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

2020 Mathematics Subject Classification: Primary 37A35; Secondary 37A44, 11J70.

On the rate of convergence of continued fraction statistics of random rationals

Ofir David Department of Mathematics, Technion, Haifa, Israel [email protected] Taehyeong Kim The Einstein Institute of Mathematics, Edmond J. Safra Campus, Givat Ram, The Hebrew University of Jerusalem, Jerusalem, 91904, Israel [email protected] Ron Mor The Einstein Institute of Mathematics, Edmond J. Safra Campus, Givat Ram, The Hebrew University of Jerusalem, Jerusalem, 91904, Israel [email protected]  and  Uri Shapira Department of Mathematics, Technion, Haifa, Israel [email protected]
Abstract.

We show that the statistics of the continued fraction expansion of a randomly chosen rational in the unit interval, with a fixed large denominator qq, approaches the Gauss-Kuzmin statistics with polynomial rate in qq. This improves on previous results giving the convergence without rate. As an application of this effective rate of convergence, we show that the statistics of a randomly chosen rational in the unit interval, with a fixed large denominator qq and prime numerator, also approaches the Gauss-Kuzmin statistics. Our results are obtained as applications of improved non-escape of mass and equidistribution statements for the geodesic flow on the space SL2()/SL2()\operatorname{SL}_{2}(\mathbb{R})/\operatorname{SL}_{2}(\mathbb{Z}).

1. Introduction

In this article we wish to revisit some of the results proved in [DS18] and strengthen them. The main novel observation in the current article is that one of the two assumptions appearing in the main theorems of [DS18] follows from the other and hence can be dropped. We then add to this by presenting a new arithmetic application of the stronger result.

1.1. Background and the main application

It is well known since ancient times that a real number x(0,1)x\in(0,1) admits a continued fraction expansion (abbreviated hereafter as c.f.e); namely, there are positive integers a1,a2,a_{1},a_{2},\dots such that

x=1a1+1a2+1.x=\frac{1}{a_{1}+\frac{1}{a_{2}+\frac{1}{\ddots}}}.

In fact, to be more precise, irrational xx’s correspond bijectively to infinite sequences (an)(a_{n})\in\mathbb{N}^{\mathbb{N}}, where the above equation is interpreted as a limit, and rational xx’s correspond to finite sequences of positive integers, where here the correspondence is not 11-11 but almost so: each rational x(0,1)x\in(0,1) can be written in exactly two ways,

x=1a1+1a2+1+1an or x=1a1+1a2+1+1(an1)+11x=\frac{1}{a_{1}+\frac{1}{a_{2}+\frac{1}{\ddots+\frac{1}{a_{n}}}}}\textrm{ or }x=\frac{1}{a_{1}+\frac{1}{a_{2}+\frac{1}{\ddots+\frac{1}{(a_{n}-1)+\frac{1}{1}}}}}

where nn\in\mathbb{N} and an2a_{n}\geq 2. As usual, we abbreviate and write x=[a1,a2,]x=\left[a_{1},a_{2},\dots\right] when xx is irrational and x=[a1,,an]x=\left[a_{1},\dots,a_{n}\right] when xx is rational. As a convention we will always use the shorter expansion for rationals, which does not end with 11. The length of the c.f.e of a rational number is defined by

(1.1) len([a1,,an])=defn.\operatorname{len}([a_{1},\ldots,a_{n}])\overset{\operatorname{def}}{=}n.

Our discussion will revolve around the continued fraction expansions of rationals with large denominators, but to motivate the phenomena discussed it is actually more natural to start with irrationals.

It is well known (see for example the introduction of [AS18]) that given any finite string of positive integers 𝐰=(w1,,wk)k\mathbf{w}=(w_{1},\dots,w_{k})\in\mathbb{N}^{k}, for Lebesgue almost any irrational x(0,1)x\in(0,1), the asymptotic frequency of appearance of 𝐰\mathbf{w} in the c.f.e of xx exists and is independent of xx. That is, the limit

(1.2) D𝐰=deflimn#{1n:𝐰=(a,a+1,,a+k1)}nD_{\mathbf{w}}\overset{\operatorname{def}}{=}\lim_{n\to\infty}\frac{\#\left\{1\leq\ell\leq n:\mathbf{w}=(a_{\ell},a_{\ell+1},\dots,a_{\ell+k-1})\right\}}{n}

exists and is independent of xx, as the notation suggests (in fact D𝐰D_{\mathbf{w}} equals 1log2I𝐰11+t𝑑t\frac{1}{\log 2}\int_{I_{\mathbf{w}}}\frac{1}{1+t}dt where I𝐰I_{\mathbf{w}} is a certain interval in (0,1)(0,1), see Remark 3.2). For example, for a randomly chosen xx, the asymptotic density of appearance of 11’s in its c.f.e is 2log3log20.4152-\frac{\log 3}{\log 2}\sim 0.415 with probability 11.

A similar density could be defined for rationals: if x=[a1,,an]x=\left[a_{1},\dots,a_{n}\right] then we can define

D𝐰(x)=def#{1nk+1:𝐰=(a,a+1,,a+k1)}nk+1.D_{\mathbf{w}}(x)\overset{\operatorname{def}}{=}\frac{\#\left\{1\leq\ell\leq n-k+1:\mathbf{w}=(a_{\ell},a_{\ell+1},\dots,a_{\ell+k-1})\right\}}{n-k+1}.

In [DS18] the authors investigated the densities D𝐰(j/q)D_{\mathbf{w}}(j/q) of rationals j/qj/q in reduced form for large denominators qq, as well as the length of the c.f.e of such rationals. In particular, the following result is proved.

Theorem 1.1 ([DS18]).

For any qq\in\mathbb{N}, let j/q(0,1)j/q\in(0,1) be a random rational with denominator qq and numerator jj co-prime to qq chosen uniformly (namely, one picks jj randomly and uniformly from the φ(q)\varphi(q) possible numerators co-prime to qq). Then for any kk\in\mathbb{N}, any string 𝐰k\mathbf{w}\in\mathbb{N}^{k}, and any ϵ>0\epsilon>0, we have

  1. (1)
    (|D𝐰(j/q)D𝐰|>ϵ)0 as q,\mathbb{P}(|D_{\mathbf{w}}(j/q)-D_{\mathbf{w}}|>\epsilon)\to 0\textrm{ as }q\to\infty,
  2. (2)
    (|len(j/q)logq12log2π2|>ϵ)0 as q.\mathbb{P}\left(|\frac{\operatorname{len}(j/q)}{\log q}-\frac{12\log 2}{\pi^{2}}|>\epsilon\right)\to 0\textrm{ as }q\to\infty.

In this paper, we are able to improve this result, and show that not only the probability in Theorem 1.1 converges to zero, but it does so at a polynomial rate.

Theorem 1.2.

For any qq\in\mathbb{N}, let j/q(0,1)j/q\in(0,1) be a random rational with denominator qq and numerator jj co-prime to qq chosen uniformly. Then, for any kk\in\mathbb{N}, any string 𝐰k\mathbf{w}\in\mathbb{N}^{k}, and any ϵ>0\epsilon>0, there exist αϵ>0\alpha_{\epsilon}>0 and αϵ,𝐰>0\alpha_{\epsilon,\mathbf{w}}>0 such that for all qq large enough:

  1. (1)
    (|D𝐰(j/q)D𝐰|>ϵ)<qαϵ,𝐰,\mathbb{P}(|D_{\mathbf{w}}(j/q)-D_{\mathbf{w}}|>\epsilon)<q^{-\alpha_{\epsilon,\mathbf{w}}},
  2. (2)
    (|len(j/q)logq12log2π2|>ϵ)<qαϵ.\mathbb{P}\left(|\frac{\operatorname{len}(j/q)}{\log q}-\frac{12\log 2}{\pi^{2}}|>\epsilon\right)<q^{-\alpha_{\epsilon}}.

Our methods do not provide any estimate of the values αϵ\alpha_{\epsilon} and αϵ,𝐰\alpha_{\epsilon,\mathbf{w}} in Theorem 1.2, only their existence. It would be interesting to find out how fast do these probabilities must decay, and in particular the dependence of these powers on ϵ\epsilon.

Theorem 1.2 is in fact equivalent to the following generalization of Theorem 1.1. We show that the probability for the c.f.e of j/qj/q to ‘behave badly’ (in the sense of Theorem 1.1) approaches 0, even if we choose jj uniformly at random from a set Λq(/q)×\Lambda_{q}\subseteq(\mathbb{Z}/q\mathbb{Z})^{\times} of smaller size qαqφ(q)q^{-\alpha_{q}}\varphi(q) with αq0\alpha_{q}\to 0, instead of choosing from (/q)×(\mathbb{Z}/q\mathbb{Z})^{\times}. Here and throughout this paper, we identify between (/q)×(\mathbb{Z}/q\mathbb{Z})^{\times} and the set {1jq:gcd(j,q)=1}\left\{1\leq j\leq q:\ \gcd(j,q)=1\right\}.

Theorem 1.3.

For any qq\in\mathbb{N}, let Λq(/q)×\Lambda_{q}\subseteq(\mathbb{Z}/q\mathbb{Z})^{\times} so that limqlog|Λq|logq=1\underset{q\to\infty}{\lim}\frac{\log|\Lambda_{q}|}{\log q}=1, and let j/qj/q be a random rational with denominator qq and numerator jΛqj\in\Lambda_{q} chosen uniformly (we denote this probability measure over Λq\Lambda_{q} by |Λq\mathbb{P}|_{\Lambda_{q}}). Then, for any kk\in\mathbb{N}, any string 𝐰k\mathbf{w}\in\mathbb{N}^{k}, and any ϵ>0\epsilon>0, we have

  1. (1)
    |Λq(|D𝐰(j/q)D𝐰|>ϵ)0 as q,\mathbb{P}|_{\Lambda_{q}}(|D_{\mathbf{w}}(j/q)-D_{\mathbf{w}}|>\epsilon)\to 0\textrm{ as }q\to\infty,
  2. (2)
    |Λq(|len(j/q)logq12log2π2|>ϵ)0 as q.\mathbb{P}|_{\Lambda_{q}}\left(|\frac{\operatorname{len}(j/q)}{\log q}-\frac{12\log 2}{\pi^{2}}|>\epsilon\right)\to 0\textrm{ as }q\to\infty.

