\{\rho _{i}'\}_{i}$
are i.i.d. Bernoulli RVs and
$t_{k}' = kN + \min \{ j : \sum _{i=1}^{j} \rho _{i}' = 2k\}$
. Hence, we have

which decays exponentially because
$\mathbb{E}[\rho _{i}'] = 0.5 m^{2}\gt 20/ \epsilon N$
. Hence, there exists
$K_{1}\gt 0$
that depends on
$m, \epsilon$
and
$N$
such that:

Let us fix the choices of
$(\rho _{i})_{i\geqslant 0}$
. This determine
$(t_{i}, t_{i}')_{i\gt 0}$
and
$\mathscr{B}(\lfloor n/2M_{0} \rfloor )$
. We then fix the data
$(\tau _{i}, \xi _{i})_{i\gt 0}$
and
$\{\eta _{i}: i \gt t_{\mathscr{B}(\lfloor n/2M_{0} \rfloor )}'\}$
. These in turn determine
$(w_{i})_{i=0}^{\mathscr{B}(\lfloor n/2M_{0} \rfloor )}$
and
$(v_{i})_{i=1}^{\mathscr{B}(\lfloor n/2M_{0} \rfloor )}$
. Furthermore,

are all independent and identically distributed according to
$\mu _{S}^{2}$
. Hence, the situation is reduced to the combinatorial model in Section 4. Corollary 5.8 asserts the following for some
$K_{2} \gt 0$
:

Combining Inequality 39 and 40, we can conclude that
$\mathbb{P}\left ( \# P_{\mathscr{B}(\lfloor n/2M_{0} \rfloor )} \leqslant (1-\epsilon ) \frac {n}{2M_{0} N} \right )$
decays exponentially.
Now, given
$\omega \in \Omega$
with
$P_{\mathscr{B}(\lfloor n/2M_{0} \rfloor )(\omega ) } (\omega ) = \{i(1) \lt i(2) \lt \ldots \}$
, we define:

We just established the estimate is Display 37 for this
$\mathcal{P}^{(n, N, \epsilon , \nu )}$
. Furthermore, note that

are Schottky axes, and that
$w_{\mathscr{B}(\lfloor n/2M_{0} \rfloor )+1, 2}^{-} = Z_{n}$
. Proposition 5.1 tells us that
$(o, \mathbf{Y}_{j(1)}, \mathbf{Y}_{j'(1)},$
$ \mathbf{Y}_{j(2)}, \mathbf{Y}_{j'(2)}, \ldots , Z_{n} o)$
is always
$D_{0}$
-semi-aligned. This settles Item (i) and (ii) in Definition 6.1.
It remains to realize the partition as in Definition 6.1 and check Item (iii) in Definition 6.1. We declare the equivalence by pivoting. More precisely, given
$\omega \in \Omega$
with
$P_{\mathscr{B}(\lfloor n/2M_{0} \rfloor }(\omega )= \{i(1) \lt i(2) \lt \ldots \}$
, we declare that another element
$\omega ' \in \Omega$
is equivalent to
$\omega$
if it has the same values of
$(\rho _{i})_{i \geqslant 0}$
(hence the same values of
$(t_{i}, t_{i}')_{i\gt 0}$
) with
$\omega$
, and if it has the same values of
$(\eta _{i}, \tau _{i}, \xi _{i})_{i \geqslant 0}$
with
$\omega$
, possibly except for

Further, we require that
$\big (\alpha _{i(k)}(\omega '), \delta _{i(k)}(\omega ') \big ) = \big (\alpha _{i(k)}(\omega ), \delta _{i(k)}(\omega ) \big )$
for each
$k$
and

Note that under this requirement,
$\omega '$
has the same values of
$(w_{i})_{i=0}^{\mathscr{B}(\lfloor n/2M_{0} \rfloor )}$
and
$\{v_{i}: i \neq i(1), \ldots , i(\#P_{\mathscr{B}(\lfloor n/2M_{0} \rfloor )})\}$
with
$\omega$
. By Lemma 5.5, we have
$P_{\mathscr{B}(\lfloor n/2M_{0} \rfloor )}(\omega ) = P_{\mathscr{B}(\lfloor n/2M_{0} \rfloor )}(\omega ')$
, and the above relation becomes an equivalence relation.
Recall that conditioned on the data
$(\rho _{i})_{i \geqslant 0}$
,
$(\beta _{j}, \gamma _{j}, v_{j})$
is distributed according to
$\mu _{S}^{2} \times \big ( \mu ^{\ast 2M_{0}} \ast (\tau ^{\ast })^{\ast ( t_{j}' - t_{j}- N - 0.5)} \big ) = \mu _{S}^{2} \times \big ( \mu ^{\ast 2M_{0}} \ast \nu ^{\ast ( t_{j}' - t_{j}- N - 0.5)} \big )$
. Now let
$\mathcal{E}$
be a pivotal equivalence class that has pivotal times
$P_{\mathscr{B}(\lfloor n/2M_{0} \rfloor )} = \{i(1) \lt \ldots \lt i(m)\}$
. Then
$(\beta _{i(l)}, \gamma _{i(l)}, v_{i(l)})$
’s are independent and distributed according to the restriction of
$\mu _{S}^{2} \times \left (\mu ^{\ast 2M_{0} N} \ast \nu ^{\ast ( t_{j}' - t_{j}- N - 1)} \right )$
onto the set of “legitimate choices”
$\tilde {\mathrm{S}}$
. To describe this, let us define a (not necessarily probability) measure

then
$(\beta _{i(l)}, \gamma _{i(l)}, v_{i(l)})$
is distributed according to the normalized version
$\mu ^{(0)}$
of
$\mu ^{(1)}$
, namely,
$\mu ^{(0)}(A) := \frac {1}{\mu ^{(1)}(S^{2} \times G)} \mu ^{(1)}(A)$
for each
$A \subseteq S^{2} \times G$
.
For each
$r \in G$
, among
$N_{0}^{2}$
choices of
$s'$
and
$s''$
in
$S$
at least
$N_{0}^{2} - 2N_{0}$
choices qualify the criterion and make
$(s', s'', r) \in \tilde {\mathrm{S}}$
by the Schottky property. (See the discussion in Display 20 and 21.) This implies the bound for each
$r \in G$
:

