\documentclass[12pt,reqno]{article} \usepackage[usenames]{color} \usepackage{amssymb} \usepackage{amsmath} \usepackage{amsthm} \usepackage{amsfonts} \usepackage{amscd} \usepackage{graphicx} \usepackage[colorlinks=true, linkcolor=webgreen, filecolor=webbrown, citecolor=webgreen]{hyperref} \definecolor{webgreen}{rgb}{0,.5,0} \definecolor{webbrown}{rgb}{.6,0,0} \usepackage{color} \usepackage{fullpage} \usepackage{float} \usepackage{psfig} \usepackage{graphics} \usepackage{latexsym} \usepackage{epsf} \usepackage{breakurl} \setlength{\textwidth}{6.5in} \setlength{\oddsidemargin}{.1in} \setlength{\evensidemargin}{.1in} \setlength{\topmargin}{-.1in} \setlength{\textheight}{8.4in} \newcommand{\seqnum}[1]{\href{https://oeis.org/#1}{\underline{#1}}} \def\modd#1 #2{#1\ \mbox{\rm (mod}\ #2\mbox{\rm)}} \begin{document} \begin{center} \epsfxsize=4in \leavevmode\epsffile{logo129.eps} \end{center} \theoremstyle{plain} \newtheorem{theorem}{Theorem} \newtheorem{corollary}[theorem]{Corollary} \newtheorem{lemma}[theorem]{Lemma} \newtheorem{proposition}[theorem]{Proposition} \theoremstyle{definition} \newtheorem{definition}[theorem]{Definition} \newtheorem{example}[theorem]{Example} \newtheorem{conjecture}[theorem]{Conjecture} \theoremstyle{remark} \newtheorem{remark}[theorem]{Remark} \begin{center} \vskip 1cm{\LARGE\bf On Second-Order Linear Sequences of \\ \vskip .1in Composite Numbers } \vskip 1cm \large Dan Ismailescu\\ Department of Mathematics \\ Hofstra University\\ Hempstead, NY 11549\\ USA\\ \href{mailto:dan.p.ismailescu@hofstra.edu}{\tt dan.p.ismailescu@hofstra.edu} \\ \ \\ Adrienne Ko\\ Ethical Culture Fieldston School\\ Bronx, NY 10471\\ USA\\ \href{mailto:19ako@ecfs.org}{\tt 19ako@ecfs.org}\\ \ \\ Celine Lee\\ Chinese International School\\ Hong Kong SAR\\ China\\ \href{mailto:celinl@student.cis.edu.hk}{\tt celinl@student.cis.edu.hk}\\ \ \\ Jae Yong Park\\ The Lawrenceville School\\ Lawrenceville, NJ 08648 \\ USA\\ \href{mailto:jpark19@lawrenceville.org}{\tt jpark19@lawrenceville.org} \end{center} \newpage \begin{abstract} We present a new proof of the following result of Somer: {\emph {Let $(a,b)\in \mathbb{Z}^2$ and let $(x_n)_{n\ge 0}$ be the sequence defined by some initial values $x_0$ and $x_1$ and the second-order linear recurrence \begin{equation*} x_{n+1}=ax_n+bx_{n-1} \end{equation*} for $n\ge 1$. Suppose that $b\neq 0$ and $(a,b)\neq (2,-1), (-2, -1)$. Then there exist two relatively prime positive integers $x_0$, $x_1$ such that $| x_n|$ is a composite integer for all $n\in \mathbb{N}$.}} The above theorem extends a result of Graham, who solved the problem when $(a,b)=(1,1)$. \end{abstract} \section{Introduction} We give a new proof of the following result of Somer: \begin{theorem}\label{maintheorem}\cite{somer, dubickas} Let $(a,b)\in \mathbb{Z}^2$ and let $(x_n)_{n\ge 0}$ be the sequence defined by some initial values $x_0$ and $x_1$ and the second-order linear recurrence \begin{equation}\label{recurrence} x_{n+1}=ax_n+bx_{n-1} \end{equation} for $n\ge 1$. Suppose that $b\neq 0$ and $(a,b)\neq (2,-1), (-2, -1)$. Then there exist two relatively prime positive integers $x_0$, $x_1$ such that $| x_n|$ is a composite integer for all $n\in \mathbb{N}$. \end{theorem} Throughout the paper we will use the following convention: a nonnegative integer $n$ is said to be \emph{composite} if $n\ne 0, 1$, and $n$ is not a prime number. Graham \cite{graham} considered the problem above in the particular case $(a,b)=(1,1)$. He found two relatively prime positive integers $x_0$ and $x_1$ such that the sequence $x_{n+1}=x_n+x_{n-1}$, consists of composite numbers only. Graham's starting pair is \begin{equation*} (x_0,x_1)=(331635635998274737472200656430763, \, 1510028911088401971189590305498785). \end{equation*} Graham's technique was successively refined by Knuth \cite{knuth}, Wilf \cite{wilf}, and Nicol \cite{nicol}, who all found smaller pairs $(x_0,x_1)$. The current record is due to Vsemirnov \cite{vsemirnov}: \begin{equation}\label{vsemirnov} (x_0, x_1)= (106276436867, 35256392432). \end{equation} The above results are based on the fact that the Fibonacci sequence is a \emph{divisibility sequence}, that is, $F_n \mid F_m$ whenever $n \mid m$, and on finding a finite \emph{covering system of congruences} $ \modd{r_i} {m_i}$, $1\le i\le t$, such that there exist distinct primes $p_1, p_2,\ldots p_t$, so that $p_i \mid F_{m_i}$ for all $i=1, 2, \ldots, t$. Somer's proof of Theorem \ref{maintheorem} is relatively short since it relies on several results of Bilu, Hanrot and Voutier \cite{bilu}, Choi \cite{choi}, and Parnami and Shorey \cite{parnami}. A few years later, unaware of Somer's article, Dubickas, Novikas and \u{S}iurys \cite{dubickas} published a new solution that, although somewhat lengthier, is essentially self-contained. In this paper, we present another free-standing proof of this theorem that, while comparable to the one in \cite{dubickas}, differs from it in several important ways. We summarize our plan for proving Theorem \ref{maintheorem} as follows. In Section 2 we prove three easy lemmata, which will be useful later on. We believe that Lemma \ref{lemma1} is of independent interest. In Section 3 we study two simple cases: $(i)\,\, a=0$ and $(ii)\,\, a^2+4b=0$. In this section we also show that the condition $(a,b)\ne (\pm 2, -1)$ is necessary. Section 4 deals with the case $| b| \ge 2$. Following \cite{dubickas}, we choose $x_1\equiv \modd{0} {|b|}$, since then \eqref{recurrence} implies that $x_n \equiv \modd{0} {|b|}$ for all $n\ge 2$. The main difficulty is to show that $x_0$ and $x_1$ can be chosen such that $x_n\neq -b, 0, b$ for every $n\ge 0$. In this case, Dubickas et al. present a mainly existential proof which relies on a series of six lemmata. In contrast, our proof is constructive, as we provide explicit expressions for $x_0$ and $x_1$ as polynomials in $a$ and $b$. In Section 5 we consider the case $| b| =1$. Dubickas, Novikas and \u{S}iurys prove that except for finitely many values of $| a|$, one can take $x_0$ and $x_1$ so that each $x_n$ is divisible by one of five distinct appropriately chosen prime numbers. We prove that, with the exception of finitely many values of $| a|$, four primes suffice. The proof in the case $| b| =1$ relies on the divisibility properties of the \emph{Lucas sequence of the first kind} $(u_n)_{n\ge0}$, defined as \begin{equation}\label{lucas} u_0=0, u_1=1 \quad \text{and} \quad u_{n+1}=au_n+bu_{n-1}, \,\,\text{for} \,\, n\ge 1. \end{equation} Finally, in Section 6 we prove that if $| a| \ge 3$ and $b=-1$, then $u_n$ is composite for all $n\ge 3$. The interesting fact is that in this case it seems likely that there is no finite set of prime numbers $p_1, p_2,\ldots p_t$ such that each $u_n$ is divisible by some $p_i$, $i=1, 2, \ldots t$. \section{Three useful lemmata} \begin{lemma}\label{lemma1} Consider the sequence $(x_n)_{n\ge 0}$ given by \eqref{recurrence}. Then \begin{equation}\label{eq1} x_{n+1}^2-ax_nx_{n+1}-bx_n^2=(-b)^n(x_1^2-ax_0x_1-bx_0^2). \end{equation} \end{lemma} \begin{proof} \begin{gather*} \begin{bmatrix} x_{n+2} & x_{n+1}\\ x_{n+1} & x_{n} \end{bmatrix} = \begin{bmatrix} ax_{n+1}+bx_n & ax_n+bx_{n-1}\\ x_{n+1} & x_{n} \end{bmatrix} = \begin{bmatrix} a & b\\ 1 & 0 \end{bmatrix} \cdot \begin{bmatrix} x_{n+1} & x_{n}\\ x_{n} & x_{n-1} \end{bmatrix} = \\ \smallskip = \begin{bmatrix} a & b\\ 1 & 0 \end{bmatrix} \cdot \begin{bmatrix} a & b\\ 1 & 0 \end{bmatrix} \cdot \begin{bmatrix} x_{n} & x_{n-1}\\ x_{n-1} & x_{n-2} \end{bmatrix} = \begin{bmatrix} a & b\\ 1 & 0 \end{bmatrix}^2 \cdot \begin{bmatrix} x_{n} & x_{n-1}\\ x_{n-1} & x_{n-2} \end{bmatrix} = \cdots = \begin{bmatrix} a & b\\ 1 & 0 \end{bmatrix}^n \cdot \begin{bmatrix} x_2 & x_1\\ x_1 & x_0 \end{bmatrix}. \end{gather*} \smallskip Taking determinants on both sides we obtain \begin{align*} &\det\begin{bmatrix} x_{n+2} & x_{n+1}\\ x_{n+1} & x_{n} \end{bmatrix} = \det\begin{bmatrix} a & b\\ 1 & 0 \end{bmatrix}^n \cdot \det\begin{bmatrix} x_2 & x_1\\ x_1 & x_0 \end{bmatrix} \implies x_{n+2}\cdot x_n-x_{n+1}^2=(-b)^n\cdot (x_2x_0-x_1^2)\\ &\implies(ax_{n+1}+bx_n)x_n-x_{n+1}^2=(-b)^n\cdot ((ax_1+bx_0)x_0-x_1^2), \,\text{which after expanding}\\ &\text{becomes}\,\, x_{n+1}^2-ax_nx_{n+1}-bx_n^2=(-b)^n(x_1^2-ax_0x_1-bx_0^2), \quad \text{as claimed.