math0211001/perm.tex
1: \documentclass[11pt]{article}
2: 
3: \usepackage{amsmath}
4: \usepackage{amssymb}
5: \usepackage{amsthm}
6: \setlength{\textwidth}{6in} \setlength{\textheight}{8in}
7: \setlength{\evensidemargin}{0in} \setlength{\oddsidemargin}{0in}
8: \setlength{\footskip}{.5in} \setlength{\topmargin}{0in}
9: \setlength{\parskip}{8pt} \baselineskip 24pt
10: 
11: \usepackage[dvips]{graphicx}
12: \usepackage{color}
13: \usepackage{epsfig}
14: \usepackage[mathscr]{eucal}
15: \newtheorem{theorem}{Theorem}
16: \renewcommand{\thetheorem}{\arabic{section}.\arabic{theorem}}
17: \newtheorem{prop}[theorem]{Proposition}
18: \newtheorem{lemma}[theorem]{Lemma}
19: \newtheorem{cor}[theorem]{Corollary}
20: \newtheorem*{claim}{Claim}
21: \newtheorem{definition}[theorem]{Definition}
22: \newenvironment{poc}{\noindent \emph{Proof of Claim.}}{\vspace*{0.5\baselineskip}}
23: \renewcommand{\qedsymbol}{\vrule height12pt width7pt
24: depth0pt}
25: 
26: \allowdisplaybreaks
27: 
28: \newcommand{\cl}[2]{\ensuremath{\langle {#1} | {#2} \rangle}}
29: 
30: \def\ispreprint{}
31: 
32: \ifx\ispreprint\undefined
33: \def\baselinestretch{2.5}
34: \else
35: \fi
36: 
37: \author{Joshua Cooper\\ \small Department of Mathematics \\ \small University of California, San Diego, La Jolla, CA \\ \small \texttt{jcooper@math.ucsd.edu}}
38: \date{\today}
39: 
40: \begin{document}
41: \title{{\huge \bf Quasirandom Permutations}}
42: \ifx\ispreprint\undefined
43: \author{{\Large Joshua N.\ Cooper}\\
44: {\it Department of Mathematics, University of California}\\
45: {\it at San Diego, La Jolla, California}\\
46: E-mail: jcooper@math.ucsd.edu\\
47: Proposed running head: Quasirandom Permutations}
48: \else
49: \author{{\Large Joshua N.\ Cooper}\\
50: {\it Department of Mathematics, University of California}\\
51: {\it at San Diego, La Jolla, California}\\
52: E-mail: jcooper@math.ucsd.edu}
53: \fi
54: 
55: \date{July 1, 2002}
56: 
57: \maketitle
58: 
59: \ifx\ispreprint\undefined \pagebreak \else \fi
60: \renewcommand{\thesection}{\Roman{section}.}
61: \maketitle
62: 
63: \begin{abstract}
64: Chung and Graham [\ref{CG2}] define quasirandom subsets of
65: $\mathbb{Z}_n$ to be those with any one of a large collection of
66: equivalent random-like properties.  We weaken their definition and
67: call a subset of $\mathbb{Z}_n$ $\epsilon$-balanced if its
68: discrepancy on each interval is bounded by $\epsilon n$.  A
69: quasirandom permutation, then, is one which maps each interval to
70: a highly balanced set.  In the spirit of previous studies of
71: quasirandomness, we exhibit several random-like properties which
72: are equivalent to this one, including the property of containing
73: (approximately) the expected number of subsequences of each
74: order-type.  We provide a few applications of these results, present a construction for a family of
75: strongly quasirandom permutations, and prove that this construction
76: is essentially optimal, using a result of W. Schmidt on the
77: discrepancy of sequences of real numbers.
78: \end{abstract}
79: 
80: \ifx\ispreprint\undefined \indent {\it Key words:} quasirandom;
81: permutation; discrepancy; van der Corput
82: \\
83: \noindent {\it Correspondence Data:}
84: 
85: \indent Joshua N. Cooper \vspace{-.25in}
86: 
87: \indent UCSD Department of Mathematics \vspace{-.25in}
88: 
89: \indent 9500 Gilman Drive 0112 \vspace{-.25in}
90: 
91: \indent La Jolla, CA 92093-0112 \vspace{-.25in}
92: 
93: \indent jcooper@math.ucsd.edu
94: 
95: \pagebreak \else \fi
96: 
97: \section{Introduction}
98: 
99: In recent years, combinatorialists have been investigating several
100: realms of random-like -- ``quasirandom'' -- objects.  For a given
101: probability space $\mathcal X$, the basic idea is to choose some
102: collection of properties that large objects in $\mathcal X$ have
103: almost surely, and define a sequence $\{X_i\}_{i=1}^{\infty}
104: \subset \mathcal X$ to be {\it quasirandom} if $X_i$ has these
105: properties in the limit.  Often, this approach amounts to choosing
106: some random variables $\eta_j$ defined on $\mathcal X$ which tend
107: to their expected values almost surely as $|X| \rightarrow
108: \infty$, and defining $X_i$ to be quasirandom when
109: $(\eta_0(X_i),\eta_1(X_i),\ldots) \rightarrow
110: (\mathbb{E}\eta_0,\mathbb{E}\eta_1,\ldots)$ sufficiently quickly.
111: The resulting definitions are explored by finding many such
112: collections of properties and showing that quasirandomness with
113: respect to any one of them is equivalent to all the rest -- often
114: rather surprisingly, since the properties may appear completely
115: unrelated to one another. Quasirandom graphs, hypergraphs, set
116: systems, subsets of $\mathbb{Z}_n$, and tournaments have all been
117: examined in this way.  Quasirandom families of permutations have
118: been defined in [\ref{MMW1}], and Gowers [\ref{G1}] has used a careful
119: quantitative analysis of strongly quasirandom
120: (``$\alpha$-uniform'', in his terminology) subsets of
121: $\mathbb{Z}_n$ as an integral component of his remarkable new
122: proof of Szemer\'edi's Theorem.  Quasirandom objects
123: also have applications in algorithms as deterministic substitutes
124: for randomly generated objects, in addition to their purely
125: theoretical uses.  In fact, specific types of random-like
126: permutations have been used already in a number of contexts.
127: Lagarias [\ref{L1}] constructed random-like permutations of a
128: $d$-dimensional array of cells in order to solve a practical
129: memory-mapping problem, and Alon [\ref{A1}] used ``pseudo-random''
130: permutations to improve on the best known deterministic
131: maximum-flow algorithm of Goldberg and Tarjan.  Quasirandom sequences of reals are also fundamental to the extensively studied ``quasi-Monte Carlo'' methods of numerical analysis ([\ref{N1}]). In this paper, we
132: add (individual) permutations to the growing list of objects for
133: which a formal notion of quasirandomness has been defined.
134: 
135: In Section 2, we discuss the concept of $\epsilon$-balance, which
136: weakens the quasirandomness of Chung and Graham.  It is shown to
137: be equivalent to several ``types'' of quasirandomness for subsets
138: of $\mathbb{Z}_n$, including an infinite family of eigenvalue
139: bounds.  Section 3 is an excursion into the realm of subsequence
140: statistics of permutations, a subject that has generated a good
141: deal of interest recently (e.g., [\ref{AF1}] and [\ref{B1}]) --
142: and whose roots go back at least as far as 1935 ([\ref{ES1}]).  In
143: Section 4, quasirandom permutations are defined as those which map
144: intervals to uniformly balanced sets, and we prove that this
145: definition is equivalent to several other random-like conditions.  Two applications are given, including a proof that random permutations have small discrepancy.
146: Section 5 contains a construction for a family of strongly
147: quasirandom permutations that generalize the classical van der Corput sequences.  We show that this construction is
148: essentially optimal, using a result of Schmidt on the discrepancy
149: of sequences of real numbers.  Finally, Section 6 concludes with
150: some open problems and directions for future work.
151: 
152: \bigskip
153: \section{Balanced Sets}
154: \setcounter{theorem}{0}
155: 
156: Throughout the following, we consider permutations, i.e., elements
157: of $S_n$, as actions on $\mathbb{Z}_n$ as well as sequences of
158: numbers $(\sigma(0), \sigma(1), \ldots , \sigma(n-1))$ (``one-line
159: notation''). When an ordering on $\mathbb{Z}_n$ is used, we mean
160: the one inherited from $[0,n-1] \subset \mathbb{Z}$.  If $f_i$,
161: $i=1,2$, is a function from a totally ordered set $A$ to a totally
162: ordered set $B_i$, we say that $f_1$ and $f_2$ are
163: \textit{isomorphic} (and write $f_1 \sim f_2$) if, for any $a_1,
164: a_2 \in A$, $f_1(a_1) < f_1(a_2)$ iff $f_2(a_1) < f_2(a_2)$.  Note
165: that this definition still makes sense when $f_1$ and $f_2$ are
166: defined on different sets $A_1$ and $A_2$, so long as $|A_1| =
167: |A_2|$ is finite and we identify them via the unique
168: order-isomorphism between them. Then, if $\sigma \in S_n$ and
169: $\tau \in S_m$, $m \leq n$, we say that $\tau$ occurs in $\sigma$
170: at the set $A = \{ a_i \}_{i=1}^{m} \subset \mathbb{Z}_n$ whenever
171: $\sigma|_A \sim \tau$.  For each $A \subset \mathbb{Z}_n$ and
172: permutation $\tau$, we write ${\bf X}^{\tau}_A(\sigma)$ for the
173: indicator random variable of the event that $\tau$ occurs in
174: $\sigma$ at $A$, and we write ${\bf X}^{\tau}(\sigma)$ for the
175: random variable that counts the number of occurrences of $\tau$ in
176: $\sigma$, i.e., ${\bf X}^{\tau}(\sigma) = \sum_A {\bf
177: X}^{\tau}_A(\sigma)$ where $A$ ranges over all subsets of
178: $\mathbb{Z}_n$ of cardinality $m$.
179: 
180: For any subset $S \subset \mathbb{Z}_n$ (or $S \subset
181: \mathbb{Z}$), there is a minimal representation of $S$ as a union
182: of intervals.  We call these intervals the \textit{components} of
183: $S$ and denote the number of them by $c(S)$.  Also, we adopt the
184: convention that the symbols for a set and the characteristic
185: function of that set be the same, so, for example, $S(x) = 1$ if
186: $x \in S$ and $S(x) = 0$ if $x \not \in S$.  Finally, for any
187: function from $\mathbb{Z}_n$ to $\mathbb{C}$, we write
188: $\tilde{f}(k)$ for the $k^{\mbox{\scriptsize th}}$ Fourier
189: coefficient of $f$, defined by
190: $$
191: \tilde{f}(k) = \sum_{x \in \mathbb{Z}_n} f(x) e^{-2 \pi ikx/n}.
192: $$
193: A well known alternative definition of the Fourier coefficients of
194: a set $S$ is the spectrum of the circulant matrix $M_S$ whose
195: $(i,j)$ entry is $S(i+j)$.
196: 
197: \bigskip
198: 
199: One would expect random permutations to ``jumble'' the elements on
200: which it acts, i.e., there should be no correlation between
201: proximity in $\mathbb{Z}_n$ and proximity in the image.  We can
202: measure proximity by means of intervals: the elements of a small
203: interval are all ``close'' to one another.  Thus, if we define an
204: interval of $\mathbb{Z}_n$ to be the projection of any interval of
205: $\mathbb{Z}$, a permutation $\sigma \in S_n$ will be called
206: ``quasirandom'' if the intersection of any interval $I$ with the
207: image of any other interval $J$ under $\sigma$ has cardinality
208: approximately $|I||J|/n$, i.e., no interval contains much more or
209: less of the image of any other interval than one would expect if
210: $\sigma$ were chosen randomly.
211: 
212: Thus, for any two sets $S, T \subset \mathbb{Z}_n$ we define the
213: {\it discrepancy} of $S$ in $T$ as
214: $$
215: D_T(S) = \left | | S \cap T | - \frac{|S||T|}{n} \right |.
216: $$
217: Note that we may apply this definition to multisets $S$ and $T$,
218: and that it is symmetric in its arguments.  Before proceeding, we
219: present a simple lemma to the effect that $D$ is subadditive:
220: 
221: \begin{lemma} \label{subadd} If $S = A \cup B$, $A$ and $B$ disjoint, then $D_S(T) \leq D_A(T) + D_B(T)$.  If $T = C \cup D$, $C$ and $D$ disjoint, then $D_S(T) \leq D_S(C) + D_S(D)$.  That is, $D$ is subadditive in both of its arguments.
222: \end{lemma}
223: \begin{proof}
224: By the triangle inequality, we have
225: \begin{align*}
226: D_T(S) & = \left | \, | \, S \cap T | - \frac{|S||T|}{n} \right | \\
227: & = \left | \, | \, A \cap T | + | \, B \cap T | - \frac{|A||T|}{n} - \frac{|B||T|}{n} \right | \\
228: & \leq \left | D_T(A) \right | + \left | D_T(B) \right |.
229: \end{align*}
230: The other statement follows by symmetry.
231: \end{proof}
232: 
233: Define $D(S)$ to be the maximum of $D_J(S)$, taken over all
234: intervals $J \subset \mathbb{Z}_n$, and call a set $S \subset
235: \mathbb{Z}_n$ $\epsilon${\it -balanced} if $D(S) < \epsilon n$.
236: This definition of quasirandomness is implied by that of Chung and
237: Graham [\ref{CG2}], according to the next proposition.
238: 
239: \begin{prop} \label{weak} If, for all $T \subset \mathbb{Z}_n$ and all but $\epsilon n$ of $x \in \mathbb{Z}_n$, $D_{T+x}(S) < \epsilon n$, then, for all intervals $J \subset \mathbb{Z}_n$, $D_J(S) < 2 \epsilon n$.
240: \end{prop}
241: \begin{proof}
242: Suppose there exists an interval $J \subset \mathbb{Z}_n$ such that
243: $$
244: \left | \, |S \cap J| - \frac{|S||J|}{n} \right | \geq 2 \epsilon
245: n.