As an application, we deduce the following prime numerator version of Theorem 1.1. It follows from the prime number theorem in combination with Theorem 1.3.

Theorem 1.4.

For any qq\in\mathbb{N}, let p/q(0,1)p/q\in(0,1) be a random rational with denominator qq and prime numerator pp co-prime to qq chosen uniformly. Then for any kk\in\mathbb{N}, any string 𝐰k\mathbf{w}\in\mathbb{N}^{k}, and any ϵ>0\epsilon>0, we have

  1. (1)
    |primes(|D𝐰(p/q)D𝐰|>ϵ)0 as q,\mathbb{P}|_{\text{primes}}(|D_{\mathbf{w}}(p/q)-D_{\mathbf{w}}|>\epsilon)\to 0\textrm{ as }q\to\infty,
  2. (2)
    |primes(|len(p/q)logq12log2π2|>ϵ)0 as q.\mathbb{P}|_{\text{primes}}\left(|\frac{\operatorname{len}(p/q)}{\log q}-\frac{12\log 2}{\pi^{2}}|>\epsilon\right)\to 0\textrm{ as }q\to\infty.

Related questions regarding the distribution of the coefficients in continued fractions of random rationals, and particularly the expected length of the expansion, were previously discussed in the literature. In [Hei69], Heilbronn computed the asymptotics of the average length over p(/q)×p\in(\mathbb{Z}/q\mathbb{Z})^{\times}, also giving an error term. This error term was improved by Porter [Por75] and then further improved by Ustinov [Ust08].

Other works considered the length question on different sets. In [Bi05], Bykovski considered for any qq the average length for the c.f.e of pq\frac{p}{q}, over all pp not necessarily coprime to qq. Others considered a larger probability space, comprised of all pairs (p,q)(p,q) where pp is coprime to qq and qQq\leq Q for some QQ. On this space, Dixon [Dix70] showed a large deviation result, namely bounded the probability that len(p/q)\operatorname{len}(p/q) deviates from the expected value by a certain amount. This result was later improved by Hensely [Hen94]. Furthermore, in [BV05] Baladi and Vallée consider a central limit theorem on this space. In comparison with these works, we consider too a large deviation question, but on a significantly smaller probability space where the denominator is considered fixed. Furthermore, we also consider the distribution of words in the c.f.e expansions rather than only its length.

For construction of normal numbers with respect to the c.f.e using rational numbers, see [AKS81, Van16].

1.2. The diagonal flow and the main result

The proof of Theorem 1.3 will be given in §3 and is based on the dynamical results we now present.

We start with the setting in which our mathematical discussion will take place, namely the space of unimodular lattices in Euclidean plane. We will use similar notation to [DS18]. Let G=SL2()G=\operatorname{SL}_{2}(\mathbb{R}), Γ=SL2()\Gamma=\operatorname{SL}_{2}(\mathbb{Z}), X=G/ΓX=G/\Gamma, and let x0Xx_{0}\in X denote the identity coset. For a point xXx\in X we let δx\delta_{x} denote the probability measure supported on the singleton {x}\left\{x\right\}. Consider the diagonal flow

at=(et/200et/2)G.a_{t}=\begin{pmatrix}e^{t/2}&0\\ 0&e^{-t/2}\end{pmatrix}\in G.

The unstable horospherical subgroup UU of GG with respect to ata_{t} is given by

U={us=(1s01):s}.U=\left\{u_{s}=\begin{pmatrix}1&s\\ 0&1\end{pmatrix}:s\in\mathbb{R}\right\}.

For a given set Λq(/q)×\Lambda_{q}\subseteq(\mathbb{Z}/q\mathbb{Z})^{\times} denote

(1.3) δΛq[0,2logq]=1|Λq|pΛq12logq02logqδatup/qx0𝑑t,\delta_{\Lambda_{q}}^{[0,2\log q]}=\frac{1}{|\Lambda_{q}|}\sum_{p\in\Lambda_{q}}\frac{1}{2\log q}\int_{0}^{2\log q}\delta_{a_{t}u_{p/q}x_{0}}dt,

that is, δΛq[0,2logq]\delta_{\Lambda_{q}}^{[0,2\log q]} is the probability measure given by the continuous average in the diagonal direction on the segment [0,2logq][0,2\log q] of the discrete counting probability measure on the points {up/qx0:pΛq}\{u_{p/q}x_{0}:p\in\Lambda_{q}\}. For an illustration of the geodesic segments on which we average, see Figure 1 and the discussion before it. Recall that we say that a sequence of probability measures (μn)n(\mu_{n})_{n\in\mathbb{N}} on XX does not exhibit escape of mass if any weak* limit of it is a probability measure. The following is one of the main results of [DS18].

Theorem 1.5.

[DS18, Theorem 1.7] Let Λq(/q)×\Lambda_{q}\subseteq(\mathbb{Z}/q\mathbb{Z})^{\times} be subsets such that

  1. (1)

    limqlog|Λq|logq=1\underset{q\to\infty}{\lim}\frac{\log|\Lambda_{q}|}{\log q}=1,

  2. (2)

    the sequence of measures δΛq[0,2logq]\delta_{\Lambda_{q}}^{[0,2\log q]} does not exhibit escape of mass.

Then δΛq[0,2logq]wμX\delta_{\Lambda_{q}}^{[0,2\log q]}\overset{\operatorname{w}^{*}}{\longrightarrow}\mu_{X}, where μX\mu_{X} is the Haar probability measure on XX.

The current paper concerns itself with proving that the second item in Theorem 1.5 is superfluous since it actually follows from the first item. Our main result is the following bound on the possible escape of mass for divergent orbits.

Theorem 1.6.

Let h[0,1]h\in[0,1]. Suppose that Λq(/q)×\Lambda_{q}\subseteq(\mathbb{Z}/q\mathbb{Z})^{\times} are given so that

lim infqlog|Λq|logqh.\liminf_{q\to\infty}\frac{\log|\Lambda_{q}|}{\log q}\geq h.

Then any weak limit μ\mu of the sequence δΛq[0,2logq]\delta_{\Lambda_{q}}^{[0,2\log q]} satisfies

μ(X)2h1.\mu(X)\geq 2h-1.

Combining Theorem 1.5 and Theorem 1.6 with h=1h=1, we have

Corollary 1.7.

Let Λq(/q)×\Lambda_{q}\subseteq(\mathbb{Z}/q\mathbb{Z})^{\times} be subsets satisfying limqlog|Λq|logq=1\lim_{q\to\infty}\frac{\log|\Lambda_{q}|}{\log q}=1, then δΛq[0,2logq]wμX\delta_{\Lambda_{q}}^{[0,2\log q]}\overset{\operatorname{w}^{*}}{\longrightarrow}\mu_{X}.

Remark 1.8.

We note here that the assumption limqlog|Λq|logq=1\lim_{q\to\infty}\frac{\log|\Lambda_{q}|}{\log q}=1 in Theorem 1.5 and Corollary 1.7 is necessary and refer the reader to §1.4 for a discussion on the matter.

To illustrate the trajectories {atup/qx0:t[0,2logq]}\{a_{t}u_{p/q}x_{0}:\ t\in[0,2\log q]\}, we show in Figure 1 their projection to the upper half plane \mathbb{H}, which is identified with the quotient K\GK\backslash G, where K=SO2()K=\operatorname{SO}_{2}(\mathbb{R}). The trajectories are shown in the standard fundamental domain, given by points zz\in\mathbb{H} with |(z)|12|\Re(z)|\leq\frac{1}{2} and |z|1|z|\geq 1. As discussed in [DS18], at times t=0,2logqt=0,2\log q, these trajectories start and end at height y=1y=1 in the upper half plane, and for t[0,2logq]t\not\in[0,2\log q] the trajectories diverge directly into the cusp. Hence the [0,2logq][0,2\log q]-portion of the geodesic trajectory is the interesting one to study.

As can be observed, there is a variance between the trajectories for different values of pp, with some comprising of just one long excursion to the cusp, and others disproportionally bounded. This is why averaging on a relatively large set is required to obtain equidistribution, and otherwise one can construct counter examples as we have in §1.4 . Furthermore, there are visible symmetries between the trajectories for different values of pp, which were proved in [DS18] and are also used in this current paper (see Lemma 2.5).

Refer to caption
Figure 1. The projection to the upper half plane, of the trajectories {atup/qx0:t[0,2logq]}\{a_{t}u_{p/q}x_{0}:\ t\in[0,2\log q]\}, for q=11q=11 and all p(/q)×p\in(\mathbb{Z}/q\mathbb{Z})^{\times}.

1.3. Prime numerator Zaremba

In this part of the introduction we wish to present a conjecture we refer to as the prime numerator Zaremba conjecture. The discussion in this subsection has a different flavour to it as we do not present any theorems but rather discuss a conjecture and present supporting data. Nevertheless, we find the conjecture interesting and think it fits in this paper as we were inspired by Theorem 1.4 when we formulated it. We hope it will inspire other people to look into it as well.