Summing this up for all
$r \in G$
, we obtain
$1-2/N_{0} \leqslant \mu ^{(1)}(S^{2} \times G) \leqslant 1$
. Combining these two estimates, we conclude the following for every
$g \in G$
:

This settles Item (iii) in Definition 6.1 as desired.
We now establish the large deviation principle for random walks.
Theorem 6.4.
Let
$(X, G, o)$
be as in Convention
2.11
and let
$(Z_{n})_{n}$
be the random walk generated by a non-elementary probability measure
$\mu$
on
$G$
. Let
$\lambda (\mu ) = \lim _{n} \frac {1}{n} \mathbb{E}[d(o, Z_{n} o)]$
be the drift of
$\mu$
. Then for each
$0 \lt L \lt \lambda (\mu )$
, the probability
$\mathbb{P}(d(o, Z_{n} o) \leqslant Ln)$
decays exponentially as
$n$
goes to infinity.
Recall that
$\lambda (\mu ) = +\infty$
when
$\mu$
has infinite first moment, by Corollary 4.12.
Proof.
Due to the subadditivity, we have
$\mathbb{E}_{\mu ^{\ast N}} [d(o, go)] \geqslant \lambda (\mu ) N$
for each
$N\gt 0$
. Since
$L$
is smaller than
$\lambda (\mu )$
, there exists
$\epsilon \gt 0$
such that

For this
$\epsilon \gt 0$
, let
$S$
be a long enough Schottky set for
$\mu$
with cardinality greater than
$100/\epsilon$
. By Proposition 6.2, there exists a non-elementary probability measure
$\nu$
, and for each sufficiently large
$N$
, a partition
$\mathscr{P}_{n, N, \epsilon }$
into
$(n, N, \epsilon , \nu )$
-pivotal equivalence classes for each
$n$
such that

decays exponentially in
$n$
. Let
$C \gt 0$
be a constant for
$\nu$
provided by Corollary 4.11: we have

for each
$g \in G$
and each
$m \gt 0$
. We now fix an
$N$
such that
.
Let
$\mathcal{E}$
be an equivalence class such that
. Then for each
$\omega \in \mathcal{E}$
,
$(o, \mathbf{Y}_{j(1)}, \ldots , \mathbf{Y}_{j'(\#\mathcal{P}/2)}, Z_{n} o)$
is
$D_{0}$
-semi-aligned. The second inequality in Item (ii) of Lemma 3.18 tells us that

Since
$r_{i}$
’s are non-negative i.i.d. with

we can apply the classical theory of large deviation. As a result, there exists
$K'\gt 0$
such that

Summing up this conditional probability, we obtain the desired exponential bound.
We now connect Theorem 6.4 with the large deviation principle. In [[Reference Bounlanger, Mathieu, Sert and SistoBMSS22], Proposition 2.3, Theorem 2.8], Boulanger, Mathieu, Sert and Sisto presented a general theory of large deviation principles on metric spaces with Schottky sets. Combining their result with Theorem A, we establish the large deviation principle for random walks on the mapping class group.
Corollary 6.5 (Large deviation principle)
Let
$(X, G, o)$
be as in Convention
2.11
and let
$(Z_{n})_{n\geqslant 0}$
be the random walk generated by a non-elementary probability measure
$\mu$
on
$G$
. Then there exists a proper convex function
$I : \mathbb{R} \rightarrow [0, +\infty ]$
, vanishing only at the drift
$\lambda (\mu )$
, such that

