Hostname: page-component-788cddb947-m6qld Total loading time: 0 Render date: 2024-10-17T20:40:51.091Z Has data issue: false hasContentIssue false

ON THE SHORTEST DISTANCE FUNCTION IN CONTINUED FRACTIONS

Published online by Cambridge University Press:  23 June 2023

SAISAI SHI
Affiliation:
Institute of Statistics and Applied Mathematics, Anhui University of Finance and Economics, Bengbu 233030, P. R. China e-mail: saisai_shi@126.com
QINGLONG ZHOU*
Affiliation:
School of Science, Wuhan University of Technology, Wuhan 430070, P. R. China
Rights & Permissions [Opens in a new window]

Abstract

Let $x\in [0,1)$ be an irrational number and let $x=[a_{1}(x),a_{2}(x),\ldots ]$ be its continued fraction expansion with partial quotients $\{a_{n}(x): n\geq 1\}$. Given a natural number m and a vector $(x_{1},\ldots ,x_{m})\in [0,1)^{m},$ we derive the asymptotic behaviour of the shortest distance function

$$ \begin{align*} M_{n,m}(x_{1},\ldots,x_{m})=\max\{k\in \mathbb{N}: a_{i+j}(x_{1})=\cdots= a_{i+j}(x_{m}) \ \text{for}~ j=1,\ldots,k \mbox{ and some } i \mbox{ with } 0\leq i \leq n-k\}, \end{align*} $$

which represents the run-length of the longest block of the same symbol among the first n partial quotients of $(x_{1},\ldots ,x_{m}).$ We also calculate the Hausdorff dimension of the level sets and exceptional sets arising from the shortest distance function.

Type
Research Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of Australian Mathematical Publishing Association Inc.

1 Introduction

Let $T:[0,1)\to [0,1)$ be the Gauss map defined by

$$ \begin{align*} T(0)= 0, \quad T(x)= \frac{1}{x} \pmod1 \quad\text{for } x \in (0,1).\end{align*} $$

Every irrational number $x\in [0,1)$ can be uniquely expanded into an infinite form

(1.1) $$ \begin{align} x:=\frac{1}{a_{1}(x)+\dfrac{1}{a_{2}(x)+\dfrac{1}{\ddots+\dfrac{1}{a_{n}+T^{n}(x)}}}} =\frac{1}{a_{1}(x)+\dfrac{1}{a_{2}(x)+\dfrac{1}{a_{3}(x)+\dfrac{1}{\ddots}}}}, \end{align} $$

where $a_{n}(x)=\lfloor 1/T^{n-1}(x)\rfloor $ are called the partial quotients of $x.$ (Here $\lfloor \cdot \rfloor $ denotes the greatest integer less than or equal to a real number and $T^0$ denotes the identity map.) For simplicity of notation, we write (1.1) as

$$ \begin{align*} x=[a_{1}(x),a_{2}(x),\ldots,a_{n}(x)+T^{n}(x)]=[a_{1}(x),a_{2}(x),a_{3}(x),\ldots]. \end{align*} $$

It is clear that the Gauss transformation T acts as the shift map on the continued fraction system. That is, for each $x=[a_1(x),a_2(x),a_3(x),\ldots ]\in [0,1)\cap \mathbb {Q}^{c},$

$$ \begin{align*}T(x)=T([a_{1}(x),a_{2}(x),a_3(x),\ldots])=[a_{2}(x),a_{3}(x),\ldots].\end{align*} $$

Gauss observed that T is measure-preserving and ergodic with respect to the Gauss measure $\mu $ defined by

$$ \begin{align*}d\mu=\frac{1}{\log 2}\frac{1}{x+1}dx.\end{align*} $$

For more information on the continued fraction expansion, see [Reference Khintchine3].

The metrical theory of continued fractions, which concerns the properties of the partial quotients for almost all $x\in [0,1),$ is one of the major themes in the study of continued fractions. Wang and Wu [Reference Wang and Wu7] considered the metrical properties of the maximal run-length function

$$ \begin{align*}R_{n}(x)=\max\{l\in \mathbb{N}: a_{i+1}(x)=\cdots= a_{i+l}(x) \ \text{for some}\ i \mbox{ with } 0\leq i \leq n-l\},\end{align*} $$

which counts the longest run of the same symbol among the first n partial quotients of x. They proved that, for $\mu $ almost all $x\in [0,1),$

$$ \begin{align*}\lim_{n\rightarrow\infty}\frac{R_{n}(x)}{\log n}=\frac{1}{2\log({(\sqrt{5}+1)}/{2})}.\end{align*} $$

Song and Zhou [Reference Song and Zhou6] gave a more subtle characterisation of the function $R_{n}(x).$ In this paper, we continue the study by considering the shortest distance function

$$ \begin{align*} M_{n,m}(x_{1},\ldots,x_{m})=\max\{k\in \mathbb{N}: a_{i+j}(x_{1}) = & \cdots= a_{i+j}(x_{m}) \ \text{for}~ j=1,\ldots,k, \\&\ \text{and some}\ i \mbox{ with } 0\leq i \leq n-k\}. \end{align*} $$

This is motivated by the behaviour of the shortest distance between two orbits,

$$ \begin{align*}S_{n,2}(x,y)=\min_{i=0,\ldots,n-1}(d(T^{i}(x),T^{i}(y)))\end{align*} $$

in the continued fraction system. Shi et al. [Reference Shi, Tan and Zhou5] proved that, for $\mu ^{2}$ almost all $(x,y)\in [0,1)\times [0,1),$

$$ \begin{align*}{H_{2}}\cdot\lim_{n\rightarrow\infty}\frac{M_{n,2}(x,y)}{\log n}=\lim_{n\rightarrow\infty}\frac{-\log S_{n,2}(x,y)}{\log n},\end{align*} $$

where $H_{2}$ is the Rényi entropy defined by (1.2). Investigating the shortest distance between two orbits amounts to estimating the longest common substrings between two sequences of partial quotients. In fact, [Reference Shi, Tan and Zhou5] focused on the asymptotics of the length of the longest common substrings in two sequences of partial quotients.

For $n\geq 1$ and $(a_{1},\ldots ,a_{n})\in \mathbb {N}^{n}$ , we call

$$ \begin{align*}I_{n}(a_{1},\ldots,a_{n})=\{x\in[0,1): a_{1}(x)=a_{1},\ldots, a_{n}(x)=a_{n}\}\end{align*} $$

an nth cylinder. For $m\geq 2$ , we define the generalised Rényi entropy with respect to the Gauss measure $\mu $ by

(1.2) $$ \begin{align} {H}_{m}=\lim_{n\rightarrow\infty}\frac{-\log \sum_{(a_1,\ldots,a_n)\in \mathbb{N}^{n}} \mu(I_{n}(a_1,\ldots,a_n))^{m}}{(m-1)n}. \end{align} $$

The existence of the limit (1.2) for the Gauss measure $\mu $ was established in [Reference Haydn and Vaienti2].

Theorem 1.1. For $\mu ^{m}$ -almost all $(x_{1},\ldots ,x_{m})\in [0,1)^{m},$

$$ \begin{align*}\lim_{n\rightarrow\infty}\frac{M_{n,m}(x_{1},\ldots,x_{m})}{\log n}=\frac{1}{(m-1)H_{m}}.\end{align*} $$

Here we use the convention that ${1}/{0}=\infty $ and ${1}/{\infty }=0.$

It is natural to study the exceptional set in this limit theorem. We define the exceptional set as

$$ \begin{align*}\widetilde{E}=\bigg\{(x_{1},\ldots,x_{m})\in [0,1)^{m}: \liminf_{n\rightarrow\infty}\frac{M_{n,m}(x_{1},\ldots,x_{m})}{\log n} <\limsup_{n\rightarrow\infty}\frac{M_{n,m}(x_{1},\ldots,x_{m})}{\log n}\bigg\}\end{align*} $$

and the level set as

$$ \begin{align*}E(\alpha)=\bigg\{(x_{1},\ldots,x_{m})\in [0,1)^{m}: \lim_{n\rightarrow\infty}\frac{M_{n,m}(x_{1},\ldots,x_{m})}{\log n}=\alpha\bigg\}.\end{align*} $$

Throughout the paper, $\dim _{H}A$ denotes the Hausdorff dimension of the set $A.$

Theorem 1.2. For any $\alpha $ with $0\leq \alpha \leq \infty ,$ $\dim _{H}\widetilde {E}=\dim _{H}E(\alpha )=m.$

In fact, Theorem 1.2 follows immediately from the following more general result. For any $0\leq \alpha \leq \beta \leq \infty ,$ set

$$ \begin{align*} E(\alpha,\beta)=\bigg\{(x_{1},\ldots,x_{m})\in [0,1)^{m}: & \liminf_{n\rightarrow\infty}\frac{M_{n,m}(x_{1},\ldots,x_{m})}{\log n}=\alpha, \\ & \limsup_{n\rightarrow\infty}\frac{M_{n,m}(x_{1},\ldots,x_{m})}{\log n}=\beta\bigg\}. \end{align*} $$

Theorem 1.3. For any $\alpha ,\beta $ with $0\leq \alpha \leq \beta \leq \infty ,$ $\dim _{H}E(\alpha ,\beta )=m.$

2 Preliminaries

In this section, we fix some notation and recall some basic properties of continued fraction expansions. A detailed account of continued fractions can be found in Khintchine’s book [Reference Khintchine3].