The Zaremba conjecture asserts that for any denominator qq, one can find j(/q)×j\in(\mathbb{Z}/q\mathbb{Z})^{\times} such that the c.f.e of j/q=[a1,,an]j/q=\left[a_{1},\dots,a_{n}\right] satisfies ai5a_{i}\leq 5 for all 1in1\leq i\leq n. Lets call such a jj, a 55-Zaremba numerator for the denominator qq. This conjecture is in direct contrast to Theorem 1.1 which implies in particular, that the set of 55-Zaremba numerators in (/q)×(\mathbb{Z}/q\mathbb{Z})^{\times} has size bounded by o(q1ϵ)o(q^{1-\epsilon}) as qq\to\infty and in particular its proportion in (/q)×(\mathbb{Z}/q\mathbb{Z})^{\times} approaches zero. Intrigued by this, we used the computer to check if one should expect the existence of 55-Zaremba prime numerators and it seems that the data suggests so. We therefore suggest:

Conjecture 1.9.

For any q>1q>1, there exists a prime residue p(/q)×p\in(\mathbb{Z}/q\mathbb{Z})^{\times} which is 55-Zaremba for qq. Moreover, for any q>1q>1 which is not one of {6,54,150}\{6,54,150\} there exists a 4-Zaremba prime numerator.

Refer to caption
Refer to caption
Figure 2. Counting the 5-Zaremba numerators and plotting the log(#numerators)log(denominator)\frac{\log(\#numerators)}{\log(denominator)}. On the left (green) all numerators, and on the right (blue) prime numerators.

The data, as can be seen in Figure 2, suggests that not only there are Zaremba numerators for any qq, but in general the number of them increases as the denominator grows. A related statement in this regard is to study the asymptotic number of Zaremba numerators for all 1qQ1\leq q\leq Q, and show it grows fast enough as a function of QQ in order for the Zaremba conjecture to hold. Indeed, without the prime restriction on the numerators, this was found by Hensley to be true [Hen89]. It would be interesting to know if it still holds with the prime numerator restriction.

One may speculate from the data in Figure 2 that the graphs converge to certain limits δ5\delta_{5} and δ5,P\delta_{5,P}, respectively, as qq\to\infty. If the limit δ5\delta_{5} exists, it easily follows from Hensley’s work that the relation δ5=2dimHE51\delta_{5}=2\dim_{H}E_{5}-1 must be satisfied, where E5E_{5} is the set of all irrationals in (0,1)(0,1) for which all digits in their c.f.e are bounded by 55. It is therefore interesting to wonder if δ5,P\delta_{5,P} is also related to the Hausdorff dimension of some set. If so, it seems reasonable to guess from our data that this dimension should be strictly smaller than dimHE5\dim_{H}E_{5}. Similar question can be asked for bounds other than 55.

1.4. Counter examples

In this section we give examples for the necessity of the assumption limqlog|Λq|logq=1\lim_{q\to\infty}\frac{\log|\Lambda_{q}|}{\log q}=1 in Theorem 1.5 and Corollary 1.7. In  §1.4.1 we show that without this assumption escape of mass can occur, and in §1.4.2 that even if there is no escape of mass, equidistribution can still fail.

1.4.1. Escape of mass

Fix any 0<α<10<\alpha<1. We consider

Λq={p(/q)×:p<qα}\Lambda_{q}=\{p\in(\mathbb{Z}/q\mathbb{Z})^{\times}:\ p<q^{\alpha}\}

and show that the measures δΛq[0,2logq]\delta_{\Lambda_{q}}^{[0,2\log q]} exhibit escape of mass. The size of these sets satisfies

limqlog|Λq|logq=α,\lim_{q\to\infty}\frac{\log|\Lambda_{q}|}{\log q}=\alpha,

which is clear for prime qq, and for general qq follows by using some inclusion-exclusion argument (as shown for example in Lemma [DS18, Lemma 2.12]).

For escape of mass, recall Mahler’s compactness theorem (c.f. §2) which says that an element gx0Xgx_{0}\in X is near the cusp if the corresponding lattice g2g\mathbb{Z}^{2} has a nontrivial short vector. More precisely, for any compact subset YXY\subset X, there is some M>1M>1 so that any gx0Ygx_{0}\in Y satisfies min0v2gv1M\underset{0\neq v\in\mathbb{Z}^{2}}{\min}\|gv\|_{\infty}\geq\frac{1}{M}.

Note that any pΛqp\in\Lambda_{q} and t(2logM,2logq(1α)2logM)t\in\big{(}2\log M,2\log q^{(1-\alpha)}-2\log M\big{)} satisfy

atup/q(01)=(et/2pqet/2)<1M,\|a_{t}u_{p/q}\begin{pmatrix}0\\ 1\end{pmatrix}\|=\|\begin{pmatrix}e^{t/2}\frac{p}{q}\\ e^{-t/2}\end{pmatrix}\|<\frac{1}{M},

i.e. the [0,2logq][0,2\log q]-trajectory of up/qx0u_{p/q}x_{0} starts with a long excursion to cusp, outside of YY, for approximately a 1α1-\alpha fraction of the total time. It follows that for any compact YXY\subset X

lim supqδΛq[0,2logq](Y)α,\limsup_{q\to\infty}\delta_{\Lambda_{q}}^{[0,2\log q]}(Y)\leq\alpha,

and therefore μ(X)α\mu(X)\leq\alpha for any weak* limit μ\mu of δΛq[0,2logq]\delta_{\Lambda_{q}}^{[0,2\log q]}, showing escape of mass and in particular preventing equidistribution.

1.4.2. Equidistribution

We show next that even if there is no escape of mass, equidistribution still may fail. For a denominator qq, consider the nn-Zaremba numerators as in  §1.3:

Λq,n={p(/q)×:all digits in the c.f.e of p/q are at most n}.\Lambda_{q,n}=\{p\in(\mathbb{Z}/q\mathbb{Z})^{\times}:\ \text{all digits in the c.f.e of $p/q$ are at most $n$}\}.

Let

Rn(Q){(p,q):pΛq,n, 1qQ}.R_{n}(Q)\coloneqq\{(p,q):\ p\in\Lambda_{q,n},\ 1\leq q\leq Q\}.

By a result of Hensely [Hen90], we have

limQ|Rn(Q)|Q2dimHEn=cn\lim_{Q\to\infty}\frac{|R_{n}(Q)|}{Q^{2\dim_{H}E_{n}}}=c_{n}

for some constant cn>0c_{n}>0, where EnE_{n} is the set of all irrationals in (0,1)(0,1) for which all digits in their c.f.e are bounded by nn. As

|Rn(Q)|=1qQ|Λq,n|Qmax1qQ|Λq,n|,|R_{n}(Q)|=\sum_{1\leq q\leq Q}|\Lambda_{q,n}|\leq Q\cdot\max_{1\leq q\leq Q}|\Lambda_{q,n}|,

it follows from Hensley’s result that there exists an increasing sequence (qi)i=1(q_{i})_{i=1}^{\infty} so that

lim infilog|Λqi,n|logqi2dimHEn1.\liminf_{i\to\infty}\frac{\log|\Lambda_{q_{i},n}|}{\log q_{i}}\geq 2\dim_{H}E_{n}-1.

It is well known [Jar29] that

limndimHEn=1.\lim_{n\to\infty}\dim_{H}E_{n}=1.

Therefore, by choosing nn large enough, the collection {Λqi,n}i=1\{\Lambda_{q_{i},n}\}_{i=1}^{\infty} can be assumed to be of logarithmic size as close to 11 as we would like.

However, due to the small digits in the c.f.e of pqi\frac{p}{q_{i}} for pΛqi,np\in\Lambda_{q_{i},n} there is some compact set which depends only on nn, on which δΛqi,n[0,2logqi]\delta_{\Lambda_{q_{i},n}}^{[0,2\log q_{i}]} is supported for all ii (c.f. the proof of Theorem 1.4 in [DS18, pg. 173]). It follows that any weak* limit of δΛqi,n[0,2logqi]\delta_{\Lambda_{q_{i}},n}^{[0,2\log q_{i}]} is a compactly supported probability measure, in particular not equal to μX\mu_{X}.

1.5. Structure of the paper

In §2 we prove Theorem 1.6, which as discussed above implies the strengthening of the equidistribution result as in Corollary 1.7. In §3 we deduce Theorems 1.2-1.4 which are the applications for the c.f.e of rationals.

Acknowledgments

We are grateful to Elon Lindenstrauss for suggesting to us to look into improving Theorem 1.5. This work has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 Research and Innovation Program, Grant agreement no. 754475. R.M. acknowledges the support by ERC 2020 grant HomDyn (grant no. 833423).

2. Non-escape of mass for divergent orbits

In this section we prove Theorem 1.6. The key input required for the proof is the relation between entropy and escape of mass, which was studied in various previous works. The G/Γ=SL2()/SL2()G/\Gamma=\operatorname{SL}_{2}(\mathbb{R})/\operatorname{SL}_{2}(\mathbb{Z}) case which we consider in this work, was studied in [ELMV12] by Einsiedler, Lindenstrauss, Michel and Venkatesh. Since then, generalizations were obtained for other rank 11 situations [EKP15, Mor22], as well as for the high rank case [EK12, KKLM17, Mor23]. We give here a brief overview of the ingredients we use and the idea of the proof.

For NN\in\mathbb{N} and η>0\eta>0, let BN,η+=n=0NanBηGanB_{N,\eta}^{+}=\bigcap_{n=0}^{N}a_{-n}B_{\eta}^{G}a_{n}, where BηGB_{\eta}^{G} is the ball in GG of radius η\eta around the identity. Define a forward Bowen (N,η)(N,\eta)-ball to be a set of the form BN,η+zB_{N,\eta}^{+}z for zG/Γz\in G/\Gamma. For Λq(/q)×\Lambda_{q}\subseteq(\mathbb{Z}/q\mathbb{Z})^{\times}, we consider the set XΛq{up/qx0:pΛq}X_{\Lambda_{q}}\coloneqq\{u_{p/q}x_{0}:\ p\in\Lambda_{q}\}. It is straightforward to show [DS18, Lemma 2.10] that the points of X(/q)×X_{(\mathbb{Z}/q\mathbb{Z})^{\times}} are well separated, in the sense that two distinct points cannot belong to the same forward Bowen (Nq,η)(N_{q},\eta)-ball, where Nq=logqN_{q}=\lfloor\log q\rfloor, assuming η\eta is small enough. Therefore, instead of studying the size of a given subset of X(/q)×X_{(\mathbb{Z}/q\mathbb{Z})^{\times}}, we study the size of its covering by balls.

