Asymptotics of the first-passage function on free and Fuchsian groups
Abstract
In this preprint we derive explicit estimates for the asymptotics of the first-passage function for a specific class of random walks on free groups and use them to prove the singularity of the hitting measure for a similarly defined class of random walks on Fuchsian groups.
1 Introduction
We continue to investigate the singularity conjecture for cocompact Fuchsian groups:
Conjecture 1.1 (Fuchsian singularity conjecture).
For every finite-range admissible random walk generated by a probability measure on a cocompact Fuchsian group , the hitting measure is singular with respect to the Lebesgue measure on .
This conjecture is usually stated for lattices in , and is attributed to Y. Guivarch, V. Kaimanovich, and F. Ledrappier. While it seems to be difficult to track down the exact source of this conjecture, we will refer to [DKN09] and [KL11] as the primary sources. See [Gui80], [KV83], [LS84], [GL93], [Le ̵08], [BHM11], [KL11], [GMM18], [DG20], [Aze+22] for related results.
In [Kos20] and [KT22] we developed techniques that allowed us to confirm Conjecture 1.1 for nearest-neighbour random walks on cocompact Fuchsian groups generated by side-pairing transformations which identify the opposite sides of a fundamental -polygon (here we assume ).
Recall the definition of the geometric distance: if , then
where is the fixed basepoint. The main approach we used consists of establishing the following two inequalities:
(1) |
(2) |
where
stands for the translation length, and denotes the Green metric. We can combine (1) and (2) to deduce the existence of a (loxodromic) generator such that
which ensures the singularity of the harmonic measure due to the following lemma which follows from [BHM11, Theorem 1.5] (or [DG20, Theorem 1.1] for non-symmetric measures):
Lemma 1.1 ([KT22], Lemma 2.3).
Consider a random walk on a Fuchsian group generated by an admissible probability measure with finite support. If there exists a loxodromic element which satisfies , then the hitting measure is singular with respect to the Lebesgue measure on .
1.1 Formulating results for non-nearest neighbour random walks
We would like to adapt this method to arbitrary admissible finite-range random walks on cocompact Fuchsian groups with centrally symmetric fundamental polygons. Recall the definition of the first-passage function for a random walk generated by a probability measure :
Then we denote
Now we are prepared to formulate an effective sufficient condition for a random walk on a Fuchsian group to satisfy Conjecture 1.1.
Proposition 1.1.
Let be a Fuchsian group with a centrally symmetric fundamental polygon equipped with side-pairing transformations identifying the opposite sides. Consider an admissible random walk on generated by a finitely supported probability measure .
If there exists a finite-range probability measure on such that , where is the natural projection defined by , and such that satisfies
(3) |
then the hitting measure is singular with respect to the Lebesgue measure on .
Proposition 1.1 suggests a reduction of the singularity conjecture to free groups. Recall the definition of cylinder sets in :
Now we formulate the main results of the paper.
Theorem 1.1.
Consider a random walk on generated by a symmetric measure such that for some . Then we have
(4) |
where is the hitting measure on corresponding to .
Remark. It is a well-known fact, established by Ancona in [Anc88], that for any on a geodesic segment in the Cayley graph of a hyperbolic group we have
for a constant which does not depend on the choice of . It implies the exponential decay of the Green/first-passage function. However, despite various improvements of this inequality for free and hyperbolic groups in [Lal93], [Led01], and [GL13], the existing methods do not provide the explicit dependence of on measures and .
Corollary 1.1.
Let be a Fuchsian group with a centrally symmetric fundamental polygon equipped with side-pairing transformations identifying the opposite sides. Consider an admissible random walk on generated by a probability measure supported on for some . Then satisfies Conjecture 1.1.
Proof.
To prove Theorem 1.1, we introduce the notions of barriers (Definition 2.1) and strong barriers (Definition 2.2) in order to establish linear relations between the measures of certain cylinder sets. It turns out that can be realized as the spectral radius of a certain irreducible matrix which corresponds to mentioned relations, and that allows us to make use of the Perron-Frobenius theorem. The proofs themselves are given in the end of Section 3.