} \end{align*} \end{proof} \begin{lemma}\label{lemma2} Consider the sequence $(x_n)_{n\ge 0}$ given by \eqref{recurrence}. Suppose that $1\le |x_0| <|x_1|$ and $|a| >|b| \ge 1$. Then the sequence $(| x_n|)_{n\ge 0}$ is strictly increasing. \end{lemma} \begin{proof} We use induction on $n$. By hypothesis, the statement is true for $n=0$. Suppose that $|x_{n-1}|<|x_n|$ for some $n\ge 1$. We intend to prove that $|x_n|<|x_{n+1}|$. Indeed \begin{align*} |x_{n+1}|&=|ax_n+bx_{n-1}|\ge |ax_n|-|bx_{n-1}| = |a||x_n|-|b||x_{n-1}|>\\ &>|a||x_n|-|b||x_n|=(|a|-|b|)|x_n|,\,\, \text{by the induction hypothesis.} \end{align*} Using $|a|-|b|\ge 1$, we obtain $|x_{n+1}|> |x_n|$, which completes the induction. \end{proof} \begin{lemma}\label{lemma3} Let $n_1$, $n_2$ and $n_3$ be three positive integers such that no prime number $p$ divides all of them. Then there exists an integer $k\ge 2$ such that $n_1$ and $n_2+k n_3$ are relatively prime. \end{lemma} \begin{proof} Let $d:=\gcd(n_2,n_3)$. Note that $\gcd(n_1,d)=1$, otherwise $d$ divides $n_1$, $n_2$ and $n_3$. By Dirichlet's theorem on arithmetic progressions there exists a $k$ such that $n_2/d+k n_3/d$ is a prime number greater than $n_1$. Then $d\,(n_2/d+k\cdot n_3/d) = n_2+k\:n_3$ is both greater and relatively prime to $n_1$. \end{proof} \section{Two simple special cases: \texorpdfstring{$(i)\,\, a=0$ and $(ii)\,\, a^2+4b=0$}{Lg}} \noindent \emph {Case (i):} Since $a=0$, it can be easily proved that $x_{2n}=b^nx_0$ and $x_{2n+1}=b^nx_1$. It suffices to take $x_0=4$ and $x_1=9$ to obtain that $x_n$ is composite for all $n\ge 0$. \bigskip \noindent \emph {Case (ii):} If $a^2+4b=0$, then $a$ must be even; $a=2c$ and therefore $b=-a^2/4=-c^2$. We divide the proof into two cases: $|b|\ge 2$ and $b=-1$. If $|b|\ge 2$, we have $|c|\ge 2$. Let us now take $x_0=4c^2-1$ and $x_1=2c^3$. Then $x_0$ and $x_1$ are relative prime positive composite integers. One immediately obtains $x_2=c^2$; thus $x_2$ is also composite. Also, it can be easily proved that $x_n=c^n\left((n-1)-(2n-4)c^2\right)$ for all $n\ge 3$. Note that for any $n\ge 3$ one cannot have $x_n = 0$; otherwise, $4\le c^2=(n-1)/(2n-4)\le 1$, which is impossible. It follows that $|x_n|$ is composite for all $n\ge 3$. If $b=-1$, then $|a|=2$. If $a=2$, then a simple induction shows that \begin{equation*} x_{n+1}=(n+1)x_1-nx_0=x_1+n(x_1-x_0) \quad \text{for all}\,\, n\ge 0, \end{equation*} that is, $(x_n)_{n\ge 0}$ is an arithmetic sequence whose first term and common difference are relatively prime. By Dirichlet's theorem on primes in arithmetic progressions, it follows that $|x_n|$ is a prime number for infinitely many values of $n$. If $a=-2$, then one can show that $x_{n+1}=(-1)^n(x_1+n(x_0+x_1))$. In this case, $(x_n)_{n\ge 0}$ is the union of two arithmetic sequences, both of which have the first term and the common difference relatively prime. Again, for any choice of $x_0$ and $x_1$ relatively prime, $|x_n|$ is a prime for infinitely many $n$. This proves the necessity of the condition $(a,b)\neq(\pm 2, -1)$. \section{ The case \texorpdfstring{$|b|\ge 2$}{Lg}} Based on the results from the previous section, from now on we can assume that \begin{equation} |a|\ge 1 \quad \text {and}\quad a^2+4b\ne 0. \end{equation} We divide the proof into three separate cases: \bigskip \noindent {\it Case I:} {$|a|>|b|$}, \\ \noindent {\it Case II:} {$1\le |a|\le |b|$ and $|b|$ is composite}, and \\ \noindent {\it Case III:} {$1\le |a|\le |b|$ and $|b|$ is a prime.} \bigskip \noindent{\it Case I: $|a|>|b|$}. To complete the proof of Theorem \ref{maintheorem}, in this case we take $x_0=b^4-1, x_1=b^4$. Clearly, $x_0$ and $x_1$ are both positive, relatively prime and composite. Moreover, $x_0|b|$ for all $n\ge 0$, and therefore $|x_n|$ is composite for all $n\ge 0 $. \bigskip \noindent{\it{Case II: $1\le |a|\le |b|$ and $|b|$ is composite}}. Choose $x_0=4b^4-1$, $x_1=2b^2$. Clearly, $\gcd(x_0,x_1)=1$, and $x_0, x_1$ are positive composite integers. It also follows that $x_n\equiv \modd{0} {b}$ for all $n\ge 1$, and since $|b|$ is composite, it follows that $|x_n|$ is composite, unless $x_n=0$ for some $n$. We will prove that this cannot happen. To get a contradiction, suppose that $x_{n+1}=0$ for some $n\ge 2$. Then, by Lemma \ref{lemma1}, we have $x_n^2=(-b)^{n-1}(x_1^2-ax_0x_1-bx_0^2)$. Since $x_0=4b^4-1$ and $x_1=2b^2$, we have \begin{equation*} x_1^2-ax_0x_1-bx_0^2=(-b)(16b^8+8ab^5-8b^4-4b^3-2ab+1),\quad \text{which implies that} \end{equation*} \begin{equation}\label{xnsquared} x_n^2=(-b)^n(16b^8+8ab^5-8b^4-4b^3-2ab+1). \end{equation} If $n$ is even, this implies that $16b^8+8ab^5-8b^4-4b^3-2ab+1$ is a perfect square. If $n$ is odd, this implies that $(-b)(16b^8+8ab^5-8b^4-4b^3-2ab+1)$ is a perfect square. We will prove that none of these are possible if $a^2+4b\ne0$. Note first that since $|a|\le |b|$, we have \begin{equation}\label{cII} (4b^4+ab-2)^2<16b^8+8ab^5-8b^4-4b^3-2ab+1<(4b^4+ab)^2. \end{equation} Indeed, these inequalities are equivalent to the following ones: \begin{equation}\label{cIIa} 8b^4-2-(4b^3+(ab-1)^2)>0, \quad \text{and}\quad 8b^4+4b^3-2+(ab+1)^2>0. \end{equation} We have \begin{equation} |a|\le |b| \implies |ab|\le b^2 \implies |ab - 1|\le |ab|+1\le b^2+1 \implies (ab- 1)^2\le b^4+2b^2+1. \end{equation} From this we obtain \begin{equation} |4b^3+(ab-1)^2|\le 4|b|^3 + (ab-1)^2 \le b^4+4|b|^3+2b^2+1. \end{equation} To prove the first inequality in \eqref{cIIa}, note that \begin{align*} & \,\, 8b^4-2-(4b^3+(ab-1)^2)\ge 8b^4-2-|4b^3+(ab-1)^2|\ge\\ \ge &\,\, 8b^4-2-(b^4+4|b|^3+2b^2+1)= 7b^4-4|b|^3-2b^2-3=\\ = & \,\, 7|b|^4-4|b|^3-2|b|^2-3=4|b|^3(|b|-1)+2|b|^2(|b|^2-1)+(|b|^4-3)\ge\\ \ge & \,\, 4|b|^3+2|b|^2+|b|^4-3\ge 4\cdot 2^3+2\cdot 2^2+2^4-3>0. \end{align*} The second inequality in \eqref{cIIa} is much easier to prove: \begin{equation*} 8b^4+4b^3-2+(ab+1)^2\ge 8b^4+4b^3-2\ge 8|b|^4-4|b|^3-2\ge 8\cdot 2^4-4\cdot 2^3-2>0. \end{equation*} Hence, the inequalities \eqref{cII} hold. If the middle term in \eqref{cII} were to be a perfect square, the only option remaining is \begin{equation*} 16b^8+8ab^5-8b^4-4b^3-2ab+1=(4b^4+ab-1)^2,\,\,\text{which implies}\,\, b^2(a^2+4b)=0. \end{equation*} However, this is impossible if $a^2+4b\ne 0$. Hence, if one assumes that $|a|\le |b|$ and $a^2+4b\ne 0$, then $16b^8+8ab^5-8b^4-4b^3-2ab+1$ cannot be a perfect square. Suppose next that $(-b)(16b^8+8ab^5-8b^4-4b^3-2ab+1)$ is a perfect square. Since the two factors are mutually prime, it follows that both have to be perfect squares. In particular, the second one has to be a perfect square, and we have already proved that it cannot be. It follows that the sequence $(x_n)$ does not contain any terms equal to $0$. This completes the proof of Theorem \ref{maintheorem} in case II. \bigskip \noindent {\it{Case III: $1\le |a|\le |b|$ and $|b|$ is a prime}}. We divide the proof of this case into three subcases: $|a|=1$, $|a|=|b|$ and $2\le |a|<|b|$. \bigskip \noindent {\it{Subcase IIIa: $|a|=1$, $|b|$ is a prime}}. If $a=1$ and $b>0$, then take $x_0=(2b^2-1)^2, x_1=b(b^2-1)$. Clearly, $x_0$ and $x_1$ are positive, relatively prime, composite integers. It follows immediately that $x_2=b^3(4b^2-3)$ is also composite. Finally, an easy induction shows that $x_n\equiv \modd{0} {b^2}$ and that $\frac{x_n}{b^2} \equiv \modd{-1} {b}$ for all $n\ge 3$. Hence, $|x_n|$ is necessarily composite for all $n\ge 3$. The other situations can be dealt with similarly. If $a=-1$ and $b<0$, take $x_0=(2b^2-1)^2, x_1=-b(b^2-1)$. Then $x_2=b^3(4b^2-3)$, and for $n\ge 3$ one can show that $x_n\equiv \modd{0} {b^2}$ and \(\frac{x_n}{b^2}\)$\equiv \modd{(-1)^{n+1}} {b}$. Again, all $|x_n|$ are composite. If $a=1$ and $b<0$, we take $x_0=(2b^2-1)^2, x_1=-b(b^2+1)$. Then $x_2=b^3(4b^2-5)$, and for all $n\ge 3$ one can prove that $x_n\equiv \modd{0} {b^2}$ and \(\frac{x_n}{b^2}\)$\equiv \modd{-1} {b}$. Hence, all $|x_n|$ are composite. Finally, if $a=-1$ and $b>0$, select $x_0=(2b^2-1)^2, x_1=b(b^2+1)$. Then $x_2=b^3(4b^2-5)$, and for all $n\ge 3$ we have that $x_n\equiv \modd{0} {b^2}$ and \(\frac{x_n}{b^2}\)$\equiv \modd{(-1)^{n+1}} {b}$. All $|x_n|$ are composite. This completes the proof of Theorem \ref{maintheorem} in subcase IIIa. \bigskip \noindent {\it{Subcase IIIb: $|a|=|b|$, $|b|$ is a prime}}. If $a=b$, take $x_0=4b^4-1, x_1=2b^2$. Then $x_2=b(4b^4+2b^2-1)$ is composite, and $x_n\equiv \modd{0} {b^2}$ for all $n\ge3$. It remains to show that $x_n\ne0$ for all $n$. As in Case II, $x_{n+1}=0$ implies that either $16b^8+8b^6-8b^4-4b^3-2b^2+1$ or $(-b)(16b^8+8b^6-8b^4-4b^3-2b^2+1)$ is a perfect square. One can use the same technique as in Case II to prove that this cannot happen if $a^2+4b=b^2+4b\ne0$. If $a=-b$, take $x_0=4b^4-1, x_1=2b^2$ (same choice). Then $x_2=b(4b^4-2b^2-1)$ is composite, and $x_n\equiv \modd{0} {b^2}$ for all $n\ge3$. It remains to show that $x_n\ne 0$ for all $n$. As in Case II, $x_{n+1}=0$ implies that either $16b^8-8b^6-8b^4-4b^3+2b^2+1$ or $(-b)(16b^8-8b^6-8b^4-4b^3+2b^2+1)$ is a perfect square. The same approach as in Case II shows that this is impossible if $a^2+4b=b^2+4b\ne0$. \bigskip \noindent{\it{Subcase IIIc: $2\le|a|<|b|$, $|b|$ is a prime}}. It follows that $\gcd(a,b)=1$. If $a>0, b>0$, take $x_0=a^3, x_1=b(b^2-a^2)$. Then $x_2=ab^3$. For $n \ge3 $, $x_n\equiv \modd{0} {b^2}$ and $\frac{x_n}{b^2} \equiv \modd{-a^{n-1}} {b}$. Since $\gcd(a,b)=1$, $x_n \neq 0$. If $a<0, b<0$, take $x_0=-a^3, x_1=-b(b^2-a^2)$. Then $x_2=-ab^3$. For $n \ge 3$, $x_n\equiv \modd{0} {b^2}$ and $\frac{x_n}{b^2} \equiv \modd{a^{n-1}} {b}$. Hence, $x_n\neq 0$ for all $n$. If $a<0, b>0$, take $x_0=-a^3, x_1=b(b^2+a^2)$. Then $x_2=ab^3$, and for $n \ge 3$, $x_n\equiv \modd{0} {b^2}$ and $\frac{x_n}{b^2}\equiv \modd{a^{n-1}} {b}$. So, $x_n\neq 0$ for all $n$. Lastly, if $a>0, b<0$, take $x_0=a^3, x_1=-b(b^2+a^2)$. Then $x_2=-ab^3$. For $n \ge 3$, $x_n\equiv \modd{0} {b^2}$ and $\frac{x_n}{b^2}\equiv \modd{-a^{n-1}} {b}$. Since $\gcd(a,b)=1$, we have $x_n \neq 0$ for all $n$. Hence, in each case one can choose $x_0$ and $x_1$ to be positive, composite and relatively prime so that $x_2$ is composite, and for all $n\ge 3$ we have that $x_n\equiv \modd{0} {b^2}$ and $x_n \neq 0$. It follows that $|x_n|$ is composite for all $n\ge 0$, and this completes the proof of Theorem \ref{maintheorem} in the case $|b|\ge 2$. \section{The case \texorpdfstring{$|b|=1$}{Lg}} The main idea behind the proof of Theorem \ref{maintheorem} in this particular case can be summarized as follows: we want to find a finite set of primes $p_1, p_2, \ldots, p_t$ such that for every $n\ge 0$ the number $|x_n|$ is divisible by at least one of these primes. We start the analysis with the simple case in which $|a|$ has at least two distinct prime factors. \begin{lemma} Let $|b|=1$, and suppose that $|a|$ has at least two distinct prime factors: $p_1$ and $p_2$, with $p_13$. In particular, $a^2+2$ is odd and therefore $\gcd(a, a^2+2)=1$. Moreover, since $a^2+2 \equiv \modd{0} {3}$ one can safely take $p_2=3$. We claim that $a^2+2$ has at least one other prime factor, $p_3$, different from $p_1$ and $p_2$. Indeed, if one assumes otherwise, then $a^2+2= p_1^{2s}+2=3^t$ for some $t\ge 2$. However, it was shown by Ljunggren \cite{ljunggren} that the more general equation $x^2+2= y^n$, $n\ge 2$ has the unique solution $x=5, y=3, n=3$. This would give $|a|=5$, which is excluded from our analysis. Finally, suppose that $p_1=3$ and therefore $|a|=3^s$ for some $s\ge 2$. Then both $a^2+1$ and $a^2+3$ are even, so one can take $p_2=2$. Note that $a^2+1=9^s+1$; hence $a^2+1\equiv \modd{10} {72}$ and $a^2+3\equiv \modd{12} {72}$. Hence, there exists a positive integer $c$ such that $a^2+1=2(36c+5)$ and $a^2+3=12(6c+1)$. Let $p_3$ be a prime factor of $36c+5$ and let $p_4$ be a prime factor of $6c+1$. It follows immediately that $p_1=3, p_2=2, p_3$, and $p_4$ are all distinct. \end{proof} We present a couple of particular instances covered by the above lemma. Suppose first that $a=8$ and $b=1$. Then $u_2=a=2^3$ while $u_4=2^4\cdot 3\cdot 11$. Thus, in this case one can take $p_1=2, p_2=3, p_3=11$. Suppose next that $a=-49$ and $b=1$. Then $u_2=a=-7^2$, while $u_4=-7^2\cdot 3^3\cdot 89$. It follows that we can choose $p_1=7, p_2=3, p_3=89$. Finally, let us assume that $a=9$ and $b=1$. Then $u_2=a=3^2$, while $a^2+1=2\cdot 41$, $a^2+3=2^2\cdot 3\cdot 7$. Hence, in this case we have $p_1=3, p_2=2, p_3=41, p_4=7$. The next lemma uses the concept of covering system introduced by Erd\H{o}s in \cite{erdos}. \begin{definition} A collection of residue classes $\modd{r_i} {m_i}$, $0\le r_i < m_i$, where $1\le i\le t$ is said to be a \emph{covering system} if every integer $n$ satisfies at least one equality $n\equiv \modd{r_i} {m_i}$. \end{definition} In the proof of the theorem when $|b|=1$, except for finitely many values of $a$ we will use one of the following two covering systems \begin{align*} &\{\modd{0} {2}, \quad \modd{1} {6},\quad \modd{3} {6},\quad \modd{5} {6}\}\quad \text{or}\\ &\{\modd{0} {2}, \quad \modd{1} {4}, \quad \modd{3} {4}\}. \end{align*} \begin{lemma}\label{last} Let $a, b$ be two integers such that $|a|\ge 2$ and $|b|=1$. Let $(u_n)_{n\ge 0}$ be the Lucas sequence defined in \eqref{lucassequence}, and suppose that there exists a finite collection of triples $(p_i, m_i, r_i)$, $1\le i\le t$ with the following properties: \begin{itemize} \item[$(i)$]{All primes $p_i$ are distinct.