Note that a-priori, an amount of ηeNq\ll_{\eta}e^{N_{q}} forward Bowen (Nq,η)(N_{q},\eta)-balls is required to cover a given compact subset of XX, as each such ball has a small diameter of ηeNq\eta e^{-N_{q}} in the unstable direction. However, the key idea that was discovered in previous works is that if we only want to cover the points which spend a large proportion of their time high up in the cusp, we can cut down the number of balls. Quantitatively, if we only consider points which spend more than κNq\kappa N_{q} time in the cusp, the required number of balls is ηe(1κ/2)Nq\ll_{\eta}e^{(1-\kappa/2)N_{q}}. We will use this idea in the proof of Theorem 1.6, where we bound the escape of mass rate, by showing that most points of XΛqX_{\Lambda_{q}} do not spend a lot of time in the cusp during the interval [0,Nq][0,N_{q}].

In our proof, we will use the version of the aforementioned covering result as given in [KKLM17]. For that, we first define the discrete average in the diagonal direction with NN steps of size tt of a measure ν\nu to be

νN,t=1Nk=0N1atkν,\nu^{N,t}=\frac{1}{N}\sum_{k=0}^{N-1}a_{tk}*\nu,

and the continuous average on the segment [T1,T2][T_{1},T_{2}]

ν[T1,T2]=1T2T1T1T2(atν)𝑑t.\nu^{[T_{1},T_{2}]}=\frac{1}{T_{2}-T_{1}}\int_{T_{1}}^{T_{2}}(a_{t}*\nu)dt.

Next, consider the distance dUd_{U} on UU induced by the absolute value on \mathbb{R}, and let BrUB_{r}^{U} be the ball in UU of radius rr centered at the identity. Given a compact subset QXQ\subset X, NN\in\mathbb{N}, κ(0,1)\kappa\in(0,1), t>0t>0, and xXx\in X, define the set of points which spend at least a fraction κ\kappa of their discrete trajectory outside QQ by

Zx(Q,N,t,κ):={uB1U:δuxN,t(XQ)κ},Z_{x}(Q,N,t,\kappa):=\left\{u\in B_{1}^{U}:\delta_{ux}^{N,t}(X\smallsetminus Q)\geq\kappa\right\},

and similarly for T>0T>0 the continuous version

Cx(Q,T,κ){uB1U:δux[0,T](XQ)κ}.C_{x}(Q,T,\kappa)\coloneqq\left\{u\in B_{1}^{U}:\delta_{ux}^{[0,T]}(X\smallsetminus Q)\geq\kappa\right\}.

Lastly, let α1:X>0\alpha_{1}:X\to\mathbb{R}_{>0} be defined by α1(x)=max{1v:vx{0}}\alpha_{1}(x)=\max\{\frac{1}{\|v\|}:v\in x\smallsetminus\{0\}\}, for \|\cdot\| the Euclidean norm, where we consider XX as the space of unimodular lattices in 2\mathbb{R}^{2}.

Then we have the following covering counting theorem based on [KKLM17, Theorem 1.5].

Theorem 2.1.

There exist t0>0t_{0}>0 such that the following holds. For any t>t0t>t_{0} there exists a compact set Q:=Q(t)Q:=Q(t) of XX such that for any NN\in\mathbb{N}, κ(0,1)\kappa\in(0,1), and xXx\in X, the set Zx(Q,N,t,κ)Z_{x}(Q,N,t,\kappa) can be covered with

α1(x)et/2(t/2)3Ne(1κ2)tN\alpha_{1}(x)e^{t/2}(t/2)^{3N}e^{(1-\frac{\kappa}{2})tN}

many balls in UU of radius et(N1)e^{-t(N-1)}.

Remark 2.2.

The original statement of [KKLM17] used slightly different notations, with ata_{t} being the geodesic flow with a doubled rate, and Zx(Q,N,t,κ)Z_{x}(Q,N,t,\kappa) defined by the requirement δatuxN,t(XQ)κ\delta_{a_{t}ux}^{N,t}(X\smallsetminus Q)\geq\kappa. However, the translation to the statement of Theorem 2.1 is straightforward as we show next.

Proof of Theorem 2.1.

Observe that

atZx(Q,N,t,κ)at{uB1U:δatuatxN,t(XQ)κ}.a_{-t}Z_{x}(Q,N,t,\kappa)a_{t}\subset\left\{u\in B_{1}^{U}:\delta_{a_{t}ua_{-t}x}^{N,t}(X\smallsetminus Q)\geq\kappa\right\}.

By [KKLM17, Theorem 1.5], the right hand side set can be covered with α1(atx)(t/2)3Ne(1κ2)tN\alpha_{1}(a_{-t}x)(t/2)^{3N}e^{(1-\frac{\kappa}{2})tN} balls in UU of radius etNe^{-tN}. Since α1(atx)et/2α1(x)\alpha_{1}(a_{-t}x)\leq e^{t/2}\alpha_{1}(x), the theorem follows by conjugating with ata_{t}. ∎

It is convenient to work with the the following standard notations for the cusp structure of XX. For any M>0M>0, denote

XM={xX:α1(x)M}andX>M={xX:α1(x)>M}.X^{\leq M}=\{x\in X:\alpha_{1}(x)\leq M\}\quad\text{and}\quad X^{>M}=\{x\in X:\alpha_{1}(x)>M\}.

While Theorem 2.1 considers discrete averages, we would like to apply it to continuous averages. For that we have the following lemma.

Lemma 2.3.
  1. (1)

    Let xXx\in X, κ(0,1)\kappa\in(0,1), M>0M>0, and T,tT,t\in\mathbb{R} with Tt>0T\geq t>0. Then

    Cx(XM,T,κ)Zx(XMet,Tt,t,κ).C_{x}(X^{\leq M},T,\kappa)\subseteq Z_{x}(X^{\leq Me^{-t}},\lfloor\frac{T}{t}\rfloor,t,\kappa).
  2. (2)

    Assume in addition that t>t0t>t_{0}, for t0t_{0} as in Theorem 2.1. Then there are M0,T0>0M_{0},T_{0}>0 which depend on tt, so that for any MM0M\geq M_{0} and TT0T\geq T_{0}, the set Cx(XM,T,κ)C_{x}(X^{\leq M},T,\kappa) can be covered by

    α1(x)exp((1κ2+3logtt)T)\alpha_{1}(x)\exp\Big{(}(1-\frac{\kappa}{2}+\frac{3\log t}{t})T\Big{)}

    many balls in UU of radius e(T2t)e^{-(T-2t)}.

Proof.

In order to prove item (1), we will show that

δxT/t,t(XMet)δx[0,T](XM),\delta_{x}^{\lfloor T/t\rfloor,t}(X^{\leq Me^{-t}})\leq\delta_{x}^{[0,T]}(X^{\leq M}),

as the desired result follows immediately from this claim and the definitions.

In order to prove the claim above, observe that for any xXx\in X,

α1(atx)et/2α1(x).\alpha_{1}(a_{t}x)\leq e^{t/2}\alpha_{1}(x).

It follows that if yXMety\in X^{\leq Me^{-t}}, then {asy:s[0,2t]}XM\{a_{s}y:\ s\in[0,2t]\}\subset X^{\leq M}. Let

I={0n<Tt:antxXMet},I=\{0\leq n<\lfloor\frac{T}{t}\rfloor:a_{nt}x\in X^{\leq Me^{-t}}\},

so that |I|T/t=δxT/t,t(XMet)\frac{|I|}{\lfloor T/t\rfloor}=\delta_{x}^{\lfloor T/t\rfloor,t}(X^{\leq Me^{-t}}). If I=I=\emptyset, then δxT/t,t(XMet)=0\delta_{x}^{\lfloor T/t\rfloor,t}(X^{\leq Me^{-t}})=0 and the claim is trivial. Otherwise, letting n0=max(I)n_{0}=\max(I) we get that

(n0nI[nt,(n+1)t))[n0t,(n0+1+{Tt})t){s[0,T]:asxXM},\left(\bigsqcup_{n_{0}\neq n\in I}\left[nt,(n+1)t\right)\right)\sqcup[n_{0}t,(n_{0}+1+\left\{\frac{T}{t}\right\})t)\subseteq\{s\in[0,T]:a_{s}x\in X^{\leq M}\},

where {y}:=yy\{y\}:=y-\lfloor y\rfloor for any yy\in\mathbb{R}. It follows that

Tδx[0,T](XM)\displaystyle T\cdot\delta_{x}^{[0,T]}(X^{\leq M}) t(|I|+{Tt})=t(Tt|I|T/t+{Tt})\displaystyle\geq t\cdot(|I|+\left\{\frac{T}{t}\right\})=t\cdot(\left\lfloor\frac{T}{t}\right\rfloor\cdot\frac{|I|}{\lfloor T/t\rfloor}+\left\{\frac{T}{t}\right\})
t(Tt+{Tt})|I|T/t=TδxT/t,t(XMet).\displaystyle\geq t\cdot(\left\lfloor\frac{T}{t}\right\rfloor+\left\{\frac{T}{t}\right\})\cdot\frac{|I|}{\lfloor T/t\rfloor}=T\cdot\delta_{x}^{\lfloor T/t\rfloor,t}(X^{\leq Me^{-t}}).

Dividing by TT proves the claim, and thus completes the proof of item 1.