Remark. We would like to remark that the idea of exploiting barriers to characterize the behaviour of random walks at infinity for hyperbolic groups goes back to the works of Derriennic ([Der75]), Lalley, and Gouëzel (see [Lal93], [GL13]).
Unfortunately, there is a wide class of finite-range random walks for which (3) does not hold at all, which does not allow us to fully settle Conjecture 1.1. Let us consider an automorphism of which is uniquely defined by . Random walks generated by measures satisfying for every will be called antisymmetric.
Theorem 1.2.
Consider an antisymmetric random walk on generated by an admissible measure satisfying the following condition: there exists such that the singleton is an -barrier for and is an -barrier (see Definition 2.1).
Then for each we have
and
Finally, we have
if and only if is an -barrier.
As a quick corollary, we get the following application:
Corollary 1.2.
For any antisymmetric random walk on supported on the set we get
and
for all . In particular, such random walks do not satisfy (3).
We provide the proof in Example 4.1. Finally, we show that there are symmetric examples which do not satisfy (3) as well, see Example 4.2.
Acknowledgments. We would like to thank Giulio Tiozzo, Kunal Chawla, and Steven Lalley for significantly improving the structure and overall presentation of the preprint, and Alexander Kalmynin for suggestions that eventually led to the proof of Lemma 3.2.
2 Prerequisites
Throughout this paper we consider (transient and irreducible) random walks on free groups generated by admissible probability measures , that is, having the support generate as a semigroup.
2.1 Notation
Let . Recall the definitions of Green function and the first-passage functions:
(Green function) | |||||
(First-passage function) |
What is known in our case is that both functions are always finite, this is equivalent to the transience of the random walks considered. Also, it is not hard to show that , therefore, it does not make a difference whether to consider the Green or first-passage function. We will be working with and , which is often referred to as the Green distance.
In particular, for admissible random walks on , almost every sample path converges to the boundary , and the respective pushforward measure is referred to as the hitting measure, and we will denote it by , or by for brevity.
Let be an element with the reduced representation , where .
where denotes the word distance in . We will call the set the -shadow.
We will also be using various restricted versions of the first-passage function.
Let , and , such that . Let us denote
If , and , then we denote
Finally, if are subsets which satisfy for all , then we inductively define
We will also denote , where denotes the word distance in with respect to .
2.2 Establishing the barrier framework
Definition 2.1.
Let . A subset is called a -barrier if for any we have .
Example 2.1.
Let us denote , and let for some with no other restrictions. Then for every element the subset is a -barrier.
Remark. Keep in mind that if the support has “holes”, in other words, for some , then this might not be a minimal barrier as the next example shows.
Example 2.2.
Suppose that . Then for every generator the geodesic segment is a minimal -barrier.
The next lemma is simple but quite important.
Lemma 2.1.
If is an -barrier, then for every we have
(5) |
Proof.
This is just an application of the full probability formula combined with the Markov property: every path either enters before hitting , or avoids altogether:
However, as is a -barrier, the term vanishes. ∎
Proposition 2.1.
Let be a -barrier. Then for every with the reduced representation we have
(6) |
Proof.
First of all, recall that there exist sets and such that
This holds because , as defined, is a -barrier. As , we can use the barrier property of and Lemma 2.1 to write
We have to treat the term carefully, though: what happens if ? Taking this into account, we rewrite the above sum as follows:
Now take the sum of both sides over , so we get
(7) | ||||
Let us treat both (double) sums on RHS separately. For we immediately get
as is a -barrier.
It is a bit trickier if , but the idea is rather similar:
as is a -barrier. This finishes our argument. ∎
Example 2.3.
Consider a symmetric random walk on supported on . Then the equations (6) for the -barrier look as follows:
(8) |
but as we have due to invariance with respect to the automorphism , we get
(9) |
Example 2.4.
Consider an admissible symmetric random walk on which is supported on . Let us consider the following barriers:
-
1.
The subset is an -barrier:
-
2.
The subset is a -barrier:
-
3.
The subset is an -barrier:
-
4.
The subset is a -barrier:
-
5.
The subset is an -barrier:
2.3 Strong barriers
In this subsection we modify the barrier definition in a way that allows us to formulate starting point-independent versions of Lemma 2.1 and Proposition 2.1.
