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

On a continued fraction algorithm in finite extensions of p\mathbb{Q}_{p} and its metrical theory

Manoj Choudhuri and Prashant J. Makadiya Department of Basic Sciences, Institute of Infrastructure, Technology, Research and Management, Near Khokhara Circle, Maninagar (East), Ahmedabad 380026, Gujarat, India. [email protected] [email protected]
Abstract.

We develop a continued fraction algorithm in finite extensions of p\mathbb{Q}_{p} generalising certain algorithms in p\mathbb{Q}_{p}, and prove the finiteness property for certain small degree extensions. We also discuss the metrical properties of the associated continued fraction maps for our algorithms using subsequence ergodic theory and moving averages.

Mathematics Subject Classification, Primary: 1111J7070, 1111J8383; Secondary: 1111J6161, 3737A4444.

Keywords: Continued fractions, finite extensions of p\mathbb{Q}_{p}, continued fraction map, exactness, metric theory.

1. Introduction

Being an indispensable tool in number theory, especially in Diophantine approximation, the study of continued fractions has attracted many mathematicians over the years. The simple or classical continued fraction expansion of a real number α\alpha is an expression of the form

(1) α=a0+1a1+1a2+1a3+,\displaystyle\alpha=a_{0}+\frac{1}{\displaystyle{a_{1}+\frac{1}{\displaystyle{a_{2}+\frac{1}{\displaystyle{a_{3}+_{\ddots}}}}}}},

which is also written as α=[a0;a1,a2,]\alpha=[a_{0};a_{1},a_{2},\dots] with aia_{i}’s being natural numbers and ai>0a_{i}>0 for i1i\geq 1 (see [15] or [19] for more details). Here, aia_{i}’s are called the partial quotients of the continued fraction expansion of α\alpha. If AnBn=[a0;a1,,an]\frac{A_{n}}{B_{n}}=[a_{0};a_{1},\dots,a_{n}], then the rational numbers AnBn\frac{A_{n}}{B_{n}} converges to α\alpha, and AnBn\frac{A_{n}}{B_{n}} is called the nnth convergent to the continued fraction of α\alpha. The classical continued fraction for real numbers has nice arithmetical properties such as rational numbers have finite continued fraction expansion; convergents are the best approximants among other rational numbers; quadratic irrationals have periodic continued fraction expansion expansions and vice versa, this fact is known as Lagrange’s theorem (see [19] for more details). The reader may look at [17] for various real continued fractions apart from the classical (simple) one. The starting of continued fraction theory for complex numbers goes back to 18871887 when A. Hurwitz ([16]) described the nearest integer continued fraction algorithm in the field of complex numbers, the partial quotients being elements of the ring of Gaussian integers. He also proved a version of lagrange’s theorem as well. See [11], [10] for more recent developments and a general approach to continued fraction theory in this setup.

It is quite natural to study continued fractions in the non-Archimedean setup as well. The reader is referred to [32] for a comprehensive introduction to the theory of continued fraction and its relation to Diophantine approximation in positive characteristics. See also the survey article [22] by Lasjaunias. For continued fraction in the field of Laurent series in one indeterminate over a finite field 𝔽q\mathbb{F}_{q}, viz. 𝔽q((X1))\mathbb{F}_{q}((X^{-1})), there is a natural choice of a set for the set of partial quotients, viz. the polynomial ring 𝔽q[X]\mathbb{F}_{q}[X]. The continued fraction in this setup is very well-behaved. For example, any element in 𝔽q(X)\mathbb{F}_{q}(X) has a finite continued fraction expansion, the convergents (which naturally belong to 𝔽q(X)\mathbb{F}_{q}(X)) provide the best approximation, in fact it is true that if α𝔽q((X1))\alpha\in\mathbb{F}_{q}((X^{-1})) and PQ𝔽q(X)\frac{P}{Q}\in\mathbb{F}_{q}(X) is such that |αPQ|<1|Q|2\left|\alpha-\frac{P}{Q}\right|<\frac{1}{|Q|^{2}}, then PQ\frac{P}{Q} is a convergent from the continued fraction expansion of α\alpha. A version of Lagrange’s theorem is true as well in this setup, see [32] or [22] for more details.

In 19401940, Mahler ([25]) initiated the study of continued fractions in the field of p-adic numbers. There are mainly two types of continued fractions in the field of pp-adic numbers, one was introduced by Schneider ([33]) in 19681968, and the other was introduced by Ruban [31] in 19701970. Rational numbers need not have finite continued fraction expansion with respect to these algorithms. See [5] for rational numbers having infinite expansion with respect to Schneider’s algorithm. In fact, Wang [34] and Laohakosol [21] independently showed that a pp-adic number α\alpha is rational if and only if the Ruban continued fraction expansion of α\alpha is either finite or periodic. In 19781978, Browkin ([3]) modified Ruban’s algorithm and proved that every rational number has a finite continued fraction expansion. Another desirable arithmetical property of any continued fraction is periodic expansion (the periodicity property) of quadratic irrational which is known as Lagrange’s theorem in the case of real numbers. In [4], Browkin modified his algorithm further and showed that m\sqrt{m} has periodic continued fraction expansion for certain positive integers for p=5p=5. Though, the same is not true for larger values of pp. So, Lagrange’s theorem is not true in this setup. See also references cited there in [4] for various work related to periodicity prior to Browkin’s ([4]) work in 20002000. Many research works have been done in recent times in which people have presented many modified algorithms to achieve the periodicity and other desirable properties of continued fractions in the field of pp-adic numbers. See for example [9], [6],[8],[27], [26], and the references cited there in. See also the survey article by Romeo ([30]) for a comprehensive history of the development of the continued fraction theory in the field of pp-adic numbers.

It is quite natural to consider continued fraction in finite extensions of p\mathbb{Q}_{p} which is left-out in the above discussion while considering continued fractions in all locally compact fields. In this article, we consider canonical extensions of the algorithms of Ruban and Browkin for p\mathbb{Q}_{p} in its finite extensions. Given any finite (necessarily simple) extension KK of p\mathbb{Q}_{p}, we consider this extension in two steps, viz. K=L(β)K=L(\beta) with K/LK/L a totally ramified extension, and L=p(γ)L=\mathbb{Q}_{p}(\gamma) with L/pL/\mathbb{Q}_{p} an unramified extension (see next section for more details). In our algorithms, the partial quotients are elements from the set [1p][γ,β]\mathbb{Z}\left[\frac{1}{p}\right][\gamma,\beta]. We show that any αK\alpha\in K has a unique continued fraction expansion, and given any sequence of partial quotients {ci}i0\{c_{i}\}_{i\geq 0}, [c0,c1,,cn][c_{0},c_{1},\ldots,c_{n}] converges to an element α\alpha of KK. For a few small degree unramified extensions of the form p(γ)\mathbb{Q}_{p}(\gamma), we show that any element of (γ)\mathbb{Q}(\gamma) has a finite continued fraction expansion.

In this article, we are also going to discuss the metrical theory of the associated continued fraction map. In the case of classical continued fraction for real numbers the Gauss map or the continued fraction map is defined as

T:(0,1)(0,1)T:(0,1)\rightarrow(0,1)
T(x)={1x},T(x)=\left\{\frac{1}{x}\right\},

where {1x}\left\{\frac{1}{x}\right\} denotes the fractional part of 1x\frac{1}{x}. It is well known that TT is ergodic with respect to the Gauss measure (see [14]). For the ergodicity of the continued fraction map for a more general class of continued fraction, see [18]. For complex continued fraction, the ergodicity of the maps associated with the nearest integer complex continued fractions over imaginary quadratic fields is discussed in a recent paper of Nakada et al. [13]. They, in fact, showed that the continued fraction map is exact (see Section 44 for definition). See also some references in [13] for some earlier works related to the metrical theory of complex continued fractions.

In the non-Archimedean settings, Berthe and Nakada [2] proved the ergodicity of the continued fraction map in positive characteristic, and as an application they obtained various metrical results regarding the averages of partial quotients, average growth rates of the denominators of the convergents, etc. In [23], Lertchoosakul and Nair proved the exactness of the continued fraction map using which they could consider more general averages concerning the partial quotients and the growth rate of the denominator of the convergents. The quantitative version of the metric theory of the continued fraction map in this setup was considered by the same authors in a subsequent paper [24]. The reader is referred to [12] for quantitative metrical results concerning real continued fraction.

In this article, we discuss metrical theory of continued fractions in finite extension of p\mathbb{Q}_{p}. We show that the associated continued fraction map is Haar measure preserving and exact. Then we obtain various metrical results analogous to the results of [23] concerning asymptotic behaviour of various quantities related to partial quotients, denominator of the convergents, etc. In these results, general averages using subsequence ergodic theory and moving averages are considered as done in [23].

2. Preliminary

For a prime number pp, the field of pp-adic numbers p\mathbb{Q}_{p} is the set of all Laurent series in pp of the form

α=jn0ajpj, where aj{0,1,,p1} and n0.\alpha=\sum\limits_{j\geq n_{0}}a_{j}p^{j},\text{ where }a_{j}\in\{0,1,\ldots,p-1\}\text{ and }n_{0}\in\mathbb{Z}.

The pp-adic valuation 𝔳p\mathfrak{v}_{p} on p\mathbb{Q}_{p} is defined as follows: if α=jn0ajpj\alpha=\sum_{j\geq n_{0}}a_{j}p^{j}, then

𝔳p(α):=inf{j:aj0}.\mathfrak{v}_{p}(\alpha):=\inf\,\{\,j\in\mathbb{Z}\ :\ a_{j}\neq 0\,\}.

Then the pp-adic absolute value of α\alpha is given by

|α|p:=p𝔳p(α)|\alpha|_{p}:=p^{-\mathfrak{v}_{p}(\alpha)}

when α0\alpha\neq 0, and |0|p=0|0|_{p}=0. The field of pp-adic numbers is the completion of \mathbb{Q} with respect to this absolute value. Let K=p(ξ)K=\mathbb{Q}_{p}(\xi) be a finite extension of p\mathbb{Q}_{p} of degree mm, i.e., [K:p]=m[\,K\,:\,\mathbb{Q}_{p}\,]=m. We may then take 𝔅={1,ξ,,ξm1}\mathfrak{B}=\{1,\xi,\ldots,\xi^{m-1}\} as a convenient vector space basis for KK over the field p\mathbb{Q}_{p}. Otherwise said, any element 𝔟K\mathfrak{b}\in K can be written uniquely as

𝔟=b0+b1ξ++bm1ξm1, where bjp for all j.\mathfrak{b}=b_{0}+b_{1}\xi+\cdots+b_{m-1}\xi^{m-1},\text{ where }b_{j}\in\mathbb{Q}_{p}\text{ for all }j.

Since every finite extension is an algebraic extension, we have for every 𝔟K\mathfrak{b}\in K that there is some monic irreducible polynomial

g(x)=xn+B1xn1++Bn1x+Bng(x)=x^{n}+B_{1}x^{n-1}+\cdots+B_{n-1}x+B_{n}

of degree at most mm and coefficients BjpB_{j}\in\mathbb{Q}_{p} such that g(𝔟)=0g(\mathfrak{b})=0 in KK. The norm map for the finite field extension K/pK/\mathbb{Q}_{p} is then defined as

NK/p(𝔟):=(1)nBn.N_{K/\mathbb{Q}_{p}}(\mathfrak{b}):=(-1)^{n}B_{n}.

Our absolute value ||p|\cdot|_{p} on p\mathbb{Q}_{p} extends uniquely to KK in the following manner (see [20] for details):

|𝔟|:=|NK/p(𝔟)|p1n,𝔟K.|\mathfrak{b}|:=\big{|}N_{K/\mathbb{Q}_{p}}(\mathfrak{b})\big{|}_{p}^{\frac{1}{n}},\quad\mathfrak{b}\in K.

Let us choose an element πK\pi\in K of maximal absolute value smaller than 1, say 0<θ:=|π|<10<\theta:=|\pi|<1. Define

𝒪K:={xK:|x|1},𝔪K:={xK:|x|<1}\mathcal{O}_{K}:=\{\,x\in K\ :\ |x|\leq 1\,\},\quad\mathfrak{m}_{K}:=\{\,x\in K\ :\ |x|<1\,\}

and

𝒪K:={xK:|x|=1}.\mathcal{O}_{K}^{*}:=\{\,x\in K\ :\ |x|=1\,\}.

We have π𝒪K=𝔪K\pi\mathcal{O}_{K}=\mathfrak{m}_{K}, and the residue field k¯=𝒪K/𝔪K\overline{k}=\mathcal{O}_{K}/\mathfrak{m}_{K} is a finite extension of 𝔽p\mathbb{F}_{p}.

Definition 1.

The residue degree of the finite extension KK of p\mathbb{Q}_{p} is the positive integer f=[k¯:𝔽p]=dim𝔽p(k¯)f=[\overline{k}:\mathbb{F}_{p}]=\text{dim}_{\mathbb{F}_{p}}(\overline{k}), where k¯\overline{k} is the residue field of KK. A finite extension KK of p\mathbb{Q}_{p} is said to be totally ramified if f=1f=1.

We also have that k¯=𝔽q\overline{k}=\mathbb{F}_{q}, where q=#(k¯)=pfq=\#\,(\overline{k})=p^{f}. This is because upto isomorphism there is exactly one finite field having qq elements.

Definition 2.

The ramification index of K/pK/\mathbb{Q}_{p} equals e=[|K|:|p|]=#(|K|/p)e=[\,|K^{*}|\,:\,|\mathbb{Q}^{*}_{p}|\,]=\#(|K^{*}|/p^{\mathbb{Z}}). A finite extension KK of p\mathbb{Q}_{p} is called unramified if e=1e=1.

For the uniformizer π𝔪K\pi\in\mathfrak{m}_{K}, we have |π|e=|p||\pi|^{e}=|p| thereby giving us

|π|=p1/e.|\pi|=p^{-1/e}.

Any element αK\alpha\in K can be represented as α=uπn\alpha=u\pi^{n} for some suitable u𝒪Ku\in\mathcal{O}^{*}_{K} and nn\in\mathbb{Z}. Then,

|α|=|π|n=pn/e.|\alpha|=|\pi|^{n}=p^{-n/e}.

The integers ee and ff given above satisfy ef=mef=m, where mm is the degree of the extension. We recall that  [29, Corollary 4-26] there exists an unramified subextension L/pL/\mathbb{Q}_{p} of degree ff such that K/LK/L is a totally ramified extension of degree ee. Also, there exists a γL\gamma\in L such that L=p(γ)L=\mathbb{Q}_{p}(\gamma) with |γ|=1|\gamma|=1 [20, § III.3]. (To boot, we may and do take γ\gamma to be some primitive (pf1)(p^{f}-1)-th root of unity)

Lemma 3.

Let K/LK/L be as above. Then, there exists some βK\beta\in K such that K=L(β)K=L(\beta) and |β|=p1/e|\beta|=p^{1/e}.

Proof.

We know that the value groups of LL^{*} and KK^{*} are \mathbb{Z} and (1/e)(1/e)\mathbb{Z}, respectively. Let us, therefore, choose some βKL\beta\in K\setminus L such that |β|=p1/e|\beta|=p^{1/e}. As K/LK/L is a finite extension, every element of KK is algebraic over the field LL. In particular, our chosen element β\beta satisfies some minimal monic polynomial

h(x)=xn+bn1xn1++b0,bjLh(x)=x^{n}+b_{n-1}x^{n-1}+\cdots+b_{0},\ b_{j}\in L

and nen\leq e. Now, we will like to show that K=L(β)K=L(\beta). It is equivalent to establishing that the minimal polynomial of β\beta has degree ee. Suppose n<en<e.

We have |β|=|b0|1/n|\beta|=|b_{0}|^{1/n} by the unique extension of the non-archimedean absolute value to KK. Here, |b0|=ps|b_{0}|=p^{s} for some ss\in\mathbb{Z} implying that |β|=ps/n=p1/e|\beta|=p^{s/n}=p^{1/e}. This is possible iff n=sen=se but 0<n<e0<n<e, a contradiction. Thus, there exists a βK\beta\in K such that K=L(β)K=L(\beta) with |β|=p1/e|\beta|=p^{1/e}. ∎

Every αK=L(β)=p(γ)(β)=p(β,γ)\alpha\in K=L(\beta)=\mathbb{Q}_{p}(\gamma)(\beta)=\mathbb{Q}_{p}(\beta,\gamma) can then be written as

(2) α=i=0e1j=0f1bi,jγjβi,bi,jp.\displaystyle\alpha=\sum_{i=0}^{e-1}\sum_{j=0}^{f-1}b_{i,j}\gamma^{j}\beta^{i},\ b_{i,j}\in\mathbb{Q}_{p}.

Now let X1X_{1} and X2X_{2} be the sets inside p\mathbb{Q}_{p} defined by:

(3) X1={j=0kajpj:k{0}andaj{0,1,,p1}for 0jk},\displaystyle X_{1}=\left\{\,\sum\limits_{j=0}^{k}\frac{a_{j}}{p^{j}}:\ k\in{\mathbb{N}}\cup\{0\}\ \text{and}\ a_{j}\in\{0,1,\ldots,p-1\}\ \text{for}\ 0\leq j\leq k\right\},

and

(4) X2={j=0kajpj:k{0}andaj{p12,,p12}for 0jk}.\displaystyle X_{2}=\left\{\,\sum\limits_{j=0}^{k}\frac{a_{j}}{p^{j}}:\ k\in{\mathbb{N}}\cup\{0\}\ \text{and}\ a_{j}\in\Big{\{}-\frac{p-1}{2},\ldots,\frac{p-1}{2}\Big{\}}\ \text{for}\ 0\leq j\leq k\right\}.

Note that the partial quotients for Ruban’s pp-adic continued fraction are elements of X1X_{1}, whereas, the partial quotients for Browkin’s algorithm are elements of X2X_{2}. In a moment, we define a set ZZ, the elements of which will be used as partial quotients for the continued fraction algorithm developed in this article. We may either use X1X_{1} or X2X_{2} while defining the set ZZ. In the first case, we get a generalization of Ruban’s algorithm in finite extensions of p\mathbb{Q}_{p}, whereas, we get a generalization of Browkin’s algorithm in the second case. From now on, we use the notation XX for both the sets X1X_{1} and X2X_{2} with the understanding that whenever we use XX, the discussion applies to both X1X_{1} and X2X_{2}. Any two distinct numbers in XX will have pp-adic distance at least one giving us that it is a 11-uniformly discrete set. Furthermore, every non-zero element has an absolute value at least one. Generalizing this observation for all finite pp-adic extensions, we have the following.

Lemma 4.

The set

Z:={i=0e1j=0f1bi,jγjβi:bi,jX}Z:=\left\{\,\sum_{i=0}^{e-1}\sum_{j=0}^{f-1}b_{i,j}\gamma^{j}\beta^{i}\ :\ \ b_{i,j}\in X\,\right\}

is 11-uniformly discrete. In particular, every non-zero element in ZZ has an absolute value at least one.

Proof.

Let z1,z2Zz_{1},z_{2}\in Z with z1z2z_{1}\neq z_{2}. This happens iff

z1z2=i=0e1j=0f1b~i,jγjβi,z_{1}-z_{2}=\sum_{i=0}^{e-1}\sum_{j=0}^{f-1}\,\widetilde{b}_{i,j}\gamma^{j}\beta^{i},

where at least one of the coefficients b~i,j\widetilde{b}_{i,j}’s (say b~k,\widetilde{b}_{k,\ell}) is a non-zero element from the set X(X)X\cup(-X). Consider

(5) y=j=0f1b~k,jγjL.y=\sum_{j=0}^{f-1}\widetilde{b}_{k,j}\gamma^{j}\in L.

Assume |y|<1|y|<1. Since γ\gamma is a primitive ff-th root of unity, it is plain that γ\gamma belongs to 𝒪L\mathcal{O}_{L}. Without loss of generality, we may assume that b~k,f1\widetilde{b}_{k,f-1} has the maximum absolute value in the representation of yy given by (5). This is because by choosing mm such that

|b~k,m|=max0jf1{|b~k,j|}\lvert\,\widetilde{b}_{k,m}\,\rvert=\max_{0\,\leq\,j\,\leq\,f-1}\big{\{}\,\lvert\,\widetilde{b}_{k,j}\,\rvert\,\big{\}}

and replacing yy with γf1my\gamma^{f-1-m}y, we can ensure that the coefficient of γf1\gamma^{f-1} has maximum absolute value amongst all the b~k,j\widetilde{b}_{k,j}’s. It then follows that

(6) γf1=(b~k,f1)1(yj=0f2b~k,jγj)\gamma^{f-1}=(\widetilde{b}_{k,f-1})^{-1}\left(y-\sum_{j=0}^{f-2}\widetilde{b}_{k,j}\gamma^{j}\right)

On reducing the above equation modulo p𝒪Lp\mathcal{O}_{L}, we get that

(7) γf1=j=0f2ajγj\gamma^{f-1}=\sum_{j=0}^{f-2}a_{j}\gamma^{j}

where aja_{j}’s are elements of the residue field p/pp\mathbb{Z}_{p}/p\mathbb{Z}_{p}. This leads to a contradiction as the degree of the residue field 𝒪L/p𝒪L\mathcal{O}_{L}/p\mathcal{O}_{L} over Zp/pZpZ_{p}/pZ_{p} is ff. Therefore, |y|1\lvert\,y\,\rvert\geq 1. We will in fact have that

z1z2=i=0e1yiβiz_{1}-z_{2}=\sum_{i=0}^{e-1}y_{i}\beta^{i}

with |yi|=pni|y_{i}|=p^{n_{i}} for some ni{0}n_{i}\in{\mathbb{N}}\cup\{0\} or |yi|=0\lvert\,y_{i}\,\rvert=0 for some ii while |β|=p1/e|\beta|=p^{1/e}. This implies that |yi1βi1||yi2βi2|\lvert\,y_{i_{1}}\beta^{i_{1}}\,\rvert\neq\lvert\,y_{i_{2}}\beta^{i_{2}}\,\rvert for any pair of indices 0i1<i2<e0\leq i_{1}<i_{2}<e with at least one yi1y_{i_{1}} or yi2y_{i_{2}} non-zero and there exists k{0,,e1}k\in\{0,\ldots,e-1\} such that

max0ie1{|yiβi|}=|ykβk|1.\max_{0\leq i\leq e-1}\{|y_{i}\beta^{i}|\}=|y_{k}\beta^{k}|\geq 1.

Therefore,

|z1z2|\displaystyle|z_{1}-z_{2}| =|i=0e1yiβi|\displaystyle=\left|\sum_{i=0}^{e-1}y_{i}\beta^{i}\right|
=max0ie1{|yiβi|}\displaystyle=\max_{0\leq i\leq e-1}\{|y_{i}\beta^{i}|\}
=|ykβk|\displaystyle=|y_{k}\beta^{k}|
1.\displaystyle\geq 1.

Hence, ZZ is a 11-uniformly discrete set. ∎

The metric balls in KK will have radius se\frac{s}{e} for ss\in\mathbb{Z}. More precisely, for αK\alpha\in K and ss\in\mathbb{Z}, let

B(α,pse)={xK:|xα|<pse}B\,(\,\alpha,\,p^{\frac{s}{e}}\,)\ =\ \{\,x\in K\ :\ |x-\alpha|<p^{\frac{s}{e}}\,\}

be the ball around α\alpha of radius psep^{\frac{s}{e}}. Let μ\mu denote the Haar measure on the local field KK (see Chapter 44 of [29] for existence of Haar measure) and it is normalized in such a way that μ(B(0,1))=1\mu\,\big{(}\,B(0,1)\,\big{)}=1. Note that

B(0,ps)=psB(0,1)={psx:xB(0,1)}B(0,p^{s})\ =\ p^{-s}B(0,1)\ =\ \{\,p^{-s}x\ :\ x\in B(0,1)\,\}

and, therefore,

(8) μ(B(0,ps))\displaystyle\mu\,\big{(}\,B(0,p^{s})\,\big{)} =μ(psB(0,1))=modK(1ps)μ(B(0,1))\displaystyle=\mu\big{(}\,p^{-s}B(0,1)\,\big{)}=\!\mod_{K}\left(\frac{1}{p^{s}}\right)\mu(B(0,1))
=modp(1ps)m=psm\displaystyle=\!\mod_{\mathbb{Q}_{p}}\left(\frac{1}{p^{s}}\right)^{m}\ =\ p^{sm}

by [29, Proposition 4-13]. Next two technical lemmas are useful for computing measure of various metric balls inside KK.

Lemma 5.

For ss\in\mathbb{Z}, the ball B(0,ps)KB(0,p^{s})\subset K is the same as the set

A:={xK:x=i=0e1j=0f1bi,jγjβi,bi,jp1sp}.A:=\big{\{}\,x\in K\ :\ x=\sum\limits_{i=0}^{e-1}\sum\limits_{j=0}^{f-1}b_{i,j}\gamma^{j}\beta^{i},\ b_{i,j}\in p^{1-s}\mathbb{Z}_{p}\,\big{\}}.
Proof.

First suppose xAx\in A. Then, x=i=0e1j=0f1bi,jγjβix\ =\ \sum_{i=0}^{e-1}\sum_{j=0}^{f-1}b_{i,j}\gamma^{j}\beta^{i} where bi,jp1spb_{i,j}\in p^{1-s}\mathbb{Z}_{p} for all i,ji,j. Since |bi,j|ps1,|γ|=1|b_{i,j}|\leq p^{s-1},\ |\gamma|=1 and |β|=p1e|\beta|=p^{\frac{1}{e}}, we obtain |x|ps1+e1e<ps|x|\leq p^{s-1+\frac{e-1}{e}}<p^{s}. Otherwise said, A{xK:|x|<ps}=B(0,ps)A\subseteq\{x\in K:|x|<p^{s}\}=B(0,p^{s}).

Conversely, let xAx\notin A so that x=i=0e1j=0f1bi,jγjβix=\sum_{i=0}^{e-1}\sum_{j=0}^{f-1}b_{i,j}\gamma^{j}\beta^{i}, where |bi,j|ps|b_{i,j}|\geq p^{s} for at least one pair (i,j)(i,j). For each 0i<e0\leq i<e and 0j<f0\leq j<f, we may write

bi,j={bi,j}+[bi,j]pb_{i,j}=\{b_{i,j}\}+[b_{i,j}]\in\mathbb{Q}_{p}

with all terms of the form l=0kaslpsl\sum_{l=0}^{k}a_{-s-l}p^{-s-l} contained in {bi,j}\{b_{i,j}\}, [bi,j]p1sp[b_{i,j}]\in p^{1-s}\mathbb{Z}_{p} and k{0}k\in{\mathbb{N}}\cup\{0\} is such that asl=0a_{-s-l}=0 for all l>kl>k. Thus, x=x1+x2x=x_{1}+x_{2} where psx1p^{s}x_{1} belongs to the set ZZ introduced in Lemma 4 while x2Ax_{2}\in A. This implies that |x1|ps\lvert\,x_{1}\,\rvert\geq p^{s} and |x2|<ps\lvert\,x_{2}\,\rvert<p^{s}. All in all, |x|ps\lvert\,x\,\rvert\geq p^{s} or equivalently, xB(0,ps)x\notin B(0,p^{s}). ∎

Lemma 6.

Let ss\in\mathbb{Z} and a{1,,e1}a\in\{1,\ldots,e-1\}. Then the ball B(0,ps1+eae)KB(0,p^{s-1+\frac{e-a}{e}})\subset K is the same as the set

A:={xK:x=i=0e1j=0f1bi,jγjβi,bi,jp1spfori<eaandbi,jp2spforeai<e}A:=\big{\{}\,x\in K:\,x=\sum\limits_{i=0}^{e-1}\sum\limits_{j=0}^{f-1}b_{i,j}\gamma^{j}\beta^{i},\begin{matrix}\ b_{i,j}\in p^{1-s}\mathbb{Z}_{p}\ \text{for}\ i<e-a\ \text{and}\\ b_{i,j}\in p^{2-s}\mathbb{Z}_{p}\ \text{for}\ e-a\leq i<e\end{matrix}\big{\}}
Proof.

First we show that AB(0,ps1+eae)A\subseteq B(0,p^{s-1+\frac{e-a}{e}}). Let xAx\in A. Then x=i=0e1j=0f1bi,jγjβix=\sum\limits_{i=0}^{e-1}\sum\limits_{j=0}^{f-1}b_{i,j}\gamma^{j}\beta^{i}, where bi,jp1spb_{i,j}\in p^{1-s}\mathbb{Z}_{p} for i<eai<e-a, and bi,jp2spb_{i,j}\in p^{2-s}\mathbb{Z}_{p} for eai<ee-a\leq i<e. Note that |bi,j|ps1|b_{i,j}|\leq p^{s-1} for all 0j<f0\leq j<f, 0i<ea0\leq i<e-a and |bi,j|ps2|b_{i,j}|\leq p^{s-2} for all 0j<f0\leq j<f, eai<ee-a\leq i<e. Also, |γ|=1|\gamma|=1 and |β|=p1e|\beta|=p^{\frac{1}{e}}, we get |x|ps1+ea1e<ps1+eae|x|\leq p^{s-1+\frac{e-a-1}{e}}<p^{s-1+\frac{e-a}{e}}. This shows that xB(0,ps1+eae)x\in B(0,p^{s-1+\frac{e-a}{e}}).

To prove the converse, suppose xAx\notin A. Then there exists a pair (i,j)(i,j) such that

bi,jp1spfor 0i<eaorbi,jp2spforeai<e.b_{i,j}\notin p^{1-s}\mathbb{Z}_{p}\ \text{for}\ 0\leq i<e-a\ \text{or}\ b_{i,j}\notin p^{2-s}\mathbb{Z}_{p}\ \text{for}\ e-a\leq i<e.

If bi,jp1spb_{i,j}\notin p^{1-s}\mathbb{Z}_{p}, then |bi,j|ps|b_{i,j}|\geq p^{s}. So, |x|psps1+eae|x|\geq p^{s}\nless p^{s-1+\frac{e-a}{e}}. And if bi,jp2spb_{i,j}\notin p^{2-s}\mathbb{Z}_{p}, then |bi,j|ps1|b_{i,j}|\geq p^{s-1}. So, |x|ps1+eaeps1+eae|x|\geq p^{s-1+\frac{e-a}{e}}\nless p^{s-1+\frac{e-a}{e}}. Thus, xB(0,ps1+eae)x\notin B(0,p^{s-1+\frac{e-a}{e}}). ∎

Now we are in a position to calculate the measure of any ball around zero inside KK.

Proposition 7.

The measure of the ball B(0,ps+ie)B\big{(}0,\ p^{s+\frac{i}{e}}\big{)} equals psm+fip^{sm+f{i}}.

Proof.

Fix some a{1,,e1}a\in\{1,\ldots,e-1\}. By (8) we know that

psm\displaystyle p^{sm}\ =μ(B(0,ps))=μ({xK:|x|<ps})\displaystyle=\ \mu\,\big{(}\,B(0,p^{s})\,\big{)}\ =\ \mu\,\big{(}\,\{\,x\in K\ :\ \lvert\,x\,\rvert<p^{s}\,\}\,\big{)}
=μ({xK:x=i=0e1j=0f1bi,jγjβi,bi,jp1sp})(by Lemma5).\displaystyle=\ \mu\,\big{(}\,\{\,x\in K\ :\ x=\sum\limits_{i=0}^{e-1}\sum\limits_{j=0}^{f-1}b_{i,j}\gamma^{j}\beta^{i},\ b_{i,j}\in p^{1-s}\mathbb{Z}_{p}\,\}\,\big{)}\ \ (\text{by Lemma}\ \ref{lem5}).

Since p1sp/p2sp({0,p1s,,(p1)p1s},+,)p^{1-s}\mathbb{Z}_{p}/p^{2-s}\mathbb{Z}_{p}\simeq(\{0,p^{1-s},\ldots,(p-1)p^{1-s}\},+,*), where

tjp1s+tjp1s:=(tj+ptj)p1sandtjp1stjp1s:=(tjptj)p1st_{j}p^{1-s}+t_{j^{\prime}}p^{1-s}:=(t_{j}+_{p}t_{j^{\prime}})p^{1-s}\ \text{and}\ t_{j}p^{1-s}*t_{j^{\prime}}p^{1-s}:=(t_{j}*_{p}t_{j^{\prime}})p^{1-s}

for all tj,tj{0,1,,p1}t_{j},t_{j^{\prime}}\in\{0,1,\ldots,p-1\}, we have a disjoint union decomposition

p1sp=p2sp(p1s+p2sp)((p1)p1s+p2sp).p^{1-s}\mathbb{Z}_{p}=p^{2-s}\mathbb{Z}_{p}\sqcup(p^{1-s}+p^{2-s}\mathbb{Z}_{p})\sqcup\ldots\sqcup((p-1)p^{1-s}+p^{2-s}\mathbb{Z}_{p}).

Therefore,

psm\displaystyle p^{sm}\ =μ({x=i=0e1j=0f1bi,jγjβi,bi,jp1spfori<eaandbi,jtjp1s+p2spforeai<e}),\displaystyle=\ \mu\left(\bigsqcup\big{\{}x=\sum\limits_{i=0}^{e-1}\sum\limits_{j=0}^{f-1}b_{i,j}\gamma^{j}\beta^{i},\begin{matrix}b_{i,j}\in p^{1-s}\mathbb{Z}_{p}\ \text{for}\ i<e-a\ \text{and}\\ b_{i,j}\in t_{j}p^{1-s}+p^{2-s}\mathbb{Z}_{p}\ \text{for}\ e-a\leq i<e\end{matrix}\big{\}}\right),

where the disjoint union is taken over all possible combinations
tj{ 0,1,,p1}t_{j}\in\{\,0,1,\ldots,p-1\,\}. By the translation invariant property of μ\mu,

μ(tjp1s+p2sp)=μ(p2sp).\mu(t_{j}p^{1-s}+p^{2-s}\mathbb{Z}_{p})=\mu(p^{2-s}\mathbb{Z}_{p}).

So,

psm\displaystyle p^{sm}\ =pfaμ({x=i=0e1j=0f1bi,jγjβi,bi,jp1spfori<eaandbi,jp2spforeai<e})\displaystyle=\ p^{fa}\mu\left(\big{\{}\,x=\sum\limits_{i=0}^{e-1}\sum\limits_{j=0}^{f-1}b_{i,j}\gamma^{j}\beta^{i},\begin{matrix}\ b_{i,j}\in p^{1-s}\mathbb{Z}_{p}\ \text{for}\ i<e-a\ \text{and}\\ b_{i,j}\in p^{2-s}\mathbb{Z}_{p}\ \text{for}\ e-a\leq i<e\end{matrix}\big{\}}\right)
=pfaμ(B( 0,ps1+eae))(by Lemma6).\displaystyle=\ p^{fa}\mu\,\big{(}\,B(\,0,\,p^{s-1+\frac{e-a}{e}}\,)\,\big{)}\ \ (\text{by Lemma}\ \ref{lem6}).

It follows that μ(B(0,ps1+eae))=psmfa\mu\,\big{(}\,B\,(0,p^{s-1+\frac{e-a}{e}})\,\big{)}=p^{sm-fa}. In general,

μ(B(0,ps+ie))=psm+fi\mu\big{(}B(0,p^{s+\frac{i}{e}})\big{)}=p^{sm+f{i}}

for all ss\in\mathbb{Z} and i{ 0,,e1}i\in\{\,0,\ldots,e-1\,\}. ∎

As ZZ is a uniformly discrete set, we can count the number of elements inside a set of elements with a fixed absolute value. This counting will be useful in some of the subsequent sections. Let YY be the set given by

Y:={j=0f1bjγj:bjX}.Y:=\left\{\sum_{j=0}^{f-1}b_{j}\gamma^{j}:b_{j}\in X\right\}.

For yYy\in Y, |y|pt|y|\leq p^{t} for some t{0}t\in{\mathbb{N}}\cup\{0\} if and only if |bj|<pt+1|b_{j}|<p^{t+1} for all j=0,,f1j=0,\ldots,f-1. Then it follows that for each n1n\geq 1,

#{yY:|y|=pn}\displaystyle\#\{y\in Y:|y|=p^{n}\} =#{yY:|y|pn}#{yY:|y|<pn}\displaystyle=\#\{y\in Y:|y|\leq p^{n}\}-\#\{y\in Y:|y|<p^{n}\}
=(pn+1)f(pn)f\displaystyle=\big{(}p^{n+1}\big{)}^{f}-\big{(}p^{n}\big{)}^{f}
=pfn(pf1).\displaystyle=p^{fn}(p^{f}-1).

Let us denote by ZZ^{*} the set of all cZc\in Z such that |c|>1|c|>1. Also let ss be a positive integer and a{0,,e1}a\in\{0,\ldots,e-1\}. Writing an element cZc\in Z^{*} as

c=i=0e1j=0f1bi,jγjβi:bi,jX,c=\sum_{i=0}^{e-1}\sum_{j=0}^{f-1}b_{i,j}\gamma^{j}\beta^{i}\ :\ \ b_{i,j}\in X,

note that if |bi,j|ps+1|b_{i,j}|\geq p^{s+1} for some 0ia& 0j<f0\leq i\leq a\ \&\ 0\leq j<f or |bi,j|ps|b_{i,j}|\geq p^{s} for some a<i<e& 0j<fa<i<e\ \&\ 0\leq j<f, then the set {cZ:|c|ps+ae}=ϕ\{c\in Z^{*}:|c|\leq p^{s+\frac{a}{e}}\}=\phi. Then it follows that

#{cZ:|c|=ps+ae}\displaystyle\#\left\{c\in Z^{*}:|c|=p^{s+\frac{a}{e}}\right\} =#{cZ:|c|ps+ae}#{cZ:|c|ps+a1e}\displaystyle=\#\left\{c\in Z^{*}:|c|\leq p^{s+\frac{a}{e}}\right\}-\#\left\{c\in Z^{*}:|c|\leq p^{s+\frac{a-1}{e}}\right\}
=(ps+1)f(a+1)(ps)f(e(a+1))(ps+1)af(ps)f(ea)\displaystyle=\big{(}p^{s+1}\big{)}^{f(a+1)}\big{(}p^{s}\big{)}^{f(e-(a+1))}\,-\,\big{(}p^{s+1}\big{)}^{af}\big{(}p^{s}\big{)}^{f(e-a)}
=pf(a+es)(pf1).\displaystyle=p^{f(a+es)}\big{(}p^{f}-1\big{)}.

Now, let a{1,,e1}a\in\{1,\ldots,e-1\}. By a similar argument, we also obtain

#{cZ:|c|=pae}=paf(pf1).\#\{c\in Z^{*}:|c|=p^{\frac{a}{e}}\}=p^{af}(p^{f}-1).

Hence, for all nn\in{\mathbb{N}},

(9) #{cZ:|c|=pne}=pfn(pf1).\#\big{\{}c\in Z^{*}:|c|=p^{\frac{n}{e}}\big{\}}=p^{fn}\big{(}p^{f}-1\big{)}.

3. Continued fraction algorithm and finiteness property

Now we describe a continued fraction algorithm for elements in any finite extension KK of p\mathbb{Q}_{p}. This algorithm generalizes Ruban’s algorithm of p\mathbb{Q}_{p} (for X=X1X=X_{1}), as well as Browkin’s algorithm (for X=X2X=X_{2}). Our algorithm is a very natural extension of Ruban’s and Browkin’s algorithm with partial quotients coming from the set ZZ.

The pp-adic floor function for Ruban’s algorithm is a function from p\mathbb{Q}_{p} to XX defined as follows: for α=jn0ajpjp\alpha=\sum\limits_{j\geq n_{0}}a_{j}p^{j}\in\mathbb{Q}_{p} with aj{0,1,,p1}a_{j}\in\{0,1,\ldots,p-1\} for X=X1X=X_{1} and aj{p12,,0,,p12}a_{j}\in\{-\frac{p-1}{2},\ldots,0,\ldots,\frac{p-1}{2}\} for X=X2X=X_{2},

αp={j=n00ajpj,if𝔳p(α)0(or|α|p1)0,otherwise.\ \lfloor\alpha\rfloor_{p}=\left\{\begin{array}[]{rcl}\displaystyle{\sum\limits_{j=n_{0}}^{0}a_{j}p^{j}},&\mbox{if}&\mathfrak{v}_{p}(\alpha)\leq 0\ (\text{or}\ |\alpha|_{p}\geq 1)\\ 0\ \ \ ,&\mbox{}&\text{otherwise.}\end{array}\right.

Using this floor function, we define a floor function on KK which is a function from KK to ZZ, as follows:

Forα=i=0e1j=0f1bi,jγjβi,bi,jp,\text{For}\ \alpha=\sum_{i=0}^{e-1}\sum_{j=0}^{f-1}b_{i,j}\gamma^{j}\beta^{i},\ b_{i,j}\in\mathbb{Q}_{p},
(10) α=i=0e1j=0f1bi,jpγjβi.\lfloor\alpha\rfloor=\sum_{i=0}^{e-1}\sum_{j=0}^{f-1}\lfloor b_{i,j}\rfloor_{p}\ \gamma^{j}\beta^{i}.

It is easy to see that

(11) |αα|1p1e<1|\alpha-\lfloor\alpha\rfloor|\leq\frac{1}{p^{\frac{1}{e}}}<1

for any αK\alpha\in K.

Following the existing literature, we call an expression of the form

c0+1c1+1\displaystyle{c_{0}+\frac{1}{\displaystyle{c_{1}+\frac{1}{\ddots}}}}

with cjZc_{j}\in Z and cjZc_{j}\in Z^{*} for j1j\geq 1, a continued fraction which is also written as [c0;c1,][c_{0};c_{1},\ldots]. It is a finite continued fraction if the sequence (cj)(c_{j}) is a finite one, otherwise, it is an infinite continued fraction. We call cjc_{j}’s the partial quotients of the continued fraction. We write,

[c0;c1,,cn]=sntn,[c_{0};c_{1},\ldots,c_{n}]=\frac{s_{n}}{t_{n}},

with sns_{n}, tn(γ,β)t_{n}\in\mathbb{Q}(\gamma,\beta), and call it the nnth convergent of the continued fraction [c0;c1,][c_{0};c_{1},\ldots]. It is easy to see that the sequence (sn)(s_{n}) and (tn)(t_{n}) satisfy the following recurrence relations:

sn=cnsn1+sn2andtn=cntn1+tn2,n2,s_{n}=c_{n}s_{n-1}+s_{n-2}\ \text{and}\ t_{n}=c_{n}t_{n-1}+t_{n-2},\ n\geq 2,

with s0=c0s_{0}=c_{0}, t0=1t_{0}=1, s1=c0c1+1s_{1}=c_{0}c_{1}+1, t1=c1t_{1}=c_{1}. The numerator and denominator of the convergents also satisfy

(12) tnsn1sntn1=(1)n,n0.t_{n}s_{n-1}-s_{n}t_{n-1}=(-1)^{n},\ n\geq 0.

Now we discuss the convergence properties of the continued fraction in our setup.

Lemma 8.

Let c0,c1,Zc_{0},c_{1},\ldots\in Z with cjZc_{j}\in Z^{*} for j1j\geq 1. Then the sequence of convergents sntn=[c0;c1,,cn]\frac{s_{n}}{t_{n}}=[c_{0};c_{1},\dots,c_{n}], converges to an element α\alpha of KK. Moreover,

|αsntn|=1|tn||tn+1|.\displaystyle{\left|\alpha-\frac{s_{n}}{t_{n}}\right|=\frac{1}{|t_{n}||t_{n+1}|}}.
Proof.

Note that,

sn+1tn+1sntn=sn+1tnsntn+1tntn+1=(1)ntn+1tnby (12).\displaystyle{\frac{s_{n+1}}{t_{n+1}}-\frac{s_{n}}{t_{n}}=\frac{s_{n+1}t_{n}-s_{n}t_{n+1}}{t_{n}t_{n+1}}=\frac{(-1)^{n}}{t_{n+1}t_{n}}}\ \ \text{by (\ref{*})}.

It is also easy to see that

|tn|=|cncn1c1|,n1|t_{n}|=|c_{n}c_{n-1}\cdots c_{1}|,\ n\geq 1

which in turn implies that |tn||t_{n}| is an increasing sequence as cjZc_{j}\in Z^{*} for j1j\geq 1. Then it follows that (|tn+1tn|)(|t_{n+1}t_{n}|) is an increasing sequence as well. Now, using the properties of ultrametric absolute value, it can be easily seen that

|smtmsntn|=1|tn+1tn|for anym>n.\left|\displaystyle{\frac{s_{m}}{t_{m}}}-\displaystyle{\frac{s_{n}}{t_{n}}}\right|=\frac{1}{|t_{n+1}t_{n}|}\ \text{for any}\ m>n.

As (|tn+1tn|)(|t_{n+1}t_{n}|) is increasing, it follows that (sntn)(\frac{s_{n}}{t_{n}}) is a Cauchy sequence, and hence converges to some αK\alpha\in K. ∎

Now, given any αK\alpha\in K, we generate its continued fraction expansion as follows:

α0=α,αn+1=(αnαn)1,cn=αn.\alpha_{0}=\alpha,\ \alpha_{n+1}=(\alpha_{n}-\lfloor\alpha_{n}\rfloor)^{-1},\ c_{n}=\lfloor\alpha_{n}\rfloor.

If αn=αn\alpha_{n}=\lfloor\alpha_{n}\rfloor for some nn, then αn+1\alpha_{n+1} is not defined and the sequences (αn)(\alpha_{n}) and (cn)(c_{n}) are finite. Otherwise, two infinite sequences are generated by the above construction. Here, cnc_{n}’s the partial quotients and αn\alpha_{n}’s the complete quotients corresponding to the continued fraction expansion of α\alpha. It is easy to see that α=[c0;c1,,cn,αn+1]\alpha=[c_{0};c_{1},\ldots,c_{n},\alpha_{n+1}].

Now, suppose αK\alpha\in K be such that the sequences (αn)(\alpha_{n}) and (cn)(c_{n}) are infinite, and let

sntn=[c0;c1,,cn].\frac{s_{n}}{t_{n}}=[c_{0};c_{1},\ldots,c_{n}].

We have cjZc_{j}\in Z^{*} for j1j\geq 1 by (11). Then it follows from Lemma 8 that the sequence of convergents sntn\frac{s_{n}}{t_{n}} converges to α\alpha.

Remark 9.

The definition of continued fraction in this article differs from the definition of continued fractions discussed by Capuano et al. in [7]. Our defining conditions of floor function are less restrictive; in fact, we do not impose a condition like the 22nd condition in Definition 3.13.1 of [7]. Also, our algorithm is less abstract which enables us to discuss the metrical theory of the associated continued fraction map. The following example shows that the 22nd condition of Definition 3.13.1 of [7] may not be satisfied in our setup.

Example 10.

Let K=p(β)=5(115)K=\mathbb{Q}_{p}(\beta)=\mathbb{Q}_{5}\left(\frac{1}{\sqrt{15}}\right). Then [K:5]=2[K:\mathbb{Q}_{5}]=2 and KK is a totally ramified extension of 5\mathbb{Q}_{5}. If αK\alpha\in K is given by α=(n=05n)115\alpha=\left(\sum\limits_{n=0}^{\infty}5^{n}\right)\frac{1}{\sqrt{15}}, then

α=(n=05n)115=115.\lfloor\alpha\rfloor=\left\lfloor\left(\sum\limits_{n=0}^{\infty}5^{n}\right)\frac{1}{\sqrt{15}}\right\rfloor=\frac{1}{\sqrt{15}}.

Now, let |.|𝔳|\ .\ |_{\mathfrak{v}*} be an ultrametric normalized absolute value on the number field (115)\mathbb{Q}\left(\frac{1}{\sqrt{15}}\right) such that 𝔳\mathfrak{v}* is a non-Archimedean place lying over the prime 33. Then

|α|𝔳=|115|𝔳=|115|312=3>1|\lfloor\alpha\rfloor|_{\mathfrak{v}*}=\left|\frac{1}{\sqrt{15}}\right|_{\mathfrak{v}*}=\left|\frac{1}{15}\right|^{\frac{1}{2}}_{3}=\sqrt{3}>1

which violates the 22nd condition of Definition 3.13.1 of [7].

One of the main difficulties for continued fractions in the pp-adic setup is that rational numbers do not necessarily have finite continued fraction expansions (also known as finiteness property) for many algorithms. In 19781978, Browkin modified Ruban’s algorithm to achieve the finiteness property for pp-adic continued fraction. The fact that Euclidean absolute value of the partial quotients in Browkin’s algorithm is less than p2\frac{p}{2} was crucially used in Browkin’s proof of finiteness. In our setup, we prove the finiteness property for some small degree extensions of p\mathbb{Q}_{p} in the case of generalization of Browkin’s algorithm, i.e., in the case X=X2X=X_{2}.

Let pp be either 33 modulo 44 or 55 modulo 1212. In the first case, we take K=p(ι)K=\mathbb{Q}_{p}(\iota), where ι\iota is the root of the polynomial X2+1=0X^{2}+1=0. In the 22nd case we take K=p(ω)K=\mathbb{Q}_{p}(\omega), where ω\omega is the root of the irreducible polynomial X2+X+1=0X^{2}+X+1=0. We show that the finiteness property holds in the cases of these extensions of p\mathbb{Q}_{p}. Note that when p1(mod12)p\equiv-1\pmod{12}, then p(ι)\mathbb{Q}_{p}(\iota) and p(ω)\mathbb{Q}_{p}(\omega) gives rise to the same extension of p\mathbb{Q}_{p}. For a cyclotomic extension p(γ)\mathbb{Q}_{p}(\gamma) of p\mathbb{Q}_{p}, where γ\gamma is some primitive nnth root of unity, we define the Galois height of field rational elements as follows: for α(γ)\alpha\in\mathbb{Q}(\gamma),

H(α):=maxσ|σ(α)|H(\alpha):=\max_{\sigma}|\sigma(\alpha)|_{\infty}

where the maximum is taken over all (distinct modulo conjugation) Galois embeddings of (γ)\mathbb{Q}(\gamma) inside {\mathbb{C}}, and |y||y|_{\infty} denotes the Euclidean norm of the complex number yy. The following lemma gives us the required bound on the Galois heights which will be useful in proving the finiteness property.

Lemma 11.

If b0,b1[1p](p2,p2)b_{0},b_{1}\ \in\ \mathbb{Z}\left[\frac{1}{p}\right]\,\cap\,\left(-\frac{p}{2},\frac{p}{2}\right), then there exists δ>0\delta>0 such that H(b0+b1ι)<p1pδH\,(\,b_{0}+b_{1}\iota\,)\ <\ p-\frac{1}{p}-\delta and H(b0+b1ω)<p1pδH\,(\,b_{0}+b_{1}\omega\,)\ <\ p-\frac{1}{p}-\delta.

Proof.

Here, |bj|<p/2\lvert\,b_{j}\,\rvert_{\infty}<p/2 for j=0,1j=0,1 giving us that

|σ(b0+b1ι)|=|b0+b1σ(ι)|max{|b0|,|b1|}2<p2\lvert\,\sigma(b_{0}+b_{1}\iota)\,\rvert_{\infty}=\lvert\,b_{0}+b_{1}\sigma(\iota)\,\rvert_{\infty}\leq\max\{\lvert\,b_{0}\,\rvert_{\infty},\lvert\,b_{1}\,\rvert_{\infty}\}\cdot\sqrt{2}<\frac{p}{\sqrt{2}}

for all σGal((ι))\sigma\in\operatorname{Gal}_{\mathbb{Q}}\big{(}\,\mathbb{Q}(\iota)\,\big{)} when p3(mod4)p\equiv 3\pmod{4}. Again,

|σ(b0+b1ω)|=|b0+b1σ(ω)|max{|b0|,|b1|}3<p32\lvert\,\sigma(b_{0}+b_{1}\omega)\,\rvert_{\infty}=\lvert\,b_{0}+b_{1}\sigma(\omega)\,\rvert_{\infty}\leq\max\{\lvert\,b_{0}\,\rvert_{\infty},\lvert\,b_{1}\,\rvert_{\infty}\}\cdot\sqrt{3}<p\cdot\frac{\sqrt{3}}{2}

for all σGal((ω))\sigma\in\operatorname{Gal}_{\mathbb{Q}}\big{(}\,\mathbb{Q}(\omega)\,\big{)} when p5(mod12)p\equiv 5\pmod{12}. Then it is clear that we can find a suitable δ>0\delta>0 such that the assertions of the lemma hold.

Proposition 12.

Let K=p(γ)K=\mathbb{Q}_{p}(\gamma) where γ=ι\gamma=\iota when p3(mod4)p\equiv 3\pmod{4} and γ=ω\gamma=\omega when p5(mod12)p\equiv 5\pmod{12}. Also let X=X2X=X_{2}, i.e., the partial quotients of the continued fraction expansion of any element of p(γ)\mathbb{Q}_{p}(\gamma) are elements of the form

b0+b1γwithb0,b1[1p](p2,p2).b_{0}+b_{1}\gamma\ \text{with}\ b_{0},b_{1}\in\mathbb{Z}\left[\frac{1}{p}\right]\cap\left(-\frac{p}{2},\frac{p}{2}\right).

Then, any α(γ)\alpha\in\mathbb{Q}(\gamma) has a finite continued fraction expansion.

Proof.

We use a suitable generalisation of the method used in Proposition 4.34.3 of [7]. As α(γ)\alpha\in\mathbb{Q}(\gamma), we can express α\alpha as

α=X0Y0\alpha=\frac{X_{0}}{Y_{0}}

with X0[1p][γ]X_{0}\in\mathbb{Z}\left[\frac{1}{p}\right][\gamma], Y0Y_{0}\in\mathbb{Z} and pY0p\not|\ Y_{0}. We define two sequences (Un)(U_{n}) and (Yn)(Y_{n}) as follows:

Un=snαtn,Yn=Y0Un.U_{n}=s_{n}-\alpha t_{n},\ Y_{n}=Y_{0}U_{n}.

Then, it is easy to see that

Un=(1)n+1j=1n+11αjU_{n}=(-1)^{n+1}\prod\limits_{j=1}^{n+1}\frac{1}{\alpha_{j}}

and, consequently,

(13) |Un|=j=1n+11|cj|as|αj|=|cj|.|U_{n}|=\prod\limits_{j=1}^{n+1}\frac{1}{|c_{j}|}\ \text{as}\ |\alpha_{j}|=|c_{j}|.

It is also easy to see that the sequence (Yn)(Y_{n}) satisfies the recurrence relation

(14) Yn=cnYn1+Yn2.Y_{n}=c_{n}Y_{n-1}+Y_{n-2}.

Clearly, Yn[1p][γ]Y_{n}\in\mathbb{Z}\left[\frac{1}{p}\right][\gamma]. Also, by definition of YnY_{n} and (13), we have Ynpnp[γ]Y_{n}\in p^{n}\mathbb{Z}_{p}[\gamma]. Hence, Yn[1p][γ]pnp[γ]=pn[γ]Y_{n}\in\mathbb{Z}\left[\frac{1}{p}\right][\gamma]\cap p^{n}\mathbb{Z}_{p}[\gamma]=p^{n}\mathbb{Z}[\gamma]. Taking the Galois height of both sides of (14), and then dividing by pnp^{n}, we have

H(Yn)pn\displaystyle\frac{H(Y_{n})}{p^{n}} H(cn)pH(Yn1)pn1+1p2H(Yn2)pn2\displaystyle\leq\frac{H(c_{n})}{p}\frac{H(Y_{n-1})}{p^{n-1}}+\frac{1}{p^{2}}\frac{H(Y_{n-2})}{p^{n-2}}
<(p1pδ)1pH(Yn1)pn1+1p2H(Yn2)pn2.\displaystyle<\left(p-\frac{1}{p}-\delta\right)\cdot\frac{1}{p}\cdot\frac{H(Y_{n-1})}{p^{n-1}}+\frac{1}{p^{2}}\cdot\frac{H(Y_{n-2})}{p^{n-2}}.

Let Tn=H(Yn)pnT_{n}=\frac{H(Y_{n})}{p^{n}}, D1=(p1pδ)1pD_{1}=\left(p-\frac{1}{p}-\delta\right)\cdot\frac{1}{p}, D2=1p2D_{2}=\frac{1}{p^{2}}. Then

Tn<D1Tn1+D2Tn2.T_{n}<D_{1}\ T_{n-1}+D_{2}\ T_{n-2}.

Since D0+D1<1D_{0}+D_{1}<1, it follows from Lemma 4.24.2 of [7] that |Tn|0|T_{n}|_{\infty}\to 0 as nn\to\infty. Hence there exists n0n_{0}\in{\mathbb{N}} such that Yn=0Y_{n}=0 nn0\forall\,n\geq n_{0} since Ynpn[γ]\frac{Y_{n}}{p^{n}}\in\mathbb{Z}[\gamma]. This means that α=sntn\alpha=\frac{s_{n}}{t_{n}} for some nn, and consequently, α\alpha has a finite continued fraction expansion. ∎

4. Exactness

Let K=p(γ,β)K=\mathbb{Q}_{p}(\gamma,\beta) be a finite extension of p\mathbb{Q}_{p}, and .\lfloor\ .\ \rfloor be the floor function defined in 10. The continued fraction map TT is defined on B(0,1)B(0,1) inside KK, as follows:

(15) T(α)=1α1α for α0 and T(0)=0,T(\alpha)=\frac{1}{\alpha}-\left\lfloor\frac{1}{\alpha}\right\rfloor\text{ for }\alpha\neq 0\text{ and }T(0)=0,

where \lfloor\cdot\rfloor is as defined in (10). In this section, we shall prove the exactness of TT, and in the subsequent section we prove various metrical results as consequences of exactness. We shall be considering the continued fraction map corresponding to the extension of Ruban’s algorithm in finite extensions of p\mathbb{Q}_{p}, though similar assumptions hold for the continued fraction map corresponding to the extension of the Browkin’s algorithm as well. Now, let αB(0,1)\alpha\in B(0,1) and α=[0;c1,c2,]\alpha=[0;c_{1},c_{2},\ldots] be the continued fraction axpansion of α\alpha. To emphasize the dependence on α\alpha, we will also use ck(α)c_{k}(\alpha) to denote the kkth partial quotient of the continued fraction expansion of α\alpha, i.e., α=[0;c1(α),c2(α),]\alpha=[0;c_{1}(\alpha),c_{2}(\alpha),\ldots]. Note that,

Tn(α)=[0;cn+1(α),cn+2(α),],andck(Tn(α))=cn+k(α)T^{n}(\alpha)=[0;c_{n+1}(\alpha),c_{n+2}(\alpha),\ldots],\ \text{and}\ c_{k}\big{(}T^{n}(\alpha)\big{)}=c_{n+k}(\alpha)

for all k1k\geq 1 and n0n\geq 0.

Recall that a measure preserving dynamical system (X,𝒞,ν,S)(X,\mathcal{C},\nu,S) is said to be exact if

n=0Sn𝒞=𝒩(modν),\bigcap\limits_{n=0}^{\infty}S^{-n}\mathcal{C}=\mathcal{N}\ (\text{mod}\ \nu),

where 𝒩\mathcal{N} is the trivial sub σ\sigma-algebra of 𝒞\mathcal{C} generated by the sets of measure 0 or 11. For nn\in{\mathbb{N}}, and c1,,cnZc_{1},\ldots,c_{n}\in Z^{*}, let Δc1,,cn\Delta_{c_{1},\ldots,c_{n}} denote the cylinder set of length nn, i.e.,

(16) Δc1,,cn={[0;c1,,cn1,cn+β]:βB(0,1)}.\Delta_{c_{1},\ldots,c_{n}}=\big{\{}[0;c_{1},\ldots,c_{n-1},c_{n}+\beta]\ :\ \beta\in B(0,1)\big{\}}.

The following lemma gives an alternate description of a cylinder set which will be helpful in calculating its measure. The proof of this lemma is similar to the proof of Lemma 22 of [23].

Lemma 13.

For any finite sequence c1,,cnZc_{1},\ldots,c_{n}\in Z^{*},

Δc1,,cn=B([0;c1,,cn],|c1cn|2).\Delta_{c_{1},\ldots,c_{n}}=B\,\big{(}\,[0;c_{1},\ldots,c_{n}],\,|c_{1}\cdots c_{n}|^{-2}\,\big{)}.

Because of the above lemma, it is not hard to see that the Borel σ\sigma-algebra on B(0,1)B(0,1) is generated by the cylinder sets described above. We denote by \mathcal{B} the Borel σ\sigma-algebra on B(0,1)B(0,1). Also, let μ\mu be the restriction of the Haar measure on B(0,1)B(0,1), and TT be the continued fraction map on B(0,1)B(0,1) defined above. We first show that TT is measure-preserving. Note that two cylinders Δc1,,cn\Delta_{c_{1},\ldots,c_{n}} and Δd1,,dn\Delta_{d_{1},\ldots,d_{n}} of the same length are disjoint if and only if cjdjc_{j}\neq d_{j} for some 1jn1\leq j\leq n.

Lemma 14.

The dynamical system (B(0,1),,μ,T)\big{(}B(0,1),\mathcal{B},\mu,T\big{)} is measure-preserving.

Proof.

Since the cylinder sets generate the Borel σ\sigma-algebra, it is enough to show that TT is measure-preserving on cylinder sets. For any cylinder set Δc1,,cn\Delta_{c_{1},\ldots,c_{n}}, there exists ss\in\mathbb{Z} and i{0,,e1}i\in\{0,\ldots,e-1\} such that |c1cn|2=ps+ie|c_{1}\cdots c_{n}|^{-2}=p^{s\,+\,\frac{i}{e}}, and consequently,

μ(Δc1,,cn)=μ(B([0,c1,,cn],|c1cn|2))=pms+fi.\mu(\Delta_{c_{1},\ldots,c_{n}})=\mu(B([0,c_{1},\ldots,c_{n}],|c_{1}\cdots c_{n}|^{-2}))=p^{ms+fi}.

The inverse image of Δc1,,cn\Delta_{c_{1},\ldots,c_{n}} under TT is given by a disjoint union as follows:

(17) T1Δc1,,cn=cZΔc,c1,,cn.T^{-1}\Delta_{c_{1},\ldots,c_{n}}=\bigcup_{c\in Z^{*}}\Delta_{c,c_{1},\ldots,c_{n}}.

Then

μ(T1Δc1,,cn)\displaystyle\mu(T^{-1}\Delta_{c_{1},\ldots,c_{n}}) =cZμ(B([c,c1,,cn],|c|2ps+ie))\displaystyle=\sum\limits_{c\in Z^{*}}\mu(B([c,c_{1},\ldots,c_{n}],|c|^{-2}p^{s+\frac{i}{e}}))
=n=1pfn(pf1)pms+fi2fn(using(9))\displaystyle=\sum\limits_{n=1}^{\infty}p^{fn}(p^{f}-1)p^{ms+fi-2fn}\ (\text{using}\ (\ref{Z*:coun}))
=(pf1)pms+fif11pf\displaystyle=(p^{f}-1)\frac{p^{ms+fi-f}}{1-\frac{1}{p^{f}}}
=pms+fi\displaystyle=p^{ms+fi}
=(ps+ie)m\displaystyle=\big{(}p^{s+\frac{i}{e}}\big{)}^{m}
=μ(Δc1,,cn).\displaystyle=\mu(\Delta_{c_{1},\ldots,c_{n}}).

The following technical lemma which is analogous to Lemma 44 of [23], is a crucial ingredient in proving exactness of the continued fraction map.

Lemma 15.

For the dynamical system (B(0,1),,μ,T)\big{(}B(0,1),\mathcal{B},\mu,T\big{)}, if EE\in\mathcal{B}, then for any natural number nn and cylinder set Δc1,,cn\Delta_{c_{1},\ldots,c_{n}}, we have

μ(Δc1,,cnTnE)=μ(Δc1,,cn)μ(E).\mu(\Delta_{c_{1},\ldots,c_{n}}\cap T^{-n}E)=\mu(\Delta_{c_{1},\ldots,c_{n}})\mu(E).
Proof.

It is enough to consider EE to be a cylinder set. Let E=Δd1,,dmE=\Delta_{d_{1},\ldots,d_{m}}. Then there exist s1,s2s_{1},s_{2}\in\mathbb{Z} and i1,i2{0,,e1}i_{1},i_{2}\in\{0,\ldots,e-1\} such that |c1cn|2=ps1+i1e|c_{1}\cdots c_{n}|^{-2}=p^{s_{1}+\frac{i_{1}}{e}} and |d1dm|2=ps2+i2e|d_{1}\cdots d_{m}|^{-2}=p^{s_{2}+\frac{i_{2}}{e}}. Now,

TnΔd1,,dm=c1,,cnZΔc1,,cn,d1,,dm.T^{-n}\Delta_{d_{1},\ldots,d_{m}}=\bigcup_{c^{\prime}_{1},\ldots,c^{\prime}_{n}\in Z^{*}}\Delta_{c^{\prime}_{1},\ldots,c^{\prime}_{n},d_{1},\ldots,d_{m}}.

Also, Δc1,,cnTnΔd1,,dm=μ(Δc1,,cn,d1,,dm)\Delta_{c_{1},\ldots,c_{n}}\cap T^{-n}\Delta_{d_{1},\ldots,d_{m}}=\mu(\Delta_{c_{1},\ldots,c_{n},d_{1},\ldots,d_{m}}). Then,

μ(Δc1,,cnTnΔd1,,dm)\displaystyle\mu(\Delta_{c_{1},\ldots,c_{n}}\cap T^{-n}\Delta_{d_{1},\ldots,d_{m}}) =μ(Δc1,,cn,d1,,dm)\displaystyle=\mu(\Delta_{c_{1},\ldots,c_{n},d_{1},\ldots,d_{m}})
=μ(B([c1,,cn,d1,,dm],|c1cnd1dm|2))\displaystyle=\mu(B([c_{1},\ldots,c_{n},d_{1},\ldots,d_{m}],|c_{1}\cdots c_{n}d_{1}\cdots d_{m}|^{-2}))
=μ(B([c1,,cn,d1,,dm],ps1+i1eps2+i2e))\displaystyle=\mu(B([c_{1},\ldots,c_{n},d_{1},\ldots,d_{m}],p^{s_{1}+\frac{i_{1}}{e}}p^{s_{2}+\frac{i_{2}}{e}}))
=pms1+fi1pms2+fi2\displaystyle=p^{ms_{1}+fi_{1}}\cdot p^{ms_{2}+fi_{2}}
=μ(Δc1,,cn)μ(Δd1,,dm).\displaystyle=\mu(\Delta_{c_{1},\ldots,c_{n}})\mu(\Delta_{d_{1},\ldots,d_{m}}).

Now we show that the continued fraction map TT is exact.

Theorem 16.

The dynamical system (B(0,1),,μ,T)\big{(}B(0,1),\mathcal{B},\mu,T\big{)} is an exact dynamical system.

Proof.

It is enough to show that n=0Tn𝒩\bigcap_{n=0}^{\infty}T^{-n}\mathcal{B}\subseteq\mathcal{N}. Let En=0TnE\in\bigcap_{n=0}^{\infty}T^{-n}\mathcal{B}. Then for each n1n\geq 1, there exists EnE_{n}\in\mathcal{B} such that E=TnEnE=T^{-n}E_{n} and μ(En)=μ(E)\mu(E_{n})=\mu(E). Now, for each cylinder set Δc1,,cn\Delta_{c_{1},\ldots,c_{n}} of length nn,

μ(EΔc1,,cn)\displaystyle\mu(E\cap\Delta_{c_{1},\ldots,c_{n}}) =μ(TnEnΔc1,,cn)\displaystyle=\mu(T^{-n}E_{n}\cap\Delta_{c_{1},\ldots,c_{n}})
=μ(E)μ(Δc1,,cn)(by Lemma 15).\displaystyle=\mu(E)\mu(\Delta_{c_{1},\ldots,c_{n}})\ (\text{by Lemma }\ \ref{E:imm}).

Then it follows from Lemma 55 of [23] that μ(E)=0\mu(E)=0 or 11, consequently, E𝒩E\in\mathcal{N}. ∎

5. Metrical results

Now we obtain results analogous to the metrical results of [23] in our setup. Since TT is exact, it is weak-mixing as well, i.e.,

1nk=1n|μ(ETkF)μ(E)μ(F)|0\frac{1}{n}\sum\limits_{k=1}^{n}\big{|}\mu(E\cap T^{-k}F)-\mu(E)\mu(F)\big{|}\to 0

as nn\to\infty for any E,FE,F\in\mathcal{B}. Weak-mixing property of the continued fraction map enables one to consider metrical results in the context of certain subsequences. This is done in [28] for continued fraction map in the case of real numbers, and in [23] in the positive characteristic setup. We do a similar study here for continued fraction map on B(0,1)B(0,1) inside KK. Before proceeding further we recall two definitions which plays crucial role in the discussion of metrical theory using subsequences.

Definition 17.

A strictly increasing sequence of positive integers (an)n=1(a_{n})_{n=1}^{\infty} is said to be L2L^{2}-good universal, if for each dynamical system (X,𝒞,ν,S)(X,\mathcal{C},\nu,S) and gL2(X,𝒞,ν)g\in L^{2}(X,\mathcal{C},\nu), the limit

limn1nj=1ng(Saj1α)\lim\limits_{n\to\infty}\frac{1}{n}\sum\limits_{j=1}^{n}g(S^{a_{j}-1}\alpha)

exists ν\nu-almost everywhere.

Definition 18.

A sequence of real numbers (xn)n=1(x_{n})_{n=1}^{\infty} is called uniformly distributed modulo 1, if for each interval I[0,1)I\subseteq[0,1), we have

limn1n#{1jn:{xj}I}=|I|,\lim\limits_{n\to\infty}\frac{1}{n}\cdot\#\{1\leq j\leq n:\{x_{j}\}\in I\}=|I|,

where |I||I| denotes the length of II and {xj}\{x_{j}\} denotes the fractional part of xjx_{j}.

Please see [23] for examples of L2L^{2}-good universal sequences. The following proposition is a consequence of weak-mixing, the proof of which can be found in [28].

Proposition 19.

Let (X,𝒞,ν,S)(X,\mathcal{C},\nu,S) be a weak-mixing dynamical system. Suppose (an)n=1(a_{n})^{\infty}_{n=1} is an L2L^{2}-good universal sequence of natural numbers such that (anγ)n=1(a_{n}\gamma)^{\infty}_{n=1} is uniformly distributed modulo 11 for any irrational number γ\gamma. Then for any gL2(X,𝒞,ν)g\in L^{2}(X,\mathcal{C},\nu),

limn1nj=1ng(Saj1α)=Xgdν\lim\limits_{n\to\infty}\frac{1}{n}\sum\limits_{j=1}^{n}g\big{(}S^{a_{j}-1}\alpha\big{)}=\int\limits_{X}g\ {\rm d}\nu

ν\nu-almost everywhere.

Proposition 20.

Let F:0F:\mathbb{R}_{\geq 0}\to\mathbb{R} be an increasing function such that

B(0,1)|F(|c1(α)|)|2dμ<.\int\limits_{B(0,1)}\big{|}F(|c_{1}(\alpha)|)\big{|}^{2}\,{\rm d}\mu<\infty.

For any natural number nn and non-negative real numbers d1,,dnd_{1},\ldots,d_{n}, let the generalized average be defined as

MF,n(d1,,dn)=F1(F(d1)++F(dn)n).M_{F,n}(d_{1},\ldots,d_{n})=F^{-1}\left(\frac{F(d_{1})+\cdots+F(d_{n})}{n}\right).

If (an)n=1(a_{n})^{\infty}_{n=1} is an L2L^{2}-good universal sequence of natural numbers such that (anγ)n=1(a_{n}\gamma)^{\infty}_{n=1} is uniformly distributed modulo 11 for any irrational number γ\gamma, then

limnMF,n(|ca1(α)|,,|can(α)|)=F1(B(0,1)F(|c1(α)|)dμ)\lim\limits_{n\to\infty}M_{F,n}(|c_{a_{1}}(\alpha)|,\ldots,|c_{a_{n}}(\alpha)|)=F^{-1}\left(\int\limits_{B(0,1)}F(|c_{1}(\alpha)|)\,{\rm d}\mu\right)

μ\mu-almost everywhere.

Proof.

Apply Proposition 19 to the function g(α)=F(|c1(α)|)g(\alpha)=F\big{(}|c_{1}(\alpha)|\big{)}. ∎

The following is also a consequence of Proposition 19, in which one considers a function from 0k\mathbb{R}^{k}_{\geq 0} to \mathbb{R}.

Proposition 21.

Suppose that H:0kH:\mathbb{R}^{k}_{\geq 0}\to\mathbb{R} is a function such that

B(0,1)|H(|c1(α)|,,|ck(α)|)|2dμ<,\int\limits_{B(0,1)}\Big{|}H\big{(}|c_{1}(\alpha)|,\ldots,|c_{k}(\alpha)|\big{)}\Big{|}^{2}\,{\rm d}\mu<\infty,

and if (an)n=1(a_{n})^{\infty}_{n=1} is an L2L^{2}-good universal sequence of natural numbers such that (anγ)n=1(a_{n}\gamma)^{\infty}_{n=1} is uniformly distributed modulo 11 for any irrational number γ\gamma. Then

limn1nj=1nH(|caj(α)|,,|caj+k1(α)|)=(i1,,ik)kH(pi1e,,pike)(pf1)kpf(i1++ik)\lim\limits_{n\to\infty}\frac{1}{n}\sum\limits_{j=1}^{n}H\big{(}|c_{a_{j}}(\alpha)|,\ldots,|c_{a_{j}+k-1}(\alpha)|\big{)}=\sum\limits_{(i_{1},\ldots,i_{k})\in\mathbb{N}^{k}}H\left(p^{\frac{i_{1}}{e}},\ldots,p^{\frac{i_{k}}{e}}\right)\frac{(p^{f}-1)^{k}}{p^{f(i_{1}+\cdots+i_{k})}}

μ\mu-almost everywhere.

Proof.

Let g(α)=H(|c1(α)|,,|ck(α)|)g(\alpha)=H\big{(}|c_{1}(\alpha)|,\ldots,|c_{k}(\alpha)|\big{)}. Then applying Proposition 19 to the function gg, we get

limnto1nj=1nH(|caj(α)|,,|caj+k1|)\displaystyle\lim\limits_{nto\infty}\frac{1}{n}\sum\limits_{j=1}^{n}H\big{(}|c_{a_{j}}(\alpha)|,\ldots,|c_{a_{j}+k-1}|\big{)} =B(0,1)H(|c1(α)|,,|ck(α)|)dμ\displaystyle=\int\limits_{B(0,1)}H\big{(}|c_{1}(\alpha)|,\ldots,|c_{k}(\alpha)|\big{)}\,{\rm d}\mu
=(i1,,ik)kH(pi1e,,pike)(pf1)kpf(i1++ik),\displaystyle=\sum\limits_{(i_{1},\ldots,i_{k})\in\mathbb{N}^{k}}H\left(p^{\frac{i_{1}}{e}},\ldots,p^{\frac{i_{k}}{e}}\right)\frac{(p^{f}-1)^{k}}{p^{f(i_{1}+\cdots+i_{k})}},

where we have used (9) to get the last equality. ∎

Now, we calculate the asymptotic frequency of partial quotients being some particular element of ZZ^{*}.

Lemma 22.

If (an)(a_{n}) is a sequence as in the above propositions, then for any zZz\in Z^{*},

limn1n#{1jn:caj(α)=z}=1|z|2m\lim\limits_{n\to\infty}\frac{1}{n}\cdot\#\big{\{}1\leq j\leq n:c_{a_{j}}(\alpha)=z\big{\}}=\frac{1}{|z|^{2m}}

almost everywhere with respect to μ\mu.

Proof.

Applying Proposition 19 with g(α)=χ{z}(c1(α))g(\alpha)=\chi_{\{z\}}\big{(}c_{1}(\alpha)\big{)}, we have

limn1n#{1jn:caj(α)=z}\displaystyle\lim\limits_{n\to\infty}\frac{1}{n}\cdot\#\big{\{}1\leq j\leq n:c_{a_{j}}(\alpha)=z\big{\}} =B(0,1)χ{z}(c1(α))dμ\displaystyle=\int\limits_{B(0,1)}\chi_{\{z\}}\big{(}c_{1}(\alpha)\big{)}\,{\rm d}\mu
=μ({αB(0,1):c1(α)=z})\displaystyle=\mu\big{(}\{\alpha\in B(0,1):c_{1}(\alpha)=z\}\big{)}
=μ(B(1z,|z|2))\displaystyle=\mu\left(B\left(\frac{1}{z},|z|^{-2}\right)\right)
=1|z|2m.\displaystyle=\frac{1}{|z|^{2m}}.

In the following two results, we assume that (an)n=1(a_{n})^{\infty}_{n=1} is an L2L^{2}-good universal sequence of natural numbers such that (anγ)n=1(a_{n}\gamma)^{\infty}_{n=1} is uniformly distributed modulo 11 for any irrational number γ\gamma. The next result is a version of Khinchin’s theorem regarding the geometric mean of the partial quotients in the case of real continued fraction.

Proposition 23.

For almost every αB(0,1)\alpha\in B(0,1) with respect to the Haar measure,

limn1nj=1n(𝔳(caj(α)))=pf(pf1)e.\lim\limits_{n\to\infty}\frac{1}{n}\sum\limits_{j=1}^{n}\big{(}-\mathfrak{v}\big{(}c_{a_{j}}(\alpha)\big{)}\big{)}=\frac{p^{f}}{(p^{f}-1)e}.
Proof.

Applying Proposition 19 to the function g(α)=logp(|c1(α)|)g(\alpha)=\log_{p}\big{(}|c_{1}(\alpha)|\big{)}, we get

limn1nj=1nlogp(|caj|)=B(0,1)logp(|c1(α)|)dμ.\lim\limits_{n\to\infty}\frac{1}{n}\sum\limits_{j=1}^{n}\log_{p}\big{(}|c_{a_{j}}|\big{)}=\int\limits_{B(0,1)}\log_{p}\big{(}|c_{1}(\alpha)|\big{)}\,{\rm d}\mu.

Now,

j=1nlogp(|caj|)=j=1nlogp(p𝔳(caj(α)))=j=1n𝔳(caj(α)).\sum\limits_{j=1}^{n}\log_{p}\big{(}|c_{a_{j}}|\big{)}=\sum\limits_{j=1}^{n}\log_{p}\Big{(}p^{-\mathfrak{v}\big{(}c_{a_{j}}(\alpha)\big{)}}\Big{)}=\sum\limits_{j=1}^{n}-\mathfrak{v}\big{(}c_{a_{j}}(\alpha)\big{)}.

Also,

B(0,1)logp(|c1(α)|)dμ\displaystyle\int\limits_{B(0,1)}\log_{p}\big{(}|c_{1}(\alpha)|\big{)}\,{\rm d}\mu =n=1neμ{αB(0,1):|α|=pne}\displaystyle=\sum\limits_{n=1}^{\infty}\frac{n}{e}\mu\left\{\alpha\in B(0,1):|\alpha|=p^{-\frac{n}{e}}\right\}
=n=1ne(p(ne+1e)mpnem)\displaystyle=\sum\limits_{n=1}^{\infty}\frac{n}{e}\Big{(}p^{\big{(}-\frac{n}{e}+\frac{1}{e}\big{)}m}-p^{-\frac{n}{e}\cdot m}\Big{)}
=n=1pnf(pf1)\displaystyle=\sum\limits_{n=1}^{\infty}p^{-nf}\big{(}p^{f}-1\big{)}
=pf1en=1npnf\displaystyle=\frac{p^{f}-1}{e}\cdot\sum\limits_{n=1}^{\infty}n\cdot p^{-nf}
=pf1epf(1pf)2\displaystyle=\frac{p^{f}-1}{e}\cdot\frac{p^{-f}}{\big{(}1-p^{-f}\big{)}^{-2}}
=pf(pf1)e\displaystyle=\frac{p^{f}}{\big{(}p^{f}-1\big{)}e}

Then the proposition follows. ∎

In the following theorem, we find the asymptotic frequency of partial quotients taking some specified absolute value (or greater or equal to some specified absolute value or absolute values in certain range).

Theorem 24.

For any positive integer ll,
(i)(i) limn1n#{1jn:|caj|=ple}=pf1pfl\lim\limits_{n\to\infty}\frac{1}{n}\cdot\#\left\{1\leq j\leq n:|c_{a_{j}}|=p^{\frac{l}{e}}\right\}=\frac{\displaystyle p^{f}-1}{\displaystyle{p^{f\,l}}}
and (ii)(ii) limn1n#{1jn:|caj|ple}=1pf(l1)\lim\limits_{n\to\infty}\frac{1}{n}\cdot\#\left\{1\leq j\leq n:|c_{a_{j}}|\geq p^{\frac{l}{e}}\right\}=\frac{\displaystyle 1}{\displaystyle{p^{f(l-1)}}} μ\mu-almost everywhere.
If kk is another positive integer with k<lk<l, then
(iii)(iii) limn1n#{1jn:pke|caj|<ple}=1pf(k1)(11pf(lk))\lim\limits_{n\to\infty}\frac{1}{n}\cdot\#\left\{1\leq j\leq n:p^{\frac{k}{e}}\leq|c_{a_{j}}|<p^{\frac{l}{e}}\right\}=\frac{\displaystyle 1}{\displaystyle{p^{f(k-1)}}}\left(1-\frac{\displaystyle 1}{\displaystyle{p^{f(l-k)}}}\right)
μ\mu-almost everywhere.

Proof.

Apply Proposition 19 with g(α)=χ{ple}(|c1(α)|)g(\alpha)=\chi_{\big{\{}p^{\frac{l}{e}}\big{\}}}\big{(}|c_{1}(\alpha)|\big{)} for the proof of (i)(i). Similarly, for the proof of (ii)(ii) and (iii)(iii), apply Proposition 19 with g(α)=χ[ple,)(|c1(α)|)g(\alpha)=\chi_{\big{[}p^{\frac{l}{e}},\infty\big{)}}\big{(}|c_{1}(\alpha)|\big{)} and g(α)=χ[pke,ple)(|c1(α)|)g(\alpha)=\chi_{\big{[}p^{\frac{k}{e}},p^{\frac{l}{e}}\big{)}}\big{(}|c_{1}(\alpha)|\big{)}, respectively. ∎

Given a measure-preserving transformation SS on a probability measure space (X,𝒞,ν)(X,\mathcal{C},\nu), we know that 1nj=0n1g(Sjx)\frac{1}{n}\sum\limits_{j=0}^{n-1}g(S^{j}x) converges almost everywhere. But what happens if we consider moving averages? This means, given a sequence of pairs of positive integers (an,bn)n=1(a_{n},b_{n})^{\infty}_{n=1} what can be said about the convergence of 1bnj=0bn1g(San+jx)\frac{1}{b_{n}}\sum\limits_{j=0}^{b_{n}-1}g(S^{a_{n}+j}x) for almost every xx. In [1], necessary and sufficient conditions were given for this kind of moving averages to converge.

Let Ω\Omega be an infinite collection of points inside ×\mathbb{Z}\times\mathbb{N}. Define

Ωh={(n,k):(n,k)Ωandkh}\displaystyle\Omega^{h}=\{(n,k):(n,k)\in\Omega\ \text{and}\ k\geq h\}
Ωαh={(z,s)2:|zy|<α(sr)for some(y,r)Ωh}\displaystyle\Omega^{h}_{\alpha}=\{(z,s)\in\mathbb{Z}^{2}:|z-y|<\alpha\,(s-r)\ \text{for some}\ (y,r)\in\Omega^{h}\}
Ωαh(λ)={n:(n,λ)Ωαh}.\displaystyle\Omega^{h}_{\alpha}(\lambda)=\{n:(n,\lambda)\in\Omega^{h}_{\alpha}\}.

Following [23], we call a sequence of pairs of natural numbers (an,bn)(a_{n},b_{n}) Stoltz if there exists a function h=h(t)h=h(t) tending to infinity with tt such that

(an,bn)n=tΩh(t)(a_{n},b_{n})^{\infty}_{n=t}\in\Omega^{h(t)}

for some Ω\Omega inside ×\mathbb{Z}\times\mathbb{N}, and \exists l,αl,\alpha and Aα>0A_{\alpha}>0 such that

|Ωαl(λ)|Aα(λ),\left|\Omega^{l}_{\alpha}(\lambda)\right|\leq A_{\alpha}(\lambda),

where |Ωαl(λ)|\left|\Omega^{l}_{\alpha}(\lambda)\right| denotes the cardinality of the set Ωαl(λ)\Omega^{l}_{\alpha}(\lambda).

The following proposition which can be considered as the moving average version of Proposition 19, is the base of the metrical results corresponding to the continued fraction map TT in the context of moving averages. The proof of the proposition is essentially contained in [1].

Proposition 25.

Let (X,𝒞,ν,S)(X,\mathcal{C},\nu,S) be an ergodic dynamical system, and let (an,bn)n=1(a_{n},b_{n})^{\infty}_{n=1} be a Stoltz sequence of natural numbers. Then for any gL1(X,𝒞,ν)g\in L^{1}(X,\mathcal{C},\nu),

limn1bnj=1bng(San+j1α)=Xgdν\lim\limits_{n\to\infty}\frac{1}{b_{n}}\sum\limits_{j=1}^{b_{n}}g\big{(}S^{a_{n}+j-1}\alpha\big{)}=\int\limits_{X}g\,{\rm d}\nu

ν\nu-almost everywhere.

The readers are referred to [23] for some examples of Stoltz sequence, and to [1] for some sequence of pairs of natural numbers for which the assumption of the above proposition fails.

Now we state the metrical results using moving averages. The results are analogous to the results mentioned above, and proofs are similar.

Proposition 26.

Let F:0F:\mathbb{R}_{\geq 0}\to\mathbb{R} and MF,nM_{F,n} be as in Proposition 20, and let (an,bn)(a_{n},b_{n}) be a Stoltz sequence of pairs of natural numbers. Then

limnMF,n(|can+1(α)|,,|can+bn(α)|)=F1(B(0,1)F(|c1(α)|)dμ)\lim\limits_{n\to\infty}M_{F,n}\big{(}|c_{a_{n}+1}(\alpha)|,\ldots,|c_{a_{n}+b_{n}}(\alpha)|\big{)}=F^{-1}\left(\ \int\limits_{B(0,1)}F\big{(}|c_{1}(\alpha)|\big{)}\ {\rm d}\mu\ \right)

almost everywhere with respect to μ\mu.

Proposition 27.

Suppose H:0kH:\mathbb{R}^{k}_{\geq 0}\to\mathbb{R} is a function as in Proposition 21, and (an,bn)(a_{n},b_{n}) be a Stoltz sequence of pairs of natural numbers. Then

limn1bnj=1bnH(|can+j(α)|,,|can+j+k1(α)|)=(i1,,ik)kH(pi1e,,pike)(pf1)kpf(i1++ik)\lim\limits_{n\to\infty}\frac{1}{b_{n}}\sum\limits_{j=1}^{b_{n}}H\big{(}|c_{a_{n}+j}(\alpha)|,\ldots,|c_{a_{n}+j+k-1}(\alpha)|\big{)}=\sum\limits_{(i_{1},\ldots,i_{k})\in\mathbb{N}^{k}}H\left(p^{\frac{i_{1}}{e}},\ldots,p^{\frac{i_{k}}{e}}\right)\frac{\big{(}p^{f}-1\big{)}^{k}}{p^{f(i_{1}+\cdots+i_{k})}}

μ\mu-almost everywhere.

We include the other results in the following theorem.

Theorem 28.

Let (an,bn)(a_{n},b_{n}) be a Stoltz sequence of pairs of natural numbers. We consider all the statements mentioned below in the almost everywhere sense with respect to the measure μ\mu.

(i)For anyzZ,\displaystyle(i)\ \text{For any}\ z\in Z^{*},
limn1bn#{1jbn:can+j(α)=z}=1|z|2m,\displaystyle\ \ \ \lim\limits_{n\to\infty}\frac{1}{b_{n}}\ \#\big{\{}1\leq j\leq b_{n}:c_{a_{n}+j}(\alpha)=z\big{\}}=\frac{1}{|z|^{2m}},
(ii)limn1bnj=1bn𝔳(can+j(α))=pf(pf1)e.\displaystyle(ii)\lim\limits_{n\to\infty}\frac{1}{b_{n}}\sum\limits_{j=1}^{b_{n}}-\mathfrak{v}\big{(}c_{a_{n}+j}(\alpha)\big{)}=\frac{p^{f}}{\big{(}p^{f}-1\big{)}e}.
(iii)For anyl,\displaystyle(iii)\ \text{For any}\ l\in{\mathbb{N}},
limn1bn#{1jbn:|can+j|=ple}=pf1pfl,\displaystyle\ \ \ \ \ \lim\limits_{n\to\infty}\frac{1}{b_{n}}\ \#\left\{1\leq j\leq b_{n}:|c_{a_{n}+j}|=p^{\frac{l}{e}}\right\}=\frac{\displaystyle p^{f}-1}{\displaystyle{p^{f\,l}}},
andlimn1bn#{1jbn:|can+j|ple}=1pf(l1).\displaystyle\ \ \ \ \ \ \text{and}\ \lim\limits_{n\to\infty}\frac{1}{b_{n}}\ \#\left\{1\leq j\leq b_{n}:|c_{a_{n}+j}|\geq p^{\frac{l}{e}}\right\}=\frac{\displaystyle 1}{\displaystyle{p^{f(l-1)}}}.
(iiii)Fork,lwithk<l,\displaystyle(iiii)\ \text{For}\ k,l\in{\mathbb{N}}\ \text{with}\ k<l,
limn1bn#{1jbn:pke|can+j|<ple}=1pf(k1)(11pf(lk)).\displaystyle\ \ \ \ \ \ \lim\limits_{n\to\infty}\frac{1}{b_{n}}\ \#\left\{1\leq j\leq b_{n}:p^{\frac{k}{e}}\leq|c_{a_{n}+j}|<p^{\frac{l}{e}}\right\}=\frac{\displaystyle 1}{\displaystyle{p^{f(k-1)}}}\left(1-\frac{\displaystyle 1}{\displaystyle{p^{f(l-k)}}}\right).
Acknowledgement .

Both the authors are thankful to Dr. L. Singhal for many helpful discussions and some insightful suggestions especially in Lemma 4. He also suggested using the bounds in Lemma 11 in the proof of Proposition 12. Prashant J. Makadiya acknowledges the support of the Government of Gujarat through the SHODH (Scheme of Developing High-Quality Research) fellowship. Prashant J. Makadiya also thanks the Council of Scientific and Industrial Research (CSIR), India for their support through the CSIR-JRF fellowship.

References

  • [1] Alexandra Bellow, Roger Jones, and Joseph Rosenblatt. Convergence for moving averages. Ergodic Theory Dynam. Systems, 10(1):43–62, 1990.
  • [2] Valérie Berthé and Hitoshi Nakada. On continued fraction expansions in positive characteristic: equivalence relations and some metric properties. Expo. Math., 18(4):257–284, 2000.
  • [3] Jerzy Browkin. Continued fractions in local fields. i. Demonstratio Math., 11(1):67–82, 1978.
  • [4] Jerzy Browkin. Continued fractions in local fields. ii. Math. Comp., 70(235):1281–1292, 2001.
  • [5] P. Bundschuh. pp-adische Kettenbrüche und Irrationalität pp-adischer Zahlen. Elem. Math., 32(2):36–40, 1977.
  • [6] Laura Capuano, Nadir Murru, and Lea Terracini. On the finiteness of 𝒫\mathcal{P}-adic continued fractions for number fields. Bull. Soc. Math. France, 150(4):743–772, 2022.
  • [7] Laura Capuano, Nadir Murru, and Lea Terracini. On the finiteness of 𝒫\mathcal{P}-adic continued fractions for number fields. Bull. Soc. Math. France, 150(4):743–772, 2022.
  • [8] Laura Capuano, Nadir Murru, and Lea Terracini. On periodicity of pp-adic Browkin continued fractions. Math. Z., 305(2):Paper No. 17, 24, 2023.
  • [9] Laura Capuano, Francesco Veneziano, and Umberto Zannier. An effective criterion for periodicity of \ell-adic continued fractions. Math. Comp., 88(318):1851–1882, 2019.
  • [10] S. G. Dani. Continued fraction expansions for complex numbers—a general approach. Acta Arith., 171(4):355–369, 2015.
  • [11] S. G. Dani and Arnaldo Nogueira. Continued fractions for complex numbers and values of binary quadratic forms. Trans. Amer. Math. Soc., 366(7):3553–3583, 2014.
  • [12] C. de Vroedt. Metrical problems concerning continued fractions. Compositio Math., 16:191–195 (1964), 1964.
  • [13] Hiromi Ei, Hitoshi Nakada, and Rie Natsui. On the ergodic theory of maps associated with the nearest integer complex continued fractions over imaginary quadratic fields. Discrete Contin. Dyn. Syst., 43(11):3883–3924, 2023.
  • [14] Manfred Einsiedler and Thomas Ward. Ergodic theory with a view towards number theory, volume 259 of Graduate Texts in Mathematics. Springer-Verlag London, Ltd., London, 2011.
  • [15] G. H. Hardy and E. M. Wright. An introduction to the theory of numbers. Oxford University Press, Oxford, sixth edition, 2008. Revised by D. R. Heath-Brown and J. H. Silverman, With a foreword by Andrew Wiles.
  • [16] A. Hurwitz. Über die Entwicklung complexer Grössen in Kettenbrüche (german). Acta Math., 11:187–200, 1887.
  • [17] Svetlana Katok and Ilie Ugarcovici. Theory of (a,b)(a,b)-continued fraction transformations and applications. Electron. Res. Announc. Math. Sci., 17:20–33, 2010.
  • [18] Svetlana Katok and Ilie Ugarcovici. Applications of (a,b)(a,b)-continued fraction transformations. Ergodic Theory Dynam. Systems, 32(2):755–777, 2012.
  • [19] A. Ya. Khinchin. Continued Fractions, volume 49 of The Mathematical Gazette. University of Chicago Press, 1964. https://doi:10.1017/S0025557200053328.
  • [20] Neal Koblitz. pp-adic Numbers, pp-adic Analysis, and Zeta-Functions, volume 58 of Graduate Texts in Mathematics. Springer-Verlag New York, NY, 1984.
  • [21] Vichian Laohakosol. A characterization of rational numbers by pp-adic Ruban continued fractions. J. Austral. Math. Soc. Ser. A, 39(3):300–305, 1985.
  • [22] A. Lasjaunias. A survey of Diophantine approximation in fields of power series. Monatsh. Math., 130(3):211–229, 2000.
  • [23] Poj Lertchoosakul and Radhakrishnan Nair. On the metric theory of continued fractions in positive characteristic. Mathematika, 60(2):307–320, 2014.
  • [24] Poj Lertchoosakul and Radhakrishnan Nair. On the quantitative metric theory of continued fractions in positive characteristic. Proc. Edinb. Math. Soc. (2), 61(1):283–293, 2018.
  • [25] Kurt Mahler. On a geometrical representation of pp-adic numbers. Ann. of Math. (2), 41:8–56, 1940.
  • [26] Nadir Murru and Giuliano Romeo. A new algorithm for pp-adic continued fractions. Math. Comp., 93:1309–1331, 2024.
  • [27] Nadir Murru, Giuliano Romeo, and Giordano Santilli. On the periodicity of an algorithm for pp-adic continued fractions. Ann. Mat. Pura Appl. (4), 202(6):2971–2984, 2023.
  • [28] R. Nair. On the metrical theory of continued fractions. Proc. Amer. Math. Soc., 120(4):1041–1046, 1994.
  • [29] Dinakar Ramakrishnan and Robert J. Valenza. Fourier analysis on number fields, volume 186 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1999.
  • [30] Giuliano Romeo. Continued fractions in the field of pp-adic numbers. Bull. Amer. Math. Soc. (N.S.), 61(2):343–371, 2024.
  • [31] A. A. Ruban. Certain metric properties of the p-adic numbers. Sibirsk. Mat. Ž., 11:222–227, 1970.
  • [32] Wolfgang M. Schmidt. On continued fractions and Diophantine approximation in power series fields. Acta Arith., 95(2):139–166, 2000.
  • [33] Th. Schneider. Über pp-adische Kettenbrüche. In Symposia Mathematica, Vol. IV (INDAM, Rome, 1968/69), pages 181–189. Academic Press, London, 1970.
  • [34] Lian Xiang Wang. pp-adic continued fractions. I, II. Sci. Sinica Ser. A, 28(10):1009–1017, 1018–1023, 1985.