Hostname: page-component-78c5997874-xbtfd Total loading time: 0 Render date: 2024-11-08T09:27:01.021Z Has data issue: false hasContentIssue false

ON p-ADIC INTERPOLATION IN TWO OF MAHLER’S PROBLEMS

Published online by Cambridge University Press:  22 September 2022

BRUNO DE PAULA MIRANDA
Affiliation:
Instituto Federal de Goiás, Avenida Saia Velha, Km 6, BR-040, s/n, Parque Esplanada V, Valparaíso de Goiás, GO 72876-601, Brazil e-mail: [email protected]
JEAN LELIS*
Affiliation:
Faculdade de Matemática, Instituto de Ciências Exatas e Naturais, Universidade Federal do Pará, Belém PA, Brazil
Rights & Permissions [Opens in a new window]

Abstract

Motivated by the p-adic approach in two of Mahler’s problems, we obtain some results on p-adic analytic interpolation of sequences of integers $(u_n)_{n\geq 0}$. We show that if $(u_n)_{n\geq 0}$ is a sequence of integers with $u_n = O(n)$ which can be p-adically interpolated by an analytic function $f:\mathbb {Z}_p\rightarrow \mathbb {Q}_p$, then $f(x)$ is a polynomial function of degree at most one. The case $u_n=O(n^d)$ with $d>1$ is also considered with additional conditions. Moreover, if X and Y are subsets of $\mathbb {Z}$ dense in $\mathbb {Z}_p$, we prove that there are uncountably many p-adic analytic injective functions $f:\mathbb {Z}_p\to \mathbb {Q}_p$, with rational coefficients, such that $f(X)=Y$.

Type
Research Article
Copyright
© The Author(s), 2022. Published by Cambridge University Press on behalf of Australian Mathematical Publishing Association Inc.

1 Introduction

In what follows, p is a prime number, $\mathbb {Q}_p$ is the field of p-adic numbers and $\mathbb {Z}_p$ is the ring of p-adic integers. Let $(u_n)_{n\geq 0}$ be a sequence of integers. If there exists a continuous function $f:\mathbb {Z}_p\rightarrow \mathbb {Q}_p$ such that $f(n)=u_n$ for all nonnegative integers n, we say that f is a p-adic interpolation of $(u_n)_{n\geq 0}$ . In addition, if f is analytic, we say that it is a p-adic analytic interpolation of this sequence. Since the set of nonnegative integers is a dense subset of $\mathbb {Z}_p$ , any given sequence of integers admits at most one such interpolation, which will only exist under certain strong conditions on the sequence (for more details, see [Reference Schikhof17]).

Many authors have studied the problem of p-adic interpolation. Bihani et al. [Reference Bihani, Sheppard and Young2] considered the problem of p-adic interpolation of the Fibonacci sequence, they proved that the sequence $(2^nF_n)_{n\geq 0}$ can be interpolated by a p-adic hypergeometric function on $\mathbb {Z}_5$ . Rowland and Yassawi in [Reference Rowland and Yassawi16] studied p-adic properties of sequences of integers (or p-adic integers) that satisfy a linear recurrence with constant coefficients. For such a sequence, they obtained an explicit approximate twisted interpolation to $\mathbb {Z}_p$ . In particular, they proved that for any prime $p\neq 2$ , there is a twisted interpolation of the Fibonacci sequence by a finite family of p-adic analytic functions with coefficients in some finite extension of $\mathbb {Q}_p$ . Inspired by the Skolem–Mahler–Lech theorem on linear recurrent sequences, Bell [Reference Bell1] proved that for a suitable choice of a p-adic analytic function f and a starting point $\overline {x}$ , the iterate-computing map $n\mapsto f^n(\overline {x})$ extends to a p-adic analytic function g defined for all $x\in \mathbb {Z}_p$ . That is, the sequence $f^n(\overline {x})$ can be interpolated by the p-adic analytic function g.

Mahler [Reference Mahler7] states that the polynomial functions

with $n\geq 0$ integer, form an orthonormal basis, called the Mahler basis, for the space of p-adic continuous functions $\mathcal {C}(\mathbb {Z}_p\rightarrow \mathbb {Q}_p)$ . More precisely, he showed that every continuous function $f:\mathbb {Z}_p\rightarrow \mathbb {Q}_p$ has a unique uniformly convergent expansion

(1.1) $$ \begin{align} f(x)=\sum_{n=0}^{\infty}a_n\binom{x}{n}, \end{align} $$

where $a_n\rightarrow 0$ and $\|f\|_{\scriptsize \mbox {sup}}=\max _{n\geq 0}\|a_n\|_p$ . Conversely, every such expansion defines a continuous function. Furthermore, if $f\in \mathcal {C}(\mathbb {Z}_p\rightarrow \mathbb {Q}_p)$ has a Mahler expansion given by (1.1), then the Mahler coefficients $a_n$ can be reconstructed from f by the inversion formula

(1.2) $$ \begin{align} a_n=\sum_{j=0}^{n}(-1)^{n-j}\binom{n}{j}f(\,j) \quad(n=0,1,2,\ldots). \end{align} $$