To prove item 2, let Q=Q(t)Q=Q(t) be as in Theorem 2.1. Let M0=M0(t)>0M_{0}=M_{0}(t)>0 large enough so that Q(t)XM0etQ(t)\subseteq X^{\leq M_{0}e^{-t}}. Then, for any MM0M\geq M_{0}, we have by definition and item 1

(2.1) Cx(XM,T,κ)Zx(Q,Tt,t,κ).{C_{x}(X^{\leq M},T,\kappa)\subseteq Z_{x}(Q,\lfloor\frac{T}{t}\rfloor,t,\kappa).}

By Theorem 2.1, the RHS of (2.1), and hence the LHS as well, can be covered by

α1(x)et/2(t/2)3Ttexp((1κ2)tTt)α1(x)exp((1κ2+3log(t/2)t+t2T)T)\alpha_{1}(x)e^{t/2}(t/2)^{3\left\lfloor\frac{T}{t}\right\rfloor}\exp\Big{(}(1-\frac{\kappa}{2})t\left\lfloor\frac{T}{t}\right\rfloor\Big{)}\leq\alpha_{1}(x)\exp\Big{(}(1-\frac{\kappa}{2}+\frac{3\log(t/2)}{t}+\frac{t}{2T})T\Big{)}

many balls in UU of radius

et(Tt1)e(T2t).e^{-t(\lfloor\frac{T}{t}\rfloor-1)}\leq e^{-(T-2t)}.

Assuming TT is large enough compared to tt, the size of this covering is also bounded by

α1(x)exp((1κ2+3logtt)T),\alpha_{1}(x)\exp\Big{(}(1-\frac{\kappa}{2}+\frac{3\log t}{t})T\Big{)},

as desired. ∎

The previous lemma handled the covering of all points in the entire open ball B1UB_{1}^{U} (with certain restrictions on the time spent in the cusp). To prove the results of this paper we need to consider instead discrete averages over rational points, as follows. Given a set Λ(/q)×\Lambda\subseteq(\mathbb{Z}/q\mathbb{Z})^{\times}, we define

δΛ=def1|Λ|pΛδup/qx0,\delta_{\Lambda}\overset{\operatorname{def}}{=}\frac{1}{|\Lambda|}\sum_{p\in\Lambda}\delta_{u_{p/q}x_{0}},

with qq implicit in the notation (note that this definition agrees with (1.3)). Using that previous covering result, we show that the measures δΛq[0,logq]\delta_{\Lambda_{q}}^{[0,\log q]} don’t exhibit escape of mass. Extending later to time 2logq2\log q will be an easier task.

Proposition 2.4.

Let h[0,1]h\in[0,1] and Λq(/q)×\Lambda_{q}\subseteq(\mathbb{Z}/q\mathbb{Z})^{\times} such that

lim infqlog|Λq|logqh.\liminf_{q\to\infty}\frac{\log|\Lambda_{q}|}{\log q}\geq h.

Then for all ϵ>0\epsilon>0 there is a compact set XϵXX_{\epsilon}\subset X so that

lim infqδΛq[0,logq](Xϵ)2h1ϵ.\liminf_{q\to\infty}\delta_{\Lambda_{q}}^{[0,\log q]}(X_{\epsilon})\geq 2h-1-\epsilon.
Proof.

Let us start, for simplicity of notation, with a general finite subset ΘU\Theta\subseteq U and time T>0T>0, and at the end specialize to Θq={up/q:pΛq}\Theta_{q}=\{u_{p/q}:\ p\in\Lambda_{q}\} and T=log(q)T=\log(q).

We begin with a general upper bound for the ‘escaping mass’

1|Θ|uΘδux0[0,T](X>M),\frac{1}{|\Theta|}\sum_{u\in\Theta}\delta_{ux_{0}}^{[0,T]}(X^{>M}),

by separating Θ\Theta into ‘good’ and ‘bad’ points according to how much time the corresponding diagonal orbit spends in the cusp. For any κ>0\kappa>0, let

Θ<κ,M{uΘ:δux0[0,T](X>M)<κ}=ΘCx0(XM,T,κ),Θκ,M{uΘ:δux0[0,T](X>M)κ}=ΘCx0(XM,T,κ).\begin{split}\Theta^{<\kappa,M}&\coloneqq\left\{u\in\Theta:\delta_{ux_{0}}^{[0,T]}(X^{>M})<\kappa\right\}=\Theta\smallsetminus C_{x_{0}}(X^{\leq M},T,\kappa),\\ \Theta^{\geq\kappa,M}&\coloneqq\left\{u\in\Theta:\delta_{ux_{0}}^{[0,T]}(X^{>M})\geq\kappa\right\}=\Theta\cap C_{x_{0}}(X^{\leq M},T,\kappa).\end{split}

Then it follows that:

1|Θ|uΘδux0[0,T](X>M)\displaystyle\frac{1}{|\Theta|}\sum_{u\in\Theta}\delta_{ux_{0}}^{[0,T]}(X^{>M})
(2.2) =\displaystyle= 1|Θ|uΘ<κ,Mδux0[0,T](X>M)+1|Θ|uΘκ,Mδux0[0,T](X>M)\displaystyle\frac{1}{|\Theta|}\sum_{u\in\Theta^{<\kappa,M}}\delta_{ux_{0}}^{[0,T]}(X^{>M})+\frac{1}{|\Theta|}\sum_{u\in\Theta^{\geq\kappa,M}}\delta_{ux_{0}}^{[0,T]}(X^{>M})
\displaystyle\leq κ+|Θκ,M||Θ|.\displaystyle\kappa+\frac{|\Theta^{\geq\kappa,M}|}{|\Theta|}.

Hence, we want to give a lower bound to |Θ||\Theta|, which will be provided by the assumption of this proposition, and an upper bound to |Θκ,M||\Theta^{\geq\kappa,M}|, which is given by Lemma 2.3 as follows. Given t>t0t>t_{0} (which will be fixed later) we fix MM0(t)M\geq M_{0}(t) as in Lemma 2.3. Note that by definition,

Θκ,MCx(XM,T,κ).\Theta^{\geq\kappa,M}\subseteq C_{x}(X^{\leq M},T,\kappa).

Therefore, we can use Lemma 2.3 to cover Θκ,M\Theta^{\geq\kappa,M} by

exp((1κ2+3logtt)T)\exp\big{(}(1-\frac{\kappa}{2}+\frac{3\log t}{t})T\big{)}

many balls in UU of radius e(T2t)e^{-(T-2t)}, for TT large enough.

Let us now apply for Θq={up/q:pΛq}\Theta_{q}=\{u_{p/q}:\ p\in\Lambda_{q}\} and T=logqT=\log q, where Λq(/q)×\Lambda_{q}\subseteq(\mathbb{Z}/q\mathbb{Z})^{\times}. Note that each ball in our covering of Θqκ,M\Theta_{q}^{\geq\kappa,M} is of radius e(T2t)=e2tqe^{-(T-2t)}=\frac{e^{2t}}{q}, hence can contain at most 2e2t+1\lfloor 2e^{2t}\rfloor+1 points from Θq\Theta_{q}. Letting ϵ>0\epsilon>0, by the proposition’s assumption, for all qq large enough we have that |Θq|=|Λq|qhϵ/4|\Theta_{q}|=|\Lambda_{q}|\geq q^{h-\epsilon/4}. Therefore, if qq is large enough, we have |Θqκ,M||Θq|qf(t,q,κ)\frac{|\Theta_{q}^{\geq\kappa,M}|}{|\Theta_{q}|}\leq q^{f(t,q,\kappa)} for

f(t,q,κ)=(1κ2+3logtt)+log(2e2t+1)logq(hϵ4).f(t,q,\kappa)=\left(1-\frac{\kappa}{2}+\frac{3\log t}{t}\right)+\frac{\log(\lfloor 2e^{2t}\rfloor+1)}{\log q}-(h-\frac{\epsilon}{4}).

Let us now prove the proposition. Since it is trivial for h12h\leq\frac{1}{2}, we may assume that h(12,1]h\in(\frac{1}{2},1]. We can then set κϵ=2(1h)+ϵ\kappa_{\epsilon}=2(1-h)+\epsilon, and note that κϵ(0,1)\kappa_{\epsilon}\in(0,1) if ϵ\epsilon is small enough. Given such ϵ\epsilon, fix t>t0t>t_{0} large enough so that 3logtt<ϵ4\frac{3\log t}{t}<\frac{\epsilon}{4} in addition. Then

limqf(t,q,κϵ)<1h+ϵ2κϵ2=0.\lim_{q\to\infty}f(t,q,\kappa_{\epsilon})<1-h+\frac{\epsilon}{2}-\frac{\kappa_{\epsilon}}{2}=0.

All together, we obtain that

|Θqκϵ,M||Θq|qf(t,q,κϵ)q0.\frac{|\Theta_{q}^{\geq\kappa_{\epsilon},M}|}{|\Theta_{q}|}\leq q^{f(t,q,\kappa_{\epsilon})}\overset{q\to\infty}{\longrightarrow}0.

Finally, note that (2) gives in fact a lower bound for δΛq[0,logq](X>M)\delta_{\Lambda_{q}}^{[0,\log q]}(X^{>M}). Then, we deduce

lim infqδΛq[0,logq](XM)1κϵ=2h1ϵ,\liminf_{q\to\infty}\delta_{\Lambda_{q}}^{[0,\log q]}(X^{\leq M})\geq 1-\kappa_{\epsilon}=2h-1-\epsilon,

concluding the proof. ∎

We now prove Theorem 1.6. To do so, we need to move to averaging over [0,2logq][0,2\log q] from the [0,logq][0,\log q] appearing in Proposition 2.4. This is done using the following symmetry that these measures enjoy.

Lemma 2.5.

[DS18, equation (2.2) page 154] Denote by τ:XX\tau:X\to X the map that sends a lattice to its dual. For p(/q)×p\in(\mathbb{Z}/q\mathbb{Z})^{\times}, let p(/q)×p^{\prime}\in(\mathbb{Z}/q\mathbb{Z})^{\times} denote the unique integer satisfying pp=1pp^{\prime}=-1 modulo qq. Then for any qq\in\mathbb{N} and any p(/q)×p\in(\mathbb{Z}/q\mathbb{Z})^{\times},

δp/q[0,2logq]=12δp/q[0,logq]+12τδp/q[0,logq].\delta_{p/q}^{[0,2\log q]}=\frac{1}{2}\delta_{p/q}^{[0,\log q]}+\frac{1}{2}\tau_{\ast}\delta_{p^{\prime}/q}^{[0,\log q]}.
Proof of Theorem 1.6.

Let Λq(/q)×\Lambda_{q}\subseteq(\mathbb{Z}/q\mathbb{Z})^{\times} be as in the statement of the theorem. Then by Proposition 2.4, for any ϵ>0\epsilon>0 there is a compact set XϵX_{\epsilon} so that

lim infqδΛq[0,logq](Xϵ)2h1ϵ.\liminf_{q\to\infty}\delta_{\Lambda_{q}}^{[0,\log q]}(X_{\epsilon})\geq 2h-1-\epsilon.

Set Λq={p:pΛq}\Lambda_{q}^{\prime}=\{p^{\prime}:\ p\in\Lambda_{q}\}. Then |Λq|=|Λq||\Lambda_{q}^{\prime}|=|\Lambda_{q}|, hence we similarly have a compact set XϵX_{\epsilon}^{\prime} so that

lim infqτδΛq[0,logq](τ(Xϵ))=lim infqδΛq[0,logq](Xϵ)2h1ϵ.\liminf_{q\to\infty}\tau_{\ast}\delta_{\Lambda_{q}^{\prime}}^{[0,\log q]}(\tau(X_{\epsilon}^{\prime}))=\liminf_{q\to\infty}\delta_{\Lambda_{q}^{\prime}}^{[0,\log q]}(X_{\epsilon}^{\prime})\geq 2h-1-\epsilon.

By Lemma 2.5 we have that

δΛq[0,2logq]=12δΛq[0,logq]+12τδΛq[0,logq].\delta_{\Lambda_{q}}^{[0,2\log q]}=\frac{1}{2}\delta_{\Lambda_{q}}^{[0,\log q]}+\frac{1}{2}\tau_{\ast}\delta_{\Lambda_{q}^{\prime}}^{[0,\log q]}.

Let Yϵ=Xϵτ(Xϵ)Y_{\epsilon}=X_{\epsilon}\cup\tau(X_{\epsilon}^{\prime}). Putting it all together, we have

lim infqδΛq[0,2logq](Yϵ)2h1ϵ.\liminf_{q\to\infty}\delta_{\Lambda_{q}}^{[0,2\log q]}(Y_{\epsilon})\geq 2h-1-\epsilon.

Note that YϵY_{\epsilon} is compact, since XϵX_{\epsilon} is compact and τ\tau is continuous, hence it follows that any weak* limit μ\mu of δΛq[0,2logq]\delta_{\Lambda_{q}}^{[0,2\log q]} satisfies

μ(X)μ(Yϵ)2h1ϵ.\mu(X)\geq\mu(Y_{\epsilon})\geq 2h-1-\epsilon.

As ϵ>0\epsilon>0 is arbitrary, the theorem follows. ∎

3. Continued fraction expansion of rationals

In this section we deduce Theorems 1.2-1.4 from Corollary 1.7. The first result we prove is Theorem 1.3, and the rest would follow with little work. We use the tight relationship between the ata_{t}-action on XX and continued fraction expansions, and a bit of the ergodicity of μX\mu_{X} with respect to a1a_{1}. We will use the analysis carried out in [DS18, §4] which in turn relies on the presentation of the aforementioned relation as in the book [EW11].

With a slight abuse of notation, for qq\in\mathbb{N} and p(/q)×p\in(\mathbb{Z}/q\mathbb{Z})^{\times} we will let δp/q=defδup/qx0\delta_{p/q}\overset{\operatorname{def}}{=}\delta_{u_{p/q}x_{0}}, where the latter is the probability measure supported on {up/qx0}\{u_{p/q}x_{0}\} as before. We will be using the following result from [DS18, §4]:

Lemma 3.1 (Lemma 4.11 and Theorem 4.12 from [DS18]).

Let QQ\subseteq\mathbb{N} be a sequence of denominators and assume that for qQq\in Q we are given pq(/q)×p_{q}\in(\mathbb{Z}/q\mathbb{Z})^{\times} satisfying δpq/q[0,2logq]μX\delta_{p_{q}/q}^{[0,2\log q]}\to\mu_{X}. Then len(pq/q)logq12log2π2\frac{\operatorname{len}(p_{q}/q)}{\log q}\to\frac{12\log 2}{\pi^{2}} and for any finite string of integers 𝐰=(w1,,wk)k\mathbf{w}=(w_{1},\dots,w_{k})\in\mathbb{N}^{k}, D𝐰(pq/q)D𝐰D_{\mathbf{w}}(p_{q}/q)\to D_{\mathbf{w}}.

Here len(p/q)\operatorname{len}(p/q) and D𝐰D_{\mathbf{w}} are defined in (1.1) and (1.2).

Remark 3.2.

We note that in [DS18, Lemma 4.11] there is no mention of the asymptotic densities D𝐰,D𝐰(p/q)D_{\mathbf{w}},D_{\mathbf{w}}(p/q). Still, the claim as we stated it follows easily; Let I𝐰I_{\mathbf{w}} denote the subinterval of the unit interval consisting of those numbers whose c.f.ec.f.e starts with 𝐰\mathbf{w}. Then, in the notations of [DS18, Lemma 4.11], it follows from the ergodicity of νGauss\nu_{\operatorname{Gauss}} with respect to the Gauss map that D𝐰=νGauss(I𝐰)D_{\mathbf{w}}=\nu_{\operatorname{Gauss}}(I_{\mathbf{w}}) (c.f. the introduction of [AS18]), and it follows straight from the definitions that D𝐰(pq/q)=len(pq/q)len(pq/q)k+1νpq/q(I𝐰)D_{\mathbf{w}}(p_{q}/q)=\frac{\operatorname{len}(p_{q}/q)}{\operatorname{len}(p_{q}/q)-k+1}\nu_{p_{q}/q}(I_{\mathbf{w}}). Since νpq/qνGauss\nu_{p_{q}/q}\to\nu_{\operatorname{Gauss}}, we have νpq/q(I𝐰)νGauss(I𝐰)\nu_{p_{q}/q}(I_{\mathbf{w}})\to\nu_{\operatorname{Gauss}}(I_{\mathbf{w}}). As len(pq/q)logq\operatorname{len}(p_{q}/q)\asymp\log q we deduce that indeed

D𝐰(pq/q)D𝐰D_{\mathbf{w}}(p_{q}/q)\to D_{\mathbf{w}}

as claimed.

Recall that in Corollary 1.7 we showed that given the condition log|Λq|logq1\frac{\log|\Lambda_{q}|}{\log q}\to 1 we obtain that δΛq[0,2logq]wμX\delta_{\Lambda_{q}}^{[0,2\log q]}\overset{\operatorname{w}^{*}}{\longrightarrow}\mu_{X}. As can be seen in Lemma  3.1, we would like to consider measures δpq/q[0,2logq]\delta_{p_{q}/q}^{[0,2\log q]} instead. This will be done by moving from averages over Λq\Lambda_{q} to a claim about ‘almost every’ pΛqp\in\Lambda_{q}. This step is actually quite general in nature. To formulate it in a more general setting, we need the following definition.

Definition 3.3.

For a finite non-empty set Ω\Omega of probability Radon measures on XX, we denote its average by

Avg(Ω)=1|Ω|νΩν.Avg(\Omega)=\frac{1}{|\Omega|}\sum_{\nu\in\Omega}\nu.

We also define Avg()Avg(\emptyset) to be the zero measure.

Definition 3.4.

If Ωq\Omega_{q} is a sequence of sets as in Definition 3.3, we say that they are uniformly almost invariant if any possible weak* limit point of Avg(Ωq)Avg(\Omega_{q}^{\prime}) is ata_{t}-invariant, for any choice of non-empty subsets ΩqΩq\Omega_{q}^{\prime}\subseteq\Omega_{q}.

Remark 3.5.

The application we have in mind for this last definition is Ωq={δp/q[0,2logq]:pΛq}\Omega_{q}=\{\delta_{p/q}^{[0,2\log q]}:\ p\in\Lambda_{q}\}, so that Avg(Ωq)=δΛq[0,2logq]Avg(\Omega_{q})=\delta_{\Lambda_{q}}^{[0,2\log q]}. The claim that Ωq\Omega_{q} is ‘uniformly almost invariant’ follows from the fact that each measure Avg(Ωq)Avg(\Omega_{q}^{\prime}) is a diagonal continuous average of a probability measure, over a time interval of length 2logq2\log q which approaches infinity.

We prove the following proposition, which is an application of the ergodicity of the Haar measure on XX.

Proposition 3.6.

Let Ωq\Omega_{q} be a sequence of uniform almost invariant sets such that Avg(Ωq)wμXAvg(\Omega_{q})\overset{\operatorname{w}^{*}}{\longrightarrow}\mu_{X}. Then there exist ΩqΩq\Omega_{q}^{\prime}\subseteq\Omega_{q} such that |Ωq||Ωq|1\frac{|\Omega_{q}^{\prime}|}{|\Omega_{q}|}\to 1 and for any choice of a sequence νqΩq\nu_{q}\in\Omega_{q}^{\prime} we have that νqwμX\nu_{q}\overset{\operatorname{w}^{*}}{\longrightarrow}\mu_{X}.

Proof.

We first claim that the conclusion of the proposition holds if for any ϵ>0\epsilon>0 and fCc(X)f\in C_{c}(X), there exists an integer q(ϵ,f)q(\epsilon,f)\in\mathbb{N} such that for all qq(ϵ,f)q\geq q(\epsilon,f),

1|Ωq|#{νΩq:|ν(f)μX(f)|>ϵ}2ϵ.\frac{1}{|\Omega_{q}|}\#\left\{\nu\in\Omega_{q}:|\nu(f)-\mu_{X}(f)|>\epsilon\right\}\leq 2\epsilon.

Indeed, since Cc(X)C_{c}(X) is a separable Banach space, there exists a countable dense subset {fi}i\{f_{i}\}_{i\in\mathbb{N}} of Cc(X)C_{c}(X). Set q0=1q_{0}=1 and for each kk\in\mathbb{N} define

qk:=max{qk1+1,q(1/k2,fi):1ik}.q_{k}:=\max\left\{q_{k-1}+1,q(1/k^{2},f_{i}):1\leq i\leq k\right\}.

It follows from the claim that for any qqkq\geq q_{k} and i=1,,ki=1,\dots,k,

(3.1) 1|Ωq|#{νΩq:|ν(fi)μX(fi)|>1k2}2k2.{\frac{1}{|\Omega_{q}|}\#\left\{\nu\in\Omega_{q}:|\nu(f_{i})-\mu_{X}(f_{i})|>\frac{1}{k^{2}}\right\}\leq\frac{2}{k^{2}}.}

For any qkq<qk+1q_{k}\leq q<q_{k+1}, we set

Ωq={νΩq:|ν(fi)μX(fi)|1k2i=1,,k}.\Omega_{q}^{\prime}=\left\{\nu\in\Omega_{q}:|\nu(f_{i})-\mu_{X}(f_{i})|\leq\frac{1}{k^{2}}\ \forall i=1,\dots,k\right\}.

By (3.1), we have |Ωq|(12k)|Ωq||\Omega_{q}^{\prime}|\geq\left(1-\frac{2}{k}\right)|\Omega_{q}|, hence |Ωq||Ωq|1\frac{|\Omega_{q}^{\prime}|}{|\Omega_{q}|}\to 1. Note that for any choice of a sequence νqΩq\nu_{q}\in\Omega_{q}^{\prime}, we have νq(fi)qμX(fi)\nu_{q}(f_{i})\overset{q\to\infty}{\to}\mu_{X}(f_{i}) for all ii\in\mathbb{N}. Since {fi}i\{f_{i}\}_{i\in\mathbb{N}} is a dense subset of Cc(X)C_{c}(X), it follows that νqwμX\nu_{q}\overset{\operatorname{w}^{*}}{\longrightarrow}\mu_{X}. This proves the reduction.

Let us now prove the claim. Assume, for the sake of contradiction, that the claim does not hold. Then there exist ϵ>0\epsilon>0, a function fCc(X)f\in C_{c}(X), and an unbounded subset QQ\subseteq\mathbb{N} of qq’s such that for all qQq\in Q

(3.2) 1|Ωq|#{νΩq:|ν(f)μX(f)|>ϵ}>2ϵ.\frac{1}{|\Omega_{q}|}\#\left\{\nu\in\Omega_{q}:|\nu(f)-\mu_{X}(f)|>\epsilon\right\}>2\epsilon.

It follows from (3.2) that there is an infinite subsequence QQQ^{\prime}\subseteq Q such that along QQ^{\prime}, at least one of the following statements hold:

  1. (1)
    1|Ωq|#{νΩq:ν(f)>μX(f)+ϵ}>ϵ.\frac{1}{|\Omega_{q}|}\#\left\{\nu\in\Omega_{q}:\nu(f)>\mu_{X}(f)+\epsilon\right\}>\epsilon.
  2. (2)
    1|Ωq|#{νΩq:ν(f)<μX(f)ϵ}>ϵ.\frac{1}{|\Omega_{q}|}\#\left\{\nu\in\Omega_{q}:\nu(f)<\mu_{X}(f)-\epsilon\right\}>\epsilon.

Let us assume without loss of generality that option (1) holds. For qQq\in Q^{\prime}, let

Aq={νΩq:ν(f)>μX(f)+ϵ},Bq=ΩqAq.A_{q}=\left\{\nu\in\Omega_{q}:\nu(f)>\mu_{X}(f)+\epsilon\right\},\qquad B_{q}=\Omega_{q}\smallsetminus A_{q}.

Since |Aq||Ωq|(ϵ,1]\frac{|A_{q}|}{|\Omega_{q}|}\in(\epsilon,1] and |Bq||Ωq|=1|Aq||Ωq|\frac{|B_{q}|}{|\Omega_{q}|}=1-\frac{|A_{q}|}{|\Omega_{q}|} are bounded, we may take an infinite subsequence Q′′QQ^{\prime\prime}\subseteq Q^{\prime} so that |Aq||Ωq|\frac{|A_{q}|}{|\Omega_{q}|} and |Bq||Ωq|\frac{|B_{q}|}{|\Omega_{q}|} converge to some numbers α,(1α)\alpha,(1-\alpha), respectively. Note that αϵ>0\alpha\geq\epsilon>0. Next, recall that the set of all Radon measures with mass at most 11, is sequentially compact in the weak* topology. Therefore, we may assume that Q′′Q^{\prime\prime} was chosen so that Avg(Aq)wν1Avg(A_{q})\overset{\operatorname{w}^{*}}{\longrightarrow}\nu_{1} and Avg(Bq)wν2Avg(B_{q})\overset{\operatorname{w}^{*}}{\longrightarrow}\nu_{2} as qq\to\infty in Q′′Q^{\prime\prime}, for some measures ν1,ν2\nu_{1},\nu_{2}. As our sets Ωq\Omega_{q} are uniformly almost invariant, these limits ν1,ν2\nu_{1},\nu_{2} are a1a_{1}-invariant measures.

Let us now present Avg(Ωq)Avg(\Omega_{q}) as a convex combination as follows:

(3.3) Avg(Ωq)=|Aq||Ωq|Avg(Aq)+|Bq||Ωq|Avg(Bq).Avg(\Omega_{q})=\frac{|A_{q}|}{|\Omega_{q}|}Avg(A_{q})+\frac{|B_{q}|}{|\Omega_{q}|}Avg(B_{q}).

Taking the limit over the sequence Q′′Q^{\prime\prime} we conclude from (3.3), using the assumed convergence Avg(Ωq)wμXAvg(\Omega_{q})\overset{\operatorname{w}^{*}}{\longrightarrow}\mu_{X}, that

μX=αν1+(1α)ν2.\mu_{X}=\alpha\nu_{1}+(1-\alpha)\nu_{2}.

If α=1\alpha=1, then μX=ν1\mu_{X}=\nu_{1}. Otherwise both α,1α>0\alpha,1-\alpha>0, so the ergodicity of μX\mu_{X} with respect to the a1a_{1}-action also implies that μX=ν1\mu_{X}=\nu_{1}. However, in either case, by definition of AqA_{q} we have Avg(Aq)(f)>μX(f)+ϵAvg(A_{q})(f)>\mu_{X}(f)+\epsilon and hence

μX(f)=ν1(f)=limqQ′′Avg(Aq)(f)μX(f)+ϵ,\mu_{X}(f)=\nu_{1}(f)=\lim_{q\in Q^{\prime\prime}}Avg(A_{q})(f)\geq\mu_{X}(f)+\epsilon,

in contradiction. This finishes the proof of the claim and with it the proposition. ∎

We can now turn to prove the main results of this paper concerning the continued fraction expansions of rationals. We start from Theorem 1.3, from which Theorem 1.2 would then easily follow.

Proof of Theorem 1.3.

Let Λq(/q)×\Lambda_{q}\subseteq(\mathbb{Z}/q\mathbb{Z})^{\times} be as in the statement of the theorem. By Corollary 1.7, δΛq[0,2logq]wμX\delta_{\Lambda_{q}}^{[0,2\log q]}\overset{\operatorname{w}^{*}}{\longrightarrow}\mu_{X}. Hence we can apply Proposition 3.6 and find a set Λ^qΛq\hat{\Lambda}_{q}\subseteq\Lambda_{q} of size |Λ^q||Λq|1\frac{|\hat{\Lambda}_{q}|}{|\Lambda_{q}|}\to 1, so that for any choices of pqΛ^qp_{q}\in\hat{\Lambda}_{q} we have that δpq/q[0,2logq]μX\delta_{p_{q}/q}^{[0,2\log q]}\to\mu_{X}. Consequently, by Lemma 3.1, for every such sequence pqp_{q} we have that

(3.4) len(pq/q)logq12log2π2,\frac{\operatorname{len}(p_{q}/q)}{\log q}\to\frac{12\log 2}{\pi^{2}},

and for any finite string of integers 𝐰k\mathbf{w}\in\mathbb{N}^{k},

(3.5) D𝐰(pq/q)D𝐰.D_{\mathbf{w}}(p_{q}/q)\to D_{\mathbf{w}}.

Assume by way of contradiction that for some ϵ>0\epsilon>0 we have

|Λq(|len(p/q)logq12log2π2|>ϵ)0,\mathbb{P}|_{\Lambda_{q}}(|\frac{\operatorname{len}(p/q)}{\log q}-\frac{12\log 2}{\pi^{2}}|>\epsilon)\nrightarrow 0,

where pΛqp\in\Lambda_{q} is chosen uniformly at random. This means that along a subsequence QQ\subseteq\mathbb{N} we have that the set

Aq={pΛq:|len(p/q)logq12log2π2|>ϵ}A_{q}=\left\{p\in\Lambda_{q}:|\frac{\operatorname{len}(p/q)}{\log q}-\frac{12\log 2}{\pi^{2}}|>\epsilon\right\}

occupies a positive proportion of Λq\Lambda_{q} (that is |Aq|/|Λq|>η|A_{q}|/|\Lambda_{q}|>\eta for some positive η\eta), and in particular, AqΛ^qA_{q}\cap\hat{\Lambda}_{q}\neq\varnothing for all large qq. We choose pqΛ^qAqp_{q}\in\hat{\Lambda}_{q}\cap A_{q} and arrive to a contradiction arising from (3.4) and the definition of AqA_{q}.