} \item[$(ii)$]{The residue classes $\modd{r_i} {m_i}$ form a covering system.} \item[$(iii)$]{$p_i \mid u_{m_i}$ for all $1\le i\le t$.} \end{itemize} Then there exist two relatively prime positive integers $x_0$ and $x_1$ such that each term of the sequence $(x_n)_{n\ge 0}$ defined in \eqref{recurrence} is composite. \end{lemma} \begin{proof} Let $P=p_1p_2\ldots p_t$. By the Chinese remainder theorem, there exist $y, z \in \{0, 1, \ldots P-1\}$ satisfying \begin{align}\label{congruences} &y\equiv u_{m_i-r_i} \pmod{p_i},\notag\\ &z\equiv u_{m_i-r_i+1} \pmod{p_i}, \end{align} \noindent for $i=1, 2,\ldots ,t $. Note that there is no prime which divides $y, z$, and $P$ simultaneously. Indeed, if such a prime $p_j$ were to exist, then there would be two consecutive terms $u_n$ and $u_{n+1}$ both divisible by $p_j$. Since $u_{n+1}=au_n+bu_{n-1}$ and $|b|=1$, then $p_j \mid u_{n-1}$. By induction, it follows that $p_j \mid u_1$ which is impossible since $u_1=1$. Let $x_0\equiv \modd{y} {P}$ and $x_1\equiv \modd{z} {P}$. Then we have $x_0\equiv \modd{u_{m_i-r_i}} {p_i}$ and $x_1\equiv \modd{u_{m_i-r_i+1}} {p_i}$ for all $i=1,2,\ldots, t$. By induction on $n$, we obtain $x_{n+1}\equiv \modd{u_{m_i-r_i+n}} {p_i}$ for every $n\ge 0$ and every $1\le i\le t$. Since the residue classes $\modd{r_i} {m_i}$ form a covering system, each nonnegative integer $n$ belongs to one of these classes, say $n=r_i+k\,m_i$ for some $k\ge 0$ and some $i\in \{1, 2,\ldots, t\}$. This implies that \begin{equation} x_{n+1}\equiv u_{m_i-r_i+n}\pmod{p_i}\equiv u_{m_i(k+1)}\pmod{p_i}\equiv 0\pmod{p_i}, \end{equation} since $p_i \mid u_{m_i}$ and $u_{m_i} \mid u_{m_i(k+1)}$. Hence, every term of the sequence $(x_n)_n$ is divisible by some prime $p_i$. It remains to choose $x_0$ and $x_1$, relatively prime positive integers $x_0\equiv \modd{y} {P}$ and $x_1\equiv \modd{z} {P}$ such that $|x_n|\ge P$ for every $n\in \mathbb{N}$. In order to achieve this, take $x_0=y+P$ and $x_1=z+kP$ where $k\ge 2$ and $\gcd(x_0, x_1)=1$. Using Lemma \ref{lemma3} with $n_1=y+P$, $n_2=z$ and $n_3=P$ shows that such a choice is always possible. Recall that we proved earlier that $\gcd(y, z, P)=1$. Such a choice implies that $0|b|=1$, Lemma \ref{lemma2} implies that $(|x_n|)_n$ is a strictly increasing sequence. It follows that $|x_n|\ge P$ for all $n\ge 0$, and therefore each such $x_n$ is composite. \end{proof} We can now prove the theorem if $|b|=1$. Suppose first that $b=-1$ and $|a|=p_1^s\ge 4$. Then, by Lemma \ref{bminus1}, there are four distinct primes $p_1, p_2, p_3, p_4$ dividing $u_2, u_6, u_6, u_6$, respectively. The theorem follows after using Lemma \ref{last} for the triples $(p_1, 2, 0), (p_2, 6, 1), (p_3, 6, 3), (p_4, 6, 5)$. As a numerical illustration, suppose that $a=-9, b=-1$. Then, as described in the paragraph following the proof of Lemma \ref{bminus1}, we have $p_1=3, p_2=2, p_3=5$ and $p_4=13$. The system \eqref{congruences} becomes \begin{align*} &y\equiv u_2\pmod{3}\quad\quad\quad z\equiv u_3\pmod{3}\\ &y\equiv u_5\pmod{2}\quad\quad\quad z\equiv u_6\pmod{2}\\ &y\equiv u_3\pmod{5}\quad\quad\quad z\equiv u_4\pmod{5}\\ &y\equiv u_1\pmod{13}\quad\quad\,\,\, z\equiv u_2\pmod{13} \end{align*} and its solution is $P=p_1p_2p_3p_4=390, y=105, z=134$. Since $y1$. Still, one can safely take $x_0=y=56$ and $x_1=z+P=129$, and now $\gcd(x_0,x_1)=1$ as desired. Then, since $01$. Still, one can safely take $x_0=y=980$ and $x_1=z+P=3333$ and now $\gcd(x_0,x_1)=1$ as desired. Moreover, since $0|b|=1$, Lemma \ref{lemma2} implies that $(| x_n|)_{n\ge 0}$ is strictly increasing. \begin{table}[htbp] \centering \begin{tabular}{c|c|c|c|c} \hline $a$ & $b$ & $\{(p_i,m_i,r_i)\}$ & $x_0$ & $x_1$\\ [0.5ex] \hline $\phantom{-}5$ & $\phantom{-}1$ & $(5, 2, 0), (2, 6, 1), (7, 6, 3), (13,6,5)$ & $495$ & $1136$ \\ $-5$ & $\phantom{-}1$ & $(5, 2, 0), (2, 6, 1), (7, 6, 3), (13,6,5)$ & $495$ & $866$ \\ $\phantom{-}4$ & $\phantom{-}1$ & $(2,2,0), (3,4,1), (7,8,3), (23,8,7)$ & $116$ & $165$ \\ $-4$ & $\phantom{-}1$ & $(2,2,0), (3,4,1), (7,8,3), (23,8,7)$ & $116$ & $801$ \\ $\phantom{-}3$ & $\phantom{-}1$ &$(3,2,0), (11,4,1), (7,8,3), (17,8,7)$ & $1803$ & $3454$ \\ $-3$ & $\phantom{-}1$ &$(3,2,0), (11,4,1), (7,8,3), (17,8,7)$ & $1803$ & $3091$ \\ $\phantom{-}2$ & $\phantom{-}1$ & $(2,2,0), (5,3,0), (3,4,1), (7,6,5), (11,12,7)$ & $260$ & $807$ \\ $-2$ & $\phantom{-}1$ & $(2,2,0), (5,3,0), (3,4,1), (7,6,5), (11,12,7)$ & $260$ & $1503$ \\ \hline $\phantom{-}3$ & $-1$ & $(3,2,0),(2,3,0),(7,4,3),(47,8,5),(23,12,5),(1103, 24, 1)$ & $7373556$ & $2006357$ \\ $-3$ & $-1$ & $(3,2,0),(2,3,0),(7,4,3),(47,8,5),(23,12,5), (1103, 24,1)$ & $7373556$ & $14686445$ \\[1ex] \hline \end{tabular} \medskip \caption{Covering triples for the cases $b=1, a=\pm 2, \pm 3, \pm 4, \pm 5$ and $b=-1, a=\pm 3$} \end{table} It remains to see what happens when $|a|=|b|=1$, as in these cases Lemma \ref{last} does not apply. If $a=-1, b=-1$, then it can be easily verified that the sequence given by the recurrence $x_{n+1}=-x_n-x_{n-1}$ has period 3. Hence, if one chooses $x_0=8$ and $x_1=27$, then $x_2=-35$, and due to the periodic behavior all terms of the sequence are composite. Similarly, if $a=1, b=-1$, then the sequence given by the recurrence $x_{n+1}=x_n-x_{n-1}$ has period 6. Again, if one chooses $x_0=8$ and $x_1=35$, then the first few terms of the sequence are $8, 35, 27, -8, -35, -37, 8, 35, 27, \ldots$; that is, $x_n$ is always composite. If $a=b=1$, then Vsemirnov's pair $v_0=106276436867$, $v_1=35256392432$ shows that all the numbers \begin{equation}\label{vv} v_n=v_{n-1}+v_{n-2}=v_1F_n+v_0F_{n-1} \end{equation} are composite. Here, $F_n$ is the $n$th Fibonacci number, where $F_{-1}=1, F_0=0, F_1=1$. For the case $a=-1, b=1$, we follow the solution in \cite{dubickas}. It can be easily checked that the general term of the sequence $x_{n+1}=-x_n+x_{n-1}$ can be written as \begin{equation}\label{x} x_n=(-1)^{n+1}x_1F_n+(-1)^nx_0F_{n-1}, \quad n\ge 0. \end{equation} We choose $x_0=v_0-v_1=71020044435$ and $x_1=v_0=106276436867$. It is easy to check that $x_0$ and $x_1$ are relatively prime composite integers. Moreover, from \eqref{vv} and \eqref{x} we obtain that \begin{align*} x_n&=(-1)^{n+1}v_0F_n+(-1)^n(v_0-v_1)F_{n-1}+= (-1)^{n+1}v_1F_{n-1}+(-1)^{n+1}v_0(F_n-F_{n-1})=\\ &=(-1)^{n+1}v_1F_{n-1}+(-1)^{n+1}v_0F_{n-2}= (-1)^{n+1}\left(v_1F_{n-1}+v_0F_{n-2}\right)=(-1)^{n-1}v_{n-1}. \end{align*} Hence, $|x_n|=v_{n-1}$ is composite for all $n\ge 0$. The proof of the theorem is now complete. \section{A surprising result} In this section we prove the following: \begin{theorem}\label{uthm} Consider the integers $a, b$ such that $|a|\ge 3$ and $b=-1$. Let $u_0=0$, $u_1=1$ and $u_{n+1}=au_n+bu_{n-1}=au_n-u_{n-1}$ be the Lucas sequence of the first kind associated with $a$ and $-1$. Then $|u_n|$ is composite for all $n\ge 3$. \end{theorem} \begin{proof} One has $u_3=a^2-1$, and $u_4=a^3-2a$, which are obviously composite. Since $|a|>|b|=1$, Lemma \ref{lemma2} implies that the sequence $(|u_n| )_{n\ge 0}$ is strictly increasing. Suppose for the sake of contradiction that there exists an $n\ge 2$ such that $|u_{n+1}|=p$, where $p$ is some prime number. Since $|u_{n+1}|\ge | u_3|=a^2-1$ it follows that necessarily $p>|a|$. Now using Lemma \ref{lemma1} for the sequence $(u_n)_{n\ge 0}$, equality \eqref{eq1} becomes \begin{equation*} u^2_{n+1}-au_nu_{n+1}+u_n^2=u_1^2-au_1u_0+u_0^2, \end{equation*} and since $u_0=0$, $u_1=1$, and $|u_{n+1}|=p$, we obtain that \begin{equation}\label{quadratic} u_n^2\pm apu_n+p^2-1=0. \end{equation} Regard the above equation as a quadratic in $u_n$. Since $u_n\in \mathbb{Z}$, it is necessary that the discriminant is a perfect square, that is, there exist a nonnegative integer $c$ such that \begin{equation}\label{eq1000} a^2p^2-4(p^2-1)=c^2 \quad \text{from which}\quad (a^2-4)p^2=c^2-4=(c-2)(c+2). \end{equation} Since $|a|\ge 3$, one can assume that $c\ge 3$. Since $p$ is a prime and $p^2$ divides $(c-2)(c+2)$, we have two possibilities. If $p$ divides both $c-2$ and $c+2$ then $p$ divides $4$, which means that $p=2$. However, this is impossible since $p>|a|\ge 3$. Otherwise, $p^2$ divides either $c-2$ or $c+2$. In either case we obtain that $c+2\ge p^2$. Using this inequality in \eqref{eq1000}, it follows that \begin{equation*} (a^2-4)p^2=(c-2)(c+2)\ge p^2(p^2-4)\implies |a|\ge p,\,\, \text{ a contradiction}. \end{equation*} This completes the proof. \end{proof} In particular, the above theorem holds if $a=4$ and $b=-1$, thus answering a question of Vos Post (see \seqnum{A001353} in \cite{oeis}). What is remarkable about this situation is that while $u_n$ is composite for every $n\ge 3$, it seems likely that there is no finite set of primes $p_1, p_2,\ldots p_t$ such that every $u_n$ is divisible by some $p_i$, $1\le i\le t$. In fact, we suspect the following is true. \begin{conjecture} Let $(u_n)_{n\ge 0}$ be the Lucas sequence of the first kind associated with some $|a|\ge 3$ and $b=-1$. Then for any two different primes $p$ and $q$, $u_p$ and $u_q$ are relatively prime. \end{conjecture} If true, this conjecture would immediately imply that there is no finite set of primes $p_1, p_2,\ldots p_t$ such that every $u_n$ is divisible by some $p_i$, $1\le i\le t$. \begin{thebibliography}{99} \bibitem{bilu} Y. Bilu, G. Hanrot, and P. M. Voutier, Existence of primitive divisors of Lucas and Lehmer numbers, \emph{ J. Reine Angew. Math.} {\bf 539} (2001), 75--122. \bibitem{choi} S. L. G. Choi, Covering the set of integers by congruence classes of distinct moduli, \emph {Math. Comp.} {\bf 25} (1971), 885--895. \bibitem{dubickas} A. Dubickas, A. Novikas, and J. \u{S}iurys, A binary linear recurrence sequence of composite numbers, \emph{J. Number Theory} {\bf 130} (2010), 1737--1749. \bibitem{erdos} P. Erd\H{o}s, On a problem concerning congruence systems, (Hungarian. Russian, English summary) \emph{Mat. Lapok} {\bf 3} (1952), 122--128. \bibitem{graham} R. L. Graham, A Fibonacci-like sequence of composite numbers, \emph{Math. Mag.} {\bf 37} (1964), 322--324. \bibitem{knuth} D. E. Knuth, A Fibonacci-like sequence of composite numbers, \emph{Math. Mag.} {\bf 63} (1990), 21--25. \bibitem{ljunggren} W. Ljunggren, \"{U}ber einige Arcustangensgleichungen die auf interessante unbestimmte Gleichungen f\"{u}hren, \emph{Ark. Mat. Astr. Fys.} {\bf 29A} (1943), 11 pp. \bibitem{catalan} P. Mih\u{a}ilescu, Primary cyclotomic units and a proof of Catalan's conjecture, \emph{J. Reine Angew. Math.} {\bf 572} (2004), 167--195. \bibitem{nagell} T. Nagell, Verallgemeinerung eines Fermatschen Satzes, \emph{Arch. Math. (Basel)} {\bf 5} (1954), 153--159. \bibitem{nicol} J. W. Nicol, A Fibonacci-like sequence of composite numbers, \emph{Electron. J. Combin.} {\bf 6} (1999), Paper R44. Available at \url{https://www.combinatorics.org/ojs/index.php/eljc/article/view/v6i1r44}. \bibitem{parnami} J. C. Parnami and T. N. Shorey, Subsequences of binary recursive sequences, \emph{Acta Arith.} {\bf 40} (1981/82), 193--196. \bibitem{somer} L. Somer, Second-order linear recurrences of composite numbers, \emph{Fibonacci Quart.} {\bf 44} (2006), 358--361. \bibitem{sury} B. Sury, On the diophantine equation $x^2+2=y^n$, \emph{Arch. Math.} (Basel) {\bf 74} (2000), 350--355. \bibitem{oeis} N. J. A. Sloane et al., \emph{The On-Line Encyclopedia of Integer Sequences}, 2019. Published electronically at \url{https://oeis.org}. \bibitem{vsemirnov} M. Vsemirnov, A new Fibonacci-like sequence of composite numbers, \emph{J. Integer Sequences} {\bf 7} (2004), \href{https://cs.uwaterloo.ca/journals/JIS/VOL7/Vsemirnov/vsem5.html}{Article 04.3.7}. \bibitem{wilf} H. S. Wilf, Letters to the editor, \emph{Math. Mag.} {\bf 63} (1990), 284. \end{thebibliography} \bigskip \hrule \bigskip \noindent 2010 {\it Mathematics Subject Classification}: Primary 11B39; Secondary 11B37, 11A07, 11Y55. \noindent \emph{Keywords:} second-order linear recurrence, Lucas sequence of the first kind, composite number, covering system of congruences. \bigskip \hrule \bigskip \noindent (Concerned with sequences \seqnum{A001906}, \seqnum{A001353}, \seqnum{A004254}, and \seqnum{A001109}.) \bigskip \hrule \bigskip \vspace*{+.1in} \noindent Received January 7 2019; revised versions received February 4 2019; September 8 2019; September 24 2019. Published in {\it Journal of Integer Sequences}, September 24 2019. \bigskip \hrule \bigskip \noindent Return to \htmladdnormallink{Journal of Integer Sequences home page}{https://cs.uwaterloo.ca/journals/JIS/}. \vskip .1in \end{document} .