Using the Mahler expansion (1.1) and the inversion formula (1.2), we conclude that the sequence $(u_n)_{n\geq 0}$ of integers can be p-adically interpolated if and only if

$$ \begin{align*} \bigg\|\sum_{j=0}^{n}(-1)^{n-j}\binom{n}{j}u_j\bigg\|_p\rightarrow0\quad\mbox{as}\ n\rightarrow\infty. \end{align*} $$

We became interested in studying the p-adic analytic interpolation of sequences of integers with polynomial growth while studying a problem about p-adic Liouville numbers. Based on the classic definition of complex Liouville numbers, Clark [Reference Clark3] called a p-adic integer $\lambda $ a p-adic Liouville number if

$$ \begin{align*} \liminf_{n\rightarrow\infty}\sqrt[n]{\|n-\lambda\|_p}=0. \end{align*} $$

It is easily seen that all p-adic Liouville numbers are transcendental p-adic numbers. Moreover, if $\lambda $ is a p-adic Liouville number and $a,b$ are integers, with $a>0$ , then $a\lambda +b$ is also a p-adic Liouville number.

In his book, Maillet [Reference Maillet10, Ch. III] discusses some arithmetic properties of complex Liouville numbers. One of them states that given a nonconstant rational function f with rational coefficients, if $\xi $ is a Liouville number, then so is $f(\xi )$ . Motivated by this fact, Mahler [Reference Mahler9] posed the following question.

Question 1.1 (Mahler [Reference Mahler9])

Are there transcendental entire functions $f:\mathbb {C}\rightarrow \mathbb {C}$ such that if $\xi $ is any Liouville number, then $f(\xi )$ is also a Liouville number?

He pointed out: ‘The difficulty of this problem lies of course in the fact that the set of all Liouville numbers is nonenumerable.’ We are interested in studying the analogous question for p-adic Liouville numbers.

Question 1.2. Are there p-adic transcendental analytic functions $f:\mathbb {Z}_p\rightarrow \mathbb {Q}_p$ such that if $\lambda $ is a p-adic Liouville number, then so is $f(\lambda )$ ?

It is important to note that the analogue of Maillet’s result is not true for p-adic Liouville numbers. In fact, Lelis and Marques [Reference Lelis and Marques5] proved that the analogue of Maillet’s result is true for a class of p-adic numbers called weak p-adic Liouville numbers, but not for all p-adic Liouville numbers.

Inspired by an argument presented by Marques and Moreira in [Reference Marques and Moreira11] and discussed by Lelis and Marques in [Reference Lelis and Marques6], we approached Question 1.2 as follows. If there were a positive integer sequence $(u_n)_{n\geq 0}$ satisfying $u_n\rightarrow \infty $ and $u_n=O(n)$ that could be interpolated by a p-adic transcendental analytic function $f:\mathbb {Z}_p\rightarrow \mathbb {Q}_p$ , then f would answer Question 1.2 affirmatively. Indeed, assuming all that is true, if we get any p-adic Liouville number $\lambda \in \mathbb {Z}_p$ , by definition there would be a sequence of integers $(n_{k})_{k\geq 0}$ such that

$$ \begin{align*} \lim_{k\rightarrow\infty}\!\sqrt[n_k]{\|n_k-\lambda\|_p}=0. \end{align*} $$

The function f being analytic would satisfy a Lipschitz condition (see [Reference Robert15, Ch. 5, Section 3]). Thus, there would be a constant $c>0$ such that

$$ \begin{align*} \|u_{n_k}-f(\lambda)\|_p=\|f(n_k)-f(\lambda)\|_p\leq c\|n_k-\lambda\|_p, \end{align*} $$

and so

$$ \begin{align*} (\|u_{n_k}-f(\lambda)\|_p)^{1/u_{n_k}}\leq (c\|n_k-\lambda\|_p)^{1/u_{n_k}}, \end{align*} $$

where $u_{n_k}\rightarrow \infty $ and $u_{n_k}=O(n_k)$ . So $f(\lambda )$ would also be a p-adic Liouville number.

In light of this, it is natural to try to characterise the p-adic analytic functions which interpolate sequences of integers $(u_n)_{n \geq 0}$ of linear growth. There are other reasons for seeking such characterisations. Indeed, one may ask whether there exists a p-adic interpolation of some arithmetic function (many of which have linear growth) or, more generally, if polynomials with integer coefficients are the only p-adic analytic functions that take positive integers into positive integers with polynomial order.

Theorem 1.3. Let $(u_n)_{n\geq 0}$ be a sequence of positive integers such that $u_n=O(n^d)$ for some fixed $d\geq 0 \ (d\in \mathbb {R})$ . Assume there exists a p-adic analytic function $f:\mathbb {Z}_p\rightarrow \mathbb {Q}_p$ which interpolates the sequence $(u_n)_{n\geq 0}$ .

  1. (i) If $d\leq 1$ , then f is a polynomial function of degree at most one.

  2. (ii) If $d>1$ and the Mahler expansion of f converges for all $x\in \mathbb {Q}_p$ , then f is a polynomial function of degree at most $\lfloor d\rfloor $ .

We remark that the condition ‘f is a p-adic analytic function on $\mathbb {Z}_p$ ’ is fundamental in the result above. Indeed, if we write $n=\sum _{i=0}^{k}a_ip^i$ in base p, then the function $f:\{0\}\cup \mathbb {N}\to \mathbb {Q}_p$ given by

$$ \begin{align*} f(n)=\begin{cases} \displaystyle\sum_{i=0}^{k-1}a_ip^i&\mbox{if}\ n\geq p,\\ n&\mbox{if}\ 0\leq n\leq p-1, \end{cases} \end{align*} $$

clearly can be extended in a unique way to a continuous function $\overline {f}:\mathbb {Z}_p\to \mathbb {Q}_p$ such that $\overline {f}(n)=O(n)$ . However, $\overline {f}$ is nonanalytic and it is clearly not a polynomial function.

Moreover, consider the p-adic function $f_d:\mathbb {Z}_p\to \mathbb {Q}_p$ defined by

$$ \begin{align*} f_d(z)=\sum_{k=0}^{\infty}a_kp^{dk}, \end{align*} $$

where $z=\sum _{k=0}^{\infty }a_kp^k$ is the p-adic expansion of $z\in \mathbb {Z}_p$ . Then it is well known that $f_d$ is a continuous function for all integers $d\geq 2$ . In fact, if $d\geq 2$ is an integer, then

$$ \begin{align*} \|f_d(x)-f_d(y)\|_p\leq\|x-y\|_p^d. \end{align*} $$

In particular, we have $f^{\prime }_d(x)=0$ for all $x\in \mathbb {Q}_p$ and $f_d\in \mathcal {C}^1(\mathbb {Z}_p\to \mathbb {Q}_p)\subset \mathcal {C}(\mathbb {Z}_p\to \mathbb {Q}_p)$ . Note that $f_d(n)=O(n^d)$ , but $f_d$ is not a polynomial function. However, since $f_d$ is not a p-adic analytic function, its Mahler expansion does not converge for all $x\in \mathbb {Q}_p$ .

Very strict conditions must be satisfied for a sequence $(u_n)_{n\geq 0}$ to be interpolated by a p-adic analytic function. However, if the set $A=\{u_0,u_1,\ldots \}\subseteq \mathbb {Z}$ is a dense subset of $\mathbb {Z}_p$ , one may ask whether there is some re-enumeration $\sigma :\{0\}\cup \mathbb {N}\rightarrow \{0\}\cup \mathbb {N}$ such that $(u_{\sigma (n)})_{n\geq 0}$ can be interpolated by a p-adic analytic function.

In the complex case, Georg [Reference Georg4] established that for each countable subset $X\subset \mathbb {C}$ and each dense subset $Y\subseteq \mathbb {C}$ , there exists a transcendental entire function f such that $f(X)\subset Y$ . In 1902, Stäckel [Reference Stäckel18] used another construction to show that there is a function $f(z)$ , analytic in a neighbourhood of the origin and with the property that both $f(z)$ and its inverse function assume, in this neighbourhood, algebraic values at all algebraic points. Based on these results, Mahler [Reference Mahler8] suggested the following question about the set of algebraic numbers $\overline {\mathbb {Q}}$ .

Question 1.4 (Mahler, [Reference Mahler8])

Are there transcendental entire functions $f(z)=\sum c_nz^n$ with rational coefficients $c_n$ and such that $f(\overline {\mathbb {Q}})\subset \overline {\mathbb {Q}}$ and $f^{-1}(\overline {\mathbb {Q}})\subset \overline {\mathbb {Q}}$ ?

This question was answered positively by Marques and Moreira [Reference Marques and Moreira12]. Moreover, in a more recent paper [Reference Marques and Moreira13], they proved that if X and Y are countable subsets of $\mathbb {C}$ satisfying some conditions necessary for analyticity, then there are uncountably many transcendental entire functions $f(z)=\sum a_nz^n$ with rational coefficients such that $f(X)\subset Y$ and $f^{-1}(Y)\subset X$ . Keeping these results in mind, we prove the following theorem.

Theorem 1.5. Let X and Y be subsets of $\mathbb {Z}$ dense in $\mathbb {Z}_p$ . Then there are uncountably many p-adic analytic injective functions $f:\mathbb {Z}_p\to \mathbb {Q}_p$ with

$$ \begin{align*} f(x)=\sum_{n=0}^{\infty} c_nx^n\in\mathbb{Q}[[x]] \end{align*} $$

such that $f(X)=Y$ .

Note that by Theorem 1.5, if $Y=\{y_0,y_1,y_2,\ldots \}\subset \mathbb {Z}$ is a dense subset of $\mathbb {Z}_p$ , that is, if Y contains a complete system of residues modulo any power of p, then there is a p-adic analytic function

$$ \begin{align*} f(x)=\sum_{n=0}^{\infty}c_nx^n,\quad c_n\in\mathbb{Q} \mbox{ for all}\ n\geq0, \end{align*} $$

and a bijection $\sigma :\{0\}\cup \mathbb {N}\rightarrow \{0\}\cup \mathbb {N}$ such that $f(n)=u_{\sigma (n)}$ , where we take $X=\{0\}\cup \mathbb {N}$ . Moreover, the series above converges for all $x\in \mathbb {Z}_p$ . Thus, if we consider the Mahler expansion, then we immediately obtain the following result.

Corollary 1.6. Let $Y=\{y_0,y_1,y_2,\ldots \}$ be a subset of $\mathbb {Z}$ dense in $\mathbb {Z}_p$ . Then there are $a_0,a_1,a_2,\ldots \in \mathbb {Z}$ and a bijection $\sigma :\{0\}\cup \mathbb {N}\rightarrow \{0\}\cup \mathbb {N}$ such that

$$ \begin{align*} \sum_{i=0}^{n}a_i\binom{i}{n}=y_{\sigma(n)}, \end{align*} $$

for all integers $n\geq 0$ , where $v_p(a_n/n!)\to \infty $ as $n\to \infty $ .

We end this section by presenting some questions which we are still unable to answer. One may ask whether Theorem 1.5 is still true if X and Y are free to contain elements outside $\mathbb {Z}$ . What could one do to guarantee rational coefficients in f in a situation like that? Moreover, if we consider the algebraic closure of $\mathbb {Q}_p$ , denoted by $\overline {\mathbb {Q}}_p$ , and its completion $\mathbb {C}_p$ , we may ask a probably more difficult question.

Question 1.7. Are there p-adic transcendental entire functions $f:\mathbb {C}_p\to \mathbb {C}_p$ given by

$$ \begin{align*} f(z)=\sum_{n=0}^{\infty} c_nz^n,\quad c_n\in\mathbb{Q} \mbox{ for all}\ n\geq0, \end{align*} $$

such that $f(\overline {\mathbb {Q}}_p)\subset \overline {\mathbb {Q}}_p$ and $f^{-1}(\overline {\mathbb {Q}}_p)\subset \overline {\mathbb {Q}}_p$ ?

Naturally, the main difficulty of this problem lies again in the fact that the set $\overline {\mathbb {Q}}_p$ is uncountable.

2 Proof of Theorem 1.3

We start by introducing the classic Strassmann’s theorem about zeros of p-adic power series. This result says that a p-adic analytic function with coefficients in $\mathbb {Q}_p$ has finitely many zeros in $\mathbb {Z}_p$ and provides a bound for the number of zeros.

Theorem 2.1 (Strassmann, [Reference Murty14])

Let $f(x)=\sum _{n=0}^{\infty }c_nx^n$ be a nonzero power series with coefficients in $\mathbb {Q}_p$ and suppose that $\lim _{n\rightarrow \infty }c_n=0$ so that $f(x)$ converges for all x in $\mathbb {Z}_p$ . Let N be the integer defined by conditions

$$ \begin{align*} \|c_N\|_p=\max\|c_n\|_p\quad\mbox{and}\quad \|c_n\|_p<\|c_N\|_p\quad\mbox{for all}\ n>N. \end{align*} $$

Then the function $f:\mathbb {Z}_p\rightarrow \mathbb {Q}_p$ defined by $x\mapsto f(x)$ has at most N zeros.

Proof of Theorem 1.3

Let $(u_n)_{n\geq 0}$ be a sequence of integers of linear or sublinear growth, that is, $u_n=O(n)$ . Suppose that $(u_n)_{n\geq 0}$ can be interpolated by some p-adic analytic function

$$ \begin{align*} f(x)=\sum_{n=0}^{\infty}c_nx^n\in\mathbb{Q}_p[[x]]. \end{align*} $$

Since $f(x)$ is a p-adic analytic function, $\lim _{n\rightarrow \infty }\|c_n\|_p=0$ . Thus, there exists an integer N defined by the conditions

$$ \begin{align*} \|c_N\|_p=\max\|c_n\|_p\quad\mbox{and}\quad \|c_n\|_p<\|c_N\|_p\quad\mbox{for all}\ n>N, \end{align*} $$

and Strassman’s theorem guarantees that the function $f:\mathbb {Z}_p\rightarrow \mathbb {Q}_p$ has at most N zeros.

By hypothesis, $u_n=O(n)$ , so there is a $C>0$ such that $0 < u_n\leq Cn$ for all $n\geq 0$ . Taking the subsequence $(u_{p^k})_{k\geq 0}$ ,

(2.1) $$ \begin{align} 0<u_{p^k}\leq Cp^k. \end{align} $$

Since f is an analytic function, it is easily seen that it satisfies the Lipschitz condition

$$ \begin{align*} \|f(x)-f(y)\|_p\leq \|x-y\|_p \end{align*} $$

for all $x,y \in \mathbb {Z}_p$ . In particular,

$$ \begin{align*} \|u_{p^k}-u_0\|_p=\|f(p^k)-f(0)\|_p\leq \|p^k\|_p, \end{align*} $$

and it follows that

(2.2) $$ \begin{align} u_{p^k}=u_0+t_kp^{k} \end{align} $$

with $t_k\in \mathbb {Z}_+$ , because $u_{p^k}$ is a positive integer. By (2.1) and (2.2), we conclude that $0\leq t_k\leq C$ . Hence, by the pigeonhole principle, there exists an integer t with $0\leq t\leq C$ such that

$$ \begin{align*} u_{p^{\kern1.5pt j}}=u_0+tp^{\,j} \end{align*} $$

for infinitely many $j\geq 0$ . Thus, the function

$$ \begin{align*} f(x)-u_0-tx=(c_1-t)x+\sum_{n=2}^{\infty}c_nx^n \end{align*} $$

has infinitely many roots and by Strassman’s theorem, we conclude that $f(x)=u_0+tx$ .

Now suppose that $u_n=O(n^d)$ for some fixed positive real number $d>1$ . Let

$$ \begin{align*} f(x)=\sum_{n=0}^{\infty}a_n\binom{x}{n} \end{align*} $$

be the Mahler expansion of f. By hypothesis, the Mahler expansion of f converges for all $x\in \mathbb {Q}_p$ , so the function $x\mapsto \sum _{n=0}^{\infty }a_n\binom {x}{n}$ is analytic on $\mathbb {C}_p$ and

$$ \begin{align*} \lim_{n\rightarrow\infty}r^n\|a_n\|_p=0 \end{align*} $$

for all real numbers $r>0$ (see [Reference Schikhof17, Ch. 3]). Taking $r=p^2$ , we find $v_p(a_n)\geq 2n$ for all n sufficiently large. Moreover, $a_n$ is an integer for all $n\geq 0$ . In fact, by the Mahler expansion,

$$ \begin{align*} a_n=\sum_{j=0}^{n}(-1)^{n-j}\binom{n}{j}f(\kern2pt j)=\sum_{j=0}^{n}(-1)^{n-j}\binom{n}{j}u_j\quad(n=0,1,2,\ldots), \end{align*} $$

where $u_j\in \mathbb {Z}_+$ for all $j\geq 0$ . Hence, either $a_n=0$ or

(2.3) $$ \begin{align} \|a_n\|_{\infty}\geq p^{2n}. \end{align} $$

However,

$$ \begin{align*} \|a_n\|_{\infty}=\bigg\|\sum_{j=0}^{n}(-1)^{n-j\,}\binom{n}{j}u_j\bigg\|_{\infty} \quad(n=0,1,2,\ldots). \end{align*} $$

Since $\|u_j\|_{\infty }\leq j^d\leq n^d$ for all $j\leq n$ , it follows that

(2.4) $$ \begin{align} \|a_n\|_{\infty}\leq Dn^d2^n, \end{align} $$

where $D>0$ is a fixed constant. It is easily seen that (2.3) and (2.4) cannot both be true for n sufficiently large. Hence, there exists an $N>0$ such that $a_n=0$ for all $n>N$ . Consequently, f is a polynomial function. Furthermore, $f(n)=O(n^d)$ , so its degree must be at most $\lfloor d\rfloor $ .

3 Proof of Theorem 1.5

Suppose that $X=\{x_0,x_1,x_2,\ldots \}$ and $Y=\{y_0,y_1,y_2,\ldots \}$ are subsets of $\mathbb {Z}$ dense in $\mathbb {Z}_p$ . Our proof consists in determining a sequence of polynomial functions $f_0,f_1,\ldots $ such that $f_n\to f$ as $n\to \infty $ , where f is a p-adic analytic injective function on $\mathbb {Z}_p$ with rational coefficients satisfying $f(X)=Y$ . In addition, we will show that there are uncountably many such functions.

To be more precise, we will construct a sequence of polynomial functions $f_0,f_1,f_2,\ldots \in \mathbb {Q}[x]$ of degrees $t_0,t_1,t_2,\ldots \in \mathbb {Z}$ , respectively, such that for all $m\geq 0$ ,

(3.1) $$ \begin{align} f_m(x)=\sum_{i=0}^{t_m}c_ix^i, \end{align} $$

where $c_0=y_0-x_0$ , $c_1=1$ and $\|c_i\|_p\leq p^{-1}$ for all $2\leq i\leq t_m$ . Furthermore, our sequence will obey the recurrence relation

(3.2) $$ \begin{align} f_{m+1}(x)=f_m(x)+x^{t_m+1}P_m(x)(\delta_m+\epsilon_m(x-x_{m+1})), \end{align} $$

where the polynomial functions $P_m\in \mathbb {Z}[x]$ are given by

(3.3) $$ \begin{align} P_m(x)=\prod_{k\in X_m\cup Y^{-1}_m}(x-k), \end{align} $$

with $X_m=\{x_0,\ldots ,x_m\}$ and $Y^{-1}_m=f_m^{-1}(\{y_0,\ldots ,y_m\})$ , and $\delta _m$ and $\epsilon _m$ are rational numbers such that

$$ \begin{align*} \max\{\|\delta_m\|_p,\|\epsilon_m\|_p\}\leq p^{-m}. \end{align*} $$

Finally, our sequence will also satisfy $f_m(x_k)\in Y$ and $f_m^{-1}(\{y_k\})\cap X \neq \emptyset $ for all ${0\leq k\leq m}$ .

We make some remarks regarding such a sequence of polynomials. First, since $f_m$ is a polynomial, $Y^{-1}_m$ must be a finite subset of $\mathbb {Z}_p$ for each m, so the polynomials $P_m$ are well defined. Second, by (3.1), $\|c_1\|_p> \|c_i\|_p$ for all $i \geq 2$ , so each $f_m$ is necessarily injective on $\mathbb {Z}_p$ by Strassmann’s theorem. Lastly, since $f_m$ is injective, there is only one $x_s \in X \cap f_m^{-1}(\{y_k\})$ . The existence of such a sequence is guaranteed by the following lemma.

Lemma 3.1. Suppose that $ f_m(x)=c_0+c_1x+\cdots +c_{t_m}x^{t_m}\in \mathbb {Q}[x] $ is a polynomial with

$$ \begin{align*} \|c_i\|_p<\|c_1\|_p\quad \mbox{for}\ 2\leq i\leq t_m\in \mathbb{Z}, \end{align*} $$

such that $f_m(X_m)\subset Y$ and $Y_m^{-1}\subset X$ . Then there exist rational numbers $\delta _m$ and $\epsilon _m$ with

$$ \begin{align*} \max\{\|\delta_m\|_p,\|\epsilon_m\|_p\}\leq p^{-m} \end{align*} $$

such that the function

$$ \begin{align*} f_{m+1}(x)=f_m(x)+x^{t_m+1}P_m(x)(\delta_m+\epsilon_m(x-x_{m+1})) \end{align*} $$

is a polynomial given by

$$ \begin{align*} f_{m+1}(x)=c_0+c_1x+\cdots+c_{t_{m+1}}x^{t_{m+1}}\in \mathbb{Q}[x] \end{align*} $$

satisfying $f_{m+1}(X_{m+1})\subset X$ and $Y_{m+1}^{-1}\subset X$ and, moreover, $\|c_i\|_p<\|c_1\|_p$ for all integers i with $2\leq i\leq t_{m+1}$ .

Proof. Suppose that for some $m\geq 0$ , there is a function $f_m$ satisfying the hypotheses of the lemma. We will show that we can choose rational numbers $\delta _m$ and $\epsilon _m$ such that

$$ \begin{align*} \max\{\|\delta_m\|_p,\|\epsilon_m\|_p\}\leq p^{-m} \end{align*} $$

in such a way that the polynomial $f_{m+1}$ in (3.2) has the desired properties.

First, we will determine $\delta _m\in \mathbb {Q}$ such that $f_{m+1}(x_{m+1})\in Y$ . Suppose that $ f_m(x_{m+1}) \in \{y_0,y_1,\ldots ,y_m\}$ . Since $P_m(x_{m+1})=0$ , we have $f_{m+1}(x_{m+1}) = f_{m}(x_{m+1})\in Y$ . Note that here we did not make direct use of $\delta _m$ to get $f_{m+1}(x_{m+1}) \in Y$ . So we are free to choose any $\delta _m \in \mathbb {Q}$ and we do so by setting $\delta _m = p^m$ . Now, suppose that $f_m(x_{m+1})\notin \{y_0,y_1,\ldots ,y_m\}$ , which implies that $P_m(x_{m+1})\neq 0$ . Since Y is a dense subset of $\mathbb {Z}_p$ , there exists $\hat {y}\in Y$ such that

$$ \begin{align*} 0<\bigg\|\frac{\hat{y}-f_m(x_{m+1})}{(x_{m+1})^{t_m+1}P_m(x_{m+1})}\bigg\|_p\leq p^{-m}. \end{align*} $$

Then, taking

$$ \begin{align*} \delta_m=\frac{\hat{y}-f_m(x_{m+1})}{(x_{m+1})^{t_m+1}P_m(x_{m+1})}, \end{align*} $$

we obtain $f_{m+1}(x_{m+1})=\hat {y}\in Y$ independently of $\epsilon _m$ . Observe that in both cases just analysed, $\|\delta _m\|_p \leq p^{-m}$ .

Now we will choose $\epsilon _m \in \mathbb {Q}$ to get $f_{m+1}(\hat {x})=y_{m+1}$ for some $\hat {x}\in X$ . Since $f_m$ is injective on $\mathbb {Z}_p$ , there is at most one $\hat {x}\in X$ such that $f_m(\hat {x})=y_{m+1}$ . If there exists $\hat {x}\in X_m$ such that $f_m(\hat {x})=y_{m+1}$ , then $P_m(\hat {x})=0$ and we obtain $f_{m+1}(\hat {x})=y_{m+1}$ . In this case, $\epsilon _m$ does not play a role and we are free to set $\epsilon _m=p^m$ . It remains to consider the case where there is no $\hat {x} \in X_m$ with $f_m(\hat {x})=y_{m+1}$ . Note that if we choose

$$ \begin{align*} \delta_m=\frac{y_{m+1}-f_m(x_{m+1})}{(x_{m+1})^{t_m+1}P_m(x_{m+1})}, \end{align*} $$

then $f_{m+1}(x_{m+1})=y_{m+1}$ and we have $\hat {x} = x_{m+1}$ . Since we again did not use $\epsilon _m$ to ensure that $f_{m+1}(x_{m+1})=y_{m+1}$ , we are free to take $\epsilon _m=p^m$ . However, if

$$ \begin{align*} \delta_m\neq\frac{y_{m+1}-f_m(x_{m+1})}{(x_{m+1})^{t_m+1}P_m(x_{m+1})}, \end{align*} $$

we consider the polynomial equation

$$ \begin{align*} f_m(x)+\delta_mx^{t_m+1}P_m(x)=y_{m+1}. \end{align*} $$

Since $\|\delta _m\|_p\leq p^{-m}$ and $\|c_i\|_p<p^{-1}$ for $i\geq 2$ ,

$$ \begin{align*} f_m(x)+\delta_mx^{t_m+1}P_m(x)-y_{m+1}\equiv y_0+x-y_{m+1}\pmod{p\mathbb{Z}_p} \end{align*} $$

for all $m\geq 2$ . Thus, the congruence

$$ \begin{align*} f_m(x)+\delta_mx^{t_m+1}P_m(x)-y_{m+1}\equiv 0\pmod{p\mathbb{Z}_p} \end{align*} $$

has a solution $\overline {x} \equiv y_{m+1}-y_0 \pmod {p\mathbb {Z}_p}$ . Moreover, taking the formal derivative,

$$ \begin{align*} [f_m(x)+\delta_mx^{t_m+1}P_m(x)-y_{m+1}]'\equiv [y_0+x-y_{m+1}]'\equiv 1 \pmod{p\mathbb{Z}_p}. \end{align*} $$

Hence, by Hensel’s lemma [Reference Murty14], there exists $b\in \mathbb {Z}_p$ such that

$$ \begin{align*} f_m(b)+\delta_mb^{t_m+1}P_m(b)=y_{m+1}. \end{align*} $$

Let $v_p(x)$ be the p-adic valuation of $x\in \mathbb {Z}_p$ and take

$$ \begin{align*} s=v_p(b^{t_m+1}P_m(b)(b-x_{m+1})). \end{align*} $$

Note that $s<+\infty $ , since $P_m(b)(b-x_{m+1})\neq 0$ . Thus, we have a Lipschitz condition on $\mathbb {Z}_p$ , namely

$$ \begin{align*} \|f_m(x)+\delta_mx^{t_m+1}P_m(x)-f_m(y)+\delta_my^{t_m+1}P_m(y)\|_p\leq \|x-y\|_p \end{align*} $$

for all $x,y\in \mathbb {Z}_p$ . Since X is a dense subset of $\mathbb {Z}_p$ , there is an integer $\hat {x}\in X$ such that

$$ \begin{align*} \|\hat{x}-b\|_p\leq\frac{1}{p^{s+m}} \end{align*} $$

and $v_p(\hat {x}^{t_m+1}P_m(\hat {x})(\hat {x}-x_{m+1}))=s$ . So,

$$ \begin{align*} \|f_m(\hat{x})+\delta_m\hat{x}^{t_m+1}P_m(\hat{x})-y_{m+1}\|_p\leq \frac{1}{p^{s+m}}. \end{align*} $$

Taking

$$ \begin{align*} \epsilon_m=\frac{y_{m+1}-f_m(\hat{x})-\delta_m\hat{x}^{t_m+1}P_m(\hat{x})}{\hat{x}^{t_m+1}P_m(\hat{x})(\hat{x}-x_{m+1})}, \end{align*} $$

we get $\epsilon _m\in \mathbb {Q}$ , $\|\epsilon _m\|_p<p^{-m}$ and $f_{m+1}(\hat {x})=y_{m+1}$ . This completes the proof of the lemma.

Proof of Theorem 1.5

If in Lemma 3.1 we start with $f_0(x)=(x-x_0)+y_0$ , we get a sequence of polynomials as described in the beginning of this section. Furthermore, in each step, we have at least two options for the choice of $\delta _m$ and $\epsilon _m$ so we get uncountably many sequences. We will fix one of these sequences and prove that $f(x)=\lim _{m\rightarrow \infty }f_m(x)$ solves Theorem 1.5. Indeed,

$$ \begin{align*} f_m(x)=y_0+(x-x_0)+\sum_{j=1}^{m-1}x^{t_j+1}P_j(x)[\delta_j+\epsilon_j(x-x_{j+1})]=\sum_{j=0}^{t_m}c_jx^j, \end{align*} $$

where $\|c_i\|_p\leq p^{-j}$ for $t_{j-1}<i\leq t_j$ and $1\leq j\leq m$ (since $\max \{\|\delta _j\|_p,\|\epsilon _j\|_p\}\leq p^{-j}$ ). Therefore, $\lim _{i\rightarrow \infty }\|c_i\|_p=0$ and

$$ \begin{align*} f(x)=\lim_{m\rightarrow\infty}f_m(x) \end{align*} $$

is a p-adic analytic function on $\mathbb {Z}_p$ .

Moreover, $f(X)= Y$ . Indeed, we are assuming that $f_k(x_k)\in Y$ . By (3.3), ${P_m(x_k)=0}$ for all $m \geq k \geq 0$ and, consequently, $f_m(x_k)=f_{m-1}(x_k)=\cdots =f_k(x_k)$ . Thus, we conclude that

$$ \begin{align*} f(x_k)=\lim_{m\rightarrow\infty}f_m(x_k)=f_k(x_k)\in Y. \end{align*} $$

However, by hypothesis, given an integer $j\geq 0$ , there exists an integer $s\geq 0$ such that $f_j(x_s)=y_j$ . Similarly,

$$ \begin{align*} f(x_s)=\lim_{m\rightarrow\infty}f_m(x_s)=f_j(x_s)=y_j\in Y \end{align*} $$

and we conclude $f(X)= Y$ .

It remains to prove that f is injective. For this, suppose that there are $a_1$ and $a_2$ in $\mathbb {Z}_p$ such that $f(a_1)=f(a_2)=b\in \mathbb {Z}_p$ and note that by (3.1), $c_1=1$ satisfies

(3.4) $$ \begin{align} \|c_1\|_p=\max\|c_j\|_p\quad\mbox{and}\quad \|c_j\|_p<\|c_1\|_p\quad\mbox{for all } j>1. \end{align} $$

Now, consider the function

$$ \begin{align*} f(x)-b=(y_0-x_0-b)+x+\sum_{n=2}^{\infty}c_nx^n. \end{align*} $$

Note that in the equation above, $c_1=1$ still satisfies the conditions in (3.4). Hence, $f(x)-b$ has at most one zero (by Strassman’s theorem), so we have $a_1=a_2$ .

References

Bell, J. P., ‘A generalised Skolem–Mahler–Lech theorem for affine varieties’, J. Lond. Math. Soc. (2) 73(2) (2006), 367379.CrossRefGoogle Scholar
Bihani, P., Sheppard, W. P. and Young, P. T., ‘ $p$ -Adic interpolation of the Fibonacci sequence via hypergeometric functions’, Fibonacci Quart. 43(3) (2005), 213226.Google Scholar
Clark, D., ‘A note on the $p$ -adic convergence of the solutions of linear differential equations’, Proc. Amer. Math. Soc. 17 (1966), 262269.Google Scholar
Georg, F., ‘Ueber arithmetische Eingenschaften analytischer Functionen’, Math. Ann. 58(4) (1904), 545557.Google Scholar
Lelis, J. and Marques, D., ‘Some results on arithmetic properties of $p$ -adic Liouville numbers’. p-Adic Numbers Ultrametric Anal. Appl. 11(3) (2019), 216222.CrossRefGoogle Scholar
Lelis, J. and Marques, D., ‘On transcendental entire functions mapping $\mathbb{Q}$ into itself’, J. Number Theory 206 (2020), 310319.CrossRefGoogle Scholar
Mahler, K., ‘An interpolation series for continuous functions of a $p$ -adic variable’, J. reine angew. Math. 199 (1958), 2334.CrossRefGoogle Scholar
Mahler, K., Lectures on Transcendental Numbers, Lecture Notes in Mathematics, 546 (Springer-Verlag, Berlin, 1976).CrossRefGoogle Scholar
Mahler, K., ‘Some suggestions for further research,’ Bull. Aust. Math. Soc. 29(1) (1984), 101108.CrossRefGoogle Scholar
Maillet, E., Introduction à la Théorie des Nombres Transcendants et des Propriétés Arithmétiques des Fonctions (Gauthier–Villars, Paris, 1906).Google Scholar
Marques, D. and Moreira, C. G., ‘On a variant of a question proposed by K. Mahler concerning Liouville numbers’, Bull. Aust. Math. Soc. 91 (2015), 2933.CrossRefGoogle Scholar
Marques, D. and Moreira, C. G., ‘A positive answer for a question proposed by K. Mahler’, Math. Ann. 368 (2017), 10591062.CrossRefGoogle Scholar
Marques, D. and Moreira, C. G., ‘On a stronger version of a question proposed by K. Mahler’, J. Number Theory 194 (2019), 372380.CrossRefGoogle Scholar
Murty, M. R., Introduction to $p$ -Adic Analytic Number Theory, Studies in Advanced Mathematics, 27 (American Mathematical Society, Providence, RI, 2009).Google Scholar
Robert, A. M., A Course in $p$ -Adic Analysis, Graduate Texts in Mathematics, 198 (Springer, New York, 2013).Google Scholar
Rowland, E. and Yassawi, E., ‘ $p$ -Adic asymptotic properties of constant-recursive sequences’, Indag. Math. (N.S.) 28(1) (2017), 205220.CrossRefGoogle Scholar
Schikhof, W. H., Ultrametric Calculus (Cambridge University Press, Cambridge, 2006).Google Scholar
Stäckel, P., ‘Arithmetische eingenschaften analytischer Functionen’, Acta Math. 25 (1902), 371383.CrossRefGoogle Scholar