246: $$
247: Then, for all $x \in \mathbb{Z}_n$,
248: $$
249: \left | \, |S \cap (J+x)| - \frac{|S||J+x|}{n} \right | \geq \left
250: | \, |S \cap J| - \frac{|S||J|}{n} \right | - |x|.
251: $$
252: Therefore, for each $x$ with $|x| \leq \epsilon n$, $D_{J+x}(S)
253: \geq \epsilon n$.  Since there are at least $\epsilon n$ such
254: $x$'s, setting $T = J$ contradicts the hypothesis of the
255: proposition.
256: \end{proof}
257: 
258: It is easy to see that the set $S = \{2x \, | \, 0 \leq x \leq n-1
259: \} \subset \mathbb{Z}_{2n}$ is, for any $\epsilon > 0$ and
260: sufficiently large $n$, $\epsilon$-bounded.  However, $S \cap
261: (S+t)$ does not have cardinality approximately $|S|^2/n$ for
262: almost all $t$, i.e., it violates ``weak translation''. Therefore,
263: $\epsilon$-boundedness is strictly weaker than quasirandomness in
264: the sense of [\ref{CG2}].
265: 
266: We use the convention that when ``little oh'' notation is used,
267: convergence in $n$ alone is intended.  (That is, the convergence
268: is uniform in any other quantities involved.)  The following is
269: the main result of this section.
270: 
271: \begin{theorem} \label{balance} For $r \in \mathbb{Z}_n$, we define $|r|$ to be the absolute value of the unique representative of $\, r$ from the interval $(-n/2,n/2]$.  Then, for any sequence of subsets $S \subset \mathbb{Z}_n$ and choice of $\alpha > 0$, the following are equivalent: \\ [.06in]
272: 
273: \noindent \begin{tabular}{p{.4in}p{4.6in}}
274: {\bf [B]}  & (Balance) $D(S) = o(n)$. \\
275: \end{tabular}
276: 
277: \noindent \begin{tabular}{p{.4in}p{4.6in}}
278: {\bf [PB]} & (Piecewise Balance) For any subset $T \subset \mathbb{Z}_n$, $D_T(S) = o(nc(T))$, where $c(T)$ denotes the number of components of $T$. \\
279: \end{tabular}
280: 
281: \noindent \begin{tabular}{p{.4in}p{4.6in}}
282: {\bf [MB]} & (Multiple Balance) Let $kS$ denote the multiset $\{ks|s \in S\}$.  Then, for any $k \in \mathbb{Z}_n \! \setminus \! \{0\}$, $D(kS) = o(n|k|)$. \\
283: \end{tabular}
284: 
285: \noindent \begin{tabular}{p{.4in}p{4.6in}}
286: {\bf [E($\frac{1}{2}$)]} & (Eigenvalue Bound $\frac{1}{2}$) For all nonzero $k \in \mathbb{Z}_n$, $\tilde{S}(k) = o(n|k|^{1/2})$. \\
287: \end{tabular}
288: 
289: \noindent \begin{tabular}{p{.4in}p{4.6in}}
290: {\bf [E($\alpha$)]} & (Eigenvalue Bound $\alpha$) For all nonzero $k \in \mathbb{Z}_n$, $\tilde{S}(k) = o(n|k|^{\alpha})$. \\
291: \end{tabular}
292: 
293: \noindent \begin{tabular}{p{.4in}p{4.6in}}
294: {\bf [S]} & (Sum) $ \sum_{r \neq 0} \left( |\tilde{S}(k)|/|k| \right )^2 = o(n^2). $ \\
295: \end{tabular}
296: 
297: \noindent \begin{tabular}{p{.4in}p{4.6in}} {\bf [T]} &
298: (Translation) For any interval $J$,
299: $$ \sum_{k \in \mathbb{Z}_n} \left ( |S \cap (J+k)| - \frac{|S||J|}{n} \right )^2 = o(n^3). $$
300: \end{tabular}
301: \end{theorem}
302: 
303: \ifx\ispreprint\undefined
304: \else
305: \begin{center}
306: \begin{figure}[ht]
307: \hspace{.7in} \input{figa.pstex_t} \caption{Diagram of
308: implications for Theorem \ref{balance}.}
309: \end{figure}
310: \end{center}
311: \fi
312: 
313: We will show that {\bf [B]} $\Rightarrow$ {\bf [PB]} $\Rightarrow$
314: {\bf [MB]} $\Rightarrow$ {\bf [E($\frac{1}{2}$)]} $\Rightarrow$
315: {\bf [E($\alpha$)]} $\Rightarrow$ {\bf [S]} $\Rightarrow$ {\bf
316: [T]} $\Rightarrow$ {\bf [B]}.  In each case, a statement involving
317: some $\epsilon$ is shown to imply the next for some $f(\epsilon)$,
318: where $f$ is a function which tends to zero as its argument does.
319: For example, Proposition \ref{division} below states that if
320: $D_T(S) < \epsilon n c(T)$ for all $T$, then $D_T(kS) < 2 \epsilon
321: n |k|$ for all k, so that $f(\epsilon)=2\epsilon$.  It appears to
322: be theoretically useful to track what happens to $\epsilon$ as we
323: pass through each implication -- see, for example, [\ref{G1}].
324: Thus, we include Figure 1 as an accompaniment to Theorem
325: \ref{balance}.  (Note that, by the proof of Proposition
326: \ref{MB2EA}, Figure 1 is only valid for $\epsilon < \pi / 8$,
327: though this is hardly a significant restriction.)  The shortcut
328: edge from {\bf [E($\frac{1}{2}$)]} to {\bf [S]} is given to
329: illustrate the (best possible) choice of $\alpha = 1/4$ in {\bf
330: [E($\alpha$)]}, and the circular arrow represents one complete
331: traversal of the cycle of implications, including the shortcut
332: edge.
333: 
334: Theorem \ref{balance} is proven in pieces, beginning with the
335: following proposition.
336: 
337: \begin{prop} {\bf [B]} $\Rightarrow$ {\bf [PB]} $\Rightarrow$ {\bf [MB]}. \label{division}
338: \end{prop}
339: \begin{proof}
340: Suppose that $D(S) < \epsilon n$.  Then, by Lemma \ref{subadd},
341: for any $T$, $D_T(S) \leq \sum D_{T_i}(S)$, where the sum is over
342: the components of $T$.  Thus, $D_T(S) < \epsilon n \, c(T)$, and
343: {\bf [B]} $\Rightarrow$ {\bf [PB]}.
344: 
345: Now, suppose {\bf [PB]} holds for $S$.  Note that, for a given $k
346: \in \mathbb{Z}_n \setminus \{0\}$ and interval $J$, the set
347: $J^\prime$ of elements $x \in \mathbb{Z}_n$ such that $kx \in J$
348: has at most $|k|$ components.  Let $J_i$ be the set of integer
349: points (viewed as elements of $\mathbb{Z}_n$) lying in $[a/k,b/k]
350: + in/k$, so that $J^\prime = \bigcup_i J_i$.  Then the cardinality
351: of $J_i$ is off from $|J|/k$ by at most $1$.  By {\bf [PB]} and
352: the triangle inequality,
353: \begin{align*}
354: D_J(kS) &= \left | \, |k(S-t) \cap J| - \frac{|I||J|}{n} \right | \\
355: & \leq \left | \, \sum_i |S \cap J_i| - \sum_i \frac{|I||J_i|}{n} \right | + \frac{|I|}{n} \left | \, \sum_i |J_i| - |J| \right | \\
356: & < \epsilon n |k| + |k| \frac{|I|}{n} \leq 2 \epsilon n |k|.
357: \end{align*}
358: since, trivially, $\epsilon \geq n^{-1}$.
359: \end{proof}
360: 
361: Now, we wish to show that Multiple Boundedness implies the first
362: eigenvalue bound.  The basic idea is to imbed the elements of $S$
363: into the unit circle via the exponential map, and then show that a
364: great deal of cancellation happens because of the relatively
365: uniform distribution of elements of $S$.
366: 
367: \begin{prop} {\bf [MB]} $\Rightarrow$ {\bf [E($\frac{1}{2}$)]}. \label{MB2EA}
368: \end{prop}
369: \begin{proof}
370: Let $\omega = e^{2 \pi i / n}$ and $J_m^j = [\frac{nj}{m},\frac{n(j+1)}{m})$, and let $\epsilon$ be the bound on $(nk)^{-1} D(k S)$.  Recall that $\epsilon \geq n^{-1}$.  First we prove the following:
371: \begin{claim} Let $m$ and $j$ be positive integers with $0 \leq j < m$, and $m \geq 2$.  If we define the multiset $S_j = k S \cap J_m^j$ and let $\gamma_j = \omega^{-n (j+1/2)/m}$, then
372: $$
373: \left | \, \sum_{x \in S_j} \omega^{-x} - \frac{|S|}{m} \gamma_j
374: \right | < \frac{\pi |S|}{m^2} + 2 \epsilon |k|n
375: $$
376: \end{claim}
377: 
378: \begin{poc} We may write the left-hand side of the above expression as
379: \begin{align*}
380: \left | \, \sum_{x \in S_j} \omega^{-x} - \frac{|S|}{m} \gamma_j \right | &= \left | \, \sum_{x \in S_j} (\omega^{-x} - \frac{|S|}{m|S_j|} \gamma_j ) \right | \\
381: & \leq \frac{|S|}{m|S_j|} \left | \, \sum_{x \in S_j} (\omega^{-x} - \gamma_j ) \right | + \left | \, \sum_{x \in S_j} \omega^{-x} (1 - \frac{|S|}{m|S_j|}) \right | \\
382: & \leq \frac{|S|}{m|S_j|} \sum_{x \in S_j} \left | \, (\omega^{-x}
383: - \gamma_j ) \right | + \sum_{x \in S_j} \left | \, \omega^{-x} (1
384: - \frac{|S|}{m|S_j|}) \right |
385: \end{align*}
386: Now, for $x \in S_j$,
387: $$
388: | \, \omega^{-x} - \gamma_j | \leq | \, \omega^{-n j / m} -
389: \omega^{-n (j+1/2) / m} | \leq \frac{n/2}{m} \cdot \frac{2 \pi}{n}
390: = \frac{\pi}{m}
391: $$
392: Plugging this expression in and applying {\bf [MB]}, we have
393: \begin{align*}
394: \left | \, \sum_{x \in S_j} \omega^{-x} - \frac{|S|}{m} \gamma_j \right | & \leq \frac{|S|}{m|S_j|} \cdot |S_j| \cdot \frac{\pi}{m} + \left | |S_j| - \frac{|S|}{m} \right | \\
395: & \leq \frac{\pi |S|}{m^2} + \left | |kS \cap J_m^j| - \frac{|I||J_m^j|}{n} \right | + \frac{|I|}{n} \left | \frac{n}{m} - |J_m^j| \right | \\
396: & < \frac{\pi |S|}{m^2} + \epsilon |k| n + \frac{|I|}{n} \\
397: & \leq \frac{\pi |S|}{m^2} + 2 \epsilon |k| n
398: \end{align*}
399: thus, proving the claim.
400: \end{poc}
401: 
402: \medskip
403: 
404: If we sum over all $j \in [0, m-1)$,
405: \begin{align*}
406: \left | \, \sum_{x \in S} \omega^{-kx} \right | &= \left | \, \sum_{j=0}^{m-1} \sum_{x \in S_j} \omega^{-x} \right | \\
407: & \leq \left | \, \sum_{j=0}^{m-1} \frac{|S|}{m} \gamma_j \right | + \sum_{j=0}^{m-1} \left | \, \sum_{x \in S_j} \omega^{-x} - \frac{|S|}{m} \gamma_j \right | \\
408: & < 0 + \frac{\pi |S|}{m} + 2 \epsilon |k| nm
409: \end{align*}
410: if we assume that $m \geq 2$.  Thus, letting $m = \left \lfloor
411: \left ( \frac{\pi |S|}{2\epsilon |k|n} \right )^{1/2} \right
412: \rfloor$, we have
413: $$
414: \left | \, \sum_{x \in S} \omega^{-kx} \right | < \sqrt{18 \pi
415: \epsilon n|k||S|} \leq n \sqrt{18 \pi \epsilon |k|}
416: $$
417: unless $m < 2$, i.e., $ \epsilon > \frac{\pi}{8} $, which is
418: eventually impossible, given {\bf [MB]}.  We may therefore
419: conclude that $| \tilde{S}(k) | = o(n|k|^{1/2})$.
420: \end{proof}
421: 
422: A small improvement to the constant in the bound above is possible
423: by letting $m$ be rational, instead of integral. However, doing so
424: adds some complexity to the proof without making any significant
425: improvements.
426: 
427: Before we proceed with the next implication, the following lemma
428: will be necessary.  It implies, surprisingly, that {\bf
429: [E($\alpha$)]} is equivalent to {\bf [E($\beta$)]} for all
430: $\alpha$ and $\beta$.
431: 
432: \begin{lemma} \label{EA2EB} {\bf [E($\alpha$)]} implies {\bf [E($\beta$)]} for any $\alpha, \beta > 0$.
433: \end{lemma}
434: \begin{proof}
435: Let $M = \left \lceil \frac{\alpha}{\beta} \right \rceil$, and
436: assume {\bf [E($\alpha$)]}.  Then
437: $$
438: \left | \tilde{S}(k) \right |^{M} = \left | \sum_{t \in
439: \mathbb{Z}_n} S(t) \omega^{-kt} \right |^{M} = \left |
440: \sum_{t_1,\ldots,t_{M}} \left [ \prod_{j=1}^{M} S(t_j) \right ]
441: \omega^{-k \sum_{i=1}^M t_i} \right |
442: $$
443: Letting $u = \sum_{i=2}^M t_i$, we have
444: \begin{align*}
445: \left | \tilde{S}(k) \right |^{M} &= \left | \sum_{t_2,\ldots,t_{M}} \left [ \prod_{j=2}^{M} S(t_j) \right ] \sum_{t_1} S(t_1) \omega^{-k (t_1 + u)} \right | \\
446: & \leq \sum_{t_2,\ldots,t_{M}} \left [ \prod_{j=2}^{M} S(t_j) \right ] \left | \sum_{t_1} S(t_1) \omega^{-kt_1} \right | \\
447: &= \sum_{t_2,\ldots,t_{M}} \left [ \prod_{j=2}^{M} S(t_j) \right ] \left | \tilde{S}(k) \right | \\
448: &< \sum_{t_2,\ldots,t_{M}} \left [ \prod_{j=2}^{M} S(t_j) \right ] \epsilon n |k|^\alpha \\
449: &= |S|^{M-1} \epsilon n |k|^\alpha \leq \epsilon n^{M} |k|^\alpha.
450: \end{align*}
451: Thus, taking the $M^{\mbox{\scriptsize th}}$ root of both sides,
452: we have
453: $$
454: \left | \tilde{S}(k) \right | < \epsilon^{1/M} n |k|^{\alpha/M}
455: \leq \epsilon^{\left \lceil \alpha/\beta \right \rceil^{-1}} n
456: |k|^{\beta}.
457: $$
458: \end{proof}
459: 
460: The following corollary is actually what is needed for Theorem
461: \ref{balance}.
462: 
463: \begin{cor} {\bf [E($\frac{1}{2}$)]} $\Rightarrow$ {\bf [E($\alpha$)]}.
464: \end{cor}
465: 
466: Note that, to proceed with the next proposition, $\alpha = 1/2$
467: would not quite be enough -- we {\it have} to reduce it by a bit
468: with Proposition \ref{EA2EB}.
469: 
470: \begin{prop} {\bf [E($\alpha$)]} $\Rightarrow$ {\bf [S]}.
471: \end{prop}
472: \begin{proof}
473: By Proposition \ref{EA2EB}, we know that $|\tilde{S}(k)| <
474: \epsilon^{\left \lceil 4\alpha \right \rceil^{-1}} n |k|^{1/4}$
475: for all $k \neq 0$. Then
476: $$
477: \sum_{r \neq 0} \left( \frac{|\tilde{S}(k)|}{|k|} \right )^2 <
478: \sum_{k \neq 0} \left( \frac{\epsilon^{\left \lceil 4\alpha \right
479: \rceil^{-1}} n |k|^{1/4}}{|k|} \right )^2 \leq \epsilon^{2 \left
480: \lceil 4\alpha \right \rceil^{-1}} n^2 \sum_{k \neq 0} |k|^{-3/2}
481: < 6 \epsilon^{2 \left \lceil 4\alpha \right \rceil^{-1}} n^2.
482: $$
483: where we have used the approximation $|\zeta(s)| < (Re(s)-1)^{-1}
484: + 1$ for $s$ with $Re(s) > 1$.
485: \end{proof}
486: 
487: We now write a cyclic sum in terms of Fourier coefficients. A proof of the following standard lemma is included for the sake of completeness.
488: 
489: \begin{lemma} \label{intb} If $J$ is an interval of $\mathbb{Z}_n$, then $\tilde{J}(k) \leq \frac{n}{2|k|}$.
490: \end{lemma}
491: \begin{proof}
492: We may write the magnitude of the $k^{\mbox{\scriptsize th}}$
493: Fourier coefficient of $J = [a+1,a+M]$ as
494: \begin{align*}
495: |\tilde{J}(k)| &= | \sum_x J(x) \omega^{-kx} | = | \sum_{x=a}^b \omega^{-kx} | = | \sum_{x=1}^M \omega^{-kx} | \\
496: & = \frac{|\omega^{-kM} - 1|}{|\omega^{-k} - 1|} \leq \frac{2}{4
497: |k| / n} = \frac{n}{2|k|}
498: \end{align*}
499: since $|e^{i \theta} - 1| \geq \frac{2 |\theta|}{\pi}$ for all
500: $\theta$.
501: \end{proof}
502: 
503: \begin{prop} {\bf [S]} $\Rightarrow$ {\bf [T]}.
504: \end{prop}
505: \begin{proof}
506: Assume that $\sum_{k \neq 0} \left( \frac{|\tilde{S}(k)|}{|k|}
507: \right )^2 < \epsilon n^2$.  We may write the ``translation'' sum
508: as
509: \begin{equation} \label{eqa1}
510: \sum_{k \in \mathbb{Z}_n} \left ( |S \cap (J+k)| -
511: \frac{|S||J|}{n} \right )^2 = \sum_{k \in \mathbb{Z}_n} |S \cap
512: (J+k)|^2 - \frac{|S|^2|J|^2}{n}
513: \end{equation}
514: Recall that $M_S$ is the $n \times n$ matrix whose $(i,j)$ entry
515: is $S(i+j)$.  Letting {\boldmath $v$} be the vector
516: $(J(0),J(1),\ldots)$, we find that $M_S \mbox{\boldmath $v$}$ is
517: the vector whose $k^{\mbox{\scriptsize th}}$ entry is $|I \cap
518: (J+k)|$.  Therefore, letting $\phi_k =
519: (1,\omega^{k},\omega^{2k},\ldots)$ be the $k^{\mbox{\scriptsize
520: th}}$ eigenvector of $M_S$,
521: \begin{align*}
522: \sum_{k \in \mathbb{Z}_n} |S \cap (J+k)|^2 &= | M_S \mbox{\boldmath $v$} |^2 = | M_S \sum_k \frac{<\mbox{\boldmath $v$},\phi_k>}{|\phi_k|^2}\,\phi_k |^2 \\
523: &= \sum_k | \tilde{S}(k)^2 \frac{<\mbox{\boldmath $v$},\phi_k>^2}{|\phi_k|^2} | \\
524: &= \sum_{k \neq 0} \left | \tilde{S}(k)
525: \frac{\tilde{J}(-k)}{\sqrt{n}} \right |^2 + \frac{|S|^2|J|^2}{n}
526: \end{align*}
527: Applying this equality, property {\bf [S]}, and Lemma \ref{intb}
528: to Equation \ref{eqa1},
529: \begin{align*}
530: \sum_{k \in \mathbb{Z}_n} \left ( |S \cap (J+k)| - \frac{|S||J|}{n} \right )^2 &= \sum_{k \neq 0} \left | \tilde{S}(k) \frac{\tilde{J}(-k)}{\sqrt{n}} \right |^2 \\
531: & \leq \frac{n}{4} \sum_{k \neq 0} \left | \frac{\tilde{S}(k)}{|k|} \right |^2 \\
532: & < \frac{\epsilon^2}{4} n^3.
533: \end{align*}
534: \end{proof}
535: 
536: To complete the circle of implications and finish the proof of
537: Theorem \ref{balance}, we show that $\epsilon$-boundedness is
538: implied by the ``translation'' property.
539: 
540: \begin{prop} {\bf [T]} $\Rightarrow$ {\bf [B]}
541: \end{prop}
542: \begin{proof}
543: Suppose that, for some interval $J \subset \mathbb{Z}_n$,
544: $$
545: \left | |S \cap J| - \frac{|S||J|}{n} \right | \geq 2
546: \epsilon^{1/3} n
547: $$
548: Then, following the line of argument given for Proposition
549: \ref{weak}, we may conclude that
550: $$
551: \left | |S \cap (J+k)| - \frac{|S||J|}{n} \right | \geq
552: \epsilon^{1/3} n
553: $$
554: whenever $|k| \leq \epsilon^{1/3} n$.  Since there are at least
555: $\epsilon^{1/3} n $ such $k$'s,
556: $$
557: \sum_k \left ||S \cap (J+k)| - \frac{|S||J|}{n} \right |^2 \geq
558: \epsilon^{1/3} n \cdot \epsilon^{2/3} n^2 = \epsilon n^3
559: $$
560: contradicting {\bf [T]}.
561: \end{proof}
562: 
563: \bigskip
564: \section{Statistics of Sub-Permutations}
565: \setcounter{theorem}{0}
566: 
567: Before we formally define quasirandom permutations, an excursion
568: into the realm of subsequence statistics is necessary. We wish to
569: relate ${\bf X}^{\tau}(\sigma)$, for $\tau \in S_m$, to the
570: quantities ${\bf X}^{\tau^\prime}(\sigma)$, with $\tau^\prime \in
571: S_{m+1}$, by counting occurrences of $\tau$ inside each occurrence
572: of $\tau^\prime$.  Define ${\bf v}_m(\sigma) \in \mathbb{Z}^{S_m}$
573: to be the vector whose $\tau$ component is ${\bf
574: X}^{\tau}(\sigma)$, and write ${\bf \tilde{v}}_m(\sigma)$ for the
575: vector
576: $$
577: {\bf v}_m(\sigma) - \mathbb{E}{\bf v}_m = {\bf v}_m(\sigma) -
578: \frac{\bf \hat{1}}{m!} \binom{n}{m}
579: $$
580: Also, let $B_m$ be the $m!$ by $(m+1)!$ matrix whose
581: $(\tau,\tau^\prime)$ entry is ${\bf X}^{\tau}(\tau^\prime)$ for
582: $\tau \in S_m$ and $\tau^\prime \in S_{m+1}$, and define $A_m$ to
583: be $B^*_m B_m$.
584: 
585: \begin{prop} For any $\sigma \in S_m$,
586: \begin{equation} \label{mainform}
587: (n-m)^2 {| {\bf \tilde{v}}_m(\sigma) |}^2 = {{\bf
588: \tilde{v}}_{m+1}(\sigma)}^* A_m {\bf \tilde{v}}_{m+1}(\sigma)
589: \end{equation}
590: \end{prop}
591: \begin{proof}
592: Let $\Gamma$ be the set of pairs $(U,V)$, with $U \subset V
593: \subset \mathbb{Z}_n$, $|U|=m$, $|V|=m+1$, and $\sigma_U \sim
594: \tau$.  Then, conditioning on the order-type of $U$ yields
595: $$
596: |\Gamma| = \sum_{\tau^\prime \in S_{m+1}} {\bf
597: X}^{\tau}(\tau^\prime) {\bf X}^{\tau^\prime}(\sigma)
598: $$
599: because each set $U$ contributes ${\bf X}^{\tau}(\sigma|_U)$ to
600: $|\Gamma|$.  If we instead condition on $V$ itself, then
601: $$
602: |\Gamma| = (n - m) {\bf X}^{\tau}(\sigma)
603: $$
604: because each subset $V$ is contained in exactly $n-m$ supersets
605: $U$.  We may therefore write
606: $$
607: \sum_{\tau^\prime \in S_{m+1}} {\bf X}^{\tau}(\tau^\prime) {\bf
608: X}^{\tau^\prime}(\sigma) = (n - m) {\bf X}^{\tau}(\sigma),
609: $$
610: i.e., $(n-m) {\bf v}_m(\sigma) = B_m {\bf v}_{m+1}(\sigma)$.  The
611: desired result then follows by linearity of expectation.
612: \end{proof}
613: 
614: Now that a numerical relationship between subsequences of length
615: $m+1$ and subsequences of length $m$ has been established, we need
616: a bound on the eigenvalues of $A_m$.
617: 
618: \begin{prop} \label{propsums} The following hold for all $m \geq 1$:
619: \begin{enumerate}
620: \item The column sums of $B_m$ are equal to $m+1$.
621: \item The row sums of $B_m$ are equal to $(m+1)^2$.
622: \item The row (and column) sums of $A_m$ are equal to $(m+1)^3$.
623: \end{enumerate}
624: \end{prop}
625: \begin{proof}
626: The proofs are all straightforward manipulations.
627: 
628: \begin{enumerate}
629: \item Let $b_{\tau \tau^\prime}$ denote the $(\tau,\tau^\prime)$ entry of the matrix $B_m$.  Then, denoting the set of subsets of $\mathbb{Z}_{m+1}$ of cardinality $m$ by $P_m$,
630: \begin{align*}
631: \sum_{\tau \in S_{m}} b_{\tau \tau^\prime} & = \sum_{\tau \in S_{m}} {\bf X}^{\tau}(\tau^\prime) = \sum_{\tau \in S_{m}} \sum_{A \in \mathcal{P}_m} {\bf X}^{\tau}_A(\tau^\prime) \\
632: & = \sum_{A \in \mathcal{P}_m} \sum_{\tau \in S_{m}} {\bf
633: X}^{\tau}_A(\tau^\prime) = \sum_{A \in \mathcal{P}_m} 1 = m+1.
634: \end{align*}
635: \item Note that, for a set $A \in P_m$, a permutation $\tau^\prime \in S_{m+1}$ is uniquely determined by its restriction to $A$.  Therefore, for a given $\tau \in S_m$, the number of $\tau^\prime \in S_{m+1}$ such that $\tau^\prime|_A \sim \tau$ is equal to the number of possible sets $\tau^\prime(A)$, i.e.,
636: $$
637: \sum_{\tau^\prime \in S_{m+1}} {\bf X}^{\tau}_A(\tau^\prime) =
638: m+1.
639: $$
640: Now, we sum over the first index:
641: \begin{align*}
642: \sum_{\tau^\prime \in S_{m+1}} b_{\tau \tau^\prime} & = \sum_{\tau^\prime \in S_{m+1}} {\bf X}^{\tau}(\tau^\prime) = \sum_{\tau^\prime \in S_{m+1}} \sum_{A \in \mathcal{P}_m} {\bf X}^{\tau}_A(\tau^\prime) \\
643: & = \sum_{A \in \mathcal{P}_m} \sum_{\tau^\prime \in S_{m+1}} {\bf
644: X}^{\tau}_A(\tau^\prime) = \sum_{A \in \mathcal{P}_m} (m+1) =
645: (m+1)^2.
646: \end{align*}
647: \item Let $a_{\tau \tau^\prime}$ denote the $(\tau,\tau^\prime)$ entry of the matrix $A_m$.  Since $A_m$ is symmetric, we need only show the result for column sums.
648: \begin{align*}
649: \sum_{\tau \in S_{m}} a_{\tau \tau^\prime} & = \sum_{\tau \in S_{m}} \sum_{\tau^{\prime \prime} \in S_{m+1}} b_{\tau \tau^{\prime \prime}} b_{\tau^\prime \tau^{\prime \prime}} = \sum_{\tau^{\prime \prime} \in S_{m+1}} (\sum_{\tau \in S_{m}} b_{\tau \tau^{\prime \prime}}) b_{\tau^\prime \tau^{\prime \prime}} \\
650: & = \sum_{\tau^{\prime \prime} \in S_{m+1}} (m+1) b_{\tau^\prime
651: \tau^{\prime \prime}} = (m+1)^3.
652: \end{align*}
653: where the third equality follows from part (1) and the fourth from
654: part (2).
655: \end{enumerate}
656: \end{proof}
657: 
658: \begin{cor} \label{eigcor} ${\bf \hat{1}}_m$ is an eigenvector of $A_m$ with eigenvalue $(m+1)^3$.
659: \end{cor}
660: 
661: \begin{proof}
662: By Proposition \ref{propsums},
663: \begin{align*}
664: A_m {\bf \hat{1}}_m = (m+1)^3 {\bf \hat{1}}_m.
665: \end{align*}
666: \end{proof}
667: 
668: \begin{prop} $(m+1)^3$ is the largest eigenvalue of $A_m$.
669: \end{prop}
670: 
671: \begin{proof}
672: By Corollary \ref{eigcor} and the Perron-Frobenius Theorem, we
673: need only show that $A_m$ is irreducible.  Consider the weighted
674: bipartite graph $G_m$ on the sets $S_m$ and $S_{m+1}$, where
675: $\sigma \in S_{m}$ is connected by an edge to $\tau \in S_{m+1}$
676: with weight ${\bf X}^\sigma (\tau)$.  (In particular, there is an
677: edge connecting $\sigma$ to $\tau$ iff $\sigma$ occurs in $\tau$.)
678: Then the adjacency matrix of $G_m$ is $B_m$, and the entries of
679: $A_m$ represent sums of weighted length-$2$ paths from $S_{m+1}$
680: to itself.  If $G_m$ is connected, then $A_m$ is irreducible.  To
681: establish connectivity, we show that there is a path from every
682: permutation $\tau \in S_{m+1}$ to the identity element of
683: $S_{m+1}$ in $G_m$.
684: 
685: \begin{claim} For each $k$, $0 \leq k \leq m$, there is a path in $G_m$ from each permutation $\tau \in S_{m+1}$ to some permutation $\tau^\prime \in S_{m+1}$ such that $\tau^\prime (i) = i$ whenever $0 \leq i \leq k$.
686: \end{claim}
687: 
688: \begin{poc}
689: We proceed by induction.  Suppose the claim is true for $k$, and
690: let $\tau$ be any element of $S_{m+1}$.  The inductive hypothesis
691: supplies us with a path from $\tau$ to a $\tau^\prime$ such that
692: $\tau^\prime (i) = i$ whenever $1 \leq i \leq k$.  Let $\sigma$ be
693: the unique permutation in $S_m$ such that $\sigma \sim \tau^\prime
694: |_{\mathbb{Z}_m \setminus \{k+1\}}$.  Then define $\tau^{\prime
695: \prime}$ as follows:
696: $$
697:   \tau^{\prime \prime}(i) = \left\{
698:                \begin{array}{ll}
699:                i & \text{if $0 \leq i \leq k+1$}, \\
700:                \sigma(i-1) + 1 & \text{if $k+1 < i \leq m$}
701:                \end{array}
702:                \right.
703: $$
704: It is easy to check that $\tau^{\prime \prime}$ is actually an
705: element of $S_{m+1}$.  Furthermore, $\tau^\prime|_{\mathbb{Z}_m
706: \setminus \{k+1\}} \sim \sigma \sim \tau^{\prime \prime}
707: |_{\mathbb{Z}_m \setminus \{k+1\}}$, so there is a path in $G_m$
708: from $\tau$ to a permutation which agrees with the identity on
709: $[k+1]$.
710: \end{poc}
711: 
712: Since every vertex of $S_{m}$ is connected to \emph{some} vertex
713: of $S_{m+1}$, and they are all connected to the identity, $G_m$ is
714: connected and $A_m$ is irreducible.
715: \end{proof}
716: 
717: We apply this result and the Courant-Fischer Theorem to Equation
718: (\ref{mainform}):
719: $$
720: (n-m)^2 {| {\bf \tilde{v}}_m(\sigma) |}^2 \leq \lambda_{max}(A_m)
721: | {\bf \tilde{v}}_{m+1}(\sigma) |^2 = (m+1)^3 | {\bf
722: \tilde{v}}_{m+1}(\sigma) |^2
723: $$
724: Thus, we have
725: \begin{cor} \label{down} For any $\sigma \in S_n$,
726: $$
727: | {\bf \tilde{v}}_{m}(\sigma) |^2 \leq \frac{(m+1)^3}{(n-m)^2} {|
728: {\bf \tilde{v}}_{m+1}(\sigma) |^2 }
729: $$
730: \end{cor}
731: 
732: Although Corollary \ref{down} is all we will need to use later, we
733: include here a short proof that $B_m$ has maximal rank.
734: 
735: \begin{definition}
736: For a permutation $\sigma \in S_{m}$, write $\sigma^\circ$ for the
737: permutation in $S_{m+1}$ such that $\sigma^\circ(0) = 0$ and
738: $\sigma^\circ(i) = \sigma(i-1) + 1$ for $1 \leq i \leq m-1$.
739: \end{definition}
740: 
741: Recall that, for two permutations $\sigma$ and $\sigma^\prime$ on
742: the totally ordered sets $S$ and $S^\prime$, respectively, where
743: $S$ and $S^\prime$ have the same cardinality, we may compare
744: $\sigma$ and $\sigma^\prime$ in the lexicographic order by
745: identifying the $i^{th}$ elements of $S$ and $S^\prime$ for each
746: $i$.
747: 
748: \begin{theorem} \label{firstoccur}
749: For a given $\sigma \in S_{m}$ the lexicographically least
750: permutation $\tau \in S_{m+1}$ such that ${\bf X}^{\sigma}(\tau)$
751: is nonzero is $\tau = \sigma^\circ$.
752: \end{theorem}
753: \begin{proof}
754: First, note that ${\bf X}^{\sigma}(\sigma^\circ) > 0$, since
755: $\sigma \sim \sigma^\circ |_S$, where $S = {\mathbb{Z}_{m+1}
756: \setminus \{0\}}$.  To show that ${\bf X}^{\sigma}(\tau) = 0$ for
757: all $\tau < \sigma^\circ$, we proceed by induction.  For $m=1$,
758: the result is trivial: $(0)$ occurs in $(01)$, and in no earlier
759: permutation, since $(01)$ is lexicographically first.  Now suppose
760: the result is true for $m-1$, but there is some $\sigma \in S_{m}$
761: and  $\tau \in S_{m+1}$ such that ${\bf X}^{\sigma}(\tau) \neq 0$
762: and $\tau < \sigma^\circ$.  We know that $\tau(0)$ must equal $0$,
763: since $\sigma^\circ$ does, and $\tau$ precedes $\sigma^\circ$.
764: Therefore, $\tau < \sigma^\circ \Rightarrow \tau |_S <
765: \sigma^\circ |_S$.  Since $\sigma^\circ |_S$ is isomorphic to
766: $\sigma$, $\tau |_S$ is lexicographically precedent to $\sigma$,
767: so $\sigma \not \sim \tau |_S$.  Thus, $\sigma \sim \tau |_T$ for
768: some $T \neq S$, i.e., a $T \in P_m$ which includes $0$.
769: Restricting both of these permutations to all but the first
770: element, we have $\sigma|_U \sim \tau|_{T \cap S} \Rightarrow {\bf
771: X}^{\sigma |_U}(\tau |_S) > 0$.  On the other hand, $\sigma \sim
772: \tau |_T \Rightarrow \sigma(0)=0$, since $0 \in T$, so the fact
773: that $\tau|_S < \sigma^\circ |_S \sim \sigma \sim
774: (\sigma|_U)^\circ$ implies that $\sigma|_U$ occurs in a
775: permutation lexicographically precedent to $(\sigma|_U)^\circ$,
776: contradicting the inductive hypothesis.
777: \end{proof}
778: 
779: \begin{cor} \label{bcor}
780: $\text{rank}(B_m) = m!$.
781: \end{cor}
782: 
783: \begin{proof}
784: If we order the columns and rows of $B_m$ by the lexicographic
785: order on their indices, then Theorem \ref{firstoccur} implies that
786: the first nonzero entry in the $i^{th}$ row occurs in the $i^{th}$
787: column, since $\sigma < \tau$ iff $\sigma^\circ < \tau^\circ$, and
788: the permutations of the form $\sigma^\circ$ precede all others.
789: \end{proof}
790: 
791: \bigskip
792: \section{Quasirandom Permutations}
793: \setcounter{theorem}{0}
794: 
795: In this section, we discuss several equivalent formulations of
796: quasirandom permutations.  The central definition is, roughly,
797: that a quasirandom permutation is one which sends each interval to
798: a highly balanced set.  Thus, we will write $D(\sigma)$ for
799: $\max(D_J(\sigma(I))$, where the maximum is taken over all
800: intervals $I$ and $J$, and a sequence of permutations ${\sigma_j}$
801: will be called {\it quasirandom} if $D(\sigma_j) = o(n)$.  The
802: following is the main result of this section.
803: 
804: \begin{theorem} \label{qrp} For any sequence of permutations $\sigma \in S_n$ and integer $m \geq 2$ with $n > m$, the following are equivalent: \\ [.06in]
805: \noindent \begin{tabular}{p{.4in}p{4.7in}}
806: {\bf [UB]} & (Uniform Balance) $D(\sigma) = o(n)$. \\
807: \end{tabular}
808: 
809: \noindent \begin{tabular}{p{.4in}p{4.7in}} {\bf [SP]} &
810: (Separability) For any intervals $I,J,K,K^\prime \subset
811: \mathbb{Z}_n$,
812: $$
813: \left | \sum_{x \in K \cap \sigma^{-1}(K^\prime)} I(x)
814: J(\sigma(x)) - \frac{1}{n} \sum_{x \in K,y \in K^\prime} I(x)J(y)
815: \right | = o(n)
816: $$ \\
817: \end{tabular}
818: 
819: \noindent \begin{tabular}{p{.4in}p{4.7in}} {\bf [mS]} &
820: (m-Subsequences) For any permutation $\tau \in S_m$ and intervals
821: $I,J \subset \mathbb{Z}_n$,
822: $$
823: {\bf X}^\tau(\sigma |_{I \cap \sigma^{-1}(J)}) = \frac{1}{m!}
824: \binom{|\sigma(I) \cap J|}{m} + o(n^m).
825: $$ \\
826: \end{tabular}
827: 
828: \noindent \begin{tabular}{p{.4in}p{4.7in}} {\bf [2S]} &
829: (2-Subsequences) For any intervals $I,J \subset \mathbb{Z}_n$,
830: $$
831: {\bf X}^{(01)}(\sigma |_{I \cap \sigma^{-1}(J)}) - {\bf
832: X}^{(10)}(\sigma |_{I \cap \sigma^{-1}(J)}) = o(n^2).
833: $$ \\
834: \end{tabular}
835: 
836: \end{theorem}
837: 
838: It follows immediately that these conditions are also equivalent
839: to each interpretation of the statement ``For all intervals $J
840: \subset \mathbb{Z}_n$, $\sigma(J)$ is $\epsilon$-balanced'' given
841: by the equivalences of Theorem \ref{balance}.  Thus, we have a
842: total of ten equivalent quasirandom properties: seven arising as
843: ``uniformly convergent'' versions of the properties in Theorem
844: \ref{balance} and three new ones, which are included with uniform
845: balance in Figure 2.
846: 
847: \ifx\ispreprint\undefined
848: \else
849: \begin{center}
850: \begin{figure}[ht]
851: \hspace{1.25in} \input{figb.pstex_t} \caption{Diagram of
852: implications for Theorem \ref{qrp}.}
853: \end{figure}
854: \end{center}
855: \fi
856: 
857: Again, we prove the theorem piece by piece, keeping track of
858: $\epsilon$ as we go.  The next result states that, if uniform
859: balance is obeyed, then the variable $x$ and its image under
860: $\sigma$ are nearly independent.
861: 
862: \begin{prop} {\bf [UB]} $\Leftrightarrow$ {\bf [SP]}.
863: \end{prop}
864: \begin{proof}
865: {\bf [UB]} holds iff, for all intervals $I,J,K,K^\prime \subset
866: \mathbb{Z}_n$,
867: $$
868: \left | |\sigma(I \cap K) \cap (J \cap K^\prime)| - \frac{1}{n} |I
869: \cap K| |J \cap K^\prime| \right | < \epsilon n.
870: $$
871: But this quantity is equal to
872: $$
873: \left | \sum_{x \in K \cap \sigma^{-1}(K^\prime)} I(x)
874: J(\sigma(x)) - \frac{1}{n} \sum_{x \in K,y \in K^\prime} I(x) J(y)
875: \right |
876: $$
877: so that {\bf [UB]} is equivalent to {\bf [SP]}.
878: \end{proof}
879: 
880: Now, we show that the separability achieved in the last
881: proposition is sufficient to imply that subsequences happen at the
882: ``right'' rate (i.e., what one would expect of truly random
883: permutations) on certain sets of indices.  A computational lemma
884: will greatly simplify the proof.
885: 
886: \begin{lemma} \label{prodest} If, for each $j$ with $1 \leq j \leq k$, $|a_j|<n^{c_j}$, $\epsilon_j > n^{-1}$, and
887: $$
888: |x_j - a_j| < \epsilon_j n^{c_j-1}
889: $$
890: then
891: $$
892: \left | \prod_{j=1}^k x_j - \prod_{j=1}^k a_j \right | < 3^{k-1}
893: n^{\sum_{j=1}^k c_j - k} \prod_{j=1}^k \epsilon_j.
894: $$
895: \end{lemma}
896: \begin{proof}
897: We show the result for $k=2$, and the general case follows by a
898: simple induction.  Thus,
899: \begin{align*}
900: \left | x_1 x_2 - a_1 a_2 \right | & \leq \left | x_1 - a_1 \right | \cdot \left | x_2 - a_2 \right | + a_1  \left | x_2 - a_2 \right | + a_2 \left | x_1 - a_1 \right | \\
901: &< \epsilon_1 \epsilon_2 n^{c_1 + c_2 - 2} + \epsilon_1 n^{c_1 + c_2 - 1} + \epsilon_2 n^{c_1 + c_2 - 1} \\
902: &< 3 \epsilon_1 \epsilon_2 n^{c_1 + c_2 - 2}.
903: \end{align*}
904: \end{proof}
905: 
906: \begin{prop} {\bf [SP]} $\Rightarrow$ {\bf [mS]}.
907: \end{prop}
908: \begin{proof}
909: Let $I,J$ be intervals, and let $K = {I \cap \sigma^{-1}(J)}$.
910: Note that we may write the number of ``occurrences'' of $\tau \in
911: S_m$ in $\sigma |_K$ as
912: $$
913: {\bf X}^\tau(\sigma |_K) = \sum_{x_1,\cdots,x_m \in K}
914: \prod_{i=0}^m \left ( \chi(x_i < x_{i+1})
915: \chi(\sigma(x_{\tau^{-1}(j)}) < \sigma(x_{\tau^{-1}(j+1)})) \right
916: )
917: $$
918: In the interest of notational compactness, we will denote
919: $\chi(x<y)$ by $\cl{x}{y}$, and define $\cl{x}{y} = 1$ if either
920: $x$ or $y$ is undefined.  Furthermore, for any subset $A \subset
921: [m]$, we will denote the following expression
922: $$
923: \sum_{\{x_i\}_{i \not \in A} \subset K} \, \sum_{\{x_k\}_{k \in A}
924: \subset I} \sum_{\{x^{\prime}_k\}_{k \in A} \subset J} \left (
925: \prod_{j=0}^m \cl{x_j}{x_{j+1}}
926: \cl{x^A_{\tau^{-1}(j)}}{x^A_{\tau^{-1}(j+1)}} \right )
927: $$
928: by $\Sigma(A)$, where $x^A_j$ means $\sigma(x_j)$ for $j \not \in
929: A$, and $x^{\prime}_k$ for $j \in A$.  Thus, ${\bf X}^\tau(\sigma
930: |_K) = \Sigma(\emptyset)$.  The proof will now proceed by
931: induction on the subsets of $[m]$, ordered by inclusion.
932: 
933: Suppose $A \subset B \subset [m]$, with $B \setminus A = \{s\}$,
934: and assume that
935: \begin{equation} \label{indstep}
936: \left | {\bf X}^\tau(\sigma |_K) - n^{-|A|} \Sigma(A) \right | <
937: |A| \epsilon n^m
938: \end{equation}
939: By {\bf [SP]}, we know that, for any $a,b,c,d \in \mathbb{Z}_n$,
940: the quantity
941: $$
942: \left | \sum_{x_s \in K} \cl{a}{x_s} \cl{x_s}{b}
943: \cl{c}{\sigma(x_s)} \cl{\sigma(x_s)}{d} - \frac{1}{n} \sum_{x_s
944: \in I, x_s^\prime \in J} \cl{a}{x_s} \cl{x_s}{b}
945: \cl{c}{x^\prime_s} \cl{x^\prime_s}{d} \right |
946: $$
947: is bounded above by $\epsilon n$.  Then, substituting $a=x_{s-1}$,
948: $b=x_{s+1}$, $c=x^A_{\tau^{-1}(\tau(s)-1)}$, and
949: $d=x^A_{\tau^{-1}(\tau(s)+1)}$ to account for all the terms
950: containing $x_s$ in the product portion of the expression
951: $\Sigma(A)$, we have (after a very messy but otherwise
952: straightforward calculation),
953: $$
954: \left | \Sigma(A) - n^{-1} \Sigma(B) \right | < \epsilon n
955: |K|^{m-|B|} |I|^{|A|} |J|^{|A|} \leq \epsilon n^{m + |A|}
956: $$
957: Applying this to the inductive hypothesis with the aid of the
958: triangle inequality yields
959: \begin{align*}
960: \left | {\bf X}^\tau(\sigma |_K) - n^{-|B|} \Sigma(B) \right | & \leq \left | {\bf X}^\tau(\sigma |_K) - n^{-|A|} \Sigma(A) \right | + n^{-|A|} \left | \Sigma(A) - n^{-1} \Sigma(B) \right | \\
961: & < |A| \epsilon n^m + n^{-|A|} \epsilon n^{m+|A|} = |B| \epsilon
962: n^m.
963: \end{align*}
964: Therefore, (\ref{indstep}) is true for all $A \subset [m]$.  In
965: particular, it is true for $A = [m]$, so that
966: $$
967: \left | {\bf X}^\tau(\sigma |_K) - n^{-m} \Sigma([m]) \right | < m
968: \epsilon n^m
969: $$
970: Since we have
971: \begin{align}
972: \nonumber \left | \sum_{\{x_j\} \subset I} \prod_{j=0}^m \cl{x_j}{x_{j+1}} - \frac{|I|^m}{m!} \right | & = \left | \binom{|I|}{m} - \frac{|I|^m}{m!} \right | \leq \frac{(|I|+m)^m - |I|^m}{m!} \\
973: \nonumber & = \frac{1}{m!} \sum_{k=1}^{m} \binom{m}{k} |I|^{m-k} m^k \leq \frac{|I|^{m-1}}{m!} \sum_{k=0}^{m} \binom{m}{k} m^k \\
974: & = \frac{(1+m)^m}{m!} |I|^{m-1} \label{approxbin}
975: \end{align}
976: and also
977: $$
978: \left | \sum_{\{x_j\} \subset J} \prod_{j=0}^m \cl{x_j}{x_{j+1}} -
979: \frac{|J|^m}{m!} \right | \leq \frac{(1+m)^m}{m!} |J|^{m-1}
980: $$
981: we may conclude that
982: $$
983: \left | \Sigma([m]) - \frac{|I|^m|J|^m}{m!^2} \right | < 3
984: n^{2m-1} \frac{(1+m)^{2m}}{m!^2}
985: $$
986: by Lemma \ref{prodest}.  Thus,
987: \begin{align*}
988: \left | {\bf X}^\tau(\sigma |_K) - \frac{1}{m!^2} \left ( \frac{|I||J|}{n} \right )^m \right | & \leq \left | {\bf X}^\tau(\sigma |_K) - \frac{1}{n^m} \Sigma([m]) \right | \\
989: & + \frac{1}{n^m} \left | \Sigma([m]) - \frac{|I|^m|J|^m}{m!^2} \right | \\
990: & < m \epsilon n^m + 3 n^{m-1} \frac{(1+m)^{2m}}{m!^2} \\
991: & < 4 \epsilon n^m \frac{(1+m)^{2m}}{m!^2}
992: \end{align*}
993: But, by {\bf [UB]} (which is equivalent to {\bf [SP]}), and Lemma
994: \ref{prodest}
995: $$
996: \left | \left ( \frac{|I||J|}{n} \right )^m - |\sigma(I) \cap J|^m \right | < 3^{m-1} \epsilon^m n^m \\
997: $$
998: Since $m \geq 2$, this gives
999: \begin{align*}
1000: \left | {\bf X}^\tau(\sigma |_K) - \frac{1}{m!^2} |\sigma(I) \cap J|^m \right | &< n^m \frac{1}{m!^2} \left (3^{m-1} \epsilon^m + 4 \epsilon (1+m)^{2m} \right ) \\
1001: & < n^m \frac{1}{m!^2} \left ((1+m)^{2m} \epsilon + 4 \epsilon (1+m)^{2m} \right) \\
1002: & = \frac{5 \epsilon (1+m)^{2m}}{m!^2} n^m
1003: \end{align*}
1004: Finally, the fact that $\epsilon \geq n^{-1}$ implies, as in
1005: (\ref{approxbin}),
1006: $$
1007: \left | \binom{|\sigma(I) \cap J|}{m} - \frac{|\sigma(I) \cap
1008: J|^m}{m!} \right | \leq \epsilon \frac{(1+m)^m}{m!} n^m
1009: $$
1010: so we may conclude
1011: $$
1012: \left | {\bf X}^\tau(\sigma |_K) - \frac{1}{m!} \binom{|\sigma(I)
1013: \cap J|}{m} \right | < \frac{6 \epsilon (1+m)^{2m}}{m!^2} n^m < 4
1014: e^{2m} \epsilon n^m
1015: $$
1016: where we have used the Stirling approximation $m! > \sqrt{2 \pi m}
1017: \left (\frac{m}{e} \right)^m$.
1018: \end{proof}
1019: 
1020: Now, we use the results of the previous section to show that {\bf
1021: [mS]} implies {\bf [2S]}.
1022: 
1023: \begin{prop} {\bf [mS]} $\Rightarrow$ {\bf [2S]}.
1024: \end{prop}
1025: \begin{proof}
1026: Let $K=\sigma(I) \cap J$ for some intervals $I,J \in
1027: \mathbb{Z}_n$.  We may assume that $n \geq 2m$, so that $n-k>n/2$
1028: for all $k<m$.  Therefore, by Corollary \ref{down}, we may write
1029: $$
1030: | {\bf \tilde{v}}_{m-1}(\sigma|_K) |^2 \leq \frac{m^3}{(n-m+1)^2}
1031: {| {\bf \tilde{v}}_{m}(\sigma|_K) | } < \frac{4 m^3}{n^2} m!
1032: \max_{\tau \in S_m} \left | X^\tau(\sigma|_K) - \binom{|K|}{m}
1033: \right |^2
1034: $$
1035: Iterating this process $m-2$ times, we find
1036: \begin{align*}
1037: | {\bf \tilde{v}}_{m-1}(\sigma) |^2 &< \frac{4^{m-5} m!^{m+1}}{n^{2m-4}} \max_{\tau \in S_m} \left | X^\tau(\sigma|_K) - \binom{|K|}{m} \right | \\
1038: & < \frac{4^{m-2} m!^{m+1}}{n^{2m-4}} \epsilon^2 n^{2m} = 4^{m-5}
1039: m!^{m+1} \epsilon^2 n^4.
1040: \end{align*}
1041: Let the quantity $d$ be defined by
1042: $$
1043: d = \left | X^{(01)}(\sigma|_K) - \binom{|K|}{2} \right | = \left
1044: | X^{(10)}(\sigma|_K) - \binom{|K|}{2} \right |.
1045: $$
1046: Then $| {\bf \tilde{v}}_{2}(\sigma) |^2 = 2d^2$, so $d < 2^{m -
1047: 11/2} m!^{(m+1)/2} \epsilon n^2$, and
1048: $$
1049: \left | X^{(01)}(\sigma|_K) - X^{(10)}(\sigma|_K) \right | < 2d <
1050: 2^m m!^{(m+1)/2} \epsilon n^2
1051: $$
1052: which implies {\bf [2S]}.
1053: \end{proof}
1054: 
1055: In what follows, we denote the complement of a set $S \in
1056: \mathbb{Z}_n$ by $\bar{S}$, and we denote by $S^*$ its projection
1057: onto $[0,n-1]$.  Also, call an interval $I \subset \mathbb{Z}_n$
1058: ``contiguous'' if $I^*$ is an interval, ``terminal'' if $\bar{I}$
1059: is contiguous, ``initial'' if it is terminal and contains $0$, and
1060: ``final'' if it is terminal and contains $n-1$.
1061: 
1062: \begin{prop} {\bf [2S]} $\Rightarrow$ {\bf [UB]}.
1063: \end{prop}
1064: \begin{proof}
1065: Suppose $\sigma$ satisfies {\bf [2S]} but not {\bf [UB]}. We claim
1066: that, for infinitely many $n$ and some $\epsilon > 0$, there are
1067: intervals $I, J \subset \mathbb{Z}_n$ with $I$ and $J$ initial,
1068: and $D_J(\sigma(I))$ at least $27\epsilon n/2$.  Since {\bf [UB]}
1069: is not true for $\sigma$, we may choose $\epsilon$ so that there
1070: are proper subintervals $I,J \subset \mathbb{Z}_n$ with
1071: $D_{J}(\sigma(I)) \geq 54 \epsilon n$.  Suppose $J$ is not
1072: contiguous.  Then $\bar{J}$ is contiguous, and, since
1073: $$
1074: \left ( \left | \sigma(I) \cap J  \, \right | - \frac{|I||J|}{n}
1075: \right ) + \left ( \left | \sigma(I) \cap \bar{J} \, \right | -
1076: \frac{|I||\bar{J}|}{n} \right ) = \left | \sigma(I) \right | -
1077: \frac{|I|n}{n} = 0
1078: $$
1079: we may replace $J$ with $\bar{J}$ and retain the property that
1080: $D_{J}(\sigma(I)) \geq 54 \epsilon n$.  Now, suppose $J$ is not
1081: terminal.  Let $J^\prime$ be a component of $(\bar{J})^*$.
1082: $J^\prime$ is terminal because $J$ is contiguous, and, since we
1083: have
1084: $$
1085: \left | \sigma(I) \cap (J \cup J^\prime)  \, \right | -
1086: \frac{|I||J \cup J^\prime |}{n} = \left ( \left | \sigma(I) \cap J
1087: \, \right | - \frac{|I||J|}{n} \right ) + \left ( \left |
1088: \sigma(I) \cap J^\prime \, \right | - \frac{|I||J^\prime|}{n}
1089: \right )
1090: $$
1091: either $D_{J \cup J^\prime}(\sigma(I)) \geq 27 \epsilon n$ or
1092: $D_{J^\prime}(\sigma(I)) \geq 27 \epsilon n$.  Thus, we may assume
1093: that $J$ is terminal (since $J^\prime$ and $J \cup J^\prime$ are),
1094: and $D_{J}(\sigma(I)) \geq 27 \epsilon n$.  If $J$ is final,
1095: taking its complement makes it initial without disturbing the
1096: discrepancy.  Apply the same process to $I$ to ensure that it is
1097: initial, with the penalty that now
1098: \begin{equation} \label{tst}
1099: D_{J}(\sigma(I)) \geq 27 \epsilon n/2.
1100: \end{equation}
1101: 
1102: For ease of notation, we will let
1103: \begin{eqnarray*}
1104: A & = I \cap \sigma^{-1}(J) & \hspace{.3in} a=|A| \\
1105: B & = I \cap \sigma^{-1}(\bar{J}) & \hspace{.3in} b=|B| \\
1106: C & = \bar{I} \cap \sigma^{-1}(J) & \hspace{.3in} c=|C| \\
1107: D & = \bar{I} \cap \sigma^{-1}(\bar{J}) & \hspace{.3in} d=|D|
1108: \end{eqnarray*}
1109: For subsets $S,T \subset \mathbb{Z}_n$, let $\partial_\sigma
1110: (S,T)$ denote the number of pairs $(x,y) \in S \times T$ such that
1111: $x < y$ and $\sigma(x) < \sigma(y)$.  Then
1112: \begin{align*}
1113: {\bf X}^{(01)}(\sigma) & = {\bf X}^{(01)}(\sigma|_I) + {\bf X}^{(01)}(\sigma|_{\bar{I}}) + \partial_\sigma(I, \bar{I}) \\
1114: & = {\bf X}^{(01)}(\sigma|_I) + {\bf X}^{(01)}(\sigma|_{\bar{I}}) + \partial_\sigma(A, C) \\
1115: & \, \, \, \, \, + \partial_\sigma(A, D) + \partial_\sigma(B, C) +
1116: \partial_\sigma(B, D)
1117: \end{align*}
1118: Now, $\partial_\sigma(B, C) = 0$ and $\partial_\sigma(A, D) = ad$,
1119: since every element of $J$ is less than every element of
1120: $\bar{J}$, and every element of $I$ is less than every element of
1121: $\bar{I}$.  Also,
1122: \begin{align*}
1123: \partial_\sigma(A, C) & = {\bf X}^{(01)}(\sigma|_{\sigma^{-1}(J)}) - {\bf X}^{(01)}(\sigma|_A) - {\bf X}^{(01)}(\sigma|_C) \\
1124: \partial_\sigma(B, D) & = {\bf X}^{(01)}(\sigma|_{\sigma^{-1}(\bar{J})}) - {\bf X}^{(01)}(\sigma|_B) - {\bf X}^{(01)}(\sigma|_D)
1125: \end{align*}
1126: Thus, we have
1127: \begin{align*}
1128: {\bf X}^{(01)}(\sigma) & = {\bf X}^{(01)}(\sigma|_I) + {\bf X}^{(01)}(\sigma|_{\bar{I}}) + {\bf X}^{(01)}(\sigma|_{\sigma^{-1}(J)}) - {\bf X}^{(01)}(\sigma|_A) \\
1129: & - {\bf X}^{(01)}(\sigma|_C) + ad + {\bf
1130: X}^{(01)}(\sigma|_{\sigma^{-1}(\bar{J})}) - {\bf
1131: X}^{(01)}(\sigma|_B) - {\bf X}^{(01)}(\sigma|_D)
1132: \end{align*}
1133: By {\bf [2S]}, for sufficiently large $n$, each term ${\bf
1134: X}^{(01)}(\sigma|_S)$ can be approximated by $\binom{|S|}{2}/2$ to
1135: within $\epsilon n^2/2$, and therefore by $|S|^2/4$ to within $3
1136: \epsilon n^2/4$, since
1137: $$
1138: \left | \binom{|S|}{2} - \frac{|S|^2}{2} \right | = \frac{|S|}{2}
1139: < \frac{\epsilon}{2} n^2.
1140: $$
1141: Therefore, rewriting and multiplying by $4$, we have that
1142: $$
1143: \left | n^2 - (a+b)^2 - (c+d)^2 - (a+c)^2 - (a+d)^2 + a^2 + b^2 +
1144: c^2 + d^2 - 4ad \right |
1145: $$
1146: is bounded above by $27 \epsilon n^2$.  Since $n = a+b+c+d$, we
1147: may simplify down to
1148: \begin{equation} \label{eq1}
1149: | bc - ad | < \frac{27 \epsilon}{2} n^2
1150: \end{equation}
1151: Let $\delta n = \left | I \cap \sigma^{-1}(J) \right | -
1152: |I||J|/n$.  Then, by (\ref{tst}),
1153: \begin{align*}
1154: | bc - ad | &= \left | \left ( \frac{|I||\bar{J}|}{n} - \delta n \right ) \left ( \frac{|\bar{I}||J|}{n} - \delta n \right ) - \left ( \frac{|I||J|}{n} + \delta n \right ) \left ( \frac{|\bar{I}||\bar{J}|}{n} + \delta n \right ) \right | \\
1155: &= |\delta| (|I||J| + |I||\bar{J}| + |\bar{I}||J| + |\bar{I}||\bar{J}|) \\
1156: &= \frac{D_J(\sigma(I))}{n} \cdot (|I|+|\bar{I}|)(|J|+|\bar{J}|)
1157: \geq \frac{27 \epsilon}{2} n^2
1158: \end{align*}
1159: contradicting (\ref{eq1}).
1160: \end{proof}
1161: 
1162: We present two simple applications of these results.  The following observation has some relevance to the investigations
1163: of [\ref{AF1}] and [\ref{B1}].
1164: 
1165: \begin{prop} If a permutation $\sigma \in S_n$ excludes $\tau \in
1166: S_m$ (in the sense that ${\bf X}^\tau(\sigma) = 0$), then
1167: $$
1168: D(\sigma) \geq \frac{n\binom{n}{m}}{4 \, e^{2m} \,m!\, n^m}.
1169: $$
1170: \end{prop}
1171: \begin{proof} Let $\epsilon = D(\sigma)/n$.  We show that, if
1172: $$
1173: \epsilon \leq \frac{\binom{n}{m}}{4 \, e^{2m} \, m! \, n^m}
1174: $$
1175: then there is at least one copy of every element $\tau$ of $S_m$
1176: in $\sigma$.  According to the implication from {\bf [UB]} to {\bf
1177: [mS]}, if $D(\sigma) < \epsilon n$, then
1178: $$
1179: \left | \textbf{X}^\tau(\sigma) - \frac{1}{m!} \binom{n}{m} \right
1180: | < 4 e^{2m} \epsilon n^m \leq \frac{1}{m!} \binom{n}{m},
1181: $$
1182: so that $\textbf{X}^\tau(\sigma) > 0$.
1183: \end{proof}
1184: 
1185: \begin{cor} There is a constant $c > 0$ so that, if $n \geq 2m$ and
1186: $\sigma \in S_n$ excludes $\tau \in S_m$, then
1187: $$
1188: \frac{D(\sigma)}{n} > \left(\frac{c}{m^2} \right)^m.
1189: $$
1190: \end{cor}
1191: 
1192: We can also use Theorem \ref{qrp} to calculate the discrepancy of a random permutation.
1193: 
1194: \begin{theorem} If a permutation $\sigma$ is chosen randomly and uniformly from $S_n$, then $D(\sigma) = O(\sqrt{n \log n})$ almost surely.
1195: \end{theorem}
1196: \begin{proof} We use $c_i$, $i \in \mathbb{N}$ to denote absolute constants throughout.  Let the random variable $\xi_n$ be the number of inversions in a randomly and uniformly chosen element of $S_n$.  Define $\eta_n$ to be the normalized random variable given by
1197: $$
1198: \eta_n = \frac{\xi_n - \mbox{\textbf{E}}(\xi_n)}{(\mbox{Var}\, \xi_n)^{1/2}} = \frac{\xi_n - {\frac{1}{2} \binom{n}{2}}}{(\mbox{Var}\, \xi_n)^{1/2}},
1199: $$
1200: and let $u_n$ denote the distribution function with an atom of mass $1/n$ at each of $0, \ldots , n-1$.  It is well known ([\ref{St1}],[\ref{St2}]) that the generating function $g_n(q)$ for the number of permutations with a given number of inversions is the $q$-factorial $[n]!$, and that its coefficients are symmetric and unimodal.  Therefore, $g_n = g_{n-1} \cdot (q^{n-1} + \ldots + 1)$, so that $\xi_n = \xi_{n-1} \ast u_n$ (the convolution product).  The unimodality of the coefficients of $g_{n-1}$ implies the concavity of the cumulative distribution function $F_n(x)$ of $\xi_n$ on the interval $[\lceil \binom{n}{2}/2 \rceil,\infty]$.  Thus, if $x \geq \binom{n-1}{2}/2+n-1$, then $F_n(x) = (F_{n-1} \ast u_n)(x) \leq F_{n-1}(x)$.  (The convolution is a finite sum, so the implicit change of order of summation is legitimate.)  Similarly, $F_n(x) = (F_{n-1} \ast u_n)(x) \geq F_{n-1}(x)$ if $x \leq \binom{n-1}{2}/2$.  We have, then,
1201: \begin{equation} \label{monotone}
1202: \mbox{\textbf{Pr}}( \left|\xi_n - \mbox{\textbf{E}}(\xi_n)\right| > \lambda) \geq \mbox{\textbf{Pr}}( \left|\xi_{m} - \mbox{\textbf{E}}(\xi_m)\right| > \lambda )
1203: \end{equation}
1204: whenever $\lambda \geq (n-1)/2$ and $m \leq n$.
1205: 
1206: 
1207: It is a theorem of Sachkov [\ref{Sa1}] that the cumulative distribution function of $\eta_n$ converges to $\Phi(0,1)$ (the c.d.f. of the standard normal distribution), and that $\sigma_n^2 = \mbox{Var}\, \xi_n = n^3/36 + O(n^2)$.  In particular, the moment generating function $M(t,n)=\mbox{\textbf{E}}(e^{t \eta_n})$ of $\eta_n$ is given by
1208: $$
1209: \log M(t,n) = \frac{t^2}{2} + \sum_{k=2}^\infty \frac{B_{2k} \, t^{2k}}{2k \, \sigma_n^{2k} \, (2k)!} \sum_{j=1}^n (j^{2k}-1)
1210: $$
1211: where $B_{2k}$ is the $2k^{\mbox{th}}$ Bernoulli number.  Let $f_n(t) = \log M(t,n) - t^2/2$.  Then, using the approximation $|B_{2k}| < 4 (2k)!/(2 \pi)^{2k}$, we have
1212: \begin{equation} \label{momest}
1213: |f_n(t)| \leq c_1 \sum_{k=2}^\infty \frac{t^{2k}}{(2 \pi)^{2k} \, 2k \, n^{3k}} \cdot n^{2k+1} \leq c_2 n \sum_{k=2}^\infty \left ( \frac{t}{2 \pi n^{1/2}} \right )^{2k} \leq \frac{c_3 t^4}{4 \pi^2 n - t^2},
1214: \end{equation}
1215: so long as $t < n^{1/2}$.
1216: 
1217: By Theorem \ref{qrp}, there exists an $\alpha>0$ so that
1218: $$
1219: \mbox{\textbf{Pr}}(D(\sigma) > \lambda \sqrt{n \log n}) \leq \mbox{\textbf{Pr}}(\max_{I,J} \left|{\bf X}^{(10)}(\sigma|_{I \cap \sigma^{-1}(J)}) - \frac{1}{2} \binom{|I \cap \sigma^{-1}(J)|}{2}\right| > c_4 \lambda n^{3/2} \sqrt{\log n}).
1220: $$
1221: By (\ref{monotone}), we may write
1222: \begin{align*}
1223: \mbox{\textbf{Pr}}(D(\sigma) > \lambda \sqrt{n \log n}) & \leq \sum_{I,J} \mbox{\textbf{Pr}}(\left|\xi_{|\sigma(I) \cap J|} - \mbox{\textbf{E}} \xi_{|\sigma(I) \cap J|}\right| > c_4 \lambda n^{3/2} \sqrt{\log n}) \\
1224: & \leq n^4 \mbox{\textbf{Pr}}(\left|\xi_n - \mbox{\textbf{E}} \xi_n \right| > c_4 \lambda n^{3/2} \sqrt{\log n})
1225: \end{align*}
1226: so long as $n$ is sufficiently large.  Furthermore, by Markov's inequality and the estimate on $\sigma_n$, for fixed $\lambda > 0$,
1227: \begin{align*}
1228: \mbox{\textbf{Pr}}(\left|\xi_n - \mbox{\textbf{E}} \xi_n \right| > c_4 \lambda n^{3/2} \sqrt{\log n}) & \leq \mbox{\textbf{Pr}}(\left|\eta_n\right| > c_5 \lambda \sqrt{\log n}) \\
1229: &= 2 \mbox{\textbf{Pr}}(e^{\eta_n} > e^{c_5 \lambda \sqrt{\log n}} ) \\
1230: &\leq 2 \mbox{\textbf{E}}(e^{t \eta_n}) e^{-t c_5 \lambda \sqrt{\log n}}.
1231: \end{align*}
1232: Setting $t = c_5 \lambda \sqrt{\log n}$ and applying the bound (\ref{momest}), we have
1233: $$
1234: \mbox{\textbf{Pr}}(\left|\xi_n - \mbox{\textbf{E}} \xi_n \right| > c_4 \lambda n^{3/2} \sqrt{\log n}) \leq e^{- c_6 \lambda^2 \log n + c_7 \log^2 n/n} \leq c_8 e^{- c_6 \lambda^2 \log n}.
1235: $$
1236: Therefore,
1237: $$
1238: \mbox{\textbf{Pr}}(D(\sigma) > \lambda \sqrt{n \log n}) \leq c_8 n^{4 - c_6 \lambda^2},
1239: $$
1240: which tends to zero if we choose $\lambda > 2 c_6^{-1/2}$.
1241: \end{proof}
1242: 
1243: 
1244: \bigskip
1245: \section{Constructions}
1246: \setcounter{theorem}{0}
1247: 
1248: In this section, we present a construction for a large class of permutations which are highly quasirandom.  We will assume throughout that $\sigma \in S_n$ and $\tau \in S_m$, unless indicated otherwise.
1249: 
1250: \begin{definition} For permutations $\sigma \in S_n$ and $\tau \in S_m$, considered as actions on $\mathbb{Z}_n$ and $\mathbb{Z}_m$, respectively, define $\sigma \otimes \tau \in S_{nm}$ by $(\sigma \otimes \tau)(x) = \tau(\lfloor \frac{x}{n} \rfloor) + m \sigma(x \! \! \mod n)$.  We will also denote the $k^{th}$ product of $\sigma$ with itself as $\sigma^{(k)}$.
1251: \end{definition}
1252: 
1253: A special case of this product appears in [\ref{D2}], where the authors define a sequence of permutations lacking ``monotone $3$-term arithmetic progressions'' by taking iterated products of the elements of $S_2$.
1254: 
1255: Note that $\sigma \otimes \tau$ has the property that $(\sigma
1256: \otimes \tau)([0,n-1])$ is the set of all elements of
1257: $\mathbb{Z}_{nm}$ congruent to $0$ mod $m$ (i.e., $m \cdot
1258: [0,n-1]$), a set which necessarily lacks the ``weak translation''
1259: property of quasirandom sets.  Thus, a sequence $\{
1260: \sigma_1,\sigma_1 \otimes \sigma_2,\sigma_1 \otimes \sigma_2
1261: \otimes \sigma_3,\ldots \}$ sends intervals to sets which are not
1262: quasirandom in the sense of [\ref{CG2}].  Nonetheless, we will
1263: prove shortly that it does satisfy ${\bf UB}$.  First, we
1264: offer a justification for the lack of parentheses in the
1265: expression for this sequence.
1266: 
1267: \begin{prop} $\otimes$ is associative.
1268: \end{prop}
1269: \begin{proof}
1270: Suppose $\sigma_i \in S_{n_i}$ for $i=1,2,3$.  Then, applying the
1271: definition of $\otimes$ twice,
1272: \begin{align*}
1273: \left [(\sigma_1 \otimes \sigma_2) \otimes \sigma_3 \right] (x) & = \sigma_3 \left( \left \lfloor \frac{x}{n_1n_2} \right \rfloor \right ) + n_3 \sigma_2 \left ( \left \lfloor \frac{x \mbox{ mod } n_1n_2}{n_1} \right \rfloor \right ) \\
1274: & \, \, \, \, \, + n_2n_3 \sigma_1(x \mbox{ mod } n_1)
1275: \end{align*}
1276: and
1277: \begin{align*}
1278: \left [\sigma_1 \otimes (\sigma_2 \otimes \sigma_3) \right ](x) & = \sigma_3 \left( \left \lfloor \frac{\left \lfloor x/n_1 \right \rfloor}{n_2} \right \rfloor \right ) + n_3 \sigma_2 \left ( \left \lfloor \frac{x}{n_1} \right \rfloor \mbox{ mod } n_2 \right ) \\
1279: & \, \, \, \, \, + n_2n_3 \sigma_1(x \mbox{ mod } n_1)
1280: \end{align*}
1281: Note that every element $x$ of $\mathbb{Z}_{n_1n_2n_3}$ can be
1282: represented uniquely as $x = an_1n_2 + bn_1 + c$, with $0 \leq a <
1283: n_3$, $0 \leq b < n_2$, and $0 \leq c < n_1$.  Using this
1284: notation, we find that
1285: \begin{align*}
1286: \left [(\sigma_1 \otimes \sigma_2) \otimes \sigma_3 \right] (x) & = \sigma_3(a) + n_3 \sigma_2 \left ( \left \lfloor \frac{bn_1 + c}{n_1} \right \rfloor \right ) + n_2n_3 \sigma_1(c) \\
1287: & = \sigma_3(a) + n_3 \sigma_2(b) + n_2n_3 \sigma_1(c)
1288: \end{align*}
1289: and
1290: \begin{align*}
1291: \left [\sigma_1 \otimes (\sigma_2 \otimes \sigma_3) \right ](x) = \,\, & \sigma_3 \left( \left \lfloor \frac{an_2 + b}{n_2} \right \rfloor \right ) + n_3 \sigma_2 \left ( (a n_2 + b) \mbox{ mod } n_2 \right ) \\
1292: & \,\, + n_2n_3 \sigma_1(c) \\
1293: = & \,\, \sigma_3(a) + n_3 \sigma_2(b) + n_2n_3 \sigma_1(c).
1294: \end{align*}
1295: \end{proof}
1296: 
1297: Define $d(\sigma)$ by
1298: $$
1299: d(\sigma) = \max_{I,J} D_J(\sigma(I))
1300: $$
1301: where $I$ is allowed to vary over all possible intervals, but $J$
1302: is restricted to initial intervals.  We denote the analogue for
1303: final intervals by $d^\prime$.  Then we have the following result:
1304: 
1305: \begin{prop} $d(\sigma \otimes \tau) \leq m - 1 + d(\sigma)$.
1306: \end{prop}
1307: \begin{proof}
1308: Let the interval $I_k=[kn,(k+1)n-1] \subset \mathbb{Z}_{nm}$.
1309: Then, any initial interval $S$ of $\mathbb{Z}_{nm}$ can, for some
1310: $l<m$, be written
1311: $$
1312: S = \bigcup_{k=0}^l I_k \cup S_0
1313: $$
1314: where $S_0$ is an initial segment of $I_{l+1}$.  For any interval
1315: $J \subset \mathbb{Z}_{nm}$, then, we may write
1316: $$
1317: D_J((\sigma \otimes \tau)(S)) \leq \sum_{k=0}^l D_J((\sigma
1318: \otimes \tau)(I_k)) + D_J((\sigma \otimes \tau)(S_0))
1319: $$
1320: by Lemma \ref{subadd}.  First, we estimate $D_J(\sigma(I_k))$.
1321: \begin{align*}
1322: D_J((\sigma \otimes \tau)(I_k)) &= \left | \, | \, (\sigma \otimes \tau)(I_k) \cap J | - \frac{|(\sigma \otimes \tau)(I_k)||J|}{nm} \right | \\
1323: &= \left | \, | \, (m[0,n-1]+k) \cap J | - \frac{n|J|}{nm} \right | \\
1324: &\leq \left | \, \frac{|J| + m - 1}{m} - \frac{|J|}{m} \right | =
1325: \frac{m-1}{m}
1326: \end{align*}
1327: Let $J_0 \subset \mathbb{Z}_n$ denote the set $\left \{ \lfloor
1328: \frac{x}{m} \rfloor | \, x \in J \right \}$, and let $S_1 \subset
1329: \mathbb{Z}_n$ be the set $S_0$ reduced mod $n$.  Then,
1330: \begin{align*}
1331: D_J((\sigma \otimes \tau)(S_0)) &= \left | \, | \, (\sigma \otimes \tau)(S_0) \cap J | - \frac{|(\sigma \otimes \tau)(S_0)||J|}{nm} \right | \\
1332: &= \left | \, | \, \sigma(S_1) \cap J_0 | - \frac{|S_1||J|}{nm} \right | \\
1333: &= \left | \, | \, \sigma(S_1) \cap J_0 | - \frac{|S_1|m|J_0|}{nm} + \frac{|S_1|}{nm} (m|J_0|-|J|) \right | \\
1334: & \leq D_{J_0}(\sigma(S_1)) + \left ( \, \frac{n}{nm} (m-1) \right
1335: ) \leq d(\sigma) + \frac{m-1}{m}
1336: \end{align*}
1337: Thus,
1338: \begin{align*}
1339: d(\sigma \otimes \tau) & \leq (m-1) \frac{m-1}{m} + \frac{m-1}{m} + d(\sigma) \\
1340: & = m - 1 + d(\sigma).
1341: \end{align*}
1342: \end{proof}
1343: 
1344: An identical result holds for $d^\prime$, by symmetry.  We use
1345: this in the next proposition, which allows us to bound
1346: discrepancies recursively.
1347: 
1348: \begin{prop} \label{inh} $D(\sigma \otimes \tau) \leq m-1 + d(\sigma) + d^\prime(\sigma)$.
1349: \end{prop}
1350: \begin{proof}
1351: Note that every interval $I$ of $\mathbb{Z}_{nm}$ is of the form
1352: $$
1353: S = \bigcup_{k \in [l,L]} I_k \cup S_0 \cup S^\prime_0
1354: $$
1355: where $[l,L]$ is an interval of $\mathbb{Z}_m$ of length no more
1356: than $m-2$, $S_0$ is an initial segment of $I_{L+1}$, and
1357: $S^\prime_0$ is a final segment of $I_{l-1}$.  Applying Lemma
1358: \ref{subadd},
1359: $$
1360: D_J((\sigma \otimes \tau)(S)) \leq \sum_{k=l}^L D_J((\sigma
1361: \otimes \tau)(I_k)) + D_J((\sigma \otimes \tau)(S_0)) +
1362: D_J((\sigma \otimes \tau)(S^\prime_0))
1363: $$
1364: By the arguments presented in the proof of the previous
1365: proposition,
1366: \begin{eqnarray*}
1367: D_J((\sigma \otimes \tau)(I_k)) & \leq & \frac{m-1}{m} \\
1368: D_J((\sigma \otimes \tau)(S_0)) & \leq & d(\sigma) + \frac{m-1}{m} \\
1369: D_J((\sigma \otimes \tau)(S^\prime_0)) & \leq & d^\prime(\sigma) +
1370: \frac{m-1}{m}
1371: \end{eqnarray*}
1372: Thus,
1373: $$
1374: D(\sigma \otimes \tau) \leq m - 1 + d(\sigma) + d^\prime(\sigma)
1375: $$
1376: \end{proof}
1377: 
1378: If we apply these results to a product of permutations,
1379: \begin{cor} \label{prodperm} If, for $i \leq k$, $\sigma_i \in S_{n_i}$, where $n_i > 1$, then
1380: $$
1381: D(\bigotimes_{i=1}^{k} \sigma_i ) \leq n_k + 2 \sum_{i=1}^{k-1}
1382: n_i - 2k + 1.
1383: $$
1384: \end{cor}
1385: \begin{proof}
1386: By Proposition \ref{inh}, $d(\bigotimes_{i=1}^{m} \sigma_i ) \leq
1387: n_m - 1 + d(\bigotimes_{i=1}^{m-1} \sigma_i )$.  Inductively,
1388: then, we find
1389: $$
1390: d(\bigotimes_{i=1}^{m} \sigma_i ) \leq \sum_{i=2}^{m} (n_i - 1) +
1391: d(\sigma_1).
1392: $$
1393: Since $d(\sigma_1) \leq n_1 - 1$, we may write
1394: $d(\bigotimes_{i=1}^{m} \sigma_i ) \leq \sum_{i=1}^{m} n_i - m$
1395: and, similarly, $d^\prime(\bigotimes_{i=1}^{m} \sigma_i ) \leq
1396: \sum_{i=1}^{m} n_i - m$.  Thus,
1397: \begin{align*}
1398: D(\bigotimes_{i=1}^{k} \sigma_i ) & \leq n_k - 1 + d(\bigotimes_{i=1}^{k-1} \sigma_i ) + d^\prime(\bigotimes_{i=1}^{k-1} \sigma_i ) \\
1399: & \leq n_k + 2 \sum_{i=1}^{k-1} n_i - 2k + 1.
1400: \end{align*}
1401: \end{proof}
1402: 
1403: Corollary \ref{prodperm} provides us with a large family of very
1404: strongly quasirandom permutations.  To see this, let
1405: $\{\sigma_i\}_{i=1}^\infty$ be a sequence of permutations with
1406: $\sigma_i \in S_{n_i}$.  Then, letting $\lambda_k = \sum_{i=1}^{k}
1407: n_i / \prod_{i=1}^{k} n_i$ and applying the corollary,
1408: $$
1409: \frac{D(\bigotimes_{i=1}^{k} \sigma_i)}{\prod_{i=1}^{k} n_i} < 2
1410: \frac{\sum_{i=1}^{k} n_i}{\prod_{i=1}^{k} n_i} = 2 \lambda_k
1411: $$
1412: But,
1413: \begin{align*}
1414: \lambda_k &= \frac{1}{n_k} \cdot \frac{\sum_{i=1}^{k-1} n_i}{\prod_{i=1}^{k-1} n_i} + \frac{n_k}{\prod_{i=1}^{k} n_i} \\
1415: &= \left(\frac{1}{n_k} + \frac{1}{\sum_{i=1}^{k-1} n_i} \right )
1416: \lambda_{k-1}
1417: \end{align*}
1418: Thus, the ratio of the discrepancy to the size of the product
1419: permutations tends to zero quickly.  In particular, $\sigma^{(k)}$
1420: is very strongly quasirandom, since $\sigma^{(k)} \in S_{n^k}$.
1421: That is, if $\sigma^{(k)} \in S_N$, then $D(\sigma^{(k)}) < 2kn =
1422: O(\log N)$.  Immediately one wonders whether permutations exist
1423: with discrepancies which grow slower than $\log N$.  A theorem of Schmidt [\ref{S1}] answers this question in the negative, implying that the $D(\sigma^{(k)})$ are, in a sense, ``maximally'' quasirandom.
1424: 
1425: \begin{theorem} [Schmidt] \label{schmidt} Let $\{x_i\}_{i=0}^{N-1} \subset [0,1)$, and define
1426: $$
1427: D(m) = \sup_{\alpha \in [0,1)} \left | |\{x_i\}_{i=0}^{m-1} \cap
1428: [0,\alpha) | - m \alpha \right |.
1429: $$
1430: Then there exists an integer $n \leq N$ so that $D(n) > \log N /
1431: 100$.
1432: \end{theorem}
1433: 
1434: We may immediately conclude that discrepancies grow at least as
1435: fast as $\log N$.
1436: 
1437: \begin{cor} \label{disclowerbound} For any $\sigma \in S_N$, $D(\sigma) > \log N / 100 - 1$.
1438: \end{cor}
1439: \begin{proof}
1440: Take $x_i = \sigma(i)/N$ in Theorem \ref{schmidt}.  Then there
1441: exists an $\alpha \in [0,1)$ and an $n \leq N$ so that
1442: $$
1443: \left | \left |\frac{\sigma([0,n-1])}{N} \cap [0,\alpha) \right |
1444: - n \alpha \right | > \frac{\log N}{100}.
1445: $$
1446: Defining $k=\lfloor \alpha N \rfloor$, we have
1447: $$
1448: \left | |\sigma([0,n-1]) \cap [0,k] | - \frac{n(k+1)}{N} \right |
1449: + \left | \frac{n(k+1)}{N} - n \alpha \right | > \frac{\log
1450: N}{100}.
1451: $$
1452: Therefore, if we let $I$ and $J$ vary over all intervals in
1453: $\mathbb{Z}_N$,
1454: $$
1455: \max_{I,J} \left | |\sigma(I) \cap J | - \frac{|I||J|}{n} \right |
1456: > \frac{\log N}{100} - 1
1457: $$
1458: so that $D(\sigma) > \log N / 100 - 1$.
1459: \end{proof}
1460: 
1461: One might expect that the algebraic properties of quasirandom
1462: permutations, such as the number of cycles, should be
1463: approximately that of random permutations (in this case, $\log
1464: n$).  However, we have the following counterexample.  Let $i_n$ be
1465: the identity permutation on $\mathbb{Z}_n$.  Then $i^{(k)}_n$ is
1466: always an involution in $S_{n^k}$ -- and the sequence $\{i^{(k)}_n/n^k\}_{i=0}^{n-1} \subset [0,1)$ is an initial segment of the van der Corput sequence.  In fact, under this interpretation, Corollary \ref{prodperm} can be considered a generalization of the classical theorem that the discrepancy of the van der Corput sequence is $O(\log N)$.  (See, for example, [\ref{F1}] for a modern version of this result.)
1467: 
1468: \begin{prop} $i^{(k)}_n(x)$ is the element of $\mathbb{Z}_{n^k}$ whose base $n$ expansion is the reverse of the base $n$ expansion of $x$.
1469: \end{prop}
1470: \begin{proof}
1471: The proof is by induction on $k$.  The case $k=1$ is obvious.
1472: Suppose it were true of $i^{(k)}_n$.  Let $x \in
1473: \mathbb{Z}_{n^{k+1}}$ have the base $n$ expansion $x_{k} x_{k-1}
1474: \cdots x_1 x_0$.  Then
1475: \begin{align*}
1476: i^{(k+1)}_n(x) & = (i^{(k)}_n \otimes i_n)(x) \\
1477: & = i_n(\lfloor \frac{x}{n^k} \rfloor) + n i^{(k)}_n(x \! \! \! \! \mod n^k) \\
1478: & = (\lfloor \frac{x}{n^k} \rfloor) + n i^{(k)}_n(x_{k-1} x_{k-2} \cdots x_1 x_0) \\
1479: & = x_k + n (x_0 x_1 \cdots x_{k-2} x_{k-1}) \\
1480: & = x_0 x_1 \cdots x_{k-2} x_{k-1} x_k.
1481: \end{align*}
1482: \end{proof}
1483: 
1484: \section{{\sc Conclusion}}
1485: \setcounter{theorem}{0}
1486: 
1487: The original motivation for this paper was a (still unanswered)
1488: question of R. L. Graham [\ref{G2}].  For a sequence of
1489: permutations $\sigma_j \in S_{n_j}$, let {\bf P(k)} be the
1490: property of \textit{asymptotic $k$-symmetry}: for each $\tau \in
1491: S_k$,
1492: $$
1493: \left | X^{\tau}(\sigma_j) - \frac{\binom{n_j}{k}}{k!} \right | =
1494: o(n_j^k).
1495: $$
1496: Note that this property is weaker than property \textbf{[kS]} of
1497: Theorem \ref{qrp}, which we will call \textit{strong} asymptotic
1498: $k$-symmetry.  Theorem \ref{qrp} says that strong asymptotic
1499: $k$-symmetry implies strong asymptotic $(k+1)$-symmetry for any $k
1500: \geq 2$.  Graham asks whether there exists an analogous $N$ so
1501: that, for all $k > N$, $\mbox{\bf P(k)} \Rightarrow \mbox{\bf
1502: P(k+1)}$?  At first it might seem like one is asking for too much.
1503: However, precisely this type of phenomenon occurs for graphs
1504: ([\ref{CGW1}]).  It turns out that, if we let {\bf G(k)} be the
1505: property that all graphs on $k$ vertices occur as subgraphs at
1506: approximately the same rate, then
1507: $$
1508: \mbox{\bf G(1)} \Leftarrow \mbox{\bf G(2)} \Leftarrow \mbox{\bf
1509: G(3)} \Leftarrow \mbox{\bf G(4)} \Leftrightarrow \mbox{\bf G(5)}
1510: \Leftrightarrow \mbox{\bf G(6)} \Leftrightarrow \cdots
1511: $$
1512: In particular, {\bf G(4)} implies quasirandomness, which in turn
1513: implies {\bf G(k)} for all $k$.
1514: 
1515: The fact that $\mbox{\bf P(1)} \not \Rightarrow \mbox{\bf P(2)}$
1516: is trivial.  To show that $\mbox{\bf P(2)} \not \Rightarrow
1517: \mbox{\bf P(3)}$, let $\sigma_n \in S_{2n}$ be the permutation
1518: which sends $x$ to $x+n$.  Then $X^{01}(\sigma_n) = 2n(2n-1)$, and
1519: $X^{10}(\sigma_n) = 4n^2$, so that $\left | X^{01}(\sigma_n) -
1520: X^{10}(\sigma_n) \right | = o((2n)^2)$.  However, the pattern
1521: $(021)$ {\it never} appears in $\sigma_n$.  We have been unable to
1522: date to provide an analogous result for any $P(k)$ with $k
1523: > 2$.
1524: 
1525: A second, very natural question is that of the existence of
1526: \textit{perfect} $m$-symmetry: the property of having all
1527: subsequence statistics \textit{precisely} equal to their expected
1528: values.  That is, for $\sigma \in S_n$,
1529: $$
1530: X^{\tau}(\sigma) = \frac{\binom{n}{k}}{k!}
1531: $$
1532: for all $\tau \in S_m$.  For this to occur, the number of
1533: permutations of length $m$ must evenly divide $\binom{n}{m}$.  Let
1534: \textbf{D($m$)} be the property of an integer $N$ that
1535: $$
1536: m! \left | \binom{n}{m} \right. \!\! .
1537: $$
1538: It is easy to see that a permutation $\sigma \in S_n$ with perfect
1539: $m$-symmetry must have perfect $m^\prime$-symmetry for any
1540: $m^\prime \leq m$, so $n$ must satisfy \textbf{D($m^\prime$)} for
1541: all such $m^\prime$.  Let $h(m)$ be the least $n$ for which this
1542: occurs.  A quick calculation reveals that $h(2)=4$, $h(3)=9$,
1543: $h(4)=64$, and $h(5)=128$.  In fact, there is a perfect
1544: $2$-symmetric permutation on $4$ symbols: 3012.  A computer search
1545: revealed that there are exactly two $3$-symmetric permutations on
1546: $9$ symbols: 650147832 and its reverse, 238741056.  No
1547: $m$-symmetric permutation is known for $m>3$, and the question of
1548: whether such permutations exist remains open.  We conjecture that
1549: an $m$-symmetric permutation on sufficiently many symbols exists
1550: for all $m$, and believe it likely that one exists on $h(m)$
1551: symbols.
1552: 
1553: Finally, the selection of intervals as the sets which measure ``proximity''
1554: in the definition of quasirandomness was a natural but somewhat arbitrary choice.
1555: It would be worth investigating the properties of ``$(\mathcal{A},\mathcal{B})$-quasirandom''
1556: permutations for families $\mathcal{A}$, $\mathcal{B} \subset 2^{\mathbb{Z}_n}$, i.e., permutations $\sigma$ such that $\max_{A,B} D_B(\sigma(A)) = o(n)$ for $A \in \mathcal{A}$ and $B \in \mathcal{B}$.
1557: 
1558: \section*{{\sc Acknowledgements}}
1559: 
1560: The author wishes to thank Fan Chung Graham and Ron Graham for
1561: their tremendous help in formulating and attacking the problems discussed above.  He also thanks Chris Dillard, Robert Ellis, and Lei Wu for helpful discussions during the development of this work.
1562: 
1563: \section*{{\sc References}}
1564: \begin{enumerate}
1565: \item \label{Ah1} {\sc L. V. Ahlfors}, ``Complex Analysis,'' McGraw-Hill Book Co., New York, 1978.
1566: \item \label{A1} {\sc N. Alon}, Generating Pseudo-Random Permutations and Maximum Flow Algorithms, {\it Inform. Process. Lett.} {\bf 35} (1990), 201--204.
1567: \item \label{AF1} {\sc N. Alon and E. Friedgut}, On the number of permutations avoiding a given pattern, {\it J. Comb. Theory Ser. A} {\bf 89} (2000), 133--140.
1568: \item \label{AS1} {\sc N. Alon and J. H. Spencer}, ``The probabilistic method,'' Wiley-Interscience Series in Discrete Mathematics and Optimization. Wiley-Interscience [John Wiley \& Sons], New York, 2000. 
1569: \item \label{B1} {\sc M. B\'{o}na}, The solution of a conjecture of Stanley and Wilf for all layered patterns, {\it J. Comb. Theory Ser. A} {\bf 85} (1999), 96--104.
1570: \item \label{Ch1} {\sc B. Chazelle}, ``The Discrepancy Method,'' Cambridge University Press, Cambridge, 2000.
1571: \item \label{CG1} {\sc F. R. K. Chung and R. L. Graham}, Quasi-random set systems, {\it J. Amer. Math. Soc.} {\bf 4} (1991), 151--196.
1572: \item \label{CG2} {\sc F. R. K. Chung and R. L. Graham}, Quasi-random subsets of $Z\sb n$, {\it J. Combin. Theory Ser. A} {\bf 61} (1992), 64--86.
1573: \item \label{CGW1} {\sc F. R. K. Chung, R. L. Graham, and R. M. Wilson}, Quasi-random graphs, {\it Combinatorica} {\bf 9} (1989), 345--362.
1574: \item \label{D1} {\sc P. J. Davis}, ``Circulant Matrices,'' Wiley, New York, 1979.
1575: \item \label{D2} {\sc J. A. Davis, R. C. Entringer, R. L. Graham, and G. J. Simmons}, On permutations containing no long arithmetic progressions, {\it Acta Arithmetica}, {\bf 34} (1977/78), 81--90.
1576: \item \label{ES1} {\sc P. Erd\H{o}s and G. Szekeres}, A combinatorial problem in geometry, {\it Compocito Math.} {\bf 2} (1935), 464--470.
1577: \item \label{F1} {\sc H. Faure}, Discr\'{e}pance quadratique de suites infinies en dimension un. Th\'{e}orie des nombres (Quebec, PQ, 1987), 207--212, de Gruyter, Berlin, 1989. 
1578: \item \label{G1} {\sc W. T. Gowers}, A new proof of Szemer\'edi's Theorem, {\it Geometric and Functional Analysis} {\bf 11} (2001), 465--588.
1579: \item \label{G2} {\sc R. L. Graham}, personal communication.
1580: \item \label{L1} {\sc J. C. Lagarias}, Well-spaced labelling of points in rectangular grids, {\it SIAM J. Discrete Math.} {\bf 13} (2000), 521--534.
1581: \item \label{N1} {\sc H. Niederreiter}, Quasi-Monte Carlo methods and pseudo-random numbers. {\it Bull. Amer. Math. Soc.} {\bf 84} (1978), 957--1041. 
1582: \item \label{MMW1} {\sc B. D. McKay, J. Morse, and H. S. Wilf}, The distributions of the entries of Young tableaux, to appear.
1583: \item \label{Sa1} {\sc V. N. Sachkov}, ``Probabilistic Methods in Combinatorial Analysis,'' Encyclopedia of Mathematics and its Applications {\bf 56}, Cambridge University Press, Cambridge, 1997.
1584: \item \label{SS1} {\sc F. W. Schmidt and R. Simion}, Restricted permutations, {\it European J. Combin.} {\bf 6} (1985), 383--406.
1585: \item \label{S1} {\sc W. M. Schmidt}, Irregularities of distribution VII, {\it Acta Arith.} {\bf 21} (1972), 45–-50.
1586: \item \label{St1} {\sc R. P. Stanley}, Log-concave and unimodal sequences in algebra, combinatorics, and geometry. Graph theory and its applications: East and West (Jinan, 1986), 500--535, Ann. New York Acad. Sci., 576, New York Acad. Sci., New York, 1989.
1587: \item \label{St2} {\sc R. P. Stanley}, Enumerative combinatorics, Vol. 1, {\it Cambridge Studies in Advanced Mathematics} {\bf 49}, Cambridge University Press, Cambridge, 1997.
1588: \end{enumerate}
1589: 
1590: \ifx\ispreprint\undefined
1591: \pagebreak
1592: \def\baselinestretch{1}
1593: \begin{center}
1594: \begin{figure}[ht]
1595: \hspace{.7in} \input{figa.pstex_t} \caption{Diagram of
1596: implications for Theorem \ref{balance}.}
1597: \end{figure}
1598: \end{center}
1599: \else
1600: \fi
1601: 
1602: \ifx\ispreprint\undefined
1603: \pagebreak
1604: \def\baselinestretch{1}
1605: \begin{center}
1606: \begin{figure}[ht]
1607: \hspace{1.25in} \input{figb.pstex_t} \caption{Diagram of
1608: implications for Theorem \ref{qrp}.}
1609: \end{figure}
1610: \end{center}
1611: \else
1612: \fi
1613: 
1614: \end{document}
1615: