cs0607089/SR2.tex
1: \documentclass[12pt]{article}
2: \usepackage{amssymb}
3: \usepackage{amsmath}
4: \usepackage{theorem}
5: %\usepackage{amsthm}
6: \usepackage{latexsym}
7: %\usepackage{showkeys}
8: \usepackage{amscd}
9: %\usepackage[small,nohug,heads=littlevee]{diagrams}
10: %\diagramstyle[labelstyle=\scriptstyle]
11: %\textwidth 7in\oddsidemargin-.7cm\evensidemargin-.7cm
12: %\textheight 8.8in\topmargin -0.5in%
13: \textwidth 6.5in\oddsidemargin 0in
14: \textheight 9in\topmargin -0.5in
15: \renewcommand{\baselinestretch}{1.1}
16: %\parindent 0mm
17: %\parskip 2ex minus 0.2ex
18: %\reversemarginpar
19: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
20: \theoremheaderfont{\bf} %\theoremstyle{break}
21: \theorembodyfont{\sl}
22: \newtheorem{theo}{Theorem}[section]
23: {\theorembodyfont{\rm} \newtheorem{defi}[theo]{Definition}}
24: {\theorembodyfont{\rm} \newtheorem{exa}[theo]{Example}}
25: {\theorembodyfont{\rm} \newtheorem{rem}[theo]{Remark}}
26: {\theorembodyfont{\rm} \newtheorem{notation}[theo]{Notation}}
27: {\theorembodyfont{\rm} \newtheorem{que}[theo]{Question}}
28: \newtheorem{prop}[theo]{Proposition}
29: \newtheorem{cor}[theo]{Corollary}
30: \newtheorem{lemma}[theo]{Lemma}
31: \newtheorem{conjecture}[theo]{Conjecture}
32: \newenvironment{completion}{{\sc Completion of the proof of
33: theorem{\footnotesize{~\ref{O}}}:}}{\mbox{}\hfill$\Box$\par}
34: \newenvironment{proof}{{\sc Proof:}}{\mbox{}\hfill$\Box$\par}
35: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
36: \newcommand{\eqr}[1]{~\mbox{$(${\rm \ref{#1}}$)$}}
37: \newcommand{\Section}[1]{\section{#1}\setcounter{equation}{0}}
38: \renewcommand{\theequation}{\thesection.\arabic{equation}}
39: \newcommand{\comment}[1]{\vspace{5mm}\hspace{15mm}
40: \begin{minipage}{5in}\baselineskip 7mm\large
41:   {{\bf Comment/Question: } #1}
42: \end{minipage}\vspace{5mm}
43: \typeout{Here is a comment line}}
44: \newcommand{\junk}[1]{}
45: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
46: \newcommand{\tx}{\textstyle}
47: \newcommand{\N}{{\mathbb N}}
48: \newcommand{\F}{{\mathbb F}}
49: \newcommand{\K}{{\mathbb K}}
50: \newcommand{\Z}{{\mathbb Z}}
51: \newcommand{\C}{{\mathcal C}}
52: \renewcommand{\H}{{\mathcal H}}
53: \newcommand{\wt}{{\rm wt}}
54: \newcommand{\rank}{{\rm rank}\,}
55: \newcommand{\spann}{\mbox{\rm span}}
56: \newcommand{\dfree}{\mbox{$d_{\mbox{\rm\tiny free}}$}}
57: \newcommand{\delay}[1]{\mbox{$\overleftarrow{#1}$}}
58: \newcommand{\T}{\mbox{$\!^{\sf T}$}}
59: \newcommand{\floor}[1]{\mbox{$\lfloor{#1}\rfloor$}}
60: \newcommand{\ceiling}[1]{\mbox{$\lceil{#1}\rceil$}}
61: \newcommand{\FDlaurent}{\mbox{$\F(\!(D)\!)$}}
62: \newcommand{\zwei}[2]{\left[ \begin{array}{c}
63:                    #1 \\ #2 \end{array} \right]}
64: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
65: \newenvironment{liste}{\begin{list}{--\hfill}{\topsep-1.4ex \labelwidth.4cm
66:    \leftmargin.5cm \labelsep.1cm \rightmargin0cm \parsep0ex \itemsep.6ex
67:    \partopsep1.4ex}}{\end{list}}
68: \newcounter{abc}
69: \newenvironment{romanlist}{\begin{list}{(\roman{abc})\hfill}{\usecounter{abc}
70:      \topsep.5ex \labelwidth.6cm \leftmargin.7cm \labelsep.1cm
71:      \rightmargin0cm \parsep0ex \itemsep.6ex
72:      \partopsep1.6ex}}{\end{list}}
73: \newenvironment{alphalist}{\begin{list}{(\alph{abc})\hfill}{\usecounter{abc}
74:      \topsep.5ex \labelwidth.6cm \leftmargin.7cm \labelsep.1cm
75:      \rightmargin0cm \parsep0ex \itemsep.6ex
76:      \partopsep1.6ex}}{\end{list}}
77: \newenvironment{arabiclist}{\begin{list}{(\arabic{abc})\hfill}{\usecounter{abc}
78:      \topsep.5ex \labelwidth.6cm \leftmargin.7cm \labelsep.1cm
79:      \rightmargin0cm \parsep0ex \itemsep.6ex
80:      \partopsep1.6ex}}{\end{list}}
81: \newenvironment{algo}{\begin{list}{{\rmfamily\bf
82:         Step~\arabic{abc}:}\hfill}{\usecounter{abc}
83:      \topsep.5ex \labelwidth1.7cm \leftmargin1.7cm \labelsep0cm
84:      \rightmargin0cm \parsep0ex \itemsep.6ex
85:      \partopsep1.6ex}}{\end{list}}
86: \newcommand{\vier}[4]{\left[ \begin{array}{cc}
87:                    #1 & #2 \\ #3 & #4 \end{array} \right]}
88: 
89: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
90: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
91: \title{Superregular Matrices and the Construction of Convolutional Codes having
92: a Maximum Distance Profile}
93:        
94: \author{
95:   Ryan Hutchinson\\
96:   {\small Mathematics Institute}\vspace{-2mm}\\
97:   {\small University of Z\"urich}\vspace{-2mm}\\
98:   {\small Z\"urich, Switzerland}\vspace{-2mm}\\
99:   {\small {\em e-mail:} rhutchin@math.unizh.ch}
100:   \and
101:   Roxana Smarandache\thanks{On leave at the University of Notre Dame, Department of Mathematics, Notre Dame, IN 46556.} \\
102:   {\small Department of Mathematics and Statistics}\vspace{-2mm}\\
103:   {\small San Diego State University}\vspace{-2mm}\\
104:   {\small San Diego, CA 92182-7720, USA}\vspace{-2mm}\\
105:   {\small {\em e-mail:} rsmarand@sciences.sdsu.edu }
106:   \and
107:   Jochen Trumpf\thanks{Currently seconded to National ICT Australia Limited, which is funded by the Australian
108:   Government's Department of Communications, Information Technology and the Arts and the Australian Research Council
109:   through Backing Australia's Ability and the ICT Centre of Excellence Program.} \\
110:   {\small Department of Information Engineering, RSISE}\vspace{-2mm}\\
111:   {\small The Australian National University}\vspace{-2mm}\\
112:   {\small Canberra ACT 0200, Australia}\vspace{-2mm}\\
113:   {\small {\em e-mail:} Jochen.Trumpf@anu.edu.au}
114: }
115: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
116: \begin{document}
117: \maketitle
118: \begin{abstract} Superregular matrices are a class of lower triangular Toeplitz
119: matrices that arise in the context of constructing convolutional codes having a
120: maximum distance profile. These matrices are characterized by the property that
121: no submatrix has a zero determinant unless it is trivially zero due to the lower triangular structure.  
122: In this paper, we discuss how superregular matrices may be used to construct
123: codes having a maximum distance profile. We also introduce group actions
124: that preserve the superregularity property and present an upper bound on the
125: minimum size a finite field must have in order that a superregular matrix of a
126: given size can exist over that field.
127: \\
128: 
129: \noindent
130: {\bf Keywords:} convolutional codes, column distances, maximum distance profile,
131: superregular matrices, partial realization problem 
132: \end{abstract}
133: 
134: \section{Introduction}
135: Convolutional codes are a class of error-correcting codes that have enjoyed
136: wide use in practical applications due to the existence of efficient
137: non-algebraic decoding algorithms. From a mathematical standpoint, however, the
138: situation is still rather unsatisfying, as there are relatively few algebraic
139: constructions of convolutional codes with provably good distance properties or
140: which can be algebraically decoded.  Recent years have seen interesting
141: developments in the algebraic theory of convolutional codes: the papers~\cite{Heide2,Heide4,Heide5,gl02u,14}
142: extend  the notion of cyclicity familiar from block code theory to
143: convolutional codes; the papers~\cite{Heide3,Heide6} investigate
144: weight enumerators and the existence of a MacWilliams Identity for convolutional
145: codes; the paper~\cite{ro99a} uses methods from
146: systems theory to construct convolutional codes having a designed distance; and the papers~\cite{Heide1,gl03r,12,ro99a1,sm01a} provide results concerning
147: convolutional codes having certain maximal distance properties.  Motivated by existence results appearing in certain members of this last set of papers, we
148: decided to investigate so-called {\em superregular matrices}.  These matrices arise when
149: one considers the problem of constructing convolutional codes having a maximum
150: distance profile.
151: 
152: The remainder of this paper is structured as follows.  In Section 2, we give a
153: brief introduction to convolutional codes, explain the maximal distance
154: properties mentioned above, and define the superregularity property.  In Section
155: 3, we discuss how superregular matrices may be used to construct codes having a
156: maximum distance profile.  In Section 4, we introduce group actions that
157: preserve the superregularity property; these actions made it possible to reduce
158: the computation time necessary for performing searches for superregular matrices.  Finally, in Section 5, we present an upper bound on
159: the minimum field size required for a superregular matrix of a given size to
160: exist.
161: 
162: \section{Preliminaries}
163: In this section, we review the theory of convolutional codes relevant to the
164: presented work.  Let $k$ and $n$ be positive integers with $k<n$.  Let $\F$ be a
165: finite field of characteristic $p$, where $p$ is a prime number.
166: \begin{defi}
167: A {\em convolutional code} $\mathcal{C}$ of {\em rate} $k/n$ is a
168: rank-$k$ submodule of the free module $(\F [s])^n$.
169: \end{defi}  
170: Since $\F [s]$ is a principal ideal domain, $\C$ is a free module and has a well-defined rank.  It follows that a convolutional code may be viewed 
171: as the column space of an $n\times k$
172: polynomial matrix $G(s)$, the columns of which form an $\F [s]$-basis for
173: $\mathcal{C}$.  As a set, we have 
174: $$
175:   \C = \{v(s)=G(s)u(s) \in (\F [s])^n \, | \, u(s)\in (\F [s])^k\}.
176: $$ $G(s)$ is called a {\em generator matrix} for $\mathcal{C}$, the
177: $u(s)$ are called {\em information vectors}, and the $v(s)$ are called
178: {\em code vectors} or {\em codewords}.  Two generator matrices
179: $G_1(s)$ and $G_2(s)$ having full column rank generate the same code
180: if and only if there exists a $k\times k$ unimodular matrix $U(s)$
181: such that $G_1(s)=G_2(s)U(s)$.
182: 
183: The columns of $G(s)$ may be thought of as polynomials with
184: coefficients in $\F ^n$; we refer to the degrees of these polynomials
185: as the {\em column degrees} of $G(s)$ and denote the degree of the
186: $i$th column by $\mu_i$.  Let $m:=\max_{1\leq i\leq k}\{ \mu _i\}$.
187: Thinking of $(\F [s])^{(n\times k)}$ as $\F ^{(n\times k)}[s]$, we may then expand $G(s)$ into a matrix polynomial,
188: $$
189:   G(s)=G_0  + G_1s + \cdots + G_ms^m,
190: $$ where $G_i$ is an $n\times k$ matrix over $\F$, the entries of
191: which are the coefficients of $s^i$ in $G(s)$.  Similarly, thinking of $(\F [s])^k$ as $\F ^k[s]$, we may expand 
192: $u(s)
193: \in (\F [s])^k$ of degree $l$ into a vector
194: polynomial:
195: $$
196:   u(s)=u_0 + u_1s + \cdots + u_ls^l.
197: $$
198: We may then represent the encoding process with the multiplication
199: $$
200:   \left[\begin{array}{c}
201:     v_0\\
202:     v_1\\
203:     \vdots \\
204:     v_{l+m}
205:   \end{array}\right]
206:   =
207:   \left[\begin{array}{ccccc}
208:     G_0 & 0 & \cdots & \cdots & 0\\
209:     G_1 & G_0 & \ddots &  & \vdots\\
210:     \vdots & G_1 & \ddots & \ddots & \vdots\\
211:     G_{m-1} & \vdots & \ddots & \ddots & 0 \\
212:     G_m & G_{m-1} &  & \ddots & G_0\\
213:     0 & G_m & \ddots &  & G_1\\
214:     \vdots & \ddots & \ddots & \ddots & \vdots\\
215:     \vdots &  & \ddots & \ddots & G_{m-1}\\
216:     0 & \cdots & \cdots & 0 &   G_m
217:   \end{array}\right]
218:   \left[\begin{array}{c}
219:   u_0\\
220:   u_1\\
221:   \vdots \\
222:   u_l
223:   \end{array}\right],
224: $$ where 0 represents the $n\times k$ matrix with all entries zero.
225: The large matrix in the middle is called a {\em sliding generator
226: matrix}.  From this representation, we see the origin of the name
227: convolutional code: the vector coefficients of the information vector
228: are convoluted with the matrix coefficients of $G(s)$ to form the
229: codewords.
230: 
231: A generator matrix $G(s)$ is called {\em basic} if the only common
232: divisors of the determinants of its $k\times k$ submatrices belong to
233: $\F\, \backslash \{0\}$.  There are several characterizations of this
234: property; more details may be found in~\cite{gl02u,14,15}.  We note
235: here two of these characterizations.  The first is that a
236: convolutional code generated by a basic generator matrix is a direct
237: summand of $\F ^n[s]$.  The second is that there
238: exists a basic $n\times (n-k)$ matrix over $\F [s]$, $H(s)$, such that
239: $\C$ is the right $\F [s]$-kernel of $H^T(s)$.  Thus, an observable
240: code may also be described as the set
241: $$
242:   \C = \{ v(s)\in (\F [s])^n \, | \, H^T(s)v(s) = [0\,0\cdots0]^T\in (\F [s])^k \}
243: $$
244: $H(s)$ is called a {\em parity check matrix} of $\C$.  As with the generator
245: matrix, one may expand H(s) as the sum
246: $$
247:   H(s) = H_0 + H_1s + \cdots + H_{m'}s^{m'},
248: $$ where $m'$ is the largest integer such that $H_{m'} \neq 0$.  A
249: convolutional code $\mathcal{C}$ is called {\em observable} if one
250: (and hence every) generator matrix of $\mathcal{C}$ is basic.
251: 
252: The {\em high-order coefficient matrix} $G_{\infty}$ of
253: $G(s)$ is a matrix, the $i$th column of which is the vector coefficient of $s^{\mu_i}$ in
254: column $i$.  
255: A basic generator matrix for which $G_{\infty}$ has full rank is
256: called {\em minimal}.  
257: An important invariant of a convolutional code is its {\em complexity}, defined as follows:
258: \begin{defi}
259: The {\em complexity} $\delta$ of a convolutional code $\mathcal{C}$ is
260: the maximum of the degrees of the (polynomial) determinants of the $k\times k$
261: submatrices of any generator matrix of $\mathcal{C}$.
262: \end{defi}
263: This definition makes sense, as the equivalence relation described above
264: preserves the degrees of these determinants.  A code of rate $k/n$ and complexity $\delta$
265: will also be referred to as an $(n,k,\delta)$-code.  If the code is observable, the
266: complexity is normally referred to as the {\em degree} of the code;
267: see~\cite{ro99a1} for the geometric motivation for this terminology.    
268: 
269: In general, $G_{\infty}$ need not have full rank; in this case,
270: $\sum_{i=1}^k \mu_i >\delta$.  It is shown in~\cite{17} that it is
271: always possible to find a unimodular matrix $U(s)$ such that
272: $G(s)U(s)$ has a full-rank high-order coefficient matrix and a set
273: $\{\nu_i\}_{i=1}^k$ of column degrees satisfying
274: $\nu_1\leq\nu_2\leq\cdots\leq\nu_k$ and $\sum_{i=1}^k \nu_i =\delta$.
275: In this case, the column degrees are invariants of the code generated
276: by $G(s)$ and are called the {\em column indices} or the {\em
277:   Kronecker indices} of the code.  If $G(s)$ is minimal, the column
278: degrees $\nu_1\leq\nu_2\leq\cdots\leq\nu_k$ mentioned above are called
279: the {\em minimal column indices} or the {\em Forney indices} of the
280: code.  For the remainder of the paper, we assume all codes to be
281: observable and generator matrices to be minimal.
282: 
283: The following truncated sliding generator matrices $G^c_j\in\F^{(j+1)n\times(j+1)k}$
284: and parity check matrices $H^c_j\in\F^{(j+1)n\times(j+1)(n-k)}$ will be of
285: importance in this work.  These are defined for each $j\in \N _0$ as
286: \begin{equation}\label{e-Gcj}
287: \begin{array}{rcl}
288:        G^c_j := &\begin{bmatrix}
289:                  G_0& 0& \cdots &0\\
290:                  G_1& G_0& \ddots &\vdots\\
291:                  \vdots & \vdots & \ddots &0\\
292:                  G_j&G_{j-1}& \cdots&G_0
293:               \end{bmatrix}\,\,\mbox{and}\,\,
294:        H^c_j := \begin{bmatrix}
295:                  H_0& H_1&\cdots&H_j      \\
296:                  0& H_0& \cdots & H_{j-1}      \\
297:                  \vdots& \ddots & \ddots &\vdots\\       
298:                  0&\cdots& 0 &H_0
299:               \end{bmatrix},
300: \end{array}
301: \end{equation}
302: where $G_j=0$ ($H_j=0$) if $j>m$ ($j>m'$).  The relation $H^T(s)G(s) =
303: 0$ immediately implies that $(H^c_j)^TG^c_j = 0$ for all $j\in \N _0$.
304: 
305: We turn now to some distance notions for convolutional codes. 
306: \begin{defi}
307: Let $x \in \F ^n$.  The {\em Hamming weight} of $x$, wt($x$), is the
308: number of nonzero components of $x$.  Let $v(s)\in \F ^n[s]$ be given
309: by $v(s):=v_0 + v_1s + \cdots + v_ls^l$ for some nonnegative integer
310: $l$.  The {\em weight} of $v(s)$, wt($v(s)$), is the sum of the
311: weights of its $\F ^n$-coefficients:
312: $$
313:   \mbox{wt}(v(s)):=\sum_{i=0}^l\mbox{wt}(v_i).
314: $$
315: \end{defi}
316: We then have
317: \begin{defi}
318: Let $\C$ be an $(n,k,\delta)$-code.  Then, the {\em free distance} of
319: $\C$, $d_{free}(\C)$, is
320: $$
321:   d_{free}(\C):=\min_{v(s)\in \mathcal{C}}\{\mbox{wt}(v(s)) \, | \, v(s) \not = 0\}.
322: $$
323: \end{defi}
324: The following theorem gives an upper bound for how large the free
325: distance of an $(n,k,\delta)$-code can be:
326: \begin{theo}\label{Singleton}
327: Let $\C$ be an $(n,k,\delta)$-code.  Then, $d_{free}(\C)$ satisfies
328: $$
329:   d_{free}(\C)\leq (n-k)\Big(\Big\lfloor\frac {\delta}{k} \Big\rfloor +1\Big) +\delta +1.
330: $$
331: \end{theo}
332: This bound is proven in~\cite{ro99a1}.  It is known as the {\em
333: generalized Singleton bound}.  The reason for this is that a
334: convolutional code of complexity 0 is a block code, as it has a
335: generator matrix with all entries in $\F$.  If we set $\delta =0$ in
336: the above expression, it reduces to the Singleton bound from the
337: theory of block codes.  In analogy with the block code case,
338: convolutional codes having a free distance meeting the generalized
339: Singleton bound are called {\em maximum distance separable} ({\em
340: MDS}).
341: 
342: The free distance is a global distance measure and determines the
343: maximum number of errors that may be introduced to a codeword without
344: jeopardizing correct decoding.  A more local distance measure, relevant 
345: to the performance of {\em sequential decoding algorithms}
346: (see, for example,~\cite{Costello}), is given by column distances.
347: These are defined as follows:
348: \begin{defi}
349: Let $\mathcal{C}$ be a convolutional code.  For $j\in \N _0$ and a
350: polynomial vector $v(s)\in \F ^n[s]$ of degree $l$, set
351: $v_{[0,j]}(s):=v_0+v_1s+\cdots +v_js^j$ (where $v_j=0$ if $j>l$).  Then, the {\em $j$th column
352: distance} of $\mathcal{C}$, $d_j^c(\mathcal{C})$, is defined as
353: $$
354: d_j^c(\mathcal{C}):=\min _{v(s)\in \mathcal{C}}\{\wt(v_{[0,j]}(s)) \,
355: | \, v_0\neq 0 \}.
356: $$
357: Because of the assumption that $\C$ is observable, the fact that $v_0 \neq 0$
358: means that the minimum is taken over codewords resulting from information
359: vectors with $u_0 \neq 0$.
360: It is easy to see that the column distances satisfy 
361: \begin{equation} 
362: \label{e-dist.inequ}
363:   d^c_0(\mathcal{C})\leq d^c_1(\mathcal{C})\leq d^c_2(\mathcal{C})\ldots 
364:   \leq \lim_{j\rightarrow\infty}d^c_j(\mathcal{C})=d_{free}(\C).
365: \end{equation}
366: \end{defi}
367: The $(m +1)$-tuple of numbers $(d^c_0(\C), d^c_1(\C),\ldots,d^c_m(\C))$ is called the
368: {\em column distance profile} of the code~\cite[p. 112]{jo99}.
369: 
370: The following theorem gives an upper bound for the $j$th column distance:
371: \begin{theo}                  \label{P-dcj.bound}
372: Let $\mathcal{C}$ be an $(n,k,\delta)$-code.  For every $j\in \N _0$, we have
373: \[
374: d_j^c(\mathcal{C})\leq(n-k)(j+1)+1.
375: \]
376: If $d_j^c(\mathcal{C})=(n-k)(j+1)+1$ for some $j$, then $d_i^c(\mathcal{C})=(n-k)(i+1)+1$ 
377: when $i\leq j$.
378: \end{theo}
379: A proof may be found in~\cite{gl03r}. By considering the bound in
380: Theorem~\ref{Singleton}, one easily sees that the largest integer for
381: which the upper bound in Theorem~\ref{P-dcj.bound} can be attained is
382: $L:=\lfloor\frac{\delta}{k}\rfloor +
383: \lfloor\frac{\delta}{n-k}\rfloor$.  An $(n,k,\delta)$-code $\C$ is
384: said to be {\em maximum distance profile} ({\em MDP}) if $d_L^c(\C) =
385: (L + 1)(n-k) + 1$; note that Theorem~\ref{P-dcj.bound} implies that
386: $d_j^c(\C) = (j + 1)(n-k) + 1$ for all $j\in\{0,1,\ldots ,L\}$.  In
387: other words, the column distances of an MDP code are maximal for as
388: long as possible.
389: 
390: We close this section by introducing superregular matrices; in Section
391: 2, we will see their relevance to the construction of MDP codes.
392: \begin{defi}\label{SRMDef}
393: Let $A$ be the $(\gamma +1) \times (\gamma +1)$ lower triangular Toeplitz
394:       matrix  
395:       $$
396:                 \begin{bmatrix}
397:           a_0    & 0      & \cdots  & 0        \\
398:           a_1    & a_0    & \ddots  & \vdots   \\
399:           \vdots & \ddots & \ddots  & 0        \\
400:           a_{\gamma}    & \cdots & a_1     & a_0
401:           \end{bmatrix}.
402:       $$
403:       Let $s\in \{1,2,\ldots ,\gamma +1\}$.  Suppose that $I:=\{ i_1,\ldots ,i_s \}$ is a set of row indices of $A$,
404:       $J:=\{ j_1,\ldots ,j_s \}$ is a set of column indices of $A$, and that that the elements of each set are
405:       ordered from least to greatest.  We denote by $A^{i_1,\ldots ,i_s}_{j_1,\ldots ,j_s}$
406:       the submatrix of $A$ formed by intersecting the columns indexed by the members of $J$ and the rows indexed by
407:       the members of $I$.  A submatrix of $A$ is said to be {\em proper} if, for each $\nu\in \{ 1,2,\dots ,s \}$, the
408:       inequality $j_\nu\leq
409:       i_\nu$ holds.  The matrix $A$ is said to be {\em superregular} if every proper submatrix of $A$ has a nonzero determinant.
410: \end{defi}
411: \begin{rem}\label{SRMRem}
412: 
413: Observe that the proper submatrices of $A$ are the
414: only ones that can possibly have a nonzero determinant: If
415: $j_{\nu}>i_{\nu}$ for some $\nu$, then, in the submatrix
416: $A^{i_1,\ldots,i_s}_{j_1,\ldots,j_s}$, the upper right block
417: consisting of the first $\nu$ rows and the last $s-\nu+1$ columns contains only zero entries.  Hence, the matrix formed from the first $\nu$ rows
418: of $A^{i_1,\ldots,i_s}_{j_1,\ldots,j_s}$ can have rank at most
419: $\nu-1$.  For example, if $\gamma \geq 4$, we can consider the
420: submatrix
421:       \begin{equation*}
422:         A^{1,2,5}_{1,3,4}=\begin{bmatrix} a_{0}&0&0\\ a_{1}&0&0\\
423:         a_{4}&a_{2}&a_{1}\end{bmatrix}.
424:       \end{equation*}
425:       This submatrix clearly has a zero determinant regardless of what
426:       $a_0,a_1,a_2$, and $a_4$ are.
427: \end{rem}
428: 
429: \section{A Construction of MDP Codes from Superregular Matrices}
430: In this section, we take a closer look at the transposes $(H^c_j)^T \in \F ^{(j+1)(n-k)\times (j+1)n}$ of the
431: matrices that were introduced in
432: (\ref{e-Gcj}):
433: $$
434:     (H^c_j)^T = \begin{bmatrix}
435:                  H_0^T& 0&\cdots&0      \\
436:                  H_1^T& H_0^T& \ddots & \vdots      \\
437:                  \vdots& \vdots & \ddots &0\\       
438:                  H_j^T&H_{j-1}^T& \cdots &H_0^T
439:               \end{bmatrix}.
440: $$
441: We then have the following definition:
442: \begin{defi}
443: $(H^c_j)^T$ is said to have the 
444: {\em maximum span property} if none of the first $n$ columns of $(H^c_j)^T$ is
445: contained in the span of any other $(j+1)(n-k)-1$ columns
446: of $(H^c_j)^T$. 
447: \end{defi}
448: We note here that, of course, each of the first $n$ columns of $(H^c_j)^T$ is in the span of some set of 
449: $(j+1)(n-k)$ columns of $(H^c_j)^T$.  We also make the observation that, if $\C$ is an $(n,k,\delta)$-code and
450: $(H^c_j)^T$ the transpose of its $j$th truncated sliding parity check matrix, then $d^c_j(\C)=(n-k)(j+1)+1$ if and only
451: if $(H^c_j)^T$ has the maximum span property.  In particular, $\C$ is an MDP code if and only if $(H^c_L)^T$ has
452: the maximum span property. 
453: 
454: Since $H(s)$ is basic, we may assume without loss of generality (for the purpose of what follows) that the
455: $(n-k)\times (n-k)$ submatrix of $H_0$ consisting of the first $n-k$ columns has full rank.  In this case, left
456: multiplication by an invertible matrix followed by a column permutation gives the matrix
457: $$
458:          \widehat {(H^c_j)^T}:=\left[\!\begin{array}{cccc|cccc}
459:                    1& 0&\cdots& 0 &\widehat H_0    & 0      & \cdots & 0 \\
460:                    0& 1&\ddots      & \vdots & \widehat H_1    &\widehat H_0   & \ddots & \vdots \\
461:                    \vdots &\ddots  &\ddots& 0 & \vdots & \ddots &\ddots & 0\\
462:                    0&\cdots  & 0     &1 &\widehat H_j& \widehat H_{j-1} &\cdots&\widehat H_0\end{array}\!\right] := \left[
463: 		   I_{(j+1)(n-k)} \, | \, \widehat H \right].
464: $$
465: Recalling Definition 3.1, we will also say that any matrix having the same form as $\widehat {(H^c_j)^T}$ has the maximum span property if none of the first $k$
466: columns of $\widehat H$ is contained in the span of any other $(n-k)(j+1)-1$ columns of $\widehat {(H^c_j)^T}$. 
