\documentclass[12pt,reqno]{article} \usepackage[usenames]{color} \usepackage{amssymb} \usepackage{graphicx} \usepackage{amscd} \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,amsmath,amssymb} \usepackage{amsthm} \usepackage{amsfonts} \usepackage{latexsym} \usepackage{epsf} \setlength{\textwidth}{6.5in} \setlength{\oddsidemargin}{.1in} \setlength{\evensidemargin}{.1in} \setlength{\topmargin}{-.1in} \setlength{\textheight}{8.4in} \newcommand{\seqnum}[1]{\href{http://oeis.org/#1}{\underline{#1}}} \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 A Generalization of the \\ \vskip .12in Digital Binomial Theorem} \vskip 1cm \large Hieu D. Nguyen\\ Department of Mathematics\\ Rowan University\\ Glassboro, NJ 08028\\ USA \\ \href{mailto:nguyen@rowan.edu}{\tt nguyen@rowan.edu} \end{center} \vskip .2 in \begin{abstract} We prove a generalization of the digital binomial theorem by constructing a one-parameter subgroup of generalized Sierpi\'{n}ski matrices. In addition, we derive new formulas for the coefficients of Prouhet-Thue-Morse polynomials and describe group relations satisfied by generating matrices defined in terms of these Sierpi\'{n}ski matrices. \end{abstract} \section{Introduction} The classical binomial theorem describes the expansion of $(x+y)^N$ in terms of binomial coefficients. More precisely, for any non-negative integer $N$, we have \begin{equation}\label{eq:binomial-theorem} (x+y)^N = \sum_{k=0}^N \binom{N}{k} x^k y^{N-k}, \end{equation} where the binomial coefficients $\binom{N}{k}$ are defined in terms of factorials: \[ \binom{N}{k}=\frac{N!}{k!(N-k)!}. \] The author \cite{N1} introduced a digital version of this theorem (based on Callan \cite{C}) where the exponents appearing in (\ref{eq:binomial-theorem}) are viewed as sums of digits. To illustrate this, consider the binomial theorem for $N=2$: \begin{equation} \label{eq:binomial-theorem-n=2} (x+y)^2 =x^2y^0+x^1y^1+x^1y^1+x^0y^2. \end{equation} It is easy to verify that (\ref{eq:binomial-theorem-n=2}) is equivalent to \begin{equation} \label{eq:sum-of-digit-expansion-n=2} (x+y)^{s(3)} = x^{s(3)}y^{s(0)}+x^{s(2)}y^{s(1)}+x^{s(1)}y^{s(2)}+x^{s(0)}y^{s(3)}, \end{equation} where $s(k)$ denotes the sum of digits of $k$ expressed in binary. For example, \[ s(3)=s(1\cdot 2^1+1\cdot 2^0)=2. \] More generally, we have \begin{theorem}[digital binomial theorem \cite{N1}] \label{th:digital-binomial-theorem} Let $n$ be a non-negative integer. Then \begin{equation} \label{eq:digital-binomial-theorem-carry-free} (x+y)^{s(n)} = \sum_{\substack{0\leq m \leq n \\ (m,n-m) \ \text{\rm carry-free}}}x^{s(m)}y^{s(n-m)}. \end{equation} \end{theorem} \noindent Here, a pair of non-negative integers $(j,k)$ is said to be {\em carry-free} if their addition involves no carries when performed in binary. For example, $(8,2)$ is carry-free since $8+2=(1\cdot 2^3+0\cdot 2^2+0\cdot 2^1+0\cdot 2^0)+1\cdot 2^1=10$ involves no carries. It is clear that $(j,k)$ is carry-free if and only if $s(j)+s(k)=s(j+k)$ \cite{BEJ, N1}. Also, observe that if $n=2^N-1$, then (\ref{eq:digital-binomial-theorem-carry-free}) reduces to (\ref{eq:binomial-theorem}). In this paper we generalize Theorem~\ref{th:digital-binomial-theorem} to any base $b\geq 2$. To state this result, we shall need to introduce a {\em digital dominance} partial order on $\mathbb{N}$ as defined by Ball, Edgar, and Juda \cite{BEJ}. Let \begin{align*} n & =n_0b^0+n_1b^1+\dots + n_{N-1}b^{N-1} \end{align*} represent the base-$b$ expansion of $n$ and denote $d_{b,i}(n):=n_i$ to be the $i$-th digit of $n$ in base $b$. We shall say that $m$ is {\em digitally} less than or equal to $n$ if $d_{b,i}(m)\leq d_{b,i}(n)$ for all $i=0,1,\ldots, N-1$. In that case, we shall write $m\preceq_b n$. We are now ready to state our result. \begin{theorem} \label{th:digital-binomial-theorem-non-binary} Let $n$ be a non-negative integer. Then for any base $b\geq 2$, we have \begin{equation} \label{eq:digital-binomial-theorem-non-binary} \prod_{i=0}^{N-1}\binom{x+y+d_{b,i}(n)-1}{d_{b,i}(n)} =\sum_{0\leq m\preceq_b n}\left( \prod_{i=0}^{N-1}\binom{x+d_{b,i}(m)-1}{d_{b,i}(m)} \prod_{i=0}^{N-1}\binom{y+d_{b,i}(n-m)-1}{d_{b,i}(n-m)} \right). \end{equation} \end{theorem} Let $\mu_j(n):=\mu_j^{(b)}(n)$ denote the multiplicity of the digit $j>0$ in the base-$b$ expansion of $n$, i.e., \[ \mu_j(n)=|\{i: d_{b,i}(n)=j\}|. \] As a corollary, we obtain \begin{corollary} \label{co:digital-binomial-theorem-non-binary} Let $n$ be a non-negative integer. Then for any base $b\geq 2$, we have \begin{align*} \prod_{j=1}^{b-1}\binom{x+y+j-1}{j}^{\mu_j(n)} & =\sum_{0\leq m\preceq_b n}\left( \prod_{j=1}^{b-1}\binom{x+j-1}{j}^{\mu_j(m)} \prod_{j=1}^{b-1}\binom{y+j-1}{j}^{\mu_j(n-m)} \right). \end{align*} \end{corollary} \noindent Observe that if $b=2$, then Corollary~\ref{co:digital-binomial-theorem-non-binary} reduces to Theorem~\ref{th:digital-binomial-theorem}. The source behind Theorem~\ref{th:digital-binomial-theorem} is a one-parameter subgroup of Sierpi\'nski matrices, investigated by Callan \cite{C}, which encodes the digital binomial theorem. To illustrate this, define a sequence of lower-triangular matrix functions $S_N(x)$ of dimension $2^N\times 2^N$ recursively by \begin{equation} S_1(x)=\left( \begin{array}{rr} 1 & 0 \\ x & 1 \\ \end{array} \right), \ \ S_{N+1}(x)=S_1(x)\otimes S_N(x), \end{equation} where $\otimes$ denotes the Kronecker product of two matrices. For example, $S_2(x)$ and $S_3(x)$ can be computed as follows: \begin{align*} S_2(x) & =S_1(x)\otimes S_1(x) = \left(\begin{array}{rr} 1\cdot S_1(x) & 0\cdot S_1(x) \\ x\cdot S_1(x) & 1\cdot S_1(x) \end{array} \right) = \left( \begin{array}{rrrr} 1 & 0 & 0 & 0 \\ x & 1 & 0 & 0 \\ x & 0 & 1 & 0 \\ x^2 & x & x & 1 \end{array} \right), \\ \\ S_3(x) & =S_1(x)\otimes S_2(x) = \left(\begin{array}{rr} 1\cdot S_2(x) & 0\cdot S_2(x) \\ x\cdot S_2(x) & 1\cdot S_2(x) \end{array} \right) =\left( \begin{array}{llllllll} 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\ x & 1 & 0 & 0 & 0 & 0 & 0 & 0 \\ x & 0 & 1 & 0 & 0 & 0 & 0 & 0 \\ x^2 & x & x & 1 & 0 & 0 & 0 & 0 \\ x & 0 & 0 & 0 & 1 & 0 & 0 & 0 \\ x^2 & x & 0 & 0 & x & 1 & 0 & 0 \\ x^2 & 0 & x & 0 & x & 0 & 1 & 0 \\ x^3 & x^2 & x^2 & x & x^2 & x & x & 1 \end{array} \right). \end{align*} Observe that when $x=1$, the infinite matrix $S=\lim_{N\rightarrow \infty}S_N(1)$ contains Sierpi\'nski's triangle. Callan \cite{C} gives a formula for the entries of $S_N(x)=(\alpha_N(j,k,x))$, $0\leq j,k\leq 2^N-1$, in terms of the sum-of-digits function: \begin{equation} \label{eq:formula-for-entries} \alpha_N(j,k,x)=\left\{ \begin{array}{cl} x^{s(j-k)}, & \mathrm{if} \ 0\leq k\leq j \leq 2^N-1 \ \mathrm{and} \ (k, j-k) \ \textrm{are} \ \textrm{carry-free;} \\ 0, & \mathrm{otherwise}. \end{array} \right. \end{equation} Moreover, Callan proved that $S_N(x)$ forms a one-parameter subgroup of $SL(2^N,\mathbb{R})$, i.e., the group of $2^N\times 2^N$ real matrices with determinant one. Namely, we have \begin{equation} \label{eq:one-parameter-matrix-product} S_N(x)S_N(y)=S_N(x+y). \end{equation} If we denote the entries of $S_N(x)S_N(y)$ by $t_N(j,k)$, then the equality \[ t_N(j,k)=\sum_{i=0}^{2^N-1}\alpha_N(j,i,x)\alpha_N(i,k,y)=\alpha_N(j,k,x+y)\] corresponds precisely to Theorem~\ref{th:digital-binomial-theorem} with $j=0$ and $k=0$. For example, if $N=2$, then (\ref{eq:one-parameter-matrix-product}) becomes \begin{align*} \left( \begin{array}{llll} 1 & 0 & 0 & 0 \\ x & 1 & 0 & 0 \\ x & 0 & 1 & 0 \\ x^2 & x & x & 1 \end{array} \right) \left( \begin{array}{llll} 1 & 0 & 0 & 0 \\ y & 1 & 0 & 0 \\ y & 0 & 1 & 0 \\ y^2 & y & y & 1 \end{array} \right) & = \left( \begin{array}{cccc} 1 & 0 & 0 & 0 \\ x+y & 1 & 0 & 0 \\ x+y & 0 & 1 & 0 \\ (x+y)^2 & x+y & x+y & 1 \end{array} \right). \end{align*} The rest of this paper is devoted to generalizing Callan's construction of Sierpi\'nski matrices to arbitrary bases and considering two applications of them. In Section~\ref{two}, we use these generalized Sierpi\'nski matrices to prove Theorem~\ref{th:digital-binomial-theorem-non-binary}. In Section~\ref{three}, we demonstrate how these matrices arise in the construction of Prouhet-Thue-Morse polynomials defined by the author \cite{N2}. In Section~\ref{four}, we describe a group presentation in terms of generators defined through these matrices and show that these generators satisfy a relation that generalizes the three-strand braid relation found by Ferrand \cite{F1}. This relation suggests that Sierpi\'nski matrices encode not only a digital binomial theorem but also an interesting group structure. %%%%%%%%%%%%% \section{Sierpi\'nski triangles} \label{two} To prove Theorem~\ref{th:digital-binomial-theorem-non-binary}, we consider the following generalization of the Sierpi\'nski matrix $S_N(x)$ in terms of binomial coefficients. Define lower-triangular matrices $S_{b,N}(x)$ of dimension $b^N\times b^N$ recursively by \begin{align*} S_{b,1}(x) & =\left( \begin{array}{cccll} 1 & 0 & 0 & \cdots & 0 \\ \binom{x}{1} & 1 & 0 & \cdots & 0 \\ \binom{x+1}{2} & \binom{x}{1} & 1 & \cdots & 0 \\ \vdots & \vdots & \vdots & \ddots & \vdots \\ \binom{x+b-2}{b-1} & \binom{x+b-3}{b-2} & \binom{x+b-4}{b-3} & \cdots & 1 \end{array} \right)= \left\{ \begin{array}{cl} \binom{x+j-k-1}{j-k}, & \textrm{if } 0\leq k \leq j \leq b-1; \\ 0, & \textrm{otherwise}, \end{array} \right. \\ \end{align*} and for $N>1$, \[ S_{b,N+1}(x) =S_{b,1}(x)\otimes S_{b,N}(x). \] \begin{example} To illustrate our generalized Sierpi\'nski matrices, we calculate $S_{3,1}(x)$ and $S_{3,2}(x)$: \begin{align*} S_{3,1}(x) & =\left( \begin{array}{ccc} 1 & 0 & 0 \\ \binom{x}{1} & 1 & 0 \\ \binom{x+1}{2} & \binom{x}{1} & 1 \end{array} \right), \\ S_{3,2}(x) & =S_{3,1}(x)\otimes S_{3,1}(x) \\ & =\left( \begin{array}{ccccccccc} 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\ \binom{x}{1} & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\ \binom{x+1}{2} & \binom{x}{1} & 1 & 0 & 0 & 0 & 0 & 0 & 0 \\ \binom{x}{1} & 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 \\ \binom{x}{1}\binom{x}{1} & \binom{x}{1} & 0 & \binom{x}{1} & 1 & 0 & 0 & 0 & 0 \\ \binom{x}{1}\binom{x+1}{2} & \binom{x}{1}\binom{x}{1} & \binom{x}{1} & \binom{x+1}{2} & \binom{x}{1} & 1 & 0 & 0 & 0 \\ \binom{x+1}{2} & 0 & 0 & \binom{x}{1} & 0 & 0 & 1 & 0 & 0 \\ \binom{x+1}{2}\binom{x}{1} & \binom{x+1}{2} & 0 & \binom{x}{1}\binom{x}{1} & \binom{x}{1} & 0 & \binom{x}{1} & 1 & 0 \\ \binom{x+1}{2}\binom{x+1}{2} & \binom{x+1}{2}\binom{x}{1} & \binom{x+1}{2} & \binom{x}{1}\binom{x+1}{2} & \binom{x}{1}\binom{x}{1} & \binom{x}{1} & \binom{x+1}{2} & \binom{x}{1} & 1 \\ \end{array} \right). \end{align*} \end{example} We now generalize Callan's result for $S_N(x)$ by presenting a formula for the entries of $S_{b,N}(x)$; see \cite{IM} for a similar generalization but along a different vein. \begin{theorem} \label{th:formula-for-entries-arbitrary-b} Let $\alpha_N(j,k):=\alpha_N(j,k,x)$ denote the $(j,k)$-entry of $S_{b,N}(x)$. Then \begin{equation} \label{eq:formula-for-entries-arbitrary-b} \alpha_N(j,k)=\left\{ \begin{array}{cl} \binom{x+d_0-1}{d_0} \binom{x+d_1-1}{d_1} \cdots \binom{x+d_{N-1}-1}{d_{N-1}}, & \mathrm{if} \ 0\leq k\leq j \leq b^N-1 \ \mathrm{and} \ k \preceq_b j; \\ \\ 0, & \mathrm{otherwise}, \end{array} \right. \end{equation} where $j-k=d_0b^0+d_1b^1+\ldots + d_{N-1}b^{N-1}$ is the base-$b$ expansion of $j-k$, assuming $j\geq k$. \end{theorem} \begin{proof} We argue by induction on $N$. It is clear that (\ref{eq:formula-for-entries-arbitrary-b}) holds for $S_{b,1}(x)$. Next, assume that (\ref{eq:formula-for-entries-arbitrary-b}) holds for $S_{b,N}(x)$ and let $\alpha_{N+1}(j,k)$ be an arbitrary entry of $S_{b,N+1}(x)$, where $pb^N\leq j \leq (p+1)b^N-1$ and $qb^N \leq k \leq (q+1)b^N-1$ for some non-negative integers $p,q\in \{0,1,\ldots,b-1\}$. Set $j'=j-pb^N$ and $k'=k-qb^N$. We consider two cases: \vskip 6pt \noindent Case 1. $p vb^N+r. \\ \end{array} \right. \] It follows that \begin{align*} t_{N+1}(j,k) &=\sum_{v=q}^{p}\sum_{r=0}^{b^{N}-1} \binom{x+p-v-1}{p-v} \binom{y+v-q-1}{v-q} \alpha_{N}(j',r,x)\alpha_{N}(r,k',y) \\ &=\left(\sum_{v=q}^{p} \binom{x+p-v-1}{p-v} \binom{y+v-q-1}{v-q}\right) \sum_{r=0}^{b^{N}-1} \alpha_{N}(j',r,x)\alpha_{N}(r,k',y) \\ & = \binom{x+y+p-q-1}{p-q} \alpha_{N}(j',k',x+y) \\ & = \alpha_{N+1}(j,k,x+y), \end{align*} where we have made use of the inductive assumption and Lemma~\ref{le:binomial-sum-of-product-2}. This proves that (\ref{eq:one-parameter}) holds. \end{proof} As a corollary, we obtain Theorem~\ref{th:digital-binomial-theorem-non-binary}, which we now prove. \begin{proof}[Proof of Theorem~\ref{th:digital-binomial-theorem-non-binary}] Let $j=n$ and $k=0$. Then the identity \[ \sum_{m=0}^{b^{N}-1} \alpha_{N}(j,m,x)\alpha_{N}(m,k,y) = \alpha_{N}(j,k,x+y), \] which follows from (\ref{eq:one-parameter}), is equivalent to (\ref{eq:digital-binomial-theorem-non-binary}). \end{proof} We end this section by describing the infinitesimal generator of $S_{b,N}(x)$. Define $X_{b,1}(x)=(\chi_{j,k})$ to be a strictly lower-triangular matrix whose entries $\chi_{j,k}$ are given by \begin{equation} \chi_{j,k}=\left\{ \begin{array}{ll} x/(j-k), & \textrm{if } j\geq k+1; \\ 0, & \textrm{otherwise.} \\ \end{array} \right. \end{equation} For $N>1$, we define matrices \[X_{b,N+1}(x)=X_{b,1}(x)\oplus X_{b,N}(x)=X_{b,1}(x)\otimes I_{b^N} +I_b\otimes X_{b,N}(x), \] where $\oplus$ denotes the Kronecker sum and $I_{b^N}$ denotes the $b^N\times b^N$ identity matrix. Observe that $X_{b,1}(x)$ has the following matrix form: \begin{equation} X_{b,1}(x)=\left( \begin{array}{ccccc} 0 & 0 & 0 & \cdots & 0 \\ x & 0 & 0 & \cdots & 0 \\ x/2 & x & 0 & \cdots & 0 \\ x/3 & x/2 & x & \cdots & 0 \\ \vdots & \vdots & \vdots & \ddots & \vdots \\ x/(b-1) & x/(b-2) & x/(b-3) & \cdots & 0 \end{array} \right) \end{equation} The following lemmas will be needed. The first states a useful identity involving the unsigned Stirling numbers of the first kind, $c(n,k)$, defined by the generating function \[ x(x+1)...(x+n-1)=\sum_{k=0}^n c(n,k) x^k. \] It is well known that $c(n,k)$ counts the number of $n$-element permutations consisting of $k$ cycles. \begin{lemma} \label{le:stirling-number-formula} Let $l$ and $n$ be positive integers with $l\geq n$. Then \begin{equation} \label{eq:stirling-number-formula} \sum_{i=1}^{l-n+1}(i-1)!\binom{l}{i}c(l-i,n-1)=n c(l,n). \end{equation} \end{lemma} \begin{proof} We give a combinatorial argument. Let $A=\{1,2,...,l\}$. We count in two different ways the number of permutations $\pi=\sigma_1\sigma_2 \cdots \sigma_n$ of $A$ consisting of $n$ cycles where we distinguish one of the cycles $\sigma_j$ of $\pi$. On the one hand, since there are $c(l,n)$ such permutations $\pi$ and $n$ ways to distinguish a cycle of $\pi$, it follows that the answer is given by $n c(l,n)$. On the other hand, we can construct $\pi$ by first choosing our distinguished cycle $\sigma_1$ consisting of $i$ elements. The number of possibilities for $\sigma_1$ is $\binom{l}{i}(i-1)!$ since there are $\binom{l}{i}$ ways to choose $i$ elements from $A$ and $(i-1)!$ ways to construct a cycle from these $i$ elements. It remains to construct the remaining cycles $\sigma_2,\dots, \sigma_n$, which we view as a permutation $\pi'=\sigma_2\cdots \sigma_n$ on $l-i$ elements consisting of $n-1$ cycles. Since there are $c(l-i,n-1)$ such possibilities for $\pi'$, it follows that the number of permutations $\pi$ with a distinguished cycle is given by \[\sum_{i=1}^{l-n+1}(i-1)!