The argument for the asymptotic frequency is identical, hence is omitted. ∎

Next, we deduce Theorem 1.2 from Theorem 1.3. We leave it to the reader to prove that the inverse deduction holds as well, so the two results are in fact equivalent.

Proof of Theorem 1.2.

We only prove item 2 concerning the length of the c.f.e. The argument for the asymptotic frequency is the same, hence is omitted. For any ϵ>0\epsilon>0 and qq\in\mathbb{N}, let

Badϵ(q)={j(/q)×:|len(j/q)logq12log2π2|>ϵ}.\text{Bad}_{\epsilon}(q)=\left\{j\in(\mathbb{Z}/q\mathbb{Z})^{\times}:\ |\frac{\operatorname{len}(j/q)}{\log q}-\frac{12\log 2}{\pi^{2}}|>\epsilon\right\}.

Note that an equivalent formulation of the theorem is the following inequality, which we now prove:

(3.6) lim supqlog(|Badϵ(q)|/φ(q))log(q)<0.{\limsup_{q\to\infty}\frac{\log\big{(}|\text{Bad}_{\epsilon}(q)|/\varphi(q)\big{)}}{\log(q)}<0.}

Since

log(φ(q))log(q)q1,\frac{\log(\varphi(q))}{\log(q)}\overset{q\to\infty}{\to}1,

(3.6) is equivalent to showing that

lim supqlog(|Badϵ(q)|)log(q)<1.\limsup_{q\to\infty}\frac{\log(|\text{Bad}_{\epsilon}(q)|)}{\log(q)}<1.

If this inequality would not hold, we could find a strictly increasing sequence (qn)n(q_{n})_{n\in\mathbb{N}} such that

log(|Badϵ(qn)|)logqn1.\frac{\log(|\text{Bad}_{\epsilon}(q_{n})|)}{\log q_{n}}{\to}1.

Define Λq(/q)×\Lambda_{q}\subseteq(\mathbb{Z}/q\mathbb{Z})^{\times} for all qq\in\mathbb{N} by

Λq={Badϵ(q)q=qn for some n(/q)×otherwise.\Lambda_{q}=\begin{cases}\text{Bad}_{\epsilon}(q)&q=q_{n}\text{ for some $n$}\\ (\mathbb{Z}/q\mathbb{Z})^{\times}&\text{otherwise}.\end{cases}

Then we can apply Theorem 1.3 and deduce that

|Λq(ΛqBadϵ(q))0 as q.\mathbb{P}|_{\Lambda_{q}}\left(\Lambda_{q}\cap\text{Bad}_{\epsilon}(q)\right)\to 0\textrm{ as }q\to\infty.

In particular,

(3.7) |Badϵ(qn)(Badϵ(qn))0 as n.{\mathbb{P}|_{\text{Bad}_{\epsilon}(q_{n})}\left(\text{Bad}_{\epsilon}(q_{n})\right)\to 0\textrm{ as }n\to\infty.}

However, the LHS in (3.7) is identically 11, in contradiction. ∎

We end up this paper with deducing Theorem 1.4, namely our application for rational numbers with prime numerators.

Proof of Theorem 1.4.

For any qq, let

Λq:={p(/q)×:p is prime}.\Lambda_{q}:=\left\{p\in(\mathbb{Z}/q\mathbb{Z})^{\times}:p\textit{ is prime}\right\}.

We will show that

(3.8) limqlog|Λq|logq=1.\lim_{q\to\infty}\frac{\log|\Lambda_{q}|}{\log q}=1.

This would allow us to apply Theorem 1.3 with this choice of Λq\Lambda_{q}, and the resulting statement is precisely Theorem 1.4.

Equation (3.8) is just a simple consequence of the prime number theorem. We include the argument for completeness. Let π(q)\pi(q) denote the number of primes q\leq q. Then

π(q)=|Λq|+#{p:p is prime, p|q}.\pi(q)=|\Lambda_{q}|+\#\left\{p:p\textit{ is prime, }p|q\right\}.

Since the number of prime divisors of qq is bounded by logq\log q we get that π(q)|Λq|+logq\pi(q)\leq|\Lambda_{q}|+\log q. In particular,

log|Λq|log(π(q)logq)=log(π(q)(1logqπ(q)))logπ(q),\log|\Lambda_{q}|\geq\log(\pi(q)-\log q)=\log(\pi(q)(1-\frac{\log q}{\pi(q)}))\sim\log\pi(q),

where the asymptotic relation follows from the fact that logqπ(q)0\frac{\log q}{\pi(q)}\to 0 by the prime number theorem. So, in order to show (3.8), it is suffices to show that logπ(q)logq1\frac{\log\pi(q)}{\log q}\to 1. Indeed, again by the prime number theorem, we know that limqπ(q)logqq=1.\displaystyle{\lim_{q\to\infty}}\frac{\pi(q)\log q}{q}=1. Taking logarithms we obtain that

limq(logπ(q)logq+loglogq)=0.\lim_{q\to\infty}\Big{(}\log\pi(q)-\log q+\log\log q\Big{)}=0.

Dividing by logq\log q we see that logπ(q)logq1\frac{\log\pi(q)}{\log q}\to 1 as needed. ∎

References

  • [AKS81] Roy Adler, Michael Keane, and Meir Smorodinsky, A construction of a normal number for the continued fraction transformation, J. Number Theory 13 (1981), no. 1, 95–105. MR 602450
  • [AS18] Menny Aka and Uri Shapira, On the evolution of continued fractions in a fixed quadratic field, J. Anal. Math. 134 (2018), no. 1, 335–397. MR 3771486
  • [Bi05] V. A. Bykovski˘i, An estimate for the dispersion of lengths of finite continued fractions, Fundam. Prikl. Mat. 11 (2005), no. 6, 15–26. MR 2204419
  • [BV05] Viviane Baladi and Brigitte Vallée, Euclidean algorithms are Gaussian, J. Number Theory 110 (2005), no. 2, 331–386. MR 2122613
  • [Dix70] John D. Dixon, The number of steps in the Euclidean algorithm, J. Number Theory 2 (1970), 414–422. MR 266889
  • [DS18] Ofir David and Uri Shapira, Equidistribution of divergent orbits and continued fraction expansion of rationals, J. Lond. Math. Soc. (2) 98 (2018), no. 1, 149–176. MR 3847236
  • [EK12] Manfred Einsiedler and Shirali Kadyrov, Entropy and escape of mass for SL3()\SL3(){\rm SL}_{3}({\mathbb{Z}})\backslash{\rm SL}_{3}({\mathbb{R}}), Israel J. Math. 190 (2012), 253–288. MR 2956241
  • [EKP15] M. Einsiedler, S. Kadyrov, and A. Pohl, Escape of mass and entropy for diagonal flows in real rank one situations, Israel J. Math. 210 (2015), no. 1, 245–295. MR 3430275
  • [ELMV12] Manfred Einsiedler, Elon Lindenstrauss, Philippe Michel, and Akshay Venkatesh, The distribution of closed geodesics on the modular surface, and Duke’s theorem, Enseign. Math. (2) 58 (2012), no. 3-4, 249–313. MR 3058601
  • [EW11] Manfred Einsiedler and Thomas Ward, Ergodic theory with a view towards number theory, Graduate Texts in Mathematics, vol. 259, Springer-Verlag London Ltd., London, 2011. MR 2723325
  • [Hei69] H. Heilbronn, On the average length of a class of finite continued fractions, Number Theory and Analysis (Papers in Honor of Edmund Landau), Plenum, New York, 1969, pp. 87–96. MR 258760
  • [Hen89] Doug Hensley, The distribution of badly approximable numbers and continuants with bounded digits, Théorie des nombres (1989), 371–385.
  • [Hen90] by same author, The distribution of badly approximable rationals and continuants with bounded digits. II, J. Number Theory 34 (1990), no. 3, 293–334. MR 1049508
  • [Hen94] by same author, The number of steps in the euclidean algorithm, J. Number Theory 49 (1994), no. 2, 142–182.
  • [Jar29] Vojtĕch Jarník, Zur metrischen theorie der diophantischen approximationen, Prace Matematyczno-Fizyczne 36 (1928-1929), no. 1, 91–106 (pol).
  • [KKLM17] S. Kadyrov, D. Kleinbock, E. Lindenstrauss, and G. A. Margulis, Singular systems of linear forms and non-escape of mass in the space of lattices, J. Anal. Math. 133 (2017), 253–277. MR 3736492
  • [Mor22] Ron Mor, Excursions to the cusps for geometrically finite hyperbolic orbifolds and equidistribution of closed geodesics in regular covers, Ergodic Theory Dynam. Systems 42 (2022), no. 12, 3745–3791. MR 4504680
  • [Mor23] by same author, Bounding entropy for one-parameter diagonal flows on SLd()/SLd()SL_{d}(\mathbb{R})/SL_{d}(\mathbb{Z}) using linear functionals, 2023, arXiv:2302.07122.
  • [Por75] J. W. Porter, On a theorem of Heilbronn, Mathematika 22 (1975), no. 1, 20–28. MR 498452
  • [Ust08] A. V. Ustinov, On the number of solutions of the congruence xyl(modq)xy\equiv l\pmod{q} under the graph of a twice continuously differentiable function, Algebra i Analiz 20 (2008), no. 5, 186–216. MR 2492364
  • [Van16] Joseph Vandehey, New normality constructions for continued fraction expansions, J. Number Theory 166 (2016), 424–451. MR 3486285