467: Note that if $(H^c_j)^T$ has the maximum span property, then so does $\widehat {(H^c_j)^T}$.  The next theorem is a
468: simplified version of~\cite[Theorem 3.5]{gl03r}; it relates the superregularity property to the maximum span
469: property:
470: \begin{theo}\label{T-superreg}
471: Let $T$ be a  $(j+1)\times (j+1)$ lower triangular Toeplitz 
472: matrix:
473: \begin{equation}\label{e-HToeplitz}
474:     T:=
475:     \begin{bmatrix} t_0    & 0      & \cdots & 0 \\
476:                     t_1    & t_0    & \ddots & \vdots  \\
477:                     \vdots & \ddots &\ddots & 0\\
478:                     t_j    & t_{j-1} &   \cdots  & t_0\end{bmatrix}
479: \end{equation}  
480:  Then, $T$ is superregular if and only if the matrix $[I_{j+1} \, | \, T]$ has the maximum span property.
481: \end{theo}
482: Given $k$, $n$, and $j$, we will now see how to use an appropriately sized superregular matrix to construct a
483: matrix which has the form of $\widehat {(H^c_j)^T}$ and the maximum span property (for more details,
484: see~\cite[Appendix C]{gl03r}):
485: 
486: \begin{theo} \label{result}
487: Let $T$ be an $l\times l$ superregular matrix, where 
488: $l\geq (j+1)(n-1)$.  
489: Let $T'$ the submatrix obtained from $T$ by intersecting the rows indexed by $\{k,
490: k+1,\ldots, n-1, n-1+k, n-1+k+1, \ldots, 2(n-1),
491: \ldots ,j(n-1)+k,j(n-1)+k+1,\ldots,j(n-1)+n-1\}$ and columns indexed by $\{1,2,\ldots, k, n-1+1,n-1+2,\ldots, n-1+k,
492: \ldots,j(n-1)+1,j(n-1)+2,\ldots,j(n-1)+k\}$. Then, $T'\in \F ^{(j+1)(n-k)\times (j+1)k}$ is a lower
493: block triangular Toeplitz matrix such that $[I_{(j+1)(n-k)} \, | \, T']$ has the maximum span property. 
494: \end{theo}
495: If we let $j=L$, we may use the resulting matrix to construct a parity check matrix of an MDP convolutional code.
496: 
497: \begin{exa}
498: In this example, we consider the following matrix over $\F _{64}$:
499: $$
500:      \begin{bmatrix}1&0&0&0&0&0&0&0\\ \omega&1&0&0&0&0&0&0\\
501:                    \omega^9&\omega&1&0&0&0&0&0\\\omega^{33}&\omega^9&\omega&1&0&0&0&0\\
502:                    \omega^{33}&\omega^{33}&\omega^9&\omega&1&0&0&0\\
503:                    \omega^9&\omega^{33}&\omega^{33}&\omega^9&\omega&1&0&0\\
504:                    \omega&\omega^9&\omega^{33}&\omega^{33}&\omega^9&\omega&1&0\\
505:                    1&\omega&\omega^9&\omega^{33}&\omega^{33}&\omega^9&\omega&1
506:      \end{bmatrix}.
507: $$
508: Here, $\omega$ is a root of $x^6+x+1 \in \F _2[x]$ and is thus a primitive field element.  According to~\cite{gl03r},
509: this matrix is superregular.  We let $k=2$, $n=3$, and $\delta =2$.  Then, $L=3$, and, since $(L+1)(n-1)=8$, we
510: may use Theorem~\ref{result} to form the matrix
511: $$
512:   \left[
513:     \begin{array}{cccc|cccccccc}
514:     
515:       1 & 0 & 0 & 0 & \omega & 1 & 0 & 0 & 0 & 0 & 0 & 0\\
516:       0 & 1 & 0 & 0 & \omega ^{33} & \omega ^{9} & \omega & 1 & 0 & 0 & 0 & 0\\
517:       0 & 0 & 1 & 0 & \omega ^9 & \omega ^{33} & \omega ^{33} & \omega ^9 & \omega & 1 & 0 & 0\\
518:       0 & 0 & 0 & 1 & 1 & \omega & \omega ^9 & \omega ^{33} & \omega ^{33} & \omega ^9 & \omega & 1
519:     \end{array}
520:   \right],
521: $$
522: which has the maximum span property.  We may then use this matrix to construct a polynomial generator matrix for
523: an MDP $(3,2,3)$ code.  We could, for example, think of the $1\times 2$ matrices making up the first block column
524: of the right-hand side of this matrix as a finite sequence of Markov parameters and compute a minimal partial
525: realization of this sequence (see, for example,~\cite{an86}).  We would then compute the corresponding transfer
526: function.  After some small algebraic manipulations and fixing an ordering for the codeword components, we would
527: arrive at a polynomial generator matrix and its corresponding
528: convolutional code (see~\cite{ro99a} for more information about the relationship between the linear systems and
529: polynomial representations of convolutional codes).  In this example, one possibility for a resulting generator matrix is
530: $$
531:   \begin{bmatrix}
532:     s^2 + \omega ^{57}s + \omega ^{62} & 0\\
533:     0                                  & s^2 + \omega ^{57}s + \omega ^{62}\\
534:     \omega s^2+\omega ^{44} s+\omega ^{54}       & s^2 +\omega ^{17} s+\omega ^{21}
535:   \end{bmatrix}.
536: $$ 
537: \end{exa}
538: \section{Group Actions Preserving the Superregularity \\Property}
539: In this section, we consider group actions that allow one to create new
540: superregular matrices from a given one.  The main utility of these actions is
541: that they reduce the number of matrices one must check when performing a computer
542: search for matrices having the superregularity property.  
543: 
544: For a prime power $p^e$ and a nonnegative integer
545: $\gamma$, denote by $SR(p^e,\gamma )$ the set of superregular matrices of
546: dimension $\gamma +1$ over $\F _{p^e}$.  The first group action is a corollary of the following result,
547: which is proven in~\cite{gl03r}:
548: \begin{theo}\label{Z2}
549: Suppose that $A\in SR(p^e,\gamma )$.  Then, $A^{-1}\in SR(p^e,\gamma )$.
550: \end{theo}
551: \begin{cor}\label{Z2Cor}
552: There is an action $*_1$ of the additive group $(\Z ,+)$ on the set $SR(p^e,\gamma )$, given by
553: $$
554:   \begin{array}{cccc}\label{Action1}
555:     *_1: & \Z \times SR(p^e,\gamma ) & \longrightarrow & SR(p^e,\gamma )\\
556:          & \hspace{-0.7cm}(x,A)                         & \longmapsto     &
557:          A^{(-1)^{x}}.
558:   \end{array}
559: $$
560: \end{cor}
561: \begin{proof}
562: By Theorem~\ref{Z2}, $A^{-1}$ is superregular, and it is clear that $*_1$ is a
563: group action.
564: \end{proof}
565: We next have the following simple result:
566: \begin{lemma}\label{SRInverse}
567: If $A\in SR(p^e,\gamma )$ and $\gamma \geq 2$, then $A\neq A^{-1}$.
568: \end{lemma}
569: \begin{proof}
570:      Suppose
571:      $$
572:           A:=
573:           \begin{bmatrix}
574:           a_0    & 0      & \cdots  & 0        \\
575:           a_1    & a_0    & \ddots  & \vdots   \\
576:           \vdots & \ddots & \ddots  & 0        \\
577:           a_{\gamma}    & \cdots & a_1     & a_0
578:           \end{bmatrix}
579:      $$
580:      is a superregular matrix, where $\gamma \geq 2$.  Suppose in addition that $A=A^{-1}$.  Then, since
581:      $A^2=I_{\gamma +1}$, we must have that $a_0a_1+a_1a_0=2a_0a_1=0$.  By
582:      hypothesis, $a_0, a_1\not =0$.  Thus, the field must have
583:      characteristic 2.  We must also have that
584:      $a_2a_0+a_1^2+a_0a_2=2a_0a_2+a_1^2=0$.  Since the field has
585:      characteristic 2, this equation reduces to $a_1^2=0$.  This is a
586:      contradiction, as $a_1\not =0$.  Thus, $A\not =A^{-1}$.  
587: \end{proof}
588: \begin{cor}\label{SREven}
589: If $\gamma \geq 2$, then $|SR(p^e,\gamma )|$ is even. 
590: \end{cor}
591: 
592: We next describe an action of the multiplicative group $\F _{p^e}^*:=\F _{p^e} \backslash \{
593: 0 \}$ of nonzero field elements of $\F _{p^e}$ on $SR(p^e,\gamma )$:
594: \begin{theo}\label{Fp*}
595:      Let $\alpha\in\F _{p^e}^*$.  Define  
596:      $$
597:           \alpha \bullet A:=
598:           \begin{bmatrix}
599:                     a_0    & 0          & \cdots        & \cdots     &
600:           0      \\
601:           \alpha    a_1    & \ddots        & \ddots        &            &
602:           \vdots \\
603:           \alpha ^2 a_2    & \ddots & \ddots           & \ddots     &
604:           \vdots \\
605:                     \vdots & \ddots     & \ddots        & \ddots     &
606:           0      \\
607:           \alpha ^{\gamma} a_{\gamma}    & \cdots     & \alpha ^2 a_2 & \alpha a_1 &
608:           a_0
609:           \end{bmatrix}
610:      $$
611:      Then, the map
612:      $$
613:        \begin{array}{cccc}\label{Action2}
614:          *_2: & \F _{p^e}^* \times SR(p^e,\gamma ) & \longrightarrow & SR(p^e,\gamma )\\
615:               & \hspace{-0.8cm}(\alpha,A)                         & \longmapsto     &
616:               \alpha \bullet A
617:        \end{array}
618:      $$
619:      is an action of $\F _{p^e}^*$ on $SR(p^e,\gamma )$.
620: \end{theo}
621: \begin{proof}
622: It is readily verified that $*_2$ is a group action.  Let 
623: $$
624:      D:=
625:      \begin{bmatrix}
626:        1       &  0       &  \cdots  &  0          \\
627:        0       &  \alpha  &  \ddots  &  \vdots     \\
628:        \vdots  &  \ddots  &  \ddots  &  0          \\
629:        0       &  \cdots  &  0       &  \alpha ^{\gamma}
630:      \end{bmatrix}.
631: $$
632: One can then describe $*_2$ through conjugation by $D$:  $\alpha \bullet A = DAD^{-1}$.  This changes the determinants of
633: submatrices of $A$ only by factors of powers of $\alpha$, and it follows that
634: $\alpha \bullet A$ is also superregular.
635: \end{proof}
636: 
637: The final group action we consider makes use of the Galois group
638: $Aut_{\F_p}\F_{p^e}$.  We recall here the fact that $Aut_{\F_p}\F_{p^e} = \langle \sigma \rangle$, where $\sigma
639: \in Aut_{\F _p}\F _{p^e}$ is the $\F _p$-automorphism defined by $\sigma (x) = x^p\,\, \forall x\in \F _{p^e}$.  
640: Consequently, $Aut_{\F _p}\F _{p^e}\cong
641: \Z/e\Z$.
642: \begin{theo}\label{ZkZ}
643:      Let $A\in SR(p^e,\gamma )$. Let $i\in \Z /e\Z$.  Define
644:      $$
645:           i \circ A:=
646:           \begin{bmatrix}
647:                     a_0^{p^i}    & 0          & \cdots        & \cdots     &
648:           0      \\
649:               a_1^{p^i}    & \ddots        & \ddots        &            &
650:           \vdots \\
651:            a_2^{p^i}    & \ddots & \ddots           & \ddots     &
652:           \vdots \\
653:                     \vdots & \ddots     & \ddots        & \ddots     &
654:           0      \\
655:            a_{\gamma}^{p^i}    & \cdots     & 
656:           a_2^{p^i} &  a_1^{p^i} &
657:           a_0^{p^i}
658:           \end{bmatrix}.
659:      $$
660:      Then, the map
661:      $$
662:        \begin{array}{cccc}\label{Action3}
663:          *_3: & \Z /e\Z \times SR(p^e,\gamma ) & \longrightarrow & SR(p^e,\gamma )\\
664:               & \hspace{-0.8cm}(i,A)                         & \longmapsto     &
665:               i \circ A
666:        \end{array}
667:      $$
668:      is an action of the additive group $\Z /e\Z$ on $SR(p^e,\gamma )$.
669: \end{theo}
670: \begin{proof}
671:      It is readily verified that $*_3$ is a group action.  As the entries of $A$ belong to a field of characteristic $p$, the
672:      determinant of a submatrix of $i \circ A$ is the determinant of the
673:      corresponding submatrix of $A$ raised to the $p^i$th power.  Thus, $i \circ A$
674:      is also superregular.
675: \end{proof}
676: 
677: \section{An Upper Bound for the Required Field Size}
678: An important question to consider in trying to better understand superregular
679: matrices is that of how large a finite field must
680: be in order that a superregular matrix of a given size can exist over that
681: field.  For example, no $3\times 3$ superregular matrix exists over the field
682: $\F _2$:  all entries in the lower triangular part of a superregular matrix must be nonzero, which means that
683: in this case all such entries would have to be 1; clearly, this does not result in a
684: superregular matrix, since the lower left $2\times 2$ submatrix has a zero determinant.  In this section, we give an upper bound on the required
685: field size.  We begin with a technical lemma:
686: \begin{lemma}\label{Bijection}
687:      Let $i$ be a nonnegative integer and $\gamma$ a positive integer.  
688:      Let $S_{i+1}$ denote the set of integer sequences $\{s_l\}_{l=0}^{i+1}$,
689:      where $s_0=0$, $s_{i+1}=\gamma$, and $s_l<s_{l+1}\, \forall \, l\in
690:      \{0,1,\ldots ,i\}$.
691:      Let $S_{i,\gamma }:=\{\{s_l\}_{l=0}^{i+1}\in S_{i+1} \,\,  with \,\,   
692:      s_j+s_{i-j+1}\leq \gamma ,\, j=0,1,\ldots ,\lceil\frac{i}{2}\rceil\}$.  
693:      Let $T_{i+1}$ denote the set of integer sequences $\{t_l\}_{l=0}^{i+1}$,
694:      where $t_0=0$, $t_{i+1}=\gamma$, and $t_l<t_{l+1}\, \forall \, l\in
695:      \{0,1,\ldots ,i\}$.  Let
696:      $T_{i,\gamma }:=\{\{t_l\}_{l=0}^{i+1}\in T_{i+1} \,\, with \,\, \sum_{l=0}^{m}
697:      (-1)^l (t_{l+1}-t_l)\geq 0, \, m=0,1,\ldots ,i\}$.
698:      Then, 
699:      $S_{i,\gamma }$ and $T_{i,\gamma }$ are finite sets, and $|S_{i,\gamma
700:      }|$=$|T_{i,\gamma }|$.    
701: \end{lemma}
702: \begin{proof}
703:      It is clear that $S_{i,\gamma }$ and $T_{i,\gamma }$ are finite
704:      sets.  We now proceed to prove the second part of the claim.  We will do this by first constructing an
705:      injective map from $S_{i,\gamma }$ to $T_{i,\gamma }$ and then an injective map from $T_{i,\gamma }$ to
706:      $S_{i,\gamma }$.  
707:      Throughout the proof, $t_{-1}$ and $s_{-1}$ are defined to be 0.
708:      Suppose first that $i$ is even.  Let $\{s_l\}_{l=0}^{i+1} \in
709:      S_{i,\gamma }$.  For $j\in \{0,1,\ldots ,\frac{i}{2} \}$, we make
710:      the recursive definitions
711:      \begin{align*}
712:      t_{2j+1}&:=t_{2j}+s_{i-j+1}-s_{i-j},\\ 
713:      t_{2j}&:=t_{2j-1}+s_{j}-s_{j-1}.  
714:      \end{align*}
715:      Note that $t_0=s_0=0$.  
716:      Then, for $m\in \{0,1,\ldots
717:      ,i\}$, we have    
718:      \begin{align*}
719:      \sum_{l=0}^{m}
720:           (-1)^l (t_{l+1}-t_l) &=t_1-(t_2-t_1)+(t_3-t_2)-\cdots
721:           +(-1)^{m}(t_{m+1}-t_m)\\
722:           &=2t_1-2t_2+2t_3-\cdots +(-1)^{m}t_{m+1}\\
723:           &=2t_1-2t_2+2t_3-\cdots
724:           -t_{2j}\\
725:           &\hspace{3cm} \mbox{or}\\
726:           &=2t_1-2t_2+2t_3-\cdots +t_{2j+1}, 
727:      \end{align*}
728:      where $j\in \{0,1,\ldots
729:      ,\frac{i}{2} \}$ is $(m-1)/2$ or $m/2$ as $m$ is odd or even, respectively.  Suppose
730:      first that $m$ is odd.  The sum then becomes
731:      \begin{align*}
732:           2t_1-2t_2+2t_3-\cdots
733:           -t_{2j}&=2t_1-2(t_1+s_1)+2t_3-\cdots
734:           -(t_{2j-1}+s_j-s_{j-1})\\
735:           &=t_{2j-1}-(s_j-s_{j-1})-2s_{j-1}\\
736:           &=t_{2j-1}-s_j-s_{j-1}.
737:      \end{align*}
738:      Using the definition of $t_{2j+1}$ with $j$ replaced by $j-1$, we may write
739:      $$
740:           t_{2j-1}-s_j-s_{j-1}=t_{2j-2}+s_{i-j+2}-s_{i-j+1}
741:           -s_j-s_{j-1}.
742:      $$    
743:      Using again the definitions above, we see that, for
744:      any integer $h\in \{0,\ldots ,j-1 \}$, 
745:      $$
746:           t_{2j-2}+s_{i-j+2}-s_{i-j+1}-s_j-s_{j-1}=
747:           t_{2(j-h)-2}+s_{i-(j-(h+1))+1}-s_{i-j+1}-s_j-s_{j-(h+1)}:
748:      $$
749:      this equation holds trivially if $h=0$, and, if $h\in \{ 0,\ldots ,j-2 \}$, we have
750:      \begin{align*}
751:           & t_{2(j-h)-2}+s_{i-(j-(h+1))+1}-s_{i-j+1}-s_j-s_{j-(h+1)}=\\
752:           &
753: t_{2(j-h-1)-1}+s_{j-h-1}+s_{(j-h-1)-1}+s_{i-(j-(h+1))+1}-s_{i-j+1}-s_j-s_{j-(h+1)}
754:           =\\&
755: t_{2(j-h-2)}+s_{i-(j-h-2)+1}-s_{i-(j-h-2)}+s_{j-h-1}-s_{(j-h-1)-1}+s_{i-(j-(h+1))+1}
756:           \\& -s_{i-j+1}-s_j-s_{j-(h+1)}=\\& t_{2(j-(h+1))-2}+s_{i-(j-((h+1)+1))+1}
757:           -s_{i-j+1}-s_j-s_{j-((h+1)+1)}.
758:      \end{align*}
759:      In particular, letting $h=j-2$ and recalling that $s_{i+1}=\gamma$ and $s_0=0$,
760:      we see that
761:      $$
762:           t_{2j-2}+s_{i-j+2}-s_{i-j+1}-s_j-s_{j-1}=\gamma -s_{i-j+1}-s_j.
763:      $$
764:      Because $\{s_l\}_{l=0}^{i+1} \in S_{i,\gamma }$, $\gamma -s_{i-j+1}-s_j\geq 0$.  
765:      
766:      Suppose now that $m$ is even.  We are then interested in the sum
767:      $$
768:           2t_1-2t_2+2t_3-\cdots +t_{2j+1}.
769:      $$
770:      From the analysis in the case of $m$ odd, we see that we may write
771:      \begin{align*}
772:           2t_1-2t_2+2t_3-\cdots
773:           +t_{2j+1}&=t_{2j-1}-s_j-s_{j-1}-t_{2j}+t_{2j+1}\\
774:           &=t_{2j-1}-s_j-s_{j-1}-t_{2j}+t_{2j}+s_{i-j+1}-s_{i-j}\\
775:           &=t_{2j-1}-s_j-s_{j-1}+s_{i-j+1}-s_{i-j}.
776:      \end{align*}
777:      Using the same method as before, we see that, for any integer
778:      $h\in \{0,\ldots ,j-1 \}$,
779:      $$
780:           t_{2j-1}-s_j-s_{j-1}+s_{i-j+1}-s_{i-j}=t_{2(j-h)-1}
781:           +s_{i-(j-h)+1}-s_{i-j}-s_j-s_{j-(h+1)}.
782:      $$
783:      In particular, letting $h=j-1$, the right side of this equality reduces to
784:      $$
785:           t_1+s_i-s_{i-j}-s_j.
786:      $$
787:      Using the definition of $t_1$, this becomes
788:      \begin{align*}
789:           s_{i+1}-s_i+s_i-s_{i-j}-s_j&=s_{i+1}-s_{i-j}-s_j\\
790:           &=\gamma -s_{i-j}-s_j.
791:      \end{align*}
792:      Since $s_{i-j}<s_{i-j+1}$ and $\gamma -s_{i-j+1}-s_j\geq 0$, it follows that $\gamma -s_{i-j}-s_j\geq 0$.
793:      
794:      It
795:      remains to be shown that $t_{i+1}=\gamma$.  We have
796:      \begin{align*}
797:           t_{2j+1}&=t_{2j}+s_{i-j+1}-s_{i-j}\\
798:           &=t_{2j-1}+s_j-s_{j-1}+s_{i-j+1}-s_{i-j}\\
799:           &=t_{2j-2}+s_{i-j+2}+s_j-s_{j-1}-s_{i-j}.
800:      \end{align*}
801:      In the same way as above, one sees that, for any integer $h\in
802:      \{1,\ldots ,j \}$, 
803:      $$
804:           t_{2j-2}+s_{i-j+2}+s_j-s_{j-1}-s_{i-j}=t_{2(j-h)}+s_{i-(j-h)+1}
805:           -s_{j-h}+s_j-s_{i-j}.
806:      $$
807:      In particular, letting $j=\frac{i}{2}$ and $h=j$, the right side of the
808:      preceding equality simplifies to $s_{i+1}$.  By hypothesis, $s_{i+1}=\gamma$. 
809:      Thus, $t_{i+1}=\gamma$.
810:      This completes the case of $i$ even.  
811:      
812:      Suppose now that $i$ is odd.  For $j\in \{0,1,\ldots
813:      ,\frac{i-1}{2} \}$, define $t_{2j+1}$ as above.  For $j\in
814:      \{0,1,\ldots ,\frac{i+1}{2} \}$, define $t_{2j}$ as above.  The
815:      proof then proceeds as in the case of $i$ even.
816:      
817:      We thus obtain a well-defined map $f$ from $S_{i,\gamma }$ to
818:      $T_{i,\gamma }$: if $\{s_l\}_{l=0}^{i+1} \in S_{i,\gamma }$, let
819:      $f(\{s_l\}_{l=0}^{i+1})=\{t_l\}_{l=0}^{i+1}$, where each $t_l$ is
820:      defined via the equations at the beginning of the proof.  It
821:      follows immediately from the definition of the $t_l$ that $f$ is
822:      injective.  Thus, to show $|S_{i,\gamma }|$=$|T_{i,\gamma }|$, it
823:      will suffice to construct a similar injective map from
824:      $T_{i,\gamma }$ to $S_{i,\gamma }$.
825:      
826:      Suppose first that
827:      $i$ is even.  Let $\{t_l\}_{l=0}^{i+1} \in T_{i,\gamma }$.  For $j\in \{0,1,\ldots
828:      ,\frac{i}{2} \}$, we make the recursive definitions
829:      \begin{align*}
830:      s_j&:=s_{j-1}
831:      +t_{2j}-t_{2j-1}\\
832:      s_{i-j}&:=s_{i-j+1}-t_{2j+1}+t_{2j}.
833:      \end{align*}
834:      Note that $s_0=t_0=0$.  
835:      Then, using these definitions, we have that
836:      \begin{align*}
837:           s_j+s_{i-j+1}&=s_{j-1}+t_{2j}-t_{2j-1}+s_{i-j+2}+t_{2j-2}-t_{2j-1}
838:           \\
839:           &= s_{j-2}+t_{2j-2}-t_{2j-3}+t_{2j}-t_{2j-1}+s_{i-j+3}+t_{2j-4}
840:           -t_{2j-3}+t_{2j-2}-t_{2j-1}\\
841:           &=\\
842:           &\hspace{0.25cm}\vdots\\
843:           &=\gamma -2t_1+2t_2-\cdots +t_{2j}.
844:      \end{align*}
845:      Because $\{t_l\}_{l=0}^{i+1} \in T_{i,\gamma }$,
846:      $2t_1-2t_2+\ldots -t_{2j}\geq 0$.  Thus, $s_j+s_{i-j+1}\leq
847:      \gamma$.  It remains to be shown that $s_{i+1}=\gamma$.  From the
848:      definition of $s_0$, it is clear that $s_0=0$.  We have that
849:      \begin{align*}
850:            s_{i+1}  & =        s_i+t_1\\
851:                     & =        s_{i-1}+t_3-t_2+t_1\\
852:                     & =        s_{i-2}+t_5-t_4+t_3-t_2+t_1\\
853:                     & \,\,\,\, \vdots   \\
854:                     & =        s_{\frac{i}{2}}+t_{i+1}-t_i+t_{i-1}-t_{i-2}+
855:                                \cdots +t_3-t_2+t_1.  
856:      \end{align*}
857:      We also have that
858:      \begin{align*}
859:            s_{\frac{i}{2}} & = s_{\frac{i-2}{2}}+t_i-t_{i-1}\\
860:                            & = s_{\frac{i-4}{2}}+t_{i-2}-t_{i-3}+t_i-t_{i-1}\\
861:                            & \,\,\,\, \vdots   \\
862:                            & = -t_1+t_2-t_3+\cdots +t_{i-2}-t_{i-3}+t_i-t_{i-1}.  
863:      \end{align*}
864:      Thus, $s_{i+1}=t_{i+1}$.  Since $t_{i+1}=\gamma$ by hypothesis, it follows that
865:      $s_{i+1}=\gamma$.
866:       
867:      Suppose now that $i$ is odd.  For $j\in \{0,1,\ldots ,\frac{i+1}{2} \}$, 
868:      define $s_j$ as above.  For $j\in \{0,1,\ldots ,\frac{i-1}{2}
869:      \}$, define $s_{i-j+1}$ as above.  The proof then proceeds as
870:      in the case of $i$ even.
871:      
872:      We thus obtain a well-defined map $g$ from $T_{i,\gamma }$ to $S_{i,\gamma }$:  if $\{t_l\}
873:      _{l=0}^{i+1} \in T_{i,\gamma }$, let $g(\{t_l\}
874:      _{l=0}^{i+1})=\{s_l\}_{l=0}^{i+1}$, where each $s_l$ is defined
875:      via the equations above.  It follows immediately from the definition of the $s_l$ that $g$ is injective.  Thus,
876:      $|S_{i,\gamma }|=|T_{i,\gamma }|$.  
877: \end{proof}
878: 
879: We need one more technical lemma before computing the upper bound:
880: \begin{lemma}\label{Count}
881:      Let $\gamma$ be a positive integer.  Then, $\prod_{1\leq i\leq j\leq \gamma
882:      }
883:      \frac{2+i+j}{i+j}=\frac{1}{\gamma +2} {2(\gamma +1)\choose \gamma +1}$.
884: \end{lemma}
885: \begin{proof}
886:      The proof is by induction.  The claim is clearly true if $\gamma =1$.
887:      Suppose that $\prod_{1\leq i\leq j\leq \gamma }
888:      \frac{2+i+j}{i+j}=\frac{1}{\gamma +2} {2(\gamma +1)\choose \gamma +1}$ for some
889:      $\gamma\geq 1$.  Then,
890:      \begin{align*}
891:           &\prod_{1\leq i\leq j\leq \gamma +1}
892:           \frac{2+i+j}{i+j}= \prod_{1\leq i\leq j\leq \gamma }
893:           \frac{2+i+j}{i+j} \prod_{1\leq i\leq \gamma +1}
894:           \frac{i+\gamma +3}{i+\gamma +1}=\\ &\frac{1}{\gamma +2} \frac{(2\gamma
895:           +2)(2\gamma +1)\cdots
896:           1}{(\gamma +1)(\gamma )\cdots 1(\gamma +1)(\gamma )\cdots 1}
897:           \frac{(\gamma +4)(\gamma +5)\cdots
898:           (2\gamma +4)}{(\gamma +2)(\gamma +3)\cdots (2\gamma +2)}=\\
899:           &\frac{(2\gamma +2)(2\gamma +1)\cdots
900:           (\gamma +3)}{(\gamma +1)(\gamma )\cdots 1} \frac{(\gamma +4)(\gamma +5)\cdots
901:           (2\gamma +4)}{(\gamma +2)(\gamma +3)\cdots (2\gamma +2)}=\\
902:           &\frac{1}{\gamma +3}
903:           \frac{(2\gamma +4)(2\gamma +3)\cdots (\gamma +3)}{(\gamma +2)(\gamma +1)\cdots 1}=
904:           \frac{1}{\gamma +3} \frac{(2\gamma +4)(2\gamma +3)\cdots 1}{(\gamma
905:           +2)(\gamma +1)\cdots
906:           1(\gamma +2)(\gamma +1)\cdots 1}=\\ &\frac{1}{\gamma +3} {2\gamma
907:           +4\choose \gamma +2}=
908:           \frac{1}{(\gamma +1)+2} {2((\gamma +1)+1)\choose (\gamma +1)+1}.