\binom{l}{i}c(l-i,n-1). \] Equating the two answers yields (\ref{eq:stirling-number-formula}) as desired. \end{proof} \begin{lemma} \label{le:infinitesimal-generator-power} Let $n$ be a positive integer with $1\leq n \leq b-1$. Then \begin{equation} \label{eq:infinitesimal-generator-power} X_{b,1}^n(x)=(\chi_n(j,k)), \end{equation} where the entries $\chi_n(j,k)$ are given by \begin{equation} \label{eq:infinitesimal-generator-entry} \chi_n(j,k)=\left\{ \begin{array}{cl} \frac{n!}{(j-k)!}c(j-k,n)x^n, & \textrm{if } j\geq k+n; \\ 0, & \textrm{otherwise.} \\ \end{array} \right. \end{equation} \end{lemma} \begin{proof} We argue by induction on $n$. It is clear that (\ref{eq:infinitesimal-generator-entry}) holds when $n=1$. Suppose $n>1$. If $j< k+n$, then $\chi_n(j,k)=0$ because $X_{b,1}^m$ is a power of strictly lower-triangular matrices. Therefore, assume $j\geq k+n$. Then \begin{align*} \chi_n(j,k) & =\sum_{i=0}^{b-1}\chi_{n-1}(j,i)\chi_1(i,k) = \sum_{i=k+1}^{j-n+1}\chi_{n-1}(j,i)\chi_1(i,k) \\ & =x^n \sum_{i=k+1}^{j-n+1}\frac{(n-1)!}{(j-i)!}c(j-i,n-1)\frac{1}{i-k} \\ & =x^n \sum_{i=k+1}^{l+k-n+1}\frac{(n-1)!}{(l+k-i)!}c(l+k-i,n-1)\frac{1}{i-k} \\ & = \frac{(n-1)! x^n}{l!} \sum_{m=1}^{l-m}\frac{l!}{m(l-m)!}c(l-m,n-1) \\ & = \frac{(n-1)! x^n}{l!} \sum_{m=1}^{l-n+1}(m-1)!\binom{l}{m}c(l-m,n-1), \end{align*} where $l=j-k$ and $m=i-k$. It follows from Lemma~\ref{le:stirling-number-formula} that \begin{align*} \chi_n(j,k)=\frac{(n-1)!x^n}{l!}\cdot n c(l,n)=\frac{n!}{(j-k)!}c(j-k,n)x^n \end{align*} as desired. \end{proof} \begin{lemma} \label{le:infinitesimal-generator-base-case} We have \begin{equation} \label{eq:infinitesimal-generator-base-case} \mathrm{exp}(X_{b,1}(x))=S_{b,1}(x). \end{equation} \end{lemma} \begin{proof} Denote the entries of $\mathrm{exp}(X_{b,1}(x))$ by $\xi(j,k)$. It is clear that $\xi(j,k)=0$ for $j1$, define \begin{equation} M_{N+1}= M_1\otimes M_N = \left( \begin{array}{rrrrrr} M_N & 0_N & 0_N & \cdots & 0_N & 0_N \\ -M_N & M_N & 0 & \cdots & 0_N & 0_N \\ & & & \ddots \\ 0_N & 0_N & 0_N & \cdots & -M_N & M_N \\ \end{array} \right), \end{equation} where $M_1 \otimes M_N$ denotes the Kronecker product. The following theorem establishes a matrix relationship between the vectors $\mathbf{a}_N$ and $\mathbf{c}_N$. \begin{theorem} \label{th:matrix-equation} Let $A=(a_0,\ldots, a_{b-1})$ be a zero-sum vector. The polynomial equation \begin{equation} \label{eq:polynomial-factorization} F_N(x;A)=P_N(x;\mathbf{c}_N)\prod_{m=0}^{N-1}(1-x^{b^m}) \end{equation} is equivalent to the matrix equation \begin{equation}\label{eq:matrix-equation} \mathbf{a}_N=M_N\mathbf{c}_N \end{equation} together with the condition $c_n=0$ for any $n$ that contains the digit $b-1$ in its base-$b$ expansion, where $0\leq n\leq b^N-1$. \end{theorem} To prove Theorem~\ref{th:matrix-equation}, we shall need the following two lemmas, which we state without proof since their results are easy to verify. \begin{lemma} \label{le:system-equations-base-case} Let $A=(a_0,a_1,\ldots, a_{b-1})$ and $\mathbf{c}_1=(c_0,\ldots,c_{b-1})^T$. Then the polynomial equation \[ \sum_{n=0}^{b-1}a_nx^n=\left(\sum_{n=0}^{b-1}c_nx^n\right)(1-x) \] is equivalent to the system of equations \begin{align*} a_0 & =c_0 \\ a_1 & =-c_0+c_1 \\ & \ \vdots \\ a_{b-1} & =-c_{b-2} +c_{b-1} \end{align*} together with the condition $c_{b-1}=0$. \end{lemma} \begin{lemma} \label{le:matrix-equation-base-case} The system of equations in Lemma~\ref{le:system-equations-base-case} can be expressed in matrix form as \[ \mathbf{a}_1=M_1\mathbf{c}_1. \] \end{lemma} \begin{proof}[Proof of Theorem~\ref{th:matrix-equation}] We argue by induction on $N$. It is clear that (\ref{eq:matrix-equation}) holds for $N=1$ since the polynomial equation $F_1(x;A)=P_1(x;\mathbf{c}_1)(1-x)$ is equivalent to $\mathbf{a}_1=M_1\mathbf{c}_1$ because of Lemmas~\ref{le:system-equations-base-case} and \ref{le:matrix-equation-base-case}. Next, assume that (\ref{eq:matrix-equation}) holds for case $N$. We shall prove that (\ref{eq:matrix-equation}) holds for case $N+1$. Define \[ P_N(x;C_N(p))=\sum_{n=0}^{b^N-1}c_{n+pb^N}x^{n+pb^N} \] to be a polynomial whose coefficients are given by the entries of $C_N(p)=(c_{pb^N},\dots,c_{(p+1)b^N-1})$, i.e., those elements in $\mathbf{c}_{N+1}$ from position $pb^N$ to position $(p+1)b^N$, so that \[ P_{N+1}(x;\mathbf{c}_{N+1}) = P_{N-1}(x;C_N(0))+\dots+P_{N-1}N(x;C_N(b-1)). \] We then expand the right-hand side of (\ref{eq:ptm-polynomial}) for case $N+1$ as follows: \begin{align*} F_{N+1}(x;A) & = \left(P_N(x;C_N(0))+\dots+P_N(x;C_N(b-1))\right)\left(\prod_{m=0}^{N-1}(1-x^{b^m})\right)(1-x^{b^N}) \\ & = \left(Q_N(x;C_N(0))+\dots+Q_N(x;C_N(b-1))\right)(1-x^{b^N}) \\ &= Q_N(x;C_N(0))+\left(Q_N(x;C_N(1))-x^{b^N}Q_N(x;C_N(0))\right)+\dots + \\ & \ \ \ \ \left(Q_N(x;C_N(b-1))-x^{b^N}Q_N(x;C_N(b-2))\right) - x^{b^N}Q_N(x;C_N(b-1)), \end{align*} where we define \[ Q_N(x;C_N(p))=P_N(x;C_N(p))\prod_{m=0}^{N-1}(1-x^{b^m}).\] Next, we equate this result with the definition of $F_{N+1}(x;A)$: \[ F_{N+1}(x;A)=\sum_{n=0}^{b^{N+1}-1}a_{u_b(n)} x^n=F_N(x;A_N(0))+\dots+F_N(x;A_N(b-1)), \] where we define \[ F_N(x;A_N(p))=\sum_{n=0}^{b^N-1}a_{u(n+pb^N)}x^{n+pb^N}. \] This leads to the system of polynomial equations \begin{align*} F_N(x;A_N(0)) & =Q_N(x;C_N(0)) \\ F_N(x;A_N(1)) & =Q_N(x;C_N(1))-x^{b^N}Q_N(x;C_N(0)) \\ & \vdots \\ F_N(x;A_N(b-1)) & =Q_N(x;C_N(b-1))-x^{b^N}Q_N(x;C_N(b-2)) \end{align*} together with the condition $Q_N(x;C_N(b-1))=0$, i.e. $c_n=0$ for all $(b-1)b^N \leq n \leq b^{N+1}-1$. It follows from Lemmas~\ref{le:system-equations-base-case} and \ref{le:matrix-equation-base-case} that this system is equivalent to the system of matrix equations \begin{align*} \mathbf{a}_N(0) & =M_N\mathbf{c}_N(0) \\ \mathbf{a}_N(1) & = -M_N\mathbf{c}_N(0) + M_N\mathbf{c}_N(1) \\ & \vdots \\ \mathbf{a}_N(b-1) & = -M_N\mathbf{c}_N(b-2) + M_N\mathbf{c}_N(b-1), \end{align*} where the first matrix equation by assumption satisfies the condition $c_n=0$ for any $0\leq n\leq b^N-1$ that contains the digit $b-1$ in its base-$b$ expansion. This in turn is equivalent to the matrix equation $\mathbf{a}_{N+1}=M_{N+1}\mathbf{c}_{N+1}$ together with the condition that $c_n=0$ for any $0\leq n \leq b^{N+1}-1$ that contains the digit $b-1$ in its base-$b$ expansion. This establishes the theorem for case $N+1$ and completes the proof. \end{proof} \begin{lemma} The matrix $M_N$ has inverse $S_N=M_N^{-1}$, where $S_N$ is given recursively by \begin{equation} S_1= \left( \begin{array}{rrrrr} 1 & 0 & 0 & \cdots & 0 \\ 1 & 1 & 0 & \cdots & 0 \\ & & & \ddots \\ 1 & 1 & 1 & \cdots & 1 \\ \end{array} \right) \end{equation} and for $N>1$, \begin{equation} S_{N+1}=S_1\otimes S_N= \left( \begin{array}{rrrrr} S_N & 0_N & 0_N & \ldots & 0_N \\ S_N & S_N & 0_N & \ldots & 0_N \\ & & & \ddots \\ S_N & S_N & S_N & \ldots & S_N \\ \end{array} \right). \end{equation} Thus, if $\mathbf{a}_N=M_N\mathbf{c}_N$, then \begin{equation} \label{eq:matrix-equation-case-N} \mathbf{c}_N=S_N\mathbf{a}_N. \end{equation} \end{lemma} \begin{proof} It is straightforward to verify directly that $S_1=M_1^{-1}$. Since $S_{N+1}=S_1\otimes S_{N}$ and $M_{N+1}=M_1\otimes M_N$, it follows that $S_N=M_N^{-1}$. \end{proof} Observe that $S_N=S_{b,N}(1)$ is the generalized Sierpi\'nski triangle defined in the previous section. We now present a formula for the coefficients $c_n$ in terms of the elements of $A$. \begin{theorem} \label{th:formula-coefficients} Let $A=(a_0,\ldots,a_{b-1})$ be a zero-sum vector. Then the polynomial equation (\ref{eq:polynomial-factorization}) has solution $P_N(x;C_N)$ whose coefficients $c_n$, $0\leq n \leq b^N-1$, are given by \begin{equation} \label{eq:formula-coefficients} c_n=\sum_{k\in I_b(n)}a_{u_b(k)}, \end{equation} where $I_b(n)=\{k\in \mathbb{N}: k\preceq_b n \}$. Moreover, $c_n=0$ for all $n$ whose base-$b$ expansion contains the digit $b-1$. \end{theorem} \begin{proof} From Theorem~\ref{th:formula-for-entries-arbitrary-b}, we know that the non-zero entries in the $n$-th row of $S_N$, which are all equal to 1, are located at $(n,k)$ where $k\preceq_b n$. Formula (\ref{eq:formula-coefficients}) now follows from (\ref{eq:matrix-equation-case-N}). It remains to show that (\ref{eq:formula-coefficients}) yields $c_n=0$ for all $n$ whose base-$b$ expansion contains the digit $b-1$. Let $n=n_0b^0+\dots+n_Lb^L+\dots + n_{N-1}b^{N-1}$ where $n_L=b-1$. Denote $I_b(n;L)$ to be the subset of $I_b(n)$ consisting of integers whose base-$b$ expansion has digit 0 at position $L$, i.e., \[ I_b(n;L)=\{k\in \mathbb{N}:k\preceq_b n, k=k_0b^0+\dots+k_Nb^N, k_L=0\}. \] Then using the fact that $A$ is a zero-sum vector, i.e., $a_0+\dots + a_{b-1}=0$, we have \begin{align*} c_n & =\sum_{k\in I_b(n;L)}a_{u_b(k)} + \sum_{k\in I_b(n;L)}a_{u_b(k+b^L)}+\dots+\sum_{k\in I_b(n;L)}a_{u_b(k+(b-1)b^L))} \\ & =\sum_{k\in I_b(n;L)}\left(a_{u_b(k)} + a_{(u_b(k)+1)_b}+\dots+a_{(u_b(k)+(b-1))_b}\right) \\ & = 0 \end{align*} as desired. \end{proof} Next, we specialize Theorem~\ref{th:formula-coefficients} to base $b=3$. Define $w(n)$ to be the sum of the digits of $n$ in its base-$3$ representation modulo 2. \begin{corollary} Suppose $b=3$. Let $n\in \mathbb{N}$ be such that its base-3 representation does not contain the digit 2. Then \begin{equation} \label{eq:formula-coefficient-base-3} c_n=(-1)^{w(n)}a_{u_3(2n)}. \end{equation} \end{corollary} \begin{proof} Let $n=n_03^0+\dots+n_{N-1}3^{N-1}$ be the base-3 representation of $n$ with no digit equal to 2 and leading digit $n_{N-1}=1$. We argue by induction on $N$. It is clear that (\ref{eq:formula-coefficient-base-3}) holds when $N=1$ since from (\ref{eq:formula-coefficients}) we have \[ c_1=a_0+a_1=-a_2=(-1)^{w(n)}a_{u_3(2n)}. \] Next, assume (\ref{eq:formula-coefficient-base-3}) holds for a given $N$. To prove that (\ref{eq:formula-coefficient-base-3}) holds for $N+1$, we consider two cases by decomposing $n=n'+3^{N}$. \\ \noindent Case 1: $n_i=0$ for all $i=0,\ldots,N-1$. Then $n=3^n$ and \begin{align*} c_{n} & =\sum_{k\in I_3(n)}a_{u_3(k)}=a_{u_3(0)}+a_{u_3(n)}=a_0+a_1 = -a_2 \\ & = (-1)^{w(n)}a_{u_3(2n)}. \end{align*} \noindent Case 2: $n_L=1$ for some $0\leq L \leq N-1$. Then set $n''=n'-3^L$. It follows that \begin{align*} c_{n} & =\sum_{k\in I_3(n)}a_{u_3(k)}=\sum_{k\in I_3(n,L)}a_{u_3(k)}+\sum_{k\in I_3(n,L)}a_{u_3(k+3^N)} \\ & =\sum_{k\in I_3(n')}a_{u_3(k)}+\sum_{k\in I_3(n')}a_{u_3(k+3^L)} \\ & =\sum_{k\in I_3(n')}a_{u_3(k)}+\sum_{k\in I_3(n',L)}a_{u_3(k+3^L)}+\sum_{k\in I_3(n',L)}a_{u_3(k+2\cdot 3^L)} +\sum_{k\in I_3(n',L)}a_{u_3(k)}-\sum_{k\in I_3(n',L)}a_{u_3(k)} \\ & =\sum_{k\in I_3(n')}a_{u_3(k)}+\sum_{k\in I_3(n',L)}\left( a_{u_3(k)}+a_{(u_3(k)+1)_3} +a_{(u_3(k)+2)_3}\right) -\sum_{k\in I_3(n'')}a_{u_3(k)} \\ & = (-1)^{w(n')}a_{u_3(2n')}- (-1)^{w(n'')}a_{u_3(2n'')} \\ & = (-1)^{w(n')}a_{u_3(2n')}+(-1)^{w(n''+3^L)}a_{u_3(2(n'-3^L))} \\ & = (-1)^{w(n')}\left( a_{u_3(2n')}+a_{(u_3(2n')-2)_3} \right) = (-1)^{w(n')}\left(-a_{(u_3(2n')-1)_3}\right) \\ & = (-1)^{w(n'+3^N)}a_{(u_3(2n')+2)_3))} = (-1)^{w(n)}a_{u_3(2n'+2\cdot 3^L)} \\ & = (-1)^{w(n)}a_{u_3(2n)}. \end{align*} Thus, (\ref{eq:formula-coefficient-base-3}) holds for $N+1$. \end{proof} %%%%%%%%%%%%%%%% \section{Group generators and relations} \label{four} In this section, we reveal further evidence of the rich structure of Sierpi\'nski matrices by describing group generators and relations defined by the matrices $S_N$ and $M_N$. Recall that $S_N$ and $M_N$ are matrices of dimension $b^N\times b^N$ for a given base $b$. Define $T_N=M_N^t$ to be the transpose of $M_N$ and $U_N=S_NT_N$, $V_N=T_NS_N$. The following lemma gives a recursive construction of $U_N$ and$V_N$. \begin{lemma} We have \begin{align*} U_{N+1} & =U_1\otimes U_N \\ V_{N+1} & = V_1\otimes V_N \end{align*} \end{lemma} \begin{proof} The result follows from the mixed-product property of the Kronecker product. We demonstrate this for $U_{N+1}$: \begin{align*} U_{N+1}=S_{N+1}T_{N+1}=(S_1\otimes S_N)(T_1\otimes T_N)=(S_1T_1)\otimes (S_NT_N)=U_1\otimes U_N. \end{align*} The calculation is the same for $V_{N+1}$. \end{proof} Observe that $U_N=(u_{i,j})$ and $V_N=(v_{i,j})$ are skew-triangular and that $V_N$ is the skew-transpose of $U_N$, i.e., $v_{i,j}=u_{b-1-j,b-1-i}$ for $i,j=0,1,\ldots,b-1$. Here are some examples of $U_N$ and $V_N$ when $b=2$: \[ U_1=\left( \begin{array}{rr} 1 & -1 \\ 1 & 0 \end{array} \right), \ \ U_2=\left( \begin{array}{rrrr} 1 & -1 & -1 & 1 \\ 1 & 0 & -1 & 0 \\ 1 & -1 & 0 & 0 \\ 1 & 0 & 0 & 0 \end{array} \right), \] \[ V_1=\left( \begin{array}{rr} 0 & -1 \\ 1 & 1 \end{array} \right), \ \ V_2=\left( \begin{array}{rrrr} 0 & 0 & 0 & 1 \\ 0 & 0 & -1 & -1 \\ 0 & -1 & 0 & -1 \\ 1 & 1 & 1 & 1 \end{array} \right). \] The next lemma establishes that the eigenvalues of $U_N$ and $V_N$ are $(b+1)$-th roots of 1 or $-1$. \begin{lemma} \label{le:eigenvalues} The set of eigenvalues of $U_1$ and $V_1$ are exactly the same and consist of all roots of the polynomial equation \[ -1+r-r^2+\dots+(-1)^{b+1} r^b=0. \] \end{lemma} \begin{proof} Let $r$ be a root of $-1+r-r^2+\dots+(-1)^{b+1} r^b=0$. We claim that $r$ is an eigenvalue of $U_1$ with eigenvector $\mathbf{v}=(v_1,\ldots, v_b)^T$, where \[ v_k=\sum_{j=1}^k (-1)^{j+1}r^j \] for $k=1,\ldots ,b$. Observe that $v_b=1$. Denote $\mathbf{w}=(U_1-rI_b)\mathbf{v}=(w_1,\ldots,w_b)^T$. It suffices to show $\mathbf{w}=0$. It is straightforward to verify that \[ U_1=(u_{j,k})= \begin{cases} 1, & \mbox{if } k=1; \\ -1, & \mbox{if } k=j+1; \\ 0, & \mbox{otherwise.} \end{cases} \] We now calculate $w_j$ by considering three cases: \vskip 6pt \noindent Case 1: $k=1$. Then \[ w_1=(u_{1,1}-r)v_1+u_{1,2}v_2=(1-r)r-(r-r^2)=0. \] Case 2: $12$. \begin{thebibliography}{9} \bibitem{AS} J.-P. Allouche and J. Shallit, The ubiquitous Prouhet-Thue-Morse sequence, in C. Ding, T. Helleseth, and H. Niederreiter, eds., \textit{Sequences and Their Applications}, Proc.~SETA'98, Springer-Verlag, 1999, pp.~1--16. \bibitem{BEJ} T. Ball, T. Edgar, and D. Juda, Dominance orders, generalized binomial coefficients, and Kummer's theorem, \textit{Math. Mag.} \textbf{87} (2014), 135--143. \bibitem{C} D. Callan, Sierpinski's triangle and the Prouhet-Thue-Morse word, preprint, \url{http://arxiv.org/abs/math/0610932}. \bibitem{F1} E. Ferrand, Pascal and Sierpinski matrices, and the three strand braid group, preprint, \url{http://webusers.imj-prg.fr/~emmanuel.ferrand/publi/PSB.ps}. \bibitem{F2} E. Ferrand, An analogue of the Thue-Morse sequence, \textit{Elect. J. Comb.} \textbf{14} (2007), \#R30. \bibitem{G} H. W. Gould, Some generalizations of Vandermonde's convolution, \textit{Amer. Math. Monthly} {\bf 63} (1956), 84--91. \bibitem{IM} A. Imani and A. R. Moghaddamfar, The inverse of the Pascal lower triangular matrix modulo $p$, \textit{Acta Math. Univ. Comenianae} \textbf{79} (2010), 135--142. \bibitem{L} D. H. Lehmer, The Tarry-Escott problem, \textit{Scripta Math.} \textbf{13} (1947), 37--41. \bibitem{N1} H. D. Nguyen, A digital binomial theorem, preprint, \\ \url{http://arxiv.org/abs/1412.3181}. \bibitem{N2} H. D. Nguyen, A new proof of the Prouhet-Tarry-Escott problem, preprint, \\ \url{http://arxiv.org/abs/1411.6168}. \bibitem{P} E. Prouhet, M\'{e}moire sur quelques relations entre les puissances des nombres, \textit{C. R. Acad. Sci. Paris}, \textbf{33} (1851), 225. \end{thebibliography} \bigskip \hrule \bigskip \noindent 2010 {\it Mathematics Subject Classification}: Primary 11C20; Secondary 05A10. \noindent \emph{Keywords: } binomial theorem, Sierpi\'nski triangle, Prouhet-Thue-Morse sequence. \bigskip \hrule \bigskip \noindent (Concerned with sequence \seqnum{A010060}.) \bigskip \hrule \bigskip \vspace*{+.1in} \noindent Received February 1 2015; revised version received April 30 2015. Published in {\it Journal of Integer Sequences}, May 26 2015. \bigskip \hrule \bigskip \noindent Return to \htmladdnormallink{Journal of Integer Sequences home page}{http://www.cs.uwaterloo.ca/journals/JIS/}. \vskip .1in \end{document} .