For any irrational number $x\in [0,1)$ with continued fraction expansion (1.1), we denote by

$$ \begin{align*}\frac{p_{n}(x)}{ q_{n}(x)} :=[a_{1}(x),\ldots,a_{n}(x)]\end{align*} $$

the nth convergent of $x.$ With the conventions

$$ \begin{align*}p_{-1}(x)=1, \quad q_{-1}(x)=0, \quad p_{0}(x)=0, \quad q_{0}(x)=1,\end{align*} $$

we have, for any $n\ge 0,$

$$ \begin{align*}p_{n+1}(x)=a_{n+1}(x) p_{n}(x)+p_{n-1}(x), \quad q_{n+1}(x)=a_{n+1}(x) q_{n}(x)+q_{n-1}(x).\end{align*} $$

Obviously, $q_{n}(x)$ is determined by the first n partial quotients $a_{1}(x),\ldots ,a_{n}(x).$ So we also write $q_{n}(a_{1}(x),\ldots ,a_{n}(x))$ in place of $q_{n}(x).$ If no confusion is likely to arise, we write $a_{n}$ and $q_{n}$ in place of $a_{n}(x)$ and $q_{n}(x),$ respectively.

Proposition 2.1 [Reference Khintchine3].

For $n\geq 1$ and $(a_{1},\ldots ,a_{n})\in \mathbb {N}^{n}$ :

  1. (1) $ q_n\ge 2^{(n-1)/2}$ and

    $$ \begin{align*} \prod^{n}_{k=1}a_{k}\leq q_n \leq \prod^{n}_{k=1}(a_{k}+1)\leq 2^{n}\prod^{n}_{k=1}a_{k};\end{align*} $$
  2. (2) the length of $I_{n}(a_{1},\ldots ,a_{n})$ satisfies

    $$ \begin{align*} \frac{1}{2q_n^2} \leq |I_{n}(a_{1},\ldots,a_{n})|=\frac{1}{(q_n+q_{n-1})q_n} \leq \frac{1}{q_n^2}. \end{align*} $$

The following $\psi $ -mixing property is essential in proving Theorem 1.1.

Lemma 2.2 [Reference Philipp4].

For any $k\geq 1,$ let $\mathbb {B}^{k}_{1}=\sigma (a_{1},\ldots ,a_{k})$ and let $\mathbb {B}^{\infty }_{k}=\sigma (a_{k},a_{k+1},\ldots )$ denote the $\sigma $ -algebras generated by the random variables $(a_{1},\ldots ,a_{k})$ and $(a_{k},a_{k+1},\ldots )$ respectively. Then, for any $E\in \mathbb {B}^{k}_{1}$ and $F\in \mathbb {B}^{\infty }_{k+n},$

$$ \begin{align*}\mu(E\cap F)=\mu(E)\cdot \mu(F)(1+\theta \rho^n),\end{align*} $$

where $|\theta |\leq K, \rho <1$ and $K, \rho $ are positive constants independent of $E, F, n$ and $k.$

To estimate the measure of a limsup set in a probability space, the following lemma is widely used.

Lemma 2.3 (Borel–Cantelli lemma).

Let $(\Omega , \mathcal {B}, \nu )$ be a finite measure space and let $\{A_n\}_{n\ge 1}$ be a sequence of measurable sets. Define $A=\bigcap _{N=1}^{\infty }\bigcup _{n=N}^{\infty }A_n.$ Then

$$ \begin{align*} \nu(A)=\begin{cases} 0 & \text{if } \displaystyle\sum_{n= 1}^{\infty}\nu(A_n)<\infty, \\[3pt] \nu(\Omega) & \displaystyle \text{if }\sum_{n= 1}^{\infty}\nu(A_n)=\infty \text{ and } \{A_n\}_{n\ge 1}\text{ are pairwise independent}. \end{cases} \end{align*} $$

Let $\mathbb {K}=\{k_{n}\}_{n\geq 1}$ be a subsequence of $\mathbb {N}$ that is not cofinite. Define a mapping $\phi _{\mathbb {K}}:[0,1)\cap \mathbb {Q}^{c}\rightarrow [0,1)\cap \mathbb {Q}^{c}$ as follows. For each $x=[a_1,a_2,\ldots ]\in [0,1)\cap \mathbb {Q}^{c},$ put $\phi _{\mathbb {K}}(x)=\overline {x}=[c_1,c_2,\ldots ],$ where $[c_1,c_2,\ldots ]$ is obtained by eliminating all the terms $a_{k_{n}}$ from the sequence $a_1, a_2,\ldots .$ Let $\{b_{n}\}_{n\geq 1}$ be a sequence with $b_{n}\in \mathbb {N},n\geq 1.$ Write

$$ \begin{align*}E(\mathbb{K},\{b_{n}\})=\{x\in[0,1)\cap \mathbb{Q}^{c}: a_{k_{n}}(x)= b_{n} \ \text{for all}~ n\geq1\}.\end{align*} $$

Lemma 2.4 [Reference Song and Zhou6].

Assume that $\{b_{n}\}_{n\geq 1}$ is bounded. If the sequence $\mathbb {K}$ is of density zero in $\mathbb {N},$ that is,

$$ \begin{align*}\lim_{n\rightarrow\infty}\frac{\sharp\{i\leq n: i\in\mathbb{K}\}}{n}=0,\end{align*} $$

where $\sharp $ denotes the number of elements in a set, then

$$ \begin{align*}\dim_{H}E(\mathbb{K},\{b_{n}\})=\dim_{H}\phi_{\mathbb{K}}(E(\mathbb{K},\{b_{n}\}))=1.\end{align*} $$

We close this section by citing Marstrand’s product theorem.

Lemma 2.5 [Reference Falconer1].

If $E,F\subset \mathbb {R}^{d}$ for some $d,$ then $\dim _{H}(E\times F)\geq \dim _{H}E+\dim _{H}F.$

3 Proof of Theorem 1.1

Theorem 1.1 can be proved from the following two propositions.

Proposition 3.1. For $\mu ^{m}$ -almost all $(x_{1},\ldots ,x_{m})\in [0,1)^{m},$

$$ \begin{align*}\limsup_{n\rightarrow\infty}\frac{M_{n,m}(x_{1},\ldots,x_{m})}{\log n}\leq\frac{1}{(m-1)H_{m}}.\end{align*} $$

Proof. We can assume that $H_{m}>0$ (the case $H_m=0$ is obvious). Fix $s_1< s_2< (m-1)H_{m}.$ By the definition of the $H_{m},$

(3.1) $$ \begin{align} \sum_{(a_{1},\ldots,a_{n})\in\mathbb{N}^{n}}\mu(I_{n}(a_{1},\ldots,a_{n}))^{m}<\exp\bigg\{-\frac{s_1+s_2}{2}n\bigg\} \end{align} $$

for sufficiently large n. Set $u_{n}=\lfloor {\log n}/{s_1}\rfloor .$ Note that, for any $(x_{1},\ldots ,x_{m})\in [0,1)^{m}$ with $M_{n,m}(x_{1},\ldots ,x_{m})=k$ , there exists i with $0\leq i\leq n-k$ such that

$$ \begin{align*} a_{i+j}(x_{1})=\cdots=a_{i+j}(x_{m})\end{align*} $$

for $j=1,\ldots ,k.$ We deduce

$$ \begin{align*} &\mu^{m}(\{(x_{1},\ldots,x_{m})\in [0,1)^{m}: M_{n,m}(x_{1},\ldots,x_{m})>u_{n}\})\\ &\quad =\sum_{k=u_{n}+1}^{\infty}\mu^{m}(\{(x_{1},\ldots,x_{m})\in [0,1)^{m}: M_{n,m}(x_{1},\ldots,x_{m})=k\})\\ &\quad \leq\sum_{k=u_{n}+1}^{\infty}\sum_{i=0}^{n-k}\mu^{m}(\{(x_{1},\ldots,x_{m})\in [0,1)^{m}: a_{i+j}(x_{1})=\cdots=a_{i+j}(x_{m}),\ j=1,\ldots,k\}). \end{align*} $$

By the invariance of $\mu $ under $T,$ it follows that

$$ \begin{align*} &\mu^{m}(\{(x_{1},\ldots,x_{m})\in [0,1)^{m}: M_{n,m}(x_{1},\ldots,x_{m})>u_{n}\})\\&\quad\leq n \sum_{k=u_{n}+1}^{\infty}\mu^{m}(\{(x_{1},\ldots,x_{m})\in [0,1)^{m}: a_{j}(x_{1})=\cdots=a_{j}(x_{m}),\ j=1,\ldots,k\})\\&\quad= n\sum_{k=u_{n}+1}^{\infty}\sum_{(a_{1},\ldots,a_{k})\in \mathbb{N}^{k}}\mu(I_{k}(a_{1},\ldots,a_{k}))^{m}\\&\quad\leq Cn\cdot\exp\bigg\{-\frac{s_1+s_2}{2}(u_{n}+1)\bigg\} \qquad (\text{by } (3.1))\\&\quad\leq Cn^{-{(s_2-s_1)}/{2s_1}}, \end{align*} $$

where $C=\sum _{k=1}^{\infty }\exp \{-{(s_1+s_2)k}/{2}\}.$ Choose an infinite subsequence of integers $\{n_{k}\}_{k\geq 1},$ where $n_{k}=k^{L}$ and $L\cdot {(s_2-s_1)}/{2s_1}>1.$ Then

$$ \begin{align*} \sum_{k=1}^{\infty}\mu^{m}(\{(x_{1},\ldots,x_{m})\in [0,1)^{m}: M_{n_{k},m}(x_{1},\ldots,x_{m})>u_{n_{k}}\})<\infty. \end{align*} $$

From the Borel–Cantelli Lemma 2.3, for almost all $(x_{1},\ldots ,x_{m})\in [0,1)^{m},$

$$ \begin{align*}M_{n_{k},m}(x_{1},\ldots,x_{m})\leq u_{n_{k}} \end{align*} $$

for sufficiently large k. Thus,

$$ \begin{align*} \limsup_{n\rightarrow\infty}\frac{M_{n,m}(x_{1},\ldots,x_{m})}{\log n} \leq&\limsup_{k\rightarrow\infty}\frac{M_{n_{k+1},m}(x_{1},\ldots,x_{m})}{\log n_{k}}\\ \leq&\limsup_{k\rightarrow\infty}\frac{M_{n_{k+1},m}(x_{1},\ldots,x_{m})}{\log n_{k+1}}\cdot\limsup_{k\rightarrow\infty}\frac{n_{k+1}}{n_{k}} \leq\frac{1}{s_{1}}. \end{align*} $$

Therefore, by the arbitrariness of $s_{1},$

$$ \begin{align*}\limsup_{n\rightarrow\infty}\frac{M_{n,m}(x_{1},\ldots,x_{m})}{\log n}\leq\frac{1}{(m-1)H_{m}}.\end{align*} $$

This completes the proof.

Proposition 3.2. For $\mu ^{m}$ -almost all $(x_{1},\ldots ,x_{m})\in [0,1)^{m},$

$$ \begin{align*}\liminf_{n\rightarrow\infty}\frac{M_{n,m}(x_{1},\ldots,x_{m})}{\log n}\geq\frac{1}{(m-1)H_{m}}.\end{align*} $$

Proof. We can assume that $H_{m}<\infty $ (the case $H_m=\infty $ is obvious). For $1\le d<n$ , set

$$ \begin{align*} M_{[d,n]}(x_{1},\ldots,x_{m})=\max\{k\in \mathbb{N}: a_{i+j}(x_{1}) & =\cdots= a_{i+j}(x_{m}) \ \text{for}~ j=1,\ldots,k, \\&\ \text{and for some } i \mbox{ with } d-1\leq i \leq n-k\}. \end{align*} $$

We denote $\{(x_{1},\ldots ,x_{m})\in [0,1)^{m}: M_{n,m}(x_{1},\ldots ,x_{m})<k\}$ by $\{M_{n,m}<k\}$ for brevity.

For any $s>(m-1)H_{m},$ by the definition of the $H_{m},$

(3.2) $$ \begin{align} \sum_{(a_{1},\ldots,a_{n}) \in\mathbb{N}^{n}}\mu(I_{n}(a_{1},\ldots,a_{n}))^{m}>\exp\bigg\{-\frac{s+(m-1)H_{m}}{2}n\bigg\} \end{align} $$

for sufficiently large n. Let $u_{n}=\lfloor {\log n}/{s}\rfloor $ and $l_{n}=\lfloor {n}/{u_{n}^{2}}\rfloor .$ Then

$$ \begin{align*} \begin{aligned} \{M_{n,m}<u_{n}\}&\subset \{M_{[iu_{n}^{2}+1,iu_{n}^{2}+u_{n}]}<u_{n} : 0\leq i <l_{n}\}\\& \subset\{M_{u_{n},m}<u_{n}\}\cap(\underbrace{T\times\cdots \times T}_{m \text{ times }})^{-u_{n}^{2}}\{M_{[iu_{n}^{2}+1,i\mu_{n}^{2}+u_{n}]}<u_{n} : 0\leq i <l_{n}-1\}. \end{aligned} \end{align*} $$

By Lemma 2.2, it follows that

$$ \begin{align*} & \mu^{m}(\{M_{n,m}<u_{n}\}) \\& \quad \leq \mu^{m}(\{M_{u_{n},m}<u_{n}\}\cap(\underbrace{T\times\cdots \times T}_{m \text{ times }})^{-u_{n}^{2}}\{M_{[iu_{n}^{2}+1,i\mu_{n}^{2}+u_{n}]}<u_{n}, 0\leq i <l_{n}-1\})\\& \quad \leq\mu^{m}(\{M_{u_{n},m}<u_{n}\})^{l_{n}}(1+\theta \rho^{u_{n}^{2}-u_{n}})^{m\cdot l_{n}}\\& \quad \leq\bigg(1-\sum_{(a_{1},\ldots,a_{u_{n}})\in \mathbb{N}^{u_{n}}} \mu(I_{u_{n}}(a_{1},\ldots,a_{u_{n}}))^{m}\bigg)^{l_{n}}(1+\theta \rho^{u_{n}^{2}-u_{n}})^{m\cdot l_{n}}\\& \quad \leq\exp\bigg\{-\frac{n}{u_{n}^{2}}\cdot n^{-{(s+(m-1)H_{m})}/{2s}}\bigg\}\cdot\exp\bigg\{\theta \rho^{u_{n}^{2}-u_{n}}\frac{m\cdot n}{u_{n}^{2}}\bigg\}\\& \quad \leq M\exp\bigg\{-\frac{n^{{(s-(m-1)H_{m})}/{2s}}}{u_{n}^{2}}\bigg\}, \end{align*} $$

where the penultimate inequality follows from (3.2) and the two facts $ (1-x)<\exp (-x)$ for $0<x<1$ and $\lim _{n\to \infty }(1+1/n)^{n}=e$ , and the last inequality follows because $\theta \rho ^{u_{n}^{2}-u_{n}}{m\cdot n}/{u_{n}^{2}}\rightarrow 0$ as $n\rightarrow \infty .$ Thus,

$$ \begin{align*} \sum_{n=1}^{\infty}\mu^{m}(\{M_{n,m}<u_{n}\})<\infty. \end{align*} $$

From the Borel–Cantelli Lemma, for $\mu ^{m}$ -almost all $(x_{1},\ldots ,x_{m})\in [0,1)^{m},$

$$ \begin{align*}\liminf_{n\rightarrow\infty}\frac{M_{n,m}(x_{1},\ldots,x_{m})}{\log n}\geq\frac{1}{(m-1)H_{m}}.\end{align*} $$

This completes the proof of Proposition 3.2 and of Theorem 1.1.

4 Proof of Theorem 1.3

This section is devoted to the proof of Theorem 1.3. Our strategy is to construct Cantor-like subsets with full Hausdorff dimension. The proof is divided into several cases according to the values of $\alpha $ and $\beta .$ We give a detailed proof for the case $0<\alpha <\beta <\infty $ and a sketch of the proof for the remaining cases.

Case 1: $0<\alpha <\beta <\infty .$

Choose two positive integer sequences $\{n_{k}\}_{k\geq 1}$ and $\{s_{k}\}_{k\geq 1}$ such that, for each $k\geq 1,$

(4.1) $$ \begin{align} n_{1}=2,\quad n_{k+1}=\lfloor n_{k}^{{\beta}/{\alpha}}\rfloor, \quad s_{k}=\lfloor \,\beta \log n_{k}\rfloor. \end{align} $$

We readily check that

(4.2) $$ \begin{align} \lim_{k\rightarrow\infty}\frac{s_{k}}{n_{k+1}-n_{k}}=0. \end{align} $$

Without loss of generality, we assume that $n_{k+1}-n_{k}>s_{k}$ for all $k\geq 1.$ Otherwise, we consider only sufficiently large k. Put

$$ \begin{align*}n_{k+1}-n_{k}=s_{k}\cdot\iota_{k}+\theta_{k},\end{align*} $$

where

$$ \begin{align*}\iota_{k}=\bigg\lfloor\frac{n_{k+1}-n_{k}}{s_{k}}\bigg\rfloor \quad \text{for } 0\leq\theta_{k}<s_{k}.\end{align*} $$

Define a marked set $\mathbb {K}$ of positive integers by

$$ \begin{align*}\mathbb{K}:=\mathbb{K}(\{n_{k}\}, \{s_{k}\})=\bigcup_{k\ge1}\{n_{k},n_{k}+1,n_{k}+2,\ldots,n_{k}+s_{k}, n_{k}+2s_{k},n_{k}+3s_{k},\ldots, n_{k}+\iota_ks_{k}\}.\end{align*} $$

Now we define m sequences as follows.

  • For $i=1,$

    $$ \begin{align*}a_{n_{k}}^{(1)}=1,\ a_{n_{k}+1}^{(1)}=\cdots=a_{n_{k}+s_{k}-1}^{(1)}=1,\ a_{n_{k}+s_{k}}^{(1)}=a_{n_{k}+2s_{k}}^{(1)}= \cdots=a_{n_{k}+\iota_ks_{k}}^{(1)}=1.\end{align*} $$
  • For $2\leq i\leq m,$

    $$ \begin{align*}a_{n_{k}}^{(i)}=i,\ a_{n_{k}+1}^{(i)}=\cdots=a_{n_{k}+s_{k}-1}^{(i)}=1,\ a_{n_{k}+s_{k}}^{(i)}=a_{n_{k}+2s_{k}}^{(i)}= \cdots=a_{n_{k}+\iota_ks_{k}}^{(i)}=i.\end{align*} $$

Then, for $i=1,2,\ldots ,m,$ write

$$ \begin{align*}E(\mathbb{K},\{a_{n}^{(i)}\}_{n\geq1})=\{x\in[0,1)\cap \mathbb{Q}^{c}: a_{n}(x)= a_{n}^{(i)} \ \text{for all}~ n\in\mathbb{K}\}.\end{align*} $$

Now we prove $\prod _{i=1}^{m}E(\mathbb {K},\{a_{n}^{(i)}\}_{n\geq 1})\subset E(\alpha ,\beta ).$ Fix $(x_{1},\ldots ,x_{m})\in \prod _{i=1}^{m}E(\mathbb {K},\{a_{n}^{(i)}\}_{n\geq 1})$ for any $n\geq n_{1}$ and let k be the integer such that $n_{k}\leq n< n_{k+1}.$ From the construction of the set $\prod _{i=1}^{m}E(\mathbb {K},\{a_{n}^{(i)}\}_{n\geq 1})$ , we see that

$$ \begin{align*} M_{n,m}(x_{1},\ldots,x_{m})=\begin{cases} s_{k-1}-1 & \text{if}\ n_{k}\leq n<n_{k}+s_{k-1},\\ n-n_{k} & \text{if}\ n_{k}+s_{k-1}\leq n<n_{k}+s_{k},\\ s_{k}-1 & \text{if}\ n_{k}+s_{k}\leq n<n_{k+1}. \end{cases} \end{align*} $$

Further, by (4.1), we deduce that

$$ \begin{align*} \begin{aligned} \liminf_{n\rightarrow\infty}\frac{M_{n,m}(x_{1},\ldots,x_{m})}{\log n}&= \liminf_{k\rightarrow\infty} \min\bigg\{\frac{M_{n_{k}+s_{k-1}-1,m}(x_{1},\ldots,x_{m})}{\log(n_{k}+s_{k-1}-1)},\frac{M_{n_{k+1}-1,m}(x_{1},\ldots,x_{m})}{\log(n_{k+1}-1)}\bigg\}\\& =\liminf_{k\rightarrow\infty} \min\bigg\{\frac{s_{k-1}-1}{\log(n_{k}+s_{k-1}-1)},\frac{s_{k}-1}{\log(n_{k+1}-1)}\bigg\}\\& =\alpha \end{aligned} \end{align*} $$

and

$$ \begin{align*} \begin{aligned} \limsup_{n\rightarrow\infty}\frac{M_{n,m}(x_{1},\ldots,x_{m})}{\log n}&= \limsup_{k\rightarrow\infty} \max\bigg\{\frac{M_{n_{k},m}(x_{1},\ldots,x_{m})}{\log(n_{k})},\frac{M_{n_{k}+s_{k}-1,m}(x_{1},\ldots,x_{m})}{\log(n_{k}+s_{k}-1)}\bigg\}\\& =\limsup_{k\rightarrow\infty} \max\bigg\{\frac{s_{k-1}-1}{\log(n_{k})},\frac{s_{k}-1}{\log(n_{k}+s_{k}-1)}\bigg\}\\& =\beta. \end{aligned} \end{align*} $$

Hence, $(x_{1},\ldots ,x_{m})\in E(\alpha ,\beta ).$

It remains to prove that the density of $\mathbb {K}\subset \mathbb {N}$ is zero. For $n_{k}\leq n<n_{k+1}$ with some $k\geq 1{:}$

  • if $n_{k}\leq n< n_{k}+s_{k},$ then $\sharp \{i\leq n : i\in \mathbb {K}\}=\sum _{j=1}^{k-1}(m_{j}+\iota _{j})+n-n_{k}+1;$

  • if $n_{k}+ls_{k}\leq n<n_{k}+(l+1)s_{k}$ for some $l$ with $0< l<\iota _{k},$ then we see that ${\sharp \{i\leq n : i\in \mathbb {K}\}=\sum _{j=1}^{k-1} (s_{j}+\iota _{j})+s_{k}+l};$

  • if $n_{k}+\iota _{k}s_{k}\leq n<n_{k+1},$ then $\sharp \{i\leq n : i\in \mathbb {K}\}=\sum _{j=1}^{k}(s_{j}+\iota _{j}).$

Consequently,

$$ \begin{align*} \begin{aligned} \limsup_{n\rightarrow\infty}\frac{\sharp\{i\leq n : i\in\mathbb{K}\}}{n}&\leq \limsup_{k\rightarrow\infty} \max_{0\leq l<\iota_{k}}\bigg\{\frac{\sum_{j=1}^{k-1}(s_{j}+\iota_{j})+s_{k}+l}{n_{k}+ls_{k}}\bigg\}\\& \leq\limsup_{k\rightarrow\infty}\bigg\{\frac{\sum_{j=1}^{k-1}(s_{j}+\iota_{j})+s_{k}+\iota_{k})}{n_{k}}\bigg\}\\& =0, \end{aligned} \end{align*} $$

where the last equality follows by the Stolz–Cesàro theorem and (4.2). By Lemmas 2.4 and 2.5,

$$ \begin{align*}\dim_{H}E_{\alpha,\beta}\geq\dim_{H}\bigg(\prod\limits_{i=1}^{m}E(\mathbb{K},\{a_{n}^{(i)}\}_{n\geq1})\bigg)\geq\sum_{i=1}^{m}\dim_{H}E(\mathbb{K},\{a_{n}^{(i)}\}_{n\geq1})=m.\end{align*} $$

Similar arguments apply to the remaining cases. We only give the constructions for the proper sequences $\{n_k\}_{k\geq 1}$ and $\{s_k\}_{k\geq 1}.$

Case 2: $0<\alpha =\beta <\infty .$ Take $n_{k}=2^{k}$ and $s_{k}=\lfloor \alpha \log n_{k}\rfloor \text { for } k\geq 1.$

Case 3: $\alpha =0<\beta <\infty .$ Take $n_{k}=2^{2^{2^{k}}}$ and $s_{k}=\lfloor \,\beta \log n_{k}\rfloor \text { for } k\geq 1.$

Case 4: $\alpha =0, \beta =\infty .$ Take $n_{k}=2^{2^{2^{k}}}$ and $s_{k}=\lfloor k\log n_{k}\rfloor ~\text { for } k\geq 1.$

Case 5: $0<\alpha <\beta =\infty .$ Take $n_{k}=2^{k!}$ and $s_{k}=\lfloor \alpha k\log n_{k}\rfloor \text { for } k\geq 1.$

Case 6: $\alpha =\beta =0.$ Take $n_{k}=2^{k}$ and $s_{k}=\lfloor \log \log n_{k}\rfloor \text { for } k\geq 1.$

Case 7: $\alpha =\beta =\infty .$ Take $n_{k}=2^{k}$ and $s_{k}=\lfloor k\log n_{k}\rfloor \text { for } k\geq 1.$

Footnotes

This work is supported by National Natural Science Foundation of China (NSFC), No. 12201476.

References

Falconer, K. J., Fractal Geometry: Mathematical Foundations and Applications, 3rd edn (Wiley, Chichester, 2014).Google Scholar
Haydn, N. and Vaienti, S., ‘The Rényi entropy function and the large deviation of short return times’, Ergod. Th. & Dynam. Sys. 30(1) (2010), 159179.CrossRefGoogle Scholar
Khintchine, A. Y., Continued Fractions (Noordhoff, Groningen, 1963), translated by Peter Wynn.Google Scholar
Philipp, W., ‘Some metrical theorems in number theory’, Pacific J. Math. 20 (1967), 109127.CrossRefGoogle Scholar
Shi, S. S., Tan, B. and Zhou, Q. L., ‘Best approximation of orbits in iterated function systems’, Discrete Contin. Dyn. Syst. 41(9) (2021), 40854104.Google Scholar
Song, T. and Zhou, Q. L., ‘On the longest block function in continued fractions’, Bull. Aust. Math. Soc. 102(2) (2020), 196206.CrossRefGoogle Scholar
Wang, B. W. and Wu, J., ‘On the maximal run-length function in continued fractions’, Ann. Univ. Sci. Budapest. Sect. Comput. 34 (2011), 247268.Google Scholar