holds for every measurable set
$E \subseteq \mathbb{R}$
.
We note the work of Corso [Reference CorsoCor21], who proved that the rate function exists and is proper for random walks involving strongly contracting isometries. Our Corollary 6.5 strengthens Corso’s result by showing that
$I(x) \neq 0$
for
$x \in [0, \lambda (\mu ))$
, which is a consequence of Theorem A.
Part II: random walks with weakly contracting isometries
In this part, we deal with groups acting on a space
$X$
and another space
$\tilde {X}$
equivariantly, where the action on
$X$
involves strong contraction and the action on
$\tilde {X}$
involves weak contraction: see Convention 7.2. After studying alignment of weakly contracting directions in Section 8, we establish limit theorems for mapping class groups in Section 9.
7. Mapping class groups and HHGs
Let
$\Sigma$
be a finite-type hyperbolic surface, let
$(\tilde {X}, \tilde {d})$
be the Cayley graph of the mapping class group
$G = \textrm {Mod}(\Sigma )$
of
$\Sigma$
, and let
$(X, d)$
be the curve complex of
$\Sigma$
or the Teichmüller space of
$\Sigma$
. The action of
$G$
on
$(X, d)$
satisfies Convention 2.11:
$G$
contains independent pseudo-Anosov mapping classes that have strongly contracting orbits on
$X$
([[Reference MinskyMin96], Contraction Theorem], [[Reference Masur and MinskyMM99], Proposition 4.6]).
Let
$\textrm {Pr} : \tilde {X} \rightarrow X$
be the orbit map:
$\textrm {Pr}(g) = go$
, where
$o \in X$
is the basepoint. Since
$G$
is finite generated and acts on
$X$
by isometries, the map
$\textrm {Pr}$
is coarsely Lipschitz and is
$G$
-equivariant. We will denote by
$\tilde {A}$
the object “in the upper space” corresponding to an object
$A$
“in the lower space”. For example, we fix basepoints
$\tilde {o} = id \in \tilde {X}$
and
$o \in X$
that satisfy
$\textrm {Pr}(\tilde {o}) = o$
.
For each subset
$\tilde {A} \subseteq \tilde {X}$
, we define the projection
$\tilde {\pi }_{\tilde {A}}$
from
$\tilde {X}$
onto
$\tilde {A}$
by referring to the closest point projection at the lower space
$X$
. Namely, for
$\tilde {x} \in \tilde {X}$
and its projection
$x := \textrm {Pr}(\tilde {x})$
, we define
$\tilde {\pi }_{\tilde {A}}(\tilde {x}) := \textrm {Pr}^{-1} \circ \pi _{A} \circ \textrm {Pr}$
by
$\tilde {a} \in \tilde {\pi }_{\tilde {A}}(\tilde {x}) \,\,\Leftrightarrow \,\, a \in \pi _{A}(x).$
Lemma 7.1.
For each
$C\gt 1$
there exists
$D\gt 1$
such that if a
$C$
-quasigeodesic
$\tilde {\gamma } : I \rightarrow \tilde {X}$
on
$\tilde {X}$
has projection
$\gamma$
onto
$X$
that is a
$C$
-contracting axis, then
$\tilde {\gamma }$
is
$D$
-weakly contracting with respect to the map
$\tilde {\pi }_{\tilde {\gamma }} := \textrm {Pr}^{-1} \circ \pi _{\gamma } \circ \textrm {Pr}$
.
Proof.
Let us first consider the case that
$\tilde {X}$
is the Cayley graph of
$\textrm {Mod}(\Sigma )$
and
$(X, d)$
is the curve complex
$\mathcal{C}(\Sigma )$
of
$\Sigma$
. Recall that there are coarsely Lipschitz projections
$\textrm {Pr}_{U} :{} \tilde {X} \rightarrow \{\textrm {uniformly bounded subsets of} \mathcal{C}U\}$
from
$\tilde {X}$
to the curve complex
$\mathcal{C}U$
of subsurfaces
$U \subseteq \Sigma$
, and
$\rho _{U}^{V} : \mathcal{C}U \rightarrow \{\textrm {uniformly bounded subsets of} \mathcal{C}V\}$
for every pair of nested subsurfaces
$V \subseteq U \subseteq \Sigma$
. Further,
$\textrm {Pr}_{U}$
and
$\rho _{\Sigma }^{U} \circ \textrm {Pr}$
are uniformly coarsely equivalent.
Since
$\tilde {\gamma }$
and
$\gamma = \Pr \circ \tilde {\gamma }$
are
$C$
-quasigeodesics,
$\{ \textrm {Pr}_{U}(\tilde {\gamma }) : U \subsetneq \Sigma \}$
have uniformly bounded diameter (depending on
$C$
). This is due to the bounded geodesic image property. Namely, given a proper subsurface
$U \subsetneq \Sigma$
, there exists a uniformly bounded neighborhood
$N$
of
$\partial U \subseteq \mathcal{C}(\Sigma )$
such that
$\gamma \setminus N$
is uniformly close to a geodesic on
$\mathcal{C}(\Sigma )$
that is disjoint from
$\partial U$
. By [[Reference Masur and MinskyMM00], Theorem 3.1],
$\rho _{\Sigma }^{U} (\gamma \setminus N)$
has bounded diameter. Since
$N$
is bounded,
$\rho _{\Sigma }^{U}(\gamma \cap N)$
is also bounded.
Given the uniform boundedness of
$\textrm {Pr}_{U}(\tilde {\gamma })$
’s, i.e., the coboundedness of
$\tilde {\gamma }$
, the weakly contracting property of
$\tilde {\gamma }$
follows from [[Reference Duchin and RafiDR09], Theorem 4.2] (cf. [[Reference BehrstockBeh06], Lemma 5.6]). More explicitly, [[Reference Duchin and RafiDR09], Theorem 4.2] guarantees a constant
$E = E(C)$
such that, for each
$\tilde {x} \in \tilde {X}$
, we have

Since the
$d$
-diameter along
$\gamma$
and
$\tilde {d}$
-diameter along
$\tilde {\gamma }$
are coarsely equivalent (as
$\gamma$
is a quasigeodesic), we conclude that
$\tilde {\gamma }$
is weakly contracting with respect to
$\tilde {\pi }_{\tilde {\gamma }}$
.
When
$(X, d)$
is the Teichmüller space of
$\Sigma$
, the strongly contracting property of
$\gamma$
implies that the Teichmüller geodesics
$[\gamma (t), \gamma (s)]$
for
$t\lt s$
are contained in a uniform neighborhood of
$\gamma$
(Corollary 3.4) and are hence uniformly thick. This in turn implies that
$\Pr _{U}(\tilde {\gamma }(t), \tilde {\gamma }(s))$
’s for
$t\lt s$
and proper subsurfaces
$U \subsetneq \Sigma$
are uniformly bounded ([[Reference RafiRaf05], Theorem 1.1], [[Reference RafiRaf14], Theorem 5.5], [[Reference Rafi and SchleimerRS09], Theorem 4.1], [[Reference Durham and TaylorDT15], Lemma 5.1]). Then we similarly deduce the weakly contracting property of
$\tilde {\gamma }$
by [[Reference Duchin and RafiDR09], Theorem 4.2].
In general, Lemma 7.1 can be generalized to the setting where
$G$
is a hierarchically hyperbolic group (HHG),
$(\tilde {X}, \tilde {d})$
is its Cayley graph and
$(X, d)$
is the top curve graph for
$G$
. This follows from [[Reference Abbott, Behrstock and DurhamABD21], Corollary 6.2] ((3)
$\Rightarrow$
(2)) and [[Reference Abbott, Behrstock and DurhamABD21], Theorem 4.4]. Note that even though Corollary 6.2 and Theorem 4.4 assumes the unbounded products of the HHG structure for
$G$
, which is not granted in general, the directions we need do not require such an assumption.
In particular, the pseudo-Anosov axes on
$\tilde {X}$
are weakly contracting. Hence, our setting is:
Convention 7.2.
We fix
$B\gt 0$
and assume that:
-
(i)
$(\tilde {X}, \tilde {d})$ ,
$(X, d)$ are geodesic metric spaces;
-
(ii)
$\textrm {Pr} : \tilde {X} \rightarrow X$ is a coarsely Lipschitz map, i.e., for all
$\tilde {x}, \tilde {y} \in \tilde {X}$
\begin{align*} d(\textrm {Pr}(\tilde {x}), \textrm {Pr}(\tilde {y})) \leqslant B\tilde {d}(\tilde {x}, \tilde {y})+ B; \end{align*}
-
(iii)
$G$ is a countable group of isometries acting on
$\tilde {X}$ and
$X$ equivariantly;
-
(iv)
$\tilde {o} \in \tilde {X}$ and
$o \in X$ are basepoints that satisfy
$\textrm {Pr}(\tilde {o}) = o$ ;
-
(v) For each
$C \gt 1$ there exists
$D\gt 1$ such that a path
$\tilde {\gamma }$ on
$\tilde {X}$ is
$D$ -weakly contracting with respect to
$\tilde {\pi }_{\tilde {\gamma }}$ whenever its projection
$\textrm {Pr}(\tilde {\gamma })$ is a
$C$ -contracting axis;
-
(vi)
$G$ contains two independent strongly contracting isometries of
$X$ .
For each object
$\tilde {A} \subseteq \tilde {X}$
, we denote by
$A$
its projection
$\textrm {Pr}(\tilde {A}) \subseteq X$
.
When Item (iii) is replaced with the coarse equivariance condition, this setting also covers HHGs acting on the top curve graph. For simplicity, we denote the word norm of
$g \in G$
by
$|g|$
. In the general case, one can replace
$|g|$
with
$\tilde {d}(\tilde {o}, g \tilde {o})$
.
8. Alignment II: weakly contracting axes
Throughout, we adopt Convention 7.2. We define the alignment among paths
$\tilde {\kappa }_{1}, \ldots , \tilde {\kappa }_{n}$
on
$\tilde {X}$
based on Definition 3.6 with respect to the projections
$\tilde {\pi }_{\tilde {\kappa }_{i}} := \textrm {Pr}^{-1} \circ \pi _{\kappa _{i}} \circ \textrm {Pr}$
.
Lemma 8.1.
For each
$K \gt 1$
there exists
$K' \gt K$
such that the following hold. Let
$\tilde {x}, \tilde {y} \in \tilde {X}$
and let
$\kappa$
be a path on
$\tilde {X}$
whose projection
$\kappa$
on
$X$
is a
$K$
-contracting axis. Then
$\tilde {\kappa }$
is a
$K'$
-quasigeodesic that is
$K'$
-weakly contracting with respect to
$\tilde {\pi }_{\tilde {\kappa }}$
. Moreover, for each
$C\gt 1$
we have the implication

Proof.
Let
$\tilde {\kappa } : I \rightarrow \tilde {X}$
and
$\kappa := \textrm {Pr} \circ \tilde {\kappa } : I \rightarrow X$
. The weakly contracting property of
$\tilde {\kappa }$
is given by Lemma 7.1. If we denote by
$F$
the coarse inverse of
$\kappa$
,
$\textrm {Pr}$
and
$\tilde {\kappa } \circ F$
are maps between
$\kappa$
and
$\tilde {\kappa }$
, and are coarse inverses of each other. This implies the coarse comparison

for all points
$\tilde {p}, \tilde {q}$
on
$\tilde {\kappa }$
, for some
$K'' = K''(K)$
. This implies the remaining items.
We now prove the main proposition of this section.
Proposition 8.2.
For each
$K, D \gt 1$
, there exist
$E, L'\gt K, D$
such that the following holds.
Let
$L \geqslant L'$
, let
$\tilde {x}, \tilde {y} \in \tilde {X}$
and let
$\tilde {\kappa }_{1}, \ldots , \tilde {\kappa }_{n}$
be paths on
$\tilde {X}$
whose domains are longer than
$L$
and such that their projections are
$K$
-contracting axes. Suppose that
$(x, \kappa _{1}, \ldots , \kappa _{n}, y)$
is
$D$
-aligned. Then there exist points
$\tilde {p}_{1}, \ldots , \tilde {p}_{n}$
on
$[\tilde {x}, \tilde {y}]$
, in order from left to right, such that

Here, we plug in
$\tilde {\kappa }_{0} = \tilde {x}$
and
$\tilde {\kappa }_{n+1} = \tilde {y}$
.
Proof.
Let
$B$
be the coarse Lipschitzness constant for
$\textrm {Pr}$
, and define the constants:
-
– let
$K_{2} = K'(K)$ be as in Lemma 8.1, which is larger than
$K\gt 1$ ;
-
– let
$K_{4} = K'(K)$ be as in Lemma 2.10, which is larger than
$K\gt 1$ ;
-
– let
$E_{1}= E(K, D)$ be as in Lemma 3.9, which is larger than
$K\gt 1$ ;
-
– let
$E_{2} = E(K, D)$ ,
$L_{0} = L(K, D)$ be as in Proposition 3.10.
Now we define constants

Then the following hold for all
$L \geqslant L'$
:

Note that
Fact 8.3.
Let
$C\gt 0$
, let
$x \in X$
and let
$\kappa$
be a
$K$
-contracting axis on
$X$
with
$L$
-long domain. If
$(x, \kappa )$
is
$C$
-aligned, then
$(\kappa , x)$
is not
$(\frac {L}{K} - C-K)$
-aligned.
Let
$L \geqslant L'$
, let
$\tilde {x}, \tilde {y} \in \tilde {X}$
and let
$\tilde {\kappa }_{1}, \ldots , \tilde {\kappa }_{n}$
be paths on
$\tilde {X}$
whose domains are longer than
$L$
and whose projections
$\kappa _{i}$
’s onto
$X$
are
$K$
-contracting axes. Recall our convention that, whenever we define
$\tilde {A} \subseteq \tilde {X}$
, we use the notation
$A := \textrm {Pr}(\tilde {A})$
.
Step 1. We prove the following for
$n \geqslant 2$
:
if
$(x, \kappa _{1})$
is
$E$
-aligned and
$(\kappa _{1}, \ldots , \kappa _{n}, y)$
is
$D$
-aligned, then
$[\tilde {x}, \tilde {y}]$
has points
$\tilde {z}_{2}, \ldots , \tilde {z}_{n}$
, in order from left to right, such that
$(\kappa _{i-1}, z_{i})$
and
$(z_{i}, \kappa _{i})$
are
$E$
-aligned (
$z_{i} = \textrm {Pr} \tilde {z}_{i}$
).
We induct on the number
$n$
of the contracting axes. First, Proposition 3.10 implies that
$(x, \kappa _{i}, y)$
is
$E_{2}$
-aligned for each
$i$
. In view of Fact 8.3,
$(\kappa _{1}, x)$
is not
$(E_{1} + E_{2} + B+4K)$
-aligned but
$(\kappa _{1}, y)$
is
$E_{2}$
-aligned. Now note that:
-
–
$\pi _{\kappa _{1}}$ is
$(1, 4K)$ -coarsely Lipschitz (Lemma 2.2),
-
–
$\textrm {Pr}$ is
$B$ -coarsely Lipschitz and hence
$\pi _{\kappa _{1}}\circ \textrm {Pr}$ is
$(B, B+4K)$ -coarsely Lipschitz, and
-
– the geodesic
$[\tilde {x}, \tilde {y}]$ is connected.
Pick the rightmost point
$\tilde {z}_{2} \in [\tilde {x}, \tilde {y}]$
such that
$(\kappa _{1}, z_{2})$
is not
$(E_{1} + E_{2})$
-aligned. Then
$(\kappa _{1}, z_{2})$
is
$(E_{1} + E_{2} + B+4K)$
-aligned, and hence
$E$
-aligned, since
$\pi _{\kappa _{1}} \circ \textrm {Pr}$
is
$B$
-coarsely Lipschitz. Since
$(\kappa _{1}, \kappa _{2})$
is
$D$
-aligned and
$(\kappa _{1}, z_{2})$
is not
$E_{1}$
-aligned, Lemma 3.9 implies that
$(z_{2}, \kappa _{1})$
is
$E_{1}$
-aligned.
When
$n = 2$
, the proof ends here. Otherwise, note that
$(z_{2}, \kappa _{2})$
is
$E$
-aligned and
$(\kappa _{2}, \ldots , \kappa _{n}, y)$
is
$D$
-aligned. By the induction hypothesis, there exist
$\tilde {z}_{3}, \ldots , \tilde {z}_{n}$
on
$[\tilde {z}_{2}, \tilde {y}]$
, in order from left to right, such that
$(\kappa _{i-1}, z_{i})$
and
$(z_{i}, \kappa _{i})$
are
$E$
-aligned for each
$i\geqslant 3$
. The claim now follows.
Step 2: Construction of
$\tilde {p}_{j}$
’s. We now assume that
$(x, \kappa _{1}, \ldots , \kappa _{n}, y)$
is
$D$
-aligned. By Step 1, we obtain points
$\tilde {z}_{2}, \ldots , \tilde {z}_{n}$
on
$[\tilde {x}, \tilde {y}]$
, in order from left to right. We let
$\tilde {z}_{1} := \tilde {x}$
and
$\tilde {z}_{n+1} := \tilde {y}$
.
Pick
$j \in \{1, \ldots , n\}$
. Then
$(z_{j}, \kappa _{j})$
is
$E$
-aligned, and hence
$L/2K$
-aligned. Meanwhile,
$(\kappa _{j}, z_{j+1})$
is
$E$
-aligned, so
$(z_{j+1}, \kappa _{j})$
is not
$(L/K - E-K)$
-aligned, and hence not
$L/2K$
-aligned. Now let
$\tilde {p}_{j}$
to be the rightmost point on
$[\tilde {z}_{j}, \tilde {z}_{j+1}]$
such that
$(p_{j}, \kappa _{j})$
is not
$L/2K$
-aligned. Then by the
$(B, B+4K)$
-Lipschitzness of
$\pi _{\kappa _{j}} \circ \textrm {Pr}$
, we have that

In particular,
$(p_{j}, \kappa _{j})$
is not
$(L/2K - (B+4K))$
-aligned. Moreover,
$(\kappa _{j} ,p_{j})$
is not
$(L/2K - (B+5K))$
-aligned by Fact 8.3. Denoting the beginning point of
$\tilde {\kappa }_{j}$
by
$q_{j}$
, Lemma 8.1 implies

For a similar reason,
$\operatorname {diam}_{\tilde {X}}\Big (\tilde {\pi }_{\tilde {\kappa }_{j}}(\tilde {z}_{j+1}) \cup \tilde {\pi }_{\tilde {\kappa }_{j}}(\tilde {p}_{j}) \Big )$
is at least
. Now Lemma 2.10 implies

Step 3: Estimating
$\tilde {d}(\tilde {p}_{j}, \tilde {\kappa }_{j})$
.
Given Inequality 42, it now suffices to prove:

for
$i = 2, \ldots , n$
. To prove this, we collect indices
$i$
that violates Inequality 43. Let
$I = \{m, m+1, \ldots , m'\}$
be a maximal 1-connected set of such indices. We aim to show that
$I$
is empty.
Suppose to the contrary that
$I$
is nonempty. Note first that
$\tilde {z}_{1}$
and
$\tilde {z}_{n+1}$
satisfy Inequality 43, i.e.,
$1, n+1 \notin I$
; hence
$m \geqslant 2$
and
$m' \leqslant n$
. We now compute
$\tilde {d}(\tilde {z}_{m-1}, \tilde {z}_{m'+1})$
in two different ways. First, using Inequality 42 we deduce

Recall that
$\tilde {d}(\tilde {\kappa }_{j-1}, \tilde {z}_{j}) + \tilde {d}(\tilde {z}_{j}, \tilde {\kappa }_{j}) \geqslant \sum _{k} 2e^{-|k - j| L/E} \operatorname {diam}_{\tilde {X}} (\tilde {\kappa }_{k-1} \cup \tilde {\kappa }_{k}) + E$
holds for
$m \leqslant j \leqslant m'$
. Moreover,
$E \cdot (\# I) \geqslant 2K_{4} \cdot (\# I + 1) + 0.5E$
because
$\# I \geqslant 1$
and
$E \geqslant 8K_{4}$
. Hence, we obtain

If we rearrange the double summation with respect to
$k$
, the RHS is at least

Next, we will obtain an upper bound of
$\tilde {d}(\tilde {z}_{m-1} , \tilde {z}_{m'+1})$
:

We know that
$6e^{-L/E} \leqslant 1$
because
$L \gt 8E$
. Having this in mind, we now make use of the fact that
$m-1$
and
$m'+1$
is not contained in
$I$
:
$\tilde {d}(\tilde {z}_{m-1}, \tilde {z}_{m' + 1} )$
is bounded from above by

which is a contradiction. Hence,
$I = \emptyset$
and Inequality 43 is established.
9. Limit laws for mapping class groups
We continue to employ the notion of Schottky sets defined in Definition 3.15. Once a Schottky set
$S$
and its element
$s$
is understood, the translates of
$\Gamma ^{\pm }(s)$
are now called Schottky axes on
$X$
, whereas the translates of
$\tilde {\Gamma }^{\pm }(s)$
are called Schottky axes on
$\tilde {X}$
.
Definition 9.1. Given a constant
$K_{0} \gt 0$
, we define:
-
–
$K_{1} = K'(K_{0})$ be as in Lemma 8.1,
-
–
$D_{0 } = D(K_{0}, K_{0})$ be as in Lemma 3.8,
-
–
$E_{0} = E(K_{0}, D_{0})$ ,
$L_{0} = L(K_{0}, D_{0})$ be as in Proposition 3.12.
-
–
$E_{1} = K'(K_{0}, E_{0})$ ,
$L_{1} = L'(K_{0}, E_{0})$ be as in Proposition 8.2.
Let
$0 \lt \epsilon \lt 1$
. If a
$K_{0}$
-Schottky set
$S \subseteq G^{M_{0}}$
consists of sequences of length

then we call
$S$
an
$\epsilon$
-constricting
$K_{0}$
-Schottky set.
Thanks to Proposition 3.19, for every non-elementary probability measure
$\mu$
on
$G$
and
$N, \epsilon \gt 0$
, there exists an
$\epsilon$
-long enough Schottky set for
$\mu$
with cardinality
$N$
. We are ready to state:
Proposition 9.2.
Let
$\mu$
be a non-elementary probability measure on the mapping class group
$G$
and
$((\check {Z}_{n})_{n}, (Z_{n})_{n})$
be the (bi-directional) random walk generated by
$\mu$
, with step sequences
$((\check {g}_{n})_{n}, (g_{n})_{n})$
. Then there exists
$K'\gt 0$
such that

holds for all
$k$
, where

Proof.
Let
$S$
and
$\check {S}$
be a long enough, large and
$(2/e)$
-constricting
$K_{0}$
-Schottky sets for
$\mu$
and
$\check {\mu }$
, respectively, for some
$K_{0}\gt 0$
. Proposition 4.2 determines a constant
$K\gt 0$
(not depending on
$k$
but only on
$\mu$
), a probability space
$(\Omega , \mathbb{P})$
for
$\mu$
and a partition of
$\Omega$
into pivotal equivalence classes that is independent of the backward steps
$(\check {g}_{n})_{n\gt 0}$
and such that

and also another partition into (backward) pivotal equivalence classes that is independent of the forward steps
$(g_{n})_{n\gt 0}$
and such that

We enumerate
$\mathcal{P}(\omega )$
by
$\{j(1) \lt j(2) \lt \ldots \}$
and
$\mathcal{P}(\check {\omega })$
by
$\{\check {j}(1) \lt \check {j}(2) \lt \ldots \}$
. Let us now define the event
$B_{k}$
in
$\Omega$
;
$(\check {\omega }, \omega ) \in B_{k}$
if:
-
(i)
$\#\mathcal{P} (\omega ) \cap \{1, \ldots , n\}\geqslant n/K$ for all
$n \geqslant k/3$ ;
-
(ii)
$\#\mathcal{P}(\check {\omega }) \cap \{1, \ldots , n\} \geqslant n/K$ for all
$n \geqslant k/3$ ;
-
(iii) for each
$n \geqslant k$ and
$m \geqslant k$ , the following are
$D_{0}$ -semi-aligned:
\begin{align*} \begin{aligned} &\big (o, \,\mathbf{Y}_{j(1)}(\omega ), \, \mathbf{Y}_{j(2)}(\omega ), \, \ldots , \, \mathbf{Y}_{j(\lfloor 2n/3K\rfloor )}(\omega ), \,Z_{n} o\big ),\\ &\big (o, \,\mathbf{Y}_{\check {j}(1)}(\check {\omega }), \, \mathbf{Y}_{\check {j}(2)}(\check {\omega }), \, \ldots , \, \mathbf{Y}_{\check {j}(\lfloor 2m/3K\rfloor )}(\check {\omega }), \,\check {Z}_{m} o\big ); \end{aligned} \end{align*}
-
(iv)
$\big (\bar {\mathbf{Y}}_{\check {j}(i)}(\check {\omega }),\, \mathbf{Y}_{j(i)}(\omega ) \big )$ is
$D_{0}$ -aligned for some
$i \leqslant k/3K$ .
In the proof of Lemma 4.10 we proved that
$\mathbb{P}(B_{k})$
decays exponentially in
$k$
. It remains to prove that
$d(id, [\check {Z}_{m}, Z_{n}]) \leqslant K'D_{k}$
for any
$n, m\gt 0$
and
$(\check {\omega }, \omega ) \in B_{k}$
, where we set
$K' \geqslant 8+ 1.5K + E_{1}$
. From now on, we fix
$k$
. When
$n \leqslant k$
, we automatically have

Similarly, the desired inequality holds when
$m \leqslant k$
. Now assume
$n, m \geqslant k$
. The sequence

is
$D_{0}$
-semi-aligned, and hence
$E_{0}$
-aligned by Proposition 3.10. Here, recall that the involved Schottky set is
$(2/e)$
-long enough and that
. Hence,
$M_{0}/E_{1} \geqslant 2$
. By Proposition 8.2, there exists
$\tilde {p} \in [\check {Z}_{m}, Z_{n} ]$
whose distance to
$\tilde {\mathbf{Y}}_{j(\lfloor k/3K \rfloor )}(\omega )$
is at most

Here, note that

Note also that
$l - \lfloor k/3K \rfloor \geqslant \frac {1}{2} l$
for
$l \gt \lfloor 2 k/3K \rfloor$
. Using these, we deduce

Since we have
$j(\lceil i/K\rceil ) \leqslant i$
for each
$i \geqslant k/3$
, this is bounded by

Moreover, since
$\operatorname {diam}(id \cup \tilde {\mathbf{Y}}_{j(\lfloor k/ 3K \rfloor )})$
is bounded by
$\sum _{i=1}^{j(\lfloor k/3K \rfloor )} |g_{i}| \leqslant \sum _{i=1}^{k} |g_{i}|$
, we conclude

Proposition 9.3.
Let
$p\gt 0$
and let
$((\check {Z}_{n})_{n}, (Z_{n})_{n})$
be the (bi-directional) random walk generated by a non-elementary probability measure
$\mu$
on
$G$
with finite
$p$
-th moment. Then there exists
$K\gt 0$
such that

In particular, for almost every sample path
$(\check {\omega }, \omega )$
, every geodesic in
$\{[\check {Z}_{m}, Z_{n}] : m, n \gt 0\}$
intersects the same bounded metric ball centered at
$id$
.
Proof.
Let
$K' \gt 0$
be as in Proposition 9.2. Let
$D_{k}$
be as defined by Equation 44 and let

Then we have

Noting that
$|x + y|^{p} \leqslant (2 \max (|x|, |y|))^{p} \leqslant |2x|^{p} + |2y|^{p}$
for
$x, y \gt 0$
, it suffices to bound
$\mathbb{E}[I_{i}^{p}]$
for:

Observe the following: when
$|g_{k}| e^{-k/2K'}$
is bounded by
$M$
for all
$k$
, we have

for
$C = (1-e^{-1/2K'})^{-1}$
. This means

For
$I_{1}$
, recall the inequality
$|t^{p} - s^{p}|\leqslant 2^{p} (|t-s|^{p} + s^{p- n_{p}} |t-s|^{n_{p}})$
for each
$t, s \geqslant 0$
and
$p \gt 0$
, where
$n_{p} = p$
for
$0 \leqslant p \leqslant 1$
and
$n_{p} = 1$
otherwise. From this, we have

Since
$\mathbb{P}\big (A_{k}^{c} \, \big | \, g_{k+1}\big ) \leqslant K' e^{-k/K'}$
by Proposition 9.2,
$\mathbb{E} \left (|g_{k+1}|^{p} 1_{A_{k}^{c}} \right ) \leqslant \left ( \mathbb{E}_{\mu }|g|^{p}\right ) \cdot K'e^{-k/K'}$
is summable. Moreover,

holds for
$c = e^{-k/2pK'}$
, which is summable for
$k$
. Hence
$\mathbb{E}[I_{1}^{p}]$
is finite. The remaining terms
$\mathbb{E}[I_{2}^{p}]$
and
$\mathbb{E}[I_{4}^{p}]$
can be handled in a similar way.
We obtain an analogous estimate for random walks with finite exponential moment. Because it follows from the proof of Corollary 4.16 given Inequality 45, we omit the proof.
Proposition 9.4.
Let
$((\check {Z}_{n})_{n\gt 0}, (Z_{n})_{n\gt 0})$
be the (bi-directional) random walk generated by a non-elementary probability measure
$\mu$
on
$G$
with finite exponential moment. Then there exists
$K\gt 0$
such that

Using Proposition 9.3, we obtain the uniform second moment deviation inequality for non-elementary probability measures on the mapping class group. Now employing Theorem 4.1 and 4.2 of [Reference Mathieu and SistoMS20] and the proof of Theorem 4.20, we establish Theorem B.
To prove Theorem C, for each
$k \geqslant 0$
and
$(\check {\omega }, \omega ) \in \check {\Omega } \times \Omega$
we define the infinite geodesic
$\Gamma _{k}(\check {\omega }, \omega )$
to be a subsequential limit of
$\{[\check {Z}_{n+k}, Z_{n-k}] : n=1, 2, \ldots \}$
(which exists by Arzela-Ascoli and the second result in Proposition 9.3). Note that
$\tilde {d}(Z_{k}, \Gamma _{0}(\check {\omega }, \omega ))$
are identically distributed with
$\tilde {d}(id, \Gamma _{k}(\check {\omega }, \omega ))$
, which are all dominated by
$\tilde {d}(id, \sup _{n, m} [\check {Z}_{n}, Z_{m}])$
. It follows that

is summable for some
$o(k^{1/p})$
-function
$g(k)$
(
$K \log k$
for some
$K\gt 0$
, resp.) when the underlying measure has finite
$p$
-th moment (finite exponential moment, resp.). By Borel-Cantelli, we deduce

We conclude this paper by establishing Theorem A. Recall that Proposition 4.2 guaranteed the alignment of Schottky axes along
$[o, Z_{n} o]$
on
$X$
, which led to the reverse triangle inequality for distances on
$X$
(Lemma 3.18). We now record the corresponding result for distances on
$\tilde {X}$
.
Lemma 9.5.
Let
$0\lt \epsilon \lt 1$
and let
$S$
be a long enough,
$\epsilon$
-constricting
$K_{0}$
-Schottky set. Let
$\tilde {x}, \tilde {y} \in \tilde {X}$
and let
$\tilde {\kappa }_{1}, \ldots , \tilde {\kappa }_{N}$
be Schottky axes on
$\tilde {X}$
. If
$(x, \kappa _{1}, \ldots , \kappa _{N}, y)$
is
$D_{0}$
-semi-aligned, then we have

Proof.
Let
$M_{0}$
be such that
$S \subseteq G^{M_{0}}$
. Note that
$4 e^{-M_{0} / 2E_{0}} \leqslant \epsilon \lt 1$
, which implies

Now let
$\tilde {\kappa }$
be an arbitrary Schottky axis on
$\tilde {X}$
. Because
$M_{0} \gt 2K_{1}E_{1}$
for
$E_{1}$
as in Definition 9.1 and
$\tilde {\kappa }$
is a
$K_{1}$
-quasigeodesics by Lemma 8.1, we have
$\operatorname {diam}_{\tilde {X}}(\tilde {\kappa }) \gt E_{1}$
.
Since
$(x, \kappa _{1}, \ldots , \ldots , \kappa _{N}, y)$
is
$D_{0}$
-semi-aligned, they are
$E_{0}$
-aligned by Proposition 3.12. Consequently,
$(\tilde {x}, \tilde {\kappa }_{1}, \ldots , \tilde {\kappa }_{N}, \tilde {y})$
is
$K_{1}E_{0}$
-aligned by Lemma 8.1. Since the domains of
$\kappa _{i}$
’s are longer than
$M_{0} \geqslant L_{1}$
, we can obtain the points
$\tilde {p}_{i}$
’s on
$[\tilde {x}, \tilde {y}]$
as described in Proposition 8.2.
We now have

Here, Proposition 8.2 tells us that

Using this, we conclude

Corollary 9.6 ([[Reference GouëzelGou22], Lemma 4.14]). Let
$\nu$
be a non-elementary probability measure on
$G$
and let
$(Z_{n})_{n}$
be the random walk generated by
$\nu$
. Then for each
$\epsilon \gt 0$
, there exists
$C\gt 0$
such that

Proof.
Let
$S$
be a large, long enough and
$\epsilon$
-constricting
$K_{0}$
-Schottky set for
$\mu$
in
$G^{M_{0}}$
, for some suitable
$K_{0}, M_{0}\gt 0$
. (This determines the constants
$K_{1}, D_{0}, \ldots$
as in Definition 9.1.)
As in the proof of Corollary 4.11, there exists
$N\gt 0$
independent of
$g$
such that

Also, when
$(g^{-1} o, \mathbf{Y}_{i}, Z_{n}o)$
is
$D_{0}$
-semi-aligned, Lemma 9.5 implies that

where
$C'' = 4( K_{1} M_{0} + K_{1})$
. This bound also holds for
$n \leqslant N$
:

Given these, the proof ends by taking large enough
$C'\gt 0$
such that

Proof of Theorem 1.1. As in the proof of Theorem 6.4, we first take
$\epsilon \gt 0$
such that

Let
$S$
be a large, long enough and
$\epsilon$
-constricting Schottky set for
$\mu$
with cardinality greater than
$10/\epsilon$
, and let

By Proposition 6.2, there exists a non-elementary probability measure
$\nu$
, and for each sufficiently large
$N$
, a partition
$\mathscr{P}_{n, N, \epsilon }$
into
$(n, N, \epsilon , \nu )$
-pivotal equivalence classes for each
$n$
such that

decays exponentially in
$n$
. Let
$C \gt 0$
be a constant for
$\nu$
provided by Corollary 9.6: we have

for each
$g \in G$
and each
$m \gt 0$
. We now fix an
$N$
such that
.
Let
$\mathcal{E}$
be an equivalence class such that
. For each
$\omega \in \mathcal{E}$
$(o, \mathbf{Y}_{j(1)}, \ldots , \mathbf{Y}_{j'(\#\mathcal{P}/2)}, Z_{n} o)$
is
$D_{0}$
-semi-aligned. Lemma 9.5 tells us that

Since
$r_{i}$
’s are i.i.d. with

we can find
$K'\gt 0$
not depending on
$n$
such that

Summing up this conditional probability, we obtain the desired exponential bound.
Acknowledgments
The author thanks Hyungryul Baik, Kunal Chawla, Ilya Gekhtman, Vivian He, Sang-hyun Kim, Joseph Maher, Hidetoshi Masai, Yulan Qing, Kasra Rafi, Samuel Taylor, Giulio Tiozzo and Wenyuan Yang for helpful discussions. The author is indebted to the anonymous referee’s helpful and careful comments. 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.
Conflicts of interest
None
Financial support
The author was supported by Samsung Science & Technology Foundation (SSTF-BA1702-01 and SSTF-BA1301-51) and by a KIAS Individual Grant (SG091901). Part of the revision was conducted while the author was supported by Mid-Career Researcher Program (RS-2023-00278510) through the NRF funded by the government of Korea. This work constitutes part of the author’s PhD thesis.
Journal information
Compositio Mathematica is owned by the Foundation Compositio Mathematica and published by the London Mathematical Society in partnership with Cambridge University Press. All surplus income from the publication of Compositio Mathematica is returned to mathematics and higher education through the charitable activities of the Foundation, the London Mathematical Society and Cambridge University Press.