Definition 2.2.
Let . A subset is a strong -barrier if
-
•
,
-
•
for any , .
Remark. It is easy to see that the second condition implies that every strong barrier is a barrier, as for any non-identity .
In order to formulate the next set of theorems, we need to establish some basic statements about shadows. Let us denote , where stands for the word distance on .
Lemma 2.2.
Let . Then if and only if .
Proof.
By the definition of word distance, if and only if the last digit of in its reduced representation does not equal to the first digit of in its reduced representation. However, the last digit of is the inverse of first digit of , and is the first digit of if and only if . ∎
Lemma 2.3.
Consider such that , and . Then .
Proof.
As , we have . We know that , therefore, Lemma 2.2 implies that . ∎
Lemma 2.4.
Let , and consider . Then .
Proof.
We will proceed by showing that and .
- •
- •
∎
We can use Lemma 2.1 to get the following statement:
Lemma 2.5.
If is a strong -barrier, then for every , we have
(10) |
Proof.
In a similar way we can generalize Proposition 2.1 as well:
Proposition 2.2.
Let be a strong -barrier. Then for every , , where is the reduced representation, we have
(11) | ||||
Proof.
Consider an anti-symmetric random walk on generated by a probability measure . For every fix a strong -barrier . Then we can define the following matrix:
Then, as a corollary from Proposition 2.2, we get the following theorem:
Theorem 2.1.
Proof.
Applying Lemma 2.5 and exploiting the antysymmetry, we get
for every . This can be rewritten as
This is equivalent to
∎
We will need the following lemma later.
Lemma 2.6.
Consider , a strong -barrier , and an element together with a set such that for every we have (this is equivalent to being a -barrier).
Then for every we have
Proof.
This is another consequence of the full-probability formula, as every path from to which avoids has to hit the barrier , as due to the definition of a strong barrier. Moreover, if lies in the intersection of and , then the term vanishes. ∎
3 Proof of Proposition 1.1 and Theorem 1.1
In this subsection we will be working in the setting of Example 2.3. We will need some algebraic lemmas in order to establish (4).
3.1 Technical linear-algebraic lemmas
This lemma follows immediately from the geometric definition of positive spans.
Lemma 3.1.
Let be positive vectors (both coordinates are positive). Then the positive linear span of ’s is the smallest cone which contains all ’s. In other words, there exist such that for all we have
Lemma 3.2.
Let be non-negative numbers, and suppose that satisfy
(12) |
Then for all .
Proof.
Suppose that
for some . Then due to being monotone increasing on . However, we can reinterpret the -th and -th equations in (12) as follows:
Moreover, as we know that for all and , due to the Lemma 3.1 applied to , there should be an index such that
But then we can repeat the argument until we get . However, this will not work, as, on the one hand, the first and second equations in (12) give
but on the other hand,
Therefore, cannot belong to the cone generated by , which leads to a contradiction. ∎
Corollary 3.1.
In the setting of the Lemma 3.2, if , we always have , where is the largest root of .
Proof.
Let us consider the matrix defined as follows:
This is an irreducible matrix, and the following identity holds:
Perron-Frobenius theorem applies here, and the Collatz-Wieland formula yields
Keep in mind that the characteristic polynomial of is precisely the polynomial in the statement of the corollary. Finally, Lemma 3.2 tells us that this minimum just equals . ∎
3.2 Proofs of the main results
Proof of Proposition 1.1.
As for any , we have for any as well, therefore,
Without loss of generality, we can assume that there exists an index such that
. In particular, there is such that . But this inequality can be rewritten as follows:
Therefore, we can apply Lemma 1.1 to finish the proof. ∎
Proof of Theorem 1.1.
First of all, let us denote , where , and rewrite the equations (9) to obtain the following system:
Corollary 3.1, applied to this system, implies that
(13) |
where is the largest root of . Now we only have to prove that
As a consequence from Lemma 2.1, for every we have
where
Here we apply the Perron-Frobeinus theorem once again. As is irreducible, we can use the min-max Collatz-Wielandt formula to get that for every we have
We know that exists, we just need to establish that the limit is, indeed, the largest eigenvalue. Denoting it by , and applying , we get
So,
(14) |
Remark. We would like to note that the equations (9) and Lemma 2.6 imply the following matrix identity corresponding to the barrier :
(15) |
where is the same operator as the one defined in Theorem 2.1, is the permutation matrix which reverses the rows. If we denote for , Lemma 2.6, applied to for various , turns out to be equivalent to the following matrix identity in for all and .
It remains to note that for the defined above matrices equal , and for we get .
4 Disproving (3) in the general case
First of all, let us prove Theorem 1.2, then we will present a random walk on which satisfies the property in Theorem 1.2, and is not supported on the generating set.
Proof of Theorem 1.2.
First of all, let us notice that if the random walk satisfies the property mentioned in the statement, then Lemma 2.1 implies that
for any , . Therefore,
Moreover, the fact that is an -barrier for all implies that by the definition of the barriers. This allows to apply Proposition 2.1 in a certain way: consider an -barrier which fully lies in . Then
Observe that for every we have : if is the reduced representation of , then we can fix the first index such that . As , is always well-defined. In this case Lemma 2.1 implies
as and (here antisymmetry implies ). Notice that the same argument implies that only if . Therefore,
is equivalent to . However,
∎
Example 4.1.
Consider an antisymmetric random walk on supported on the set , and assume that , . Then it is easy to see that is an -barrier for , and is still an -barrier, as there is no way to enter directly from . Therefore, we get
due to being disjoint.
Example 4.2.
We would like to show that (3) breaks in the symmetric case as well. Consider Example 2.4. Due to being a -barrier, and being an -barrier, we have
due to Lemma 2.1. Therefore, , and .
To finish the argument, we consider an automorphism , defined by . Denoting and using the invariance of the first-passage functions w.r.t. automorphisms, we get
As the support of is , the first-passage function is still multiplicative, and we have an exact identity for the measures of the cylinder sets for , so we get
5 Further directions
-
•
As pointed out to the author by Kunal Chawla, the existing results on continuity of the drift and entropy, in particular, [GMM18, Proposition 2.3] and [GMM18, Theorem 2.9], are applicable due to the random walks considered being finitely supported. Therefore, the dimension also continuously depends on due to [BHM11] and [Tan19], which allows us to extend Corollary 1.1 to sufficiently small perturbations of measures supported strictly on the powers of generators. Also, see [Aze+22] for the most recent results in this direction.
-
•
At the same time, we would like to note that Theorem 1.2 does not yield any actual counterexamples to the Fuchsian version of (3): while (3) does not hold for every random walk on free groups, it can still hold when we pass to Fuchsian groups as the corresponding first-passage functions are strictly larger for Fuchsian groups.
Besides, even if (3) is not true for arbitrary finite range random walks on Fuchsian groups, it still doesn’t necessarly contradict the singularity conjecture itself, as is expected to hold for an element other than a power of a side-pairing generator of .
-
•
Finally, we would like to remark that Theorem 2.1 proved to be not enough by itself to establish Theorem 1.1 and, consequently, (3). Essentially, the proof required us to represent the respective operator as a power of an operator with a much simpler structure (see the remark after the proof of Theorem 1.1). It seems reasonable that Lemma 2.6 yields similar (to (15)) decompositions for barrier operators in the general case, and we hope that they will, potentially, suggest the correct generalization of (3) for arbitrary random walks on free groups.
References
- [Anc88] Alano Ancona “Positive harmonic functions and hyperbolicity” In Potential Theory Surveys and Problems Berlin, Heidelberg: Springer Berlin Heidelberg, 1988, pp. 1–23
- [Aze+22] Aitor Azemar et al. “Random walk speed is a proper function on Teichmüller space” arXiv, 2022 DOI: 10.48550/ARXIV.2212.06581
- [BHM11] Sébastien Blachère, Peter Haïssinsky and Pierre Mathieu “Harmonic measures versus quasiconformal measures for hyperbolic groups” In Ann. Sci. Éc. Norm. Supér. (4) 44.4, 2011, pp. 683–721 DOI: 10.24033/asens.2153
- [Der75] Yves Derriennic “Marche aléatoire sur le groupe libre et frontière de Martin” In Z. Wahrscheinlichkeitstheorie und Verw. Gebiete 32.4, 1975, pp. 261–276 DOI: 10.1007/BF00535840
- [DG20] Matthieu Dussaule and Ilya Gekhtman “Entropy and drift for word metrics on relatively hyperbolic groups” In Groups Geom. Dyn. 14.4, 2020, pp. 1455–1509
- [DKN09] Bertrand Deroin, Victor Kleptsyn and Andrés Navas “On the question of ergodicity for minimal group actions on the circle” In Mosc. Math. J. 9.2, 2009, pp. 263–303\bibrangessepback matter DOI: 10.17323/1609-4514-2009-9-2-263-303
- [GL13] Sébastien Gouëzel and Steven P. Lalley “Random walks on co-compact fuchsian groups” In Annales scientifiques de l’École Normale Supérieure Ser. 4, 46.1 Société mathématique de France, 2013, pp. 131–175 DOI: 10.24033/asens.2186
- [GL93] Y. Guivarc’h and Y. Le Jan “Asymptotic winding of the geodesic flow on modular surfaces and continuous fractions” In Annales scientifiques de l’École Normale Supérieure Ser. 4, 26.1 Elsevier, 1993, pp. 23–50 DOI: 10.24033/asens.1666
- [GMM18] Sébastien Gouëzel, Frédéric Mathéus and François Maucourant “Entropy and drift in word hyperbolic groups” In Invent. Math. 211.3, 2018, pp. 1201–1255 DOI: 10.1007/s00222-018-0788-y
- [Gui80] Y. Guivarc’h “Sur la loi des grands nombres et le rayon spectral d’une marche aléatoire” In Conference on Random Walks (Kleebach, 1979) (French) 74, Astérisque Soc. Math. France, Paris, 1980, pp. 47–98\bibrangessep3
- [KL11] Vadim A. Kaimanovich and Vincent Le Prince “Matrix random products with singular harmonic measure” In Geom. Dedicata 150, 2011, pp. 257–279 DOI: 10.1007/s10711-010-9504-9
- [Kos20] Petr Kosenko “Fundamental Inequality for Hyperbolic Coxeter and Fuchsian Groups Equipped with Geometric Distances” In International Mathematics Research Notices 2021.6, 2020, pp. 4709–4728 DOI: 10.1093/imrn/rnaa213
- [KT22] Petr Kosenko and Giulio Tiozzo “The fundamental inequality for cocompact Fuchsian groups” In Forum of Mathematics, Sigma 10 Cambridge University Press, 2022, pp. e102 DOI: 10.1017/fms.2022.94
- [KV83] V.. Kaimanovich and A.. Vershik “Random walks on discrete groups: boundary and entropy” In Ann. Probab. 11.3, 1983, pp. 457–490
- [Lal93] Steven P. Lalley “Finite Range Random Walk on Free Groups and Homogeneous Trees” In The Annals of Probability 21.4 Institute of Mathematical Statistics, 1993, pp. 2087–2130 URL: http://www.jstor.org/stable/2244712
- [Le ̵08] Vincent Le Prince “A relation between dimension of the harmonic measure, entropy and drift for a random walk on a hyperbolic space” In Electron. Commun. Probab. 13, 2008, pp. 45–53 DOI: 10.1214/ECP.v13-1350
- [Led01] François Ledrappier “Some asymptotic properties of random walks on free groups” In Topics in probability and Lie groups: boundary theory 28, CRM Proc. Lecture Notes Amer. Math. Soc., Providence, RI, 2001, pp. 117–152 DOI: 10.1090/crmp/028/05
- [LS84] Terry Lyons and Dennis Sullivan “Function theory, random paths and covering spaces” In J. Differential Geom. 19.2, 1984, pp. 299–323 URL: http://projecteuclid.org/euclid.jdg/1214438681
- [Tan19] Ryokichi Tanaka “Dimension of harmonic measures in hyperbolic spaces” In Ergodic Theory Dynam. Systems 39.2, 2019, pp. 474–499 DOI: 10.1017/etds.2017.23