909:      \end{align*} 
910:      This proves the claim.
911: \end{proof}
912: 
913: We are now ready to prove the main result of this section:
914: \begin{theo}\label{Bound}
915:      Let $\gamma\in \N$.  Let $C_{\gamma}$ denote the $\gamma$th Catalan number: 
916:      $C_{\gamma}:=\frac{1}{\gamma +1}
917:      {2\gamma\choose \gamma }$.  Let $\F$ be a finite field such
918:      that $|\F|>\frac{1}{2}\big (C_{\gamma -1} +{\gamma -1\choose \lfloor
919:      \frac{\gamma -1}{2} \rfloor} \big )$.  Then, there exists a $\gamma\times
920:      \gamma$
921:      superregular matrix over $\F$.
922: \end{theo}
923: \begin{proof}
924:      We first upper bound the number of submatrices of a
925:      $\gamma\times \gamma$ lower triangular Toeplitz matrix that possess a certain property and show that this 
926:      upper bound is given by the expression in the
927:      statement of the theorem.  We then observe that the number thus
928:      obtained is actually an upper bound on the minimal size a finite field $\F$ must
929:      have in order that a $\gamma\times \gamma$ superregular matrix over $\F$ can exist.
930: 
931:      For convenience, we drop the matrix entries with index 0.  Thus, a
932:      $\gamma\times \gamma$ lower triangular Toeplitz matrix $X$ with indeterminate
933:      entries is now defined by a first column of the form 
934:      $$
935:           [x_1\,\,\,\,\,\, x_2\,\,\,\,\,\, \cdots \,\,\,\,\,\, x_{\gamma}]^T.
936:      $$
937:      The determinants
938:      of the proper square submatrices of such a matrix are given by nonzero polynomials in
939:      these indeterminates.  Notice that $x_{\gamma}$ can appear to at most the first
940:      power
941:      in any of these polynomials; in other words, each of these polynomials has
942:      the property that either it is linear in $x_{\gamma}$, or $x_{\gamma}$ does not appear in
943:      any of its terms.  We are interested in those proper
944:      square submatrices whose determinants are linear in $x_{\gamma}$, and we denote the
945:      set of such submatrices by $L_{\gamma}$.  
946: 
947:      First, we observe that $L_{\gamma}$ consists of the single entry
948:      $x_{\gamma}=X_{1}^{\gamma}$ and 
949:      the submatrices $X_{1,j_1,\ldots
950:      ,j_{s-1}}^{i_1,i_2,\ldots ,i_{s-1},\gamma }$, where $s\in \{ 2,\ldots ,\gamma -1
951:      \}$, where
952:      $j_{\nu}\leq
953:      i_{\nu}$ for all $\nu\in \{1,2,\ldots ,s-1\}$.  That $X_{1}^{\gamma}$ is the only $1\times
954:      1$ submatrix in $L_{\gamma}$ is obvious; to see the second part, evaluate
955:      det$X_{1,j_1,\ldots ,j_{s-1}}^{i_1,i_2,\ldots ,i_{s-1},\gamma }$ by doing a
956:      cofactor expansion along the first column of $X_{1,j_1,\ldots
957:      ,j_{s-1}}^{i_1,i_2,\ldots ,i_{s-1},\gamma }$.  The indeterminate
958:      $x_{\gamma}$ appears if and only if det$X_{j_1,j_2,\ldots
959:      ,j_{s-1}}^{i_1,i_2,\ldots ,i_{s-1}}\not =0$.  This is the
960:      case if and only if $X_{j_1,j_2,\ldots
961:      ,j_{s-1}}^{i_1,i_2,\ldots ,i_{s-1}}$ is a proper submatrix, and
962:      $X_{j_1,j_2,\ldots
963:      ,j_{s-1}}^{i_1,i_2,\ldots ,i_{s-1}}$ is a proper submatrix if and
964:      only if $j_{\nu}\leq
965:      i_{\nu}$ for all $\nu\in \{1,2,\ldots ,s-1\}$.
966: 
967:      Next, we observe that $L_{\gamma }$ is closed with respect to
968:      transpose about the antidiagonal of $X$.  In order to see this,
969:      note that the transpose $\bar X_{1,j_1,\ldots
970:      ,j_{s-1}}^{i_1,i_2,\ldots ,i_{s-1},\gamma }$ of any submatrix
971:      $X_{1,j_1,\ldots ,j_{s-1}}^{i_1,i_2,\ldots ,i_{s-1},\gamma }$
972:      about the antidiagonal is given by $\bar X_{1,j_1,\ldots
973:      ,j_{s-1}}^{i_1,i_2,\ldots ,i_{s-1},\gamma }=X_{1,\gamma
974:      -i_{s-1}+1,\ldots ,\gamma - i_1+1}^{\gamma -j_{s-1}+1,\gamma
975:      -j_{s-2}+1,\ldots ,\gamma }$.  Clearly, $\gamma -i_{\nu}+1\leq
976:      \gamma -j_{\nu}+1 \iff j_{\nu}\leq i_{\nu}$.  Of course,
977:      $X_{1,j_1,\ldots ,j_{s-1}}^{i_1,i_2,\ldots ,i_{s-1},\gamma }\in
978:      L_{\gamma}$ is symmetric with respect to the antidiagonal of $X$
979:      if and only if $X_{1,j_1,\ldots ,j_{s-1}}^{i_1,i_2,\ldots
980:      ,i_{s-1},\gamma }=\bar X_{1,j_1,\ldots ,j_{s-1}}^{i_1,i_2\ldots
981:      ,i_{s-1},\gamma }$.  Let $L'_{\gamma}\subset L_{\gamma}$ denote
982:      those elements of $L_{\gamma}$ that are symmetric with respect to
983:      the antidiagonal.
984: 
985:      Finally, because $X$ is symmetric with respect to the antidiagonal, taking 
986:      the transpose $\bar X_{1,j_1,\ldots
987:      ,j_{s-1}}^{i_1,i_2,\ldots ,i_{s-1},\gamma }$ of a square submatrix $X_{1,j_1,\ldots
988:      ,j_{s+1}}^{i_1,i_2,\ldots ,i_{s-1},\gamma }$ about the antidiagonal of $X$ amounts to
989:      taking the transpose of $X_{1,j_1,\ldots
990:      ,j_{s-1}}^{i_1,i_2\ldots ,i_{s-1},\gamma }$ about its own antidiagonal.  Since the
991:      determinant of a matrix is the same as the determinant of its transpose about
992:      its antidiagonal, we have that
993:      det$X_{1,j_1,\ldots
994:      ,j_{s-1}}^{i_1,i_2,\ldots ,i_{s-1},\gamma }$= det$\bar X_{1,j_1,\ldots
995:      ,j_{s-1}}^{i_1,i_2,\ldots ,i_{s-1},\gamma }$.  
996: 
997:      By definition, the determinants of the elements of $L_{\gamma}$ are polynomials 
998:      linear in $x_{\gamma}$.  We wish to determine how large a finite field must
999:      be in order to guarantee that nonzero field
1000:      elements may be substituted for $x_1,x_2,\ldots ,x_{\gamma}$ in such a
1001:      way that none of these determinants is zero.  We do this by
1002:      computing an upper bound $N_{\gamma}$ on the number of distinct polynomials giving the
1003:      determinants of the elements of $L_{\gamma}$.  As long as the order of
1004:      the field is bigger than $N_{\gamma}$, it is clearly possible to make all
1005:      of these determinants nonzero.  From the above
1006:      observations, we see that, once we have computed $|L_{\gamma}|$ and
1007:      $|L'_{\gamma}|$, we
1008:      may take as an upper bound 
1009:      $N_{\gamma}:=\frac{1}{2}(|L_{\gamma}|+|L'_{\gamma}|)$.
1010: 
1011:      As seen above, the $s\times s$ submatrices in $L_{\gamma}$, $s\in
1012:      \{ 2,\ldots ,\gamma -1 \}$, are described by sets
1013:      $\{i_1,i_2,\ldots ,i_{s-1},j_1,j_2,\ldots ,j_{s-1}\}$ of indices,
1014:      where $j_{\nu}\leq i_{\nu}$ for all $\nu\in \{1,2,\ldots ,s-1\}$,
1015:      $1<i_1<\ldots <i_{s-1}<\gamma$, and $1<j_1<\ldots
1016:      <j_{s-1}<\gamma$ (and the $1\times 1$ submatrix $x_{\gamma}$ is
1017:      associated with an empty set of indices).  Each nonempty index
1018:      set defines a generalized Young tableau with columns having
1019:      height 2 and integer entries in $\{ 2,\ldots ,\gamma -1 \}$;
1020:      conversely the entries of such a tableau constitute a set of
1021:      indices corresponding to a submatrix in $L_{\gamma}$.  The empty
1022:      set of indices corresponds to a tableau with all columns having
1023:      height 0.  In~\cite{CatVien}, it is shown that the number of such
1024:      tableaux is given by $\prod_{1\leq i\leq j\leq \gamma -2}
1025:      \frac{2+i+j}{i+j}$.  By Lemma~\ref{Count}, this product is
1026:      $C_{\gamma -1}$.  Consequently, $|L_{\gamma}|=C_{\gamma -1}$.
1027: 
1028:      We compute $|L'_{\gamma}|$ as follows.  If the submatrix
1029:      $X_{1,j_1,\ldots ,j_{\gamma}}^{i_1,i_2\ldots ,i_{s-1},\gamma }\in
1030:      L'_{\gamma}$, where $s\in \{ 2,\ldots ,\gamma -1 \}$, then, since
1031:      $X_{1,j_1,\ldots ,j_{s-1}}^{i_1,i_2,\ldots ,i_{s-1},\gamma }=\bar
1032:      X_{1,j_1,\ldots ,j_{s-1}}^{i_1,i_2\ldots ,i_{s-1},\gamma }$, it
1033:      must be that $\gamma -i_{\nu}=j_{s-\nu }-1$ for all $\nu\in
1034:      \{1,2,\ldots ,s-1\}$.  Thus, this submatrix is completely
1035:      determined by the $s-1$ integers $w_1=j_1-1, w_2=j_2-1,\ldots
1036:      ,w_{s-1}=j_{s-1}-1$.  Since $X_{1,j_1,\ldots
1037:      ,j_{s-1}}^{i_1,i_2,\ldots ,i_{s-1},\gamma }\in L_{\gamma}$, we
1038:      have $j_l=w_l+1\leq \gamma -w_{s-l}=i_l$, $l\in \{ 1,\ldots
1039:      ,\lceil\frac{s-1}{2}\rceil \}$.  These inequalities can be
1040:      rewritten as $w_l+w_{s-1-l}\leq \gamma -1$, $l\in \{ 1,\ldots
1041:      ,\lceil\frac{s-1}{2}\rceil \}$.  In other words, $\{ 0,w_1,\ldots
1042:      ,w_{s-1},\gamma -1 \}$ is a sequence that belongs to
1043:      $S_{s-1,\gamma -1}$.  Thus, to each $s\times s$ submatrix in
1044:      $L'_{\gamma}$, where $s\in \{ 2,\ldots ,\gamma -1 \}$, we have
1045:      associated a unique sequence in $S_{s-1,\gamma -1}$.  Similarly,
1046:      to each sequence in $S_{s-1,\gamma -1}$, we can associate a
1047:      unique $s\times s$ submatrix in $L'_{\gamma}$.  If $s=1$, there
1048:      is only the submatrix $X_{\gamma}^1$ to consider, and we have
1049:      already associated this submatrix with the empty sequence, which
1050:      in turn corresponds with the sequence $\{ 0,\gamma -1 \}$, the
1051:      sole member of $S_{0,\gamma -1}$.  Thus,
1052:      $|L'_{\gamma}|=\sum_{y=0}^{\gamma -2} |S_{y,\gamma -1}|$.  From
1053:      Lemma~\ref{Bijection}, we know that $\sum_{y=0}^{\gamma -2}
1054:      |S_{y,\gamma -1}|=\sum_{y=0}^{\gamma -2} |T_{y,\gamma -1}|$.  It
1055:      is sufficient, then, to compute $\sum_{y=0}^{\gamma -2}
1056:      |T_{y,\gamma -1}|$.
1057:      
1058:      Suppose $\{t_l\}_{l=0}^{s} \in T_{s-1,\gamma -1}$.  To this 
1059:      sequence, we can associate a nonnegative
1060:      planar walk of length $\gamma -1$ with $s+1$ vertices.  The walk begins at the
1061:      origin, and steps are
1062:      given by the
1063:      vectors $(1,1)$ and $(1,-1)$.  The vertices are the origin, the
1064:      endpoint of the walk, and the points where the
1065:      direction of the walk changes from up to down or from down to up.  We make
1066:      the association in the following way:  let $t_i$ be the $x$-coordinate
1067:      of the $i$th vertex.  It is clear that this walk is nonnegative, as the
1068:      condition defining membership in $T_{s-1,\gamma -1}$ guarantees nonnegativity of
1069:      the associated walk.  Conversely, the $x$-coordinates of
1070:      the $s+1$ vertices in a nonnegative planar walk of length $\gamma -1$ form a
1071:      sequence in $T_{s-1,\gamma -1}$.  Therefore, this association
1072:      gives a bijective correspondence between sequences in
1073:      $\cup_{y=0}^{\gamma -2} T_{y,\gamma -1}$ and nonnegative planar walks of
1074:      length $\gamma -1$.  It is a fact (see, for example,~\cite{Feller})
1075:      that the number of nonnegative planar walks of length $\gamma -1$ is
1076:      given by ${\gamma -1\choose \lfloor \frac{\gamma -1}{2} \rfloor }$.  This
1077:      means that $\sum_{y=0}^{\gamma -2}
1078:      |T_{y,\gamma -1}|={\gamma -1\choose \lfloor \frac{\gamma -1}{2} \rfloor }$.
1079:      Consequently, $|L'_{\gamma}|={\gamma -1\choose \lfloor \frac{\gamma -1}{2} \rfloor
1080:      }$.
1081: 
1082:      It remains to show that, in fact, a field of order bigger than 
1083:      $N_{\gamma}$ elements is large enough so that
1084:      a superregular $\gamma\times \gamma$ matrix can exist over it.  This
1085:      may 
1086:      be easily
1087:      seen in the following way:  the determinant of each submatrix in
1088:      $S_l$, $l\in \{ 1,\ldots ,\gamma \}$, is linear in $x_l$.  $N_l$ increases with
1089:      $l$.  Thus, if we work over a field of
1090:      order bigger than $N_{\gamma}$ elements, it is
1091:      possible to successively substitute nonzero field elements for 
1092:      $x_1,x_2,\ldots
1093:      ,x_{\gamma}$ in such a way that the 
1094:      determinant of each
1095:      submatrix in $S_l$, $l\in \{ 1,\ldots ,\gamma \}$, is nonzero.  This completes
1096:      the proof.  
1097: \end{proof}
1098: 
1099: Using computer searches, we were able to determine the exact minimum
1100: field size for small values of $\gamma$.  These may be seen in Table
1101: 1.  We observe that the upper bound $N_{\gamma}+1$ appears to grow
1102: much more quickly than necessary:
1103: \begin{table}[h]
1104:   \begin{center}
1105:       \begin{tabular}{|l|c|r|}
1106:         \hline
1107:         $\gamma$ & Minimum Field Size Required & Upper Bound ($N_{\gamma}$ +1) \\
1108:         \hline
1109:         3 & 3 & 3 \\
1110:         4 & 5 & 5 \\
1111:         5 & 7 & 11 \\
1112:         6 & 11 & 27 \\
1113:         7 & 17 & 77 \\
1114:         8 & 31 & 233 \\
1115:         9 & 59 & 751 \\
1116:         10 & $\leq 127$ & 2495 \\
1117:         \hline
1118:      \end{tabular}
1119:      \caption{Comparison of Actual Required Field Size and $N_{\gamma}+1$}
1120:   \end{center}
1121: \end{table}
1122: 
1123: \noindent It is still an open problem as to how this bound may be refined. 
1124: Computer searches lead us to make the following conjecture; if true, it would offer a
1125: significant improvement to the bound given here:
1126: \begin{conjecture}\label{ConjBound}
1127: For $\gamma \geq 5$, there exists a $\gamma\times \gamma$ superregular matrix over the field
1128: $\F_{2^{\gamma-2}}$.
1129: \end{conjecture}
1130: 
1131: \section{For Future Research:  Finding a Construction}
1132: At this point, little is understood about how to construct
1133: superregular matrices.  The problem of finding constructions appears
1134: to be a very hard one.  One must find a way of guaranteeing that all
1135: proper submatrices with any number of zeroes above the diagonal have a nonzero determinant and do so with additional
1136: constraints coming from the Toeplitz structure.  In~\cite{ToepArray}, a method is given for
1137: constructing, for any prime number $p$, a triangular Toeplitz array of
1138: dimension $p$ having the property that all full square submatrices
1139: have a nonzero determinant.  Unfortunately, there is no way of
1140: extending this construction to the much more general situation we
1141: consider here.
1142: 
1143: In~\cite{gl03r}, the following result is proven:
1144: \begin{theo}\label{TotPos}      
1145: Let $X$ be the $\gamma\times \gamma$ matrix given by
1146:       \begin{equation*}
1147:         X:=\left[\begin{array}{cccccc}
1148:         1&0&\cdots&\cdots&\cdots&0\\
1149:         1&1&\ddots&&&\vdots\\
1150:         0&1&1&\ddots&&\vdots\\
1151:         \vdots&\ddots&\ddots&\ddots &\ddots&\vdots\\
1152:         \vdots&&\ddots&1&1&0\\
1153:         0&\cdots&\cdots&0&1&1\\
1154:         \end{array}\right].
1155:       \end{equation*}
1156:             Then, 
1157:       \begin{equation*}
1158:         X^{\gamma -1}=\left[\begin{array}{cccccc}
1159:         1&0&\cdots&\cdots&\cdots&0\\
1160:         {\gamma -1\choose 1} &1&\ddots&&&\vdots\\
1161:         {\gamma -1\choose  2} &{\gamma -1\choose 1}&1&\ddots&&\vdots\\
1162:         \vdots&\vdots&\ddots&\ddots &\ddots&\vdots\\
1163:         {\gamma -1\choose \gamma -2}&\gamma -1\choose {\gamma -3}&\cdots&{\gamma
1164:         -1\choose 1}&1&0\\ 
1165:         1&\gamma -1\choose {\gamma -2}&\cdots&\cdots&{\gamma -1\choose 1}&1
1166:         \end{array}\right],
1167:       \end{equation*}
1168:       where ${\gamma\choose \omega }$=$\frac{\gamma !}{\omega !(\gamma -\omega )!}$,
1169:       is totally positive over the real numbers.  Thus, for a sufficiently
1170:       large prime number $p$, taking the entries of this matrix modulo $p$ gives
1171:       a superregular matrix.
1172: \end{theo}
1173: This result gives a construction insofar as one knows that, modulo a large enough
1174: prime number, the matrix $X^{\gamma -1}$ is superregular.  It is not clear,
1175: however, how one may give a good bound as to how large $p$ must be for a given
1176: $\gamma$.
1177:  
1178: \noindent
1179:     
1180: 
1181: \bibliographystyle{Bib}
1182: \bibliography{SR2}
1183: 
1184: \end{document}
1185: