1: \documentclass[11pt]{amsart}
2: \usepackage{amsmath, amssymb, amsthm, amscd, graphicx}
3: %\usepackage{epstopdf}
4:
5: \textwidth=14.5cm
6: \textheight=21cm
7:
8: % for 100% printing
9: %\hoffset-1.2cm
10: %\voffset+0.5cm
11:
12: %for 110% printing
13: \hoffset-1.0cm
14: \voffset-0.6cm
15:
16: \setlength{\unitlength}{1mm}
17:
18: \theoremstyle{plain}
19: \newtheorem{prop}{Proposition}[section]
20: \newtheorem{coro}[prop]{Corollary}
21: \newtheorem{conj}[prop]{Conjecture}
22: \newtheorem{lemm}[prop]{Lemma}
23: \newtheorem{ques}[prop]{Question}
24: \newtheorem{thrm}[prop]{Theorem}
25:
26: \theoremstyle{definition}
27: \newtheorem{defi}[prop]{Definition}
28: \newtheorem*{propA}{Property A}
29: \newtheorem*{propC}{Property C}
30: \newtheorem{nota}[prop]{Notation}
31: \newtheorem{exam}[prop]{Example}
32: \newtheorem{rema}[prop]{Remark}
33:
34: \numberwithin{equation}{section}
35:
36: % bibliography
37:
38: \def\Reff#1; #2; #3; #4; #5; #6; #7\par{%
39: \bibitem{#1} #2, {\it #3}, #4 {\bf #5} (#6) #7}
40:
41: \def\Ref#1; #2; #3; #4\par{%
42: \bibitem{#1} #2, {\it #3}, #4}
43:
44: % Specific macros of this text $$$$
45:
46: \catcode`\¨=\active\def¨{~}
47:
48: \renewcommand{\aa}[3]{a_{#1,#2,#3}}%_Entry_of_matrix
49: \newcommand{\aab}[2]{\overline{a}_{#1,#2}}%_Entry_of_matrix
50: \newcommand{\aas}[3]{\widehat a_{#1, #2,#3}}
51:
52: \renewcommand{\Bar}[1]{\overline{\vline width0pt
53: height5.5pt #1}}
54: \newcommand{\bb}[2]{b_{#1,#2}}
55: \newcommand{\bbb}[3]{b_{#1,#2}(#3)}
56: \newcommand{\bbbb}[3]{b'_{#1,#2}(#3)}
57: \newcommand{\bs}{x}%_A_simple
58: \newcommand{\bt}{y}%_Another_simple
59: \newcommand{\bx}{x}%_A_braid
60: \newcommand{\by}{y}
61:
62: \newcommand{\card}{\mathtt{\#}}
63: \newcommand{\charpol}[1]{P_{#1}(x)}
64: \newcommand{\cl}[1]{[#1]}
65: \newcommand{\comp}[2]{[#2]_{#1}}
66: \newcommand{\concat}{\mathbin{{}^{\scriptscriptstyle\frown}}}
67:
68: \newcommand{\D}{\Delta}
69: \newcommand{\DD}[2]{\D_{#1}^{#2}}
70: \newcommand{\dd}{d}%_Degree
71: \renewcommand{\div}{\prec}
72: \newcommand{\Div}{\mathrm{Div}}
73: \newcommand{\dive}{\preccurlyeq}
74: \newcommand{\DivL}{D_{\!L}}
75: \newcommand{\DivR}{D_{\!R}}
76: \newcommand{\dual}[1]{{}^*#1}
77: \newcommand{\pz}{z}
78:
79: \newcommand{\eg}{{\it e.g.}}
80: \newcommand{\etc}{{\it etc.}}
81: \newcommand{\ev}{\rho}%_Eigenvalue
82:
83: \newcommand{\ff}{f}
84: \newcommand{\flip}[1]{\phi_{#1}}
85: \newcommand{\first}[1]{(#1)_{\!_1}}
86:
87: \newcommand{\g}{\gamma}
88: \let\ge=\geqslant
89: \renewcommand{\gg}{g}
90:
91: \newcommand{\ie}{{\it i.e.}}
92: \newcommand{\II}{I}
93: \let\ince=\subseteq
94: \newcommand{\ind}{i}
95: \newcommand{\Ind}{\mathrm{ind}}
96: \newcommand{\indd}{j}
97: \newcommand{\Int}[1]{[\![1, #1]\!]}
98: \newcommand{\inv}{^{-1}}
99: \newcommand{\Inv}{{}^{-1}}
100: \newcommand{\ik}{i}
101: \newcommand{\ip}{p}
102: \newcommand{\iq}{q}
103: \newcommand{\ir}{k}
104: \newcommand{\is}{\ell}
105:
106: \newcommand{\JJ}{J}
107:
108: \newcommand{\kk}{k}
109: \newcommand{\KK}{K}
110:
111: \renewcommand{\l}{\lambda}
112: \let\le=\leqslant
113: \renewcommand{\ll}{\ell}
114:
115: \newcommand{\m}{\mu}
116: \newcommand{\Mat}[1]{M_{#1}}%_Transition_matrix
117: \newcommand{\Matt}[1]{M'_{#1}}
118: \newcommand{\Mattt}[1]{\overline M_{#1}}
119: \newcommand{\mm}{m}
120:
121: \newcommand{\nbpart}[1]{p(#1)}% number of partitions
122: \newcommand{\nn}{n}
123: \newcommand{\NN}{N}
124: \newcommand{\nno}{{\nn-1}}
125: \newcommand{\NNN}{N}
126:
127: \newcommand{\op}{\cdot}
128:
129: \renewcommand{\part}[2]{\{#2\}_{#1}}
130: \newcommand{\perm}[1]{\pi(#1)}
131: \newcommand{\pp}{p}
132: \newcommand{\ppp}{p}
133: \newcommand{\pppp}{p}
134: \newcommand{\Pw}{\mathfrak{P}}
135:
136: \newcommand{\qq}{q}
137: \newcommand{\QQ}{\mathbb{Q}}
138:
139: \newcommand{\resp}{{\it resp.{¨}}}
140: \newcommand{\rr}{r}
141: \newcommand{\RR}{\mathbb{R}}
142: \newcommand{\rrr}{r}
143:
144: \let\s=\sigma
145: \newcommand{\set}[1]{\widetilde{#1}}
146: \newcommand{\sh}{\mathrm{sh}}
147: \renewcommand{\sp}[1]{\tau_{#1}}
148: \renewcommand{\ss}[1]{\sigma_{#1}}
149: \renewcommand{\SS}{S}
150: \newcommand{\sx}{x}%_A_simple_factor_of_x
151: \newcommand{\Sym}[1]{\mathfrak{S}_{#1}}
152:
153: \newcommand{\uu}{u}
154:
155: \newcommand{\var}{u}
156: \newcommand{\vs}{\emph{vs.} }
157: \newcommand{\vv}{v}
158:
159: \newcommand{\ww}{w}
160:
161: \newcommand{\xx}{x}
162: \newcommand{\XX}{X}
163:
164: \newcommand{\yy}{y}
165:
166: \newcommand{\zz}{z}
167:
168: \hyphenation{mon-oid mon-oids Lem-ma}
169: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
170: \begin{document}
171:
172: \author{Patrick DEHORNOY}
173: \address{Laboratoire de Math\'ematiques Nicolas
174: Oresme UMR 6139\\ Universit\'e de Caen,
175: 14032~Caen, France}
176: \email{dehornoy@math.unicaen.fr}
177: \urladdr{//www.math.unicaen.fr/\textasciitilde dehornoy}
178:
179: \title{Combinatorics of normal sequences of
180: braids}
181:
182: \keywords{braid group; normal form; fundamental
183: braid; counting; generating function}
184:
185: \subjclass{20F36, 05A05}
186:
187: \begin{abstract}
188: Many natural counting problems arise in
189: connection with the normal form of braids---and
190: seem to have not been much considered so far. Here
191: we solve some of them by analysing the normality
192: condition in terms of the associated
193: permutations, their descents and the
194: corresponding partitions. A number of different
195: induction schemes appear in that framework.
196: \end{abstract}
197:
198: \maketitle
199:
200: Ubiquitous and connected with a number of domains,
201: Artin's braid groups have received much attention
202: in the recent years. However, not so many works
203: are devoted to a purely combinatorial study of
204: braids, presumably because counting arguments did
205: not prove so far to be much helpful for
206: investigating braids. Nevertheless, although
207: braid groups are infinite, they admit several
208: filtrations leading to finite sets and,
209: therefore, to natural enumeration problems.
210:
211: For each presentation of braid groups, (at least) two
212: natural counting poblems arise, namely, on the one hand,
213: counting how many braids admit an expression of
214: a given length, in particular evaluating the associated
215: growth rate---we shall refer to this as Question¨1 in the
216: sequel---and, on the other hand, counting, for a given
217: braid, how many words represent that specific braid, a
218: relevant question when the number is finite,
219: typically when we discard the inverses of the
220: generators and only consider positive expressions,
221: \ie, when we restrict to some submonoid of~$B_\nn$---we
222: shall refer to that as Question¨2.
223:
224: In the case of the Artin generators¨$\ss i$, both types of
225: questions have been addressed, and at least partially
226: solved: Question¨1, actually not for¨$B_\nn$ but for the
227: submonoid¨$B_\nn^+$ of¨$B_\nn$ generated by the¨$\ss i$'s, was
228: investigated in~\cite{Xua}, and completely solved
229: in~\cite{Bro}. As for Question~2, it is natural in this context to
230: address it for the particular elements¨$\D_n^d$, where
231: $\D_n$ is Garside's fundamental braid \cite{Gar}. It was
232: investigated and solved for~$n = 3$ in~\cite{CrH}.
233:
234: In this paper, we address Question~1 for another natural
235: generating set, namely the so-called simple braids, also called
236: the Garside generators below~\cite{Gar}.
237: These generators, which are the divisors of~$\D_n$ in
238: the monoid~$B_\nn^+$, are in one-to-one
239: correspondence with permutations of $n$~objects,
240: and they give rise to a remarkable unique
241: decomposition for each braid, usually called its normal
242: form \cite{Dlg, Adj, Eps, ElM}. Because of its
243: uniqueness and of its many nice properties,
244: expressed in particular in the existence of a bi-automatic
245: structure, the normal form
246: of braids is the preferred way of specifying
247: braids in many recent developments, in particular
248: those of algorithmic or cryptographical nature \cite{Geb, KoL}.
249:
250: In the case of the Garside generators and the associated
251: normal form(s), what we called Question¨2 above is not
252: relevant, as each braid has one unique distinguished
253: decomposition in terms of the generators, and counting the non-normal expressions would appear artificial. But Question¨1, namely counting the
254: number of braids with a normal form of a given length, is
255: quite natural. The question was briefly considered
256: by R.\,Charney in¨\cite{Chb}: she observed that, because
257: normal words can be recognized by a finite state automaton,
258: the number of braids with length¨$\dd$ obeys a linear
259: induction rule, and the associated generating function is
260: rational. Explicit values are given in the case of
261: $3$¨and $4$¨strand braids.
262:
263: The aim of this paper is to go further in the investigation
264: of counting problems connected with the normal form of
265: braids. We consider the case of the monoid¨$B_\nn^+$, and
266: mainly study the number of positive $n$¨strand braids
267: with a normal form of length (at most)¨$\dd$, \ie, we
268: consider for Garside generators the problem solved
269: in¨\cite{Bro} in the case of Artin generators. The main
270: difference is that, in the case of
271: Garside generators, the length is no longer an additive
272: parameter: for instance, multiplying two simple braids may
273: result in a simple braid, \ie, multiplying two braids of
274: length¨$1$ may result in a braid of length¨$1$. This makes
275: the current study much more uneasy.
276:
277: Let $\bb\nn\dd$ denote the number of positive $n$¨strand
278: braids with a normal form of length at most¨$\dd$, \ie, the
279: number of divisors of¨$\DD\nn\dd$ in¨$B_\nn^+$. We establish
280: various results about the numbers¨$\bb\nn\dd$,
281: and about the connected numbers¨$\bbb\nn\dd\bs$ that count,
282: for $\bs$ a simple braid, the positive $n$¨strand braid
283: with a normal form of length at most¨$\dd$ whose
284: $\dd$th factor is precisely¨$\bs$. Two types of results are established,
285: namely results for fixed braid index¨$n$, and results for fixed degree¨$\dd$.
286: When $n$ is fixed and $\dd$ varies, as was recalled above,
287: the numbers¨$\bb\nn\dd$ and $\bbb\nn\dd\bs$ obey a linear
288: induction rule associated with a certain $n! \times n!$
289: incidence matrix¨$\Mat\nn$. Here we show that $\Mat\nn$
290: can be replaced with a smaller matrix of size $p(\nn)
291: \times p(\nn)$, where $p(\nn)$ is the number of partitions
292: of¨$\nn$. The result relies on analysing the descents
293: of the permutations associated with simple braids, and it
294: is connected with a classical result by Solomon¨\cite{Sol}.
295: It is then easy to deduce the numerical value of¨$\bb\nn\dd$
296: for small¨$\nn, \dd$, as well as explicit formulas, at least
297: for $n \le 4$. We are also led to several
298: conjectures about the eigenvalues of the matrix¨$\Mat\nn$
299: that seem to have never been considered so far. The most
300: puzzling one claims that the characteristic polynomial
301: of¨$\Mat{\nno}$ divides that of¨$\Mat\nn$. It holds at
302: least for $\nn \le 10$.
303:
304: When $\dd$ is fixed and $\nn$ varies, quite different
305: induction rules appear. Everything is trivial for $\dd =
306: 1$, and an explicit formula for¨$\bb\nn2$ can be deduced from
307: the results of¨\cite{CSV, CSV2}. It seems difficult to
308: go further in general, but new results (and new
309: induction schemes) appear when we consider the numbers
310: $\bbb\nn\dd{\D_{\nn-\rrr}}$ with $1 \le \rr \le \nn$,
311: typically in the (non-trivial) case $\rr = 1$, and, more
312: generally, when $\rr$ is fixed. In particular, we
313: obtain explicit values for $\bbb\nn3{\D_{\nno}}$,
314: $\bbb\nn3{\D_{\nn-2}}$, and $\bbb\nn4{\D_{\nno}}$.
315:
316: The specific questions investigated in this paper, in
317: particular that of the value
318: of¨$\bbb\nn\dd{\D_{\nn-\rrr}}$, arose in¨\cite{Dhh}. There
319: exists a distinguished linear ordering of braids¨$<$ relying
320: on the notion of a
321: $\s$-positive braid word¨\cite{Dgr}, and the aim
322: of¨\cite{Dhh} is to develop a new approach to that ordering
323: based on the study of its connection with the Garside
324: structure. It turns out that certain parameters describing the
325: restriction of¨$<$ to positive $n$-braids of degree at
326: most¨$\dd$ can be expressed in terms of the
327: numbers¨$\bbb\nn\dd{\D_\rr}$, an initial
328: motivation for our current study of these numbers. However,
329: we think that the formulas and methods developed in the
330: current paper go beyond the above specific applications. In
331: particular, the great diversity of the induction
332: schemes appearing in connection with various
333: specializations of the general problem is remarkable. At the
334: least, the current study should demonstrate the richness of
335: the combinatorics underlying the normal form of braids.
336:
337: Still other presentations of the braid groups are known, in
338: particular the one involving the so-called dual
339: monoid¨\cite{Bes, BKL}, which gives rise to an alternative
340: Garside structure, and, therefore, to an alternative
341: normal form analogous to that considered here, where the
342: role of simple braids is played by elements that are in
343: one-to-one correspondence with non-crossing partitions.
344: All questions considered in the current paper could be
345: similarly addressed for the dual structure, and, more
346: generally, for the many presentations of¨$B_\nn$ known to
347: date. Similarly, Artin's braid groups~$B_\nn$ belong to larger
348: families of groups, typically Artin-Tits groups of spherical
349: type and, more generally, Garside groups~\cite{Dfx, Dgk, McC}.
350: Once again, all
351: questions considered here extend to such frameworks
352: naturally. However, mainly because of the specific
353: applications mentioned above, we find it interesting to
354: consider here the specific framework of braids and
355: permutations, and we leave the extensions for further
356: investigation.
357:
358: The paper is organized as follows. Section¨\ref{S:Back}
359: sets the framework and the basic definitions. In
360: Section¨\ref{S:Adj} we introduce the incidence
361: matrix¨$\Mat\nn$ that controls the sequences¨$\bb\nn\dd$ for
362: fixed¨$\nn$ and show how to reduce their size
363: from¨$\nn!$ to¨$2^\nno$. In Section¨\ref{S:Part}, we
364: show how to further reduce the size to¨$p(\nn)$, and solve
365: the induction for small values of¨$\nn$. Finally, in
366: Section¨\ref{S:Deg}, we turn to the cases when the
367: degree is fixed and the braid index varies.
368:
369: \section*{Acknowledgment}
370:
371: The author thanks C.\,Hohlweg, F.\,Hivert and
372: J.C.\,Novelli for interesting discussions about the topics
373: investigated in this paper, in particular Remark¨\ref{R:Hohlweg}.
374:
375: \section{Background and preliminary results}
376: \label{S:Back}
377:
378: Our notation is standard, and we refer to textbooks
379: like¨\cite{Bir} or¨\cite{Eps} for basic results about
380: braid groups. We recall that the $\nn$¨strand braid group¨$B_\nn$
381: is defined for $\nn \ge 1$ by the presentation
382: \begin{equation} \label{E:Present}
383: B_\nn = \left<\ss1, \dots, \ss{\nno} \,;\,
384: \begin{array}{cl}
385: \ss i \ss j = \ss j \ss i
386: &\text{\quad for $\vert
387: i - j\vert \ge 2$}\\
388: \ \ss i \ss j \ss i = \ss j \ss i \ss j
389: &\text{\quad for $\vert i - j\vert = 1$}
390: \end{array} \right>.
391: \end{equation}
392: So, $B_1$ is a trivial group¨$\{1\}$, while $B_2$
393: is the free group generated by¨$\ss1$. The elements of¨$B_\nn$
394: are called $\nn$¨strand braids, or simply {\it $\nn$-braids}.
395: We use $B_\infty$ for the group generated by an infinite
396: sequence of¨$\ss i$'s subject to the relations
397: of¨\eqref{E:Present}, \ie, the direct limit of all¨$B_\nn$'s
398: under the inclusion of¨$B_\nn$ into¨$B_{n+1}$.
399:
400: By definition, every $\nn$-braid¨$\bx$ admits (infinitely
401: many) expressions in terms of the generators¨$\ss i$, $1
402: \le i < n$. Such a expression is called an $\nn$¨strand {\it
403: braid word}. Two braid words¨$\ww, \ww'$ representing the
404: same braid are said to be {\it equivalent}; the braid represented
405: by a braid word¨$\ww$ is denoted¨$\cl\ww$.
406:
407: It is standard to associate with every $\nn$¨strand braid
408: word¨$\ww$ an $\nn$¨strand braid {\it diagram} by stacking
409: elementary diagrams associated with the successive letters
410: according to the rules
411: \begin{center}
412: \begin{picture}(83,13)(0,0)
413: \put(10,0){\includegraphics{Sigma.eps}}
414: \put(-4,3){$\ss i\mapsto$}
415: \put(9.5,9){$1$}
416: \put(19.5,9){$2$}
417: \put(50,9){$i$}
418: \put(58,9){$i{+}1$}
419: \put(27,3){\dots}
420: \put(77,3){\dots}
421: \end{picture}
422: \end{center}
423: \begin{center}
424: \begin{picture}(83,10)(0,0)
425: \put(10,0){\includegraphics{SigmaInv.eps}}
426: \put(-4,3){$\ss i\inv \mapsto$}
427: \put(27,3){\dots}
428: \put(77,3){\dots}
429: \end{picture}
430: \end{center}
431: Then two braid words are
432: equivalent if and only if the diagrams they encode are the
433: projections of ambient isotopic figures in¨$\RR^3$, \ie, one
434: can deform one diagram into the other without allowing the
435: strands to cross or moving the endpoints.
436:
437: \subsection{The monoid¨$B_\nn^+$ and the braids¨$\D_\nn$}
438:
439: Let $B_\nn^+$ be the monoid admitting
440: the presentation¨\eqref{E:Present}. The elements of¨$B_\nn^+$
441: are called {\it positive} $\nn$-braids.
442:
443: \begin{defi}
444: For $\bx, \by$ in¨$B_\nn^+$, we say that $\bx$ is a {\it
445: left divisor} of¨$\by$, denoted $\bx \dive \by$, or,
446: equivalently, that $\by$ is a {\it right multiple}
447: of¨$\bx$, if $\by = \bx \pz$ holds for some¨$\pz$
448: in¨$B_\nn^+$. We denote by¨$\Div(\by)$ the (finite) set of
449: all left divisors of¨$\by$ in¨$B_\nn^+$.
450: \end{defi}
451:
452: As $B_\nn^+$ is not commutative for $\nn \ge 3$, there are
453: the symmetric notions of a right divisor and a left
454: multiple---but we shall mostly use left divisors here.
455: Note that $\bx$ is a (left) divisor of¨$\by$ in the sense
456: of¨$B_\nn^+$ if and only if it is a (left) divisor in the
457: sense of¨$B_\infty^+$, so there is no need to
458: always specify the index¨$\nn$.
459:
460: Wih respect to left divisibility, $B_\nn^+$ has the structure
461: of a lattice \cite{Gar}: any two positive $\nn$-braids¨$\bx,
462: \by$ admit a greatest common left divisor, denoted
463: $\gcd(\bx, \by)$, and a least common right multiple. A
464: special role is played by the lcm of the elements¨$\ss1$,
465: \dots, $\ss{\nno}$, traditionally denoted¨$\D_\nn$, which we
466: recall is inductively defined by
467: \begin{equation} \label{E:Delta}
468: \D_1 = 1, \qquad
469: \D_\nn = \ss1 \ss2 \dots \ss{\nno} \, \D_{\nno}.
470: \end{equation}
471: It is well known that $\D_\nn^2$ belongs to the
472: centre of¨$B_\nn$ (and even generates it for $\nn \ge 3$), and
473: that the inner automorphism¨$\flip n$ of¨$B_\nn$ corresponding
474: to conjugation by¨$\D_\nn$ exchanges¨$\ss i$ and¨$\ss{n-i}$
475: for $1 \le i \le \nno$.
476:
477: \subsection{The normal form}
478:
479: In¨$B_\nn^+$, the left and the right divisors of¨$\D_\nn$
480: coincide, and they make a finite sublattice of¨$(B_\nn^+,
481: \dive)$ with $\nn!$¨elements. These braids will be called
482: {\it simple} in the sequel. Geometrically, simple braids
483: are those positive braids that can be represented by a
484: braid diagram in which any two strands cross at most once.
485:
486: For each positive $\nn$-braid¨$\bx$ distinct of¨$1$,
487: the simple braid $\gcd(\bx, \D_\nn)$ is the maximal
488: simple left divisor of¨$\bx$, and we obtain a
489: distinguished expression $\bx = \sx_1 \bx'$
490: with¨$\sx_1$ simple. By decomposing¨$\bx'$ in the
491: same way and iterating, we obtains the so-called
492: normal expression¨\cite{ElM, Eps}.
493:
494: \begin{defi}
495: A sequence $(\sx_1, \dots, \sx_\dd)$ of simple
496: $\nn$-braids is said to be {\it normal} if, for
497: each¨$\kk$, one has $\sx_\kk =
498: \gcd(\D_\nn, \sx_\kk \dots \sx_\dd)$.
499: \end{defi}
500:
501: Clearly, each positive braid admits a unique normal
502: expression. It will be convenient here to consider the
503: normal expression as unbounded on the right by
504: completing it with as many trivial factors¨$1$ we need. In
505: this way, we can speak of the {\it $\dd$th factor} (in the
506: normal form) of¨$\bx$ for each positive braid¨$\bx$. We
507: say that a positive braid has {\it degree¨$\dd$} if $\dd$ is
508: the largest integer such that the $\dd$th factor of¨$\bx$ is
509: not¨$1$. It is well known that the positive $\nn$-braids of
510: degree at most¨$\dd$ coincide with the (left or oright)
511: divisors of¨$\DD\nn\dd$.
512:
513: The only properties of the normal form we shall use here
514: are as follows:
515:
516: \begin{lemm} \cite{ElM} \label{L:Normal}
517: Assume that $(\sx_1, \dots, \sx_\dd)$ is a sequence
518: of simple
519: $\nn$-braids. Then the following are equivalent:
520:
521: $(i)$ The sequence $(\sx_1, \dots, \sx_r)$ is normal;
522:
523: $(ii)$ For $1 \le \kk < \dd$, the sequence
524: $(\sx_\kk, \sx_{\kk+1})$ is normal;
525:
526: $(iii)$ For $1 \le \kk < \dd$, every¨$\ss i$
527: dividing¨$\sx_{\kk+1}$ on the left divides¨$\sx_\kk$
528: on the right.
529: \end{lemm}
530:
531: \begin{defi}
532: For¨$\sx$ a simple $\nn$-braid, we define
533: $\DivL(\sx)$ (\resp $\DivR(\sx)$) to be the set of
534: all¨$i$'s such that $\ss i$ is a left (\resp
535: right) divisor of¨$\sx$.
536: \end{defi}
537:
538: The example of¨$\ss2$ and $\ss2\ss1\ss3\ss2$,
539: for which both¨$\DivL$ and¨$\DivR$ is¨$\{2\}$,
540: shows that these sets do not determine a
541: simple braid. However, as far as normal sequences are
542: concerned, they contain all needed
543: information, as Lemma¨\ref{L:Normal}
544: can be restated as:
545:
546: \begin{lemm} \label{L:NormalBis}
547: A sequence of simple $\nn$-braids $(\sx_1, \dots,
548: \sx_\dd)$ is normal if and only if, for each¨$\kk <
549: \dd$, we have $\DivR(\sx_\kk)
550: \supseteq \DivL(\sx_{\kk+1})$.
551: \end{lemm}
552:
553: \subsection{Connection with permutations}
554:
555: Everywhere in the sequel, we write¨$\Int\nn$ for $\{1,
556: \dots, \nn\}$. By mapping¨$\ss i$ to the transposition¨$(i,
557: i+1)$, one defines a surjective homomorphism¨$\pi$
558: of¨$B_\nn$ onto the symmetric group¨$\Sym\nn$. The
559: restriction of¨$\pi$ to simple braids is a bijection:
560: for every permutation¨$f$ of¨$\Int\nn$, there exists exactly
561: one simple braid¨$\bs$ satisfying $\perm\bs = f$.
562:
563: The Exchange Lemma for Coxeter groups connects
564: the sets¨$\DivL(\sx)$ and¨$\DivR(\sx)$ with the
565: permutation associated with¨$\sx$ and their descents.
566: For¨$\ff$ a permutation, use $\ell(\ff)$ for the minimal
567: number of factors occurring in a decomposition of¨$\ff$
568: as a product of transpositions. The precise statement is
569:
570: \begin{lemm} \label{L:Exchange}
571: Let $\bs$ be a simple $\nn$-braid¨$\bs$. For $1 \le i
572: < \nn$, the following are equivalent:
573:
574: $(i)$ The braid¨$\ss i$ is a left divisor of¨$\bs$
575: in¨$B_\nn^+$, \ie, $i$ belongs to¨$\DivL(\bs)$;
576:
577: $(ii)$ The strands starting at positions¨$i$ and¨$i+1$
578: cross in any (positive) diagram for¨$\bs$;
579:
580: $(iii)$ We have $\pi(\bs)\inv(i) > \pi(\bs)\inv(i+1)$;
581:
582: $(iv)$ We have $\ell(\pi(\ss i\bs)) < \ell(\perm\bs)$, \ie, $i$ is
583: a descent of¨$\perm\bs\inv$.\\
584: Symmetrically, the following are equivalent:
585:
586: $(i')$ The braid¨$\ss i$ is a right divisor of¨$\bs$
587: in¨$B_\nn^+$, \ie, $i$ belongs to¨$\DivR(\bs)$;
588:
589: $(ii')$ The strands finishing at positions¨$i$ and¨$i+1$
590: cross in any (positive) diagram for¨$\bs$;
591:
592: $(iii')$ We have $\pi(\bs)(i) > \pi(\bs)(i+1)$;
593:
594: $(iv')$ We have $\ell(\pi(\bs\ss i)) < \ell(\perm\bs)$, \ie, $i$ is
595: a descent of¨$\perm\bs$.
596: \end{lemm}
597:
598: So, for¨$\bs$ a simple braid, the indices¨$i$ such that $\ss i$ is a right
599: divisor of¨$\bs$ are the descents of the associated permutation¨$\perm\bs$,
600: while those such that $\ss i$ is a left divisor of¨$\bs$ are the
601: descents of¨$\perm\bs\inv$.
602:
603: \subsection{The numbers $\bb\nn\dd$ and $\bbb\nn\dd\bs$}
604:
605: Our aim in this paper is to solve various counting problems
606: involving the normal form of positive braids. The main
607: numbers we investigate are as follows:
608:
609: \begin{defi}
610: For $\nn, \dd \ge 1$, we denote by¨$\bb\nn\dd$ the number
611: of positive $\nn$¨strand braids of degree at most¨$\dd$,
612: \ie, the number of divisors of¨$\DD\nn\dd$ in the braid
613: monoid.
614: \end{defi}
615:
616: By Lemma¨\ref{L:NormalBis}, $\bb\nn\dd$ is the number of
617: normal sequences of length¨$\dd$, \ie, the number of
618: sequences $(\sx_1, \dots, \sx_\dd)$ where all¨$\sx_\kk$ are
619: simple braids and $\DivL(\sx_\kk) \supseteq
620: \DivR(\sx_{\kk+1})$ holds for $\kk < \dd$. By
621: Lemma¨\ref{L:Exchange}, it is also the number of sequences
622: of permutations $(\ff_1, \dots, \ff_\dd)$ such that, for
623: each¨$\kk < \dd$, the descents of¨$\ff_{\kk+1}\inv$ are
624: included in those of¨$\ff_\kk$.
625:
626: For¨$\dd = 1$, the bijection between simple $\nn$¨strand
627: braids and permutations of¨$\Int\nn$ immediately gives
628: \begin{equation}
629: \bb\nn1 = n!,
630: \end{equation}
631: which implies for all¨$\nn, \dd$
632: \begin{equation} \label{E:Upper}
633: \bb\nn\dd \le (\nn!)^\dd.
634: \end{equation}
635:
636: In the sequel, we shall have to count normal sequences
637: satisfying some constraints. So we introduce one more
638: notation.
639:
640: \begin{defi}
641: For $\nn, \dd \ge 1$ and $\bs$ a simple $\nn$-braid, we
642: denote by¨$\bbb\nn\dd\bs$ the number of positive
643: $\nn$¨strand braids of degree at most¨$\dd$ with $\dd$th
644: factor equal to¨$\bs$.
645: \end{defi}
646:
647: In other words, $\bbb\nn\dd\bs$ is the number of normal
648: sequences of the form $(\bs_1, \dots, \bs_{\dd-1}, \bs)$.
649: Some connections are obvious:
650:
651: \begin{prop} \label{P:Total}
652: For all¨$\nn, \dd$, we have
653: \begin{equation} \label{E:Last}
654: \bb\nn\dd = \sum_{\mbox{\Small\rm $\bs$ simple}} \bbb\nn\dd\bs =
655: \bbb\nn{\dd+1}1.
656: \end{equation}
657: \end{prop}
658:
659: \begin{proof}
660: The first equality is obvious. The second one follows from
661: the fact that $(\bs_1, \dots, \bs_\dd)$ is normal if and
662: only if $(\bs_1, \dots, \bs_\dd, 1)$ is: indeed, $1$ has no
663: left divisor but itself, so, by Lemma¨\ref{L:Normal}, every
664: sequence $(\bs, 1)$ is normal.
665: \end{proof}
666:
667: \section{Adjacence matrices}
668: \label{S:Adj}
669:
670: In this section, we study the numbers $\bb\nn\dd$ and
671: $\bbb\nn\dd\bs$ when $\nn$ is fixed and $\dd$ varies. By
672: Lemma¨\ref{L:Normal}, normal sequence of simple braids are
673: characterized by a purely local criterion that only involves
674: adjacent entries. It follows that the set of all normal
675: sequences can be recognized by a finite state automaton
676: \cite{Eps}, and, as a consequence, the associated counting
677: numbers obey a linear induction rule specified by a certain
678: incidence matrix¨\cite{Eps2}. In this section, we define
679: the matrix involved in the current situation, and show how
680: its size, which is originally¨$\nn!$, can be lowered
681: to¨$2^\nno$.
682:
683: \subsection{Enumeration of simple braids}
684:
685: Below we consider matrices whose entries are indexed
686: by simple braids (or, equivalently, permutations). Fixing an
687: enumeration of simple braids is not important at a
688: conceptual level, but this is necessary when the objects
689: are to be specified explicitly. We shall use the
690: restriction of the canonical linear ordering of braids
691: denoted¨$<^{\scriptscriptstyle\phi}$
692: in¨\cite{Dgr}---which gives for each¨$\nn$ a well-ordering of
693: ordinal type¨$\omega^{\omega^{\nn-2}}$ on¨$B_\nn^+$.
694: The corresponding increasing enumeration of simple
695: $\nn$-braids can be constructed directly using induction
696: on¨$\nn$. We start from the following easy remark:
697:
698: \begin{lemm} \label{L:Factor}
699: For $1 \le i \le \nn$, write $\ss{i, \nn}$ for $\ss i
700: \ss{i+1} \dots \ss{\nno}$ (so that $\ss{\nn, \nn}$
701: is¨$1$). Then every simple $\nn$-braid¨$\bs$ admits a unique
702: decomposition $\bs = \ss{i, \nn} \, \bt$ with $1
703: \le i \le \nn$ and $\bt$ a simple $(\nno)$-braid.
704: \end{lemm}
705:
706: \begin{proof}
707: (Figure¨\ref{F:Factor})
708: Let $i = \perm\bs(\nn)$. Then we can realize¨$\bs$ by a
709: diagram in which the $i$th strand is first sent to the
710: rightmost position, and it remains a simple $(\nno)$-braid.
711: Conversely, we have $i = \perm{\ss{i, \nn}\bt}(\nn)$, so the
712: decomposition is unique.
713: \end{proof}
714:
715: \begin{figure} [htb]
716: \begin{picture}(35,20)(0, 0)
717: \put(0,0){\includegraphics{Factor.eps}}
718: \put(34,9){$\bs$}
719: \put(11.5,5){$\bt$}
720: \put(12,19.5){$i$}
721: \end{picture}
722: \caption{\smaller Proof of Lemma¨\ref{L:Factor}}
723: \label{F:Factor}
724: \end{figure}
725:
726: \begin{defi}
727: We inductively define an enumeration¨$S_\nn$ of simple
728: $\nn$-braids by
729: \begin{equation}
730: S_1:= (1), \quad
731: S_\nn:= S_{\nno} \concat \ss{\nno, \nn} S_{\nno} \concat
732: \dots \concat \ss{1, \nn} S_{\nno},
733: \end{equation}
734: where $\concat$ stands for list concatenation, and $\bs S$ is
735: the list obtained from¨$S$ by multipliying all entries
736: by¨$\bs$ on the left. The $\kk$th element in¨$\bigcup_\nn
737: S_\nn$ is denoted¨$\sp\kk$.
738: \end{defi}
739:
740: The first¨$\sp\kk$'s are, in increasing order,
741: $$\sp1 = 1\ ,
742: \sp2 = \ss1\ ,
743: \sp3 = \ss2\ ,
744: \sp4 = \ss2\ss1\ ,
745: \sp5 = \ss1\ss2\ ,
746: \sp6 = \ss1\ss2\ss1\ ,
747: \sp7 = \ss3, \dots$$
748: Lemma¨\ref{L:Factor} guarantees that all simple braids
749: occur in the above enumeration. Note that, for every¨$\nn$,
750: we have $\D_\nn = \sp{\nn!}$.
751:
752: It is easy to check that
753: the ordering of simple braids we use corresponds to a
754: reversed antilexicographic ordering of the inverses of the
755: associated permutations: $\bs$ occurs before¨$\bt$ if and
756: only if we have $\perm\bs\inv < \perm\bt\inv$, where
757: $\ff < \gg$ is said to hold if we have $\ff(i) > \gg(i)$ for
758: the largest¨$i$ for which $\ff$ and $\gg$ do not agree.
759:
760: \subsection{The matrix¨$\Mat\nn$}
761:
762: Everywhere in the sequel, we write $(M)_{\xx, \yy}$ for the
763: $(\xx, \yy)$-entry of a matrix¨$M$.
764:
765: \begin{defi}
766: For $\nn \ge 1$, we define $\Mat\nn$ to be the $\nn! \times
767: \nn!$ matrix satisfying
768: $$
769: (\Mat\nn)_{\kk, \ll} =
770: \begin{cases}
771: 1 & \mbox{if $(\sp\kk, \sp\ll)$ is normal,}\\
772: 0 & \mbox{otherwise.}
773: \end{cases}$$
774: \end{defi}
775:
776: Instead of referring to integer entries, it will be often
777: convenient to think of the entries of¨$\Mat\nn$ as directly
778: indexed by simple braids; for $\bs, \bt$ simple braids, we
779: simply write $(\Mat\nn)_{\bs, \bt}$ for the
780: corresponding entry.
781:
782: \begin{exam}
783: The first 3 matrices¨$\Mat\nn$ are
784: $$
785: \Mat1 = 1, \quad
786: \Mat2 =
787: \left(\begin{matrix}1&0\\1&1\end{matrix}\right), \quad
788: \Mat3 = \left(
789: \begin{matrix}
790: 1&0&0&0&0&0\\
791: 1&1&0&0&1&0\\
792: 1&0&1&1&0&0\\
793: 1&1&0&0&1&0\\
794: 1&0&1&1&0&0\\
795: 1&1&1&1&1&1
796: \end{matrix}
797: \right).$$
798: \end{exam}
799:
800: The construction of the matrix¨$\Mat\nn$ immediately
801: implies the following results:
802:
803: \begin{lemm} \label{L:Stack}
804: $(i)$ The first column and the last row of¨$\Mat\nn$
805: contain only¨$1$'s; the first row, its first entry
806: excepted, and the last column, its last entry excepted,
807: contain only¨$0$'s.
808:
809: $(ii)$ The first $(\nno)!$ columns of¨$\Mat\nn$ consist of
810: $\nn$¨stacked copies of¨$\Mat{\nno}$.
811:
812: $(iii)$ If $\DivR(\sp\kk) = \DivR(\sp{\kk'})$ holds, then
813: the $\kk$th and the $\kk'$th rows in¨$\Mat\nn$ coincide.
814: Similarly, if $\DivL(\sp\ll) = \DivL(\sp{\ll'})$ holds, then
815: the $\ll$th and the $\ll'$th columns in¨$\Mat\nn$ coincide.
816: \end{lemm}
817:
818: \begin{proof}
819: $(i)$ By construction, we have $\sp1 = 1$ and
820: $\sp{\nn!} = \DD\nn{}$. Now $(\bs, 1)$ is always normal,
821: and so is $(\DD\nn{}, \bs)$. On the other hand, $(1 , \bs)$
822: is normal only for $\bs = 1$, and $(\bs, \DD\nn{})$ is
823: normal only for $\bs = \DD\nn{}$.
824:
825: $(ii)$ Assume $\kk \le (\nno)!$ and $\kk' = \kk + (\nn-i)
826: \cdot (\nno)!$. Our enumeration of simple braids implies
827: $\sp{\kk'} = \ss{i, \nn} \sp\kk$. Then
828: Figure¨\ref{F:Factor} makes the equality
829: $\DivR(\sp{\kk'}) \cap \Int{\nno} = \DivR(\sp\kk)$ clear.
830: For every simple $(\nno)$-braid¨$\bt$, the
831: set¨$\DivL(\bt)$ is included in¨$\Int{\nno}$, and it
832: follows that $(\sp{\kk'}, \bt)$ is normal if and only
833: if¨$(\sp\kk, \bt)$ is. In other words, we have
834: $(\Mat\nn)_{\kk', \ll} = (\Mat\nn)_{\kk, \ll}$ for $\ll \le
835: (\nno)!$.
836:
837: $(iii)$ By Lemma¨\ref{L:Normal}, the value of
838: $(\Mat\nn)_{\kk, \ll}$ only depends on¨$\DivR(\sp\kk)$ and
839: on¨$\DivL(\sp\ll)$.
840: \end{proof}
841:
842: The connection between the numbers¨$\bbb\nn\dd\bs$ and the
843: matrix¨$\Mat\nn$ is straightforward:
844:
845: \begin{lemm} \label{L:Comput}
846: For every simple¨$\bt$ and every
847: $\dd \ge 1$, we have
848: \begin{equation} \label{E:Number}
849: \bbb\nn\dd{\bt} = ((1, 1, \dots, 1) \,
850: \Mat\nn\!\!\!{}^{\dd-1})_\bt.
851: \end{equation}
852: \end{lemm}
853:
854: \begin{proof}
855: Induction on¨$\dd$. For $\dd=1$, and for each simple
856: $\nn$-braid¨$\bs$, there is exactly one braid of
857: degree at most¨$1$ whose first factor is¨$\bt$,
858: namely $\bt$¨itself, and we have $\bbb\nn1\bt = 1$.
859: Assume $\dd \ge 2$. By Lemma¨\ref{L:Normal}, $(\bs_1,
860: \dots, \bs_{\dd-1}, \bt)$ is normal if and only if $(\bs_1,
861: \dots, \bs_{\dd-1})$ and $(\bs_{\dd-1}, \bt)$ are normal,
862: so we get
863: \begin{equation*} \label{E:Comput}
864: \bbb\nn\dd{\bt}
865: = \sum_{(\bs, \bt)\mbox{ \Small normal}}
866: \bbb\nn{\dd-1}{\bs} = \sum_\bs
867: \bbb\nn{\dd-1}{\bs}
868: \, (\Mat\nn)_{\bs, \bt},
869: \end{equation*}
870: and \eqref{E:Number} follows inductively.
871: \end{proof}
872:
873: \begin{rema}
874: As last row of¨$\Mat\nn$ is $(1, \dots, 1)$, we have
875: $(1, \dots, 1) = (0, \dots, 0, 1) \Mat\nn$, and we can
876: replace¨\eqref{E:Number} with
877: \begin{equation} \label{E:NumberBis}
878: \bbb\nn\dd{\bt} = ((0, \dots, 0, 1) \,
879: \Mat\nn\!\!\!{}^{\dd})_\bt.
880: \end{equation}
881: \end{rema}
882:
883: \begin{exam}
884: Using the value of¨$\Mat2$, we immediately
885: find $\bbb2\dd1 = \dd$, $\bbb2\dd{\ss1} = 1$, as
886: could be expected: there are $\dd+1$ braids of degree
887: at most¨$\dd$, namely the braids¨$\ss1^\kk$ with
888: $\kk < \dd$, whose $\dd$th factor is¨$1$, and
889: $\ss1^\dd$, whose $\dd$th factor is¨$\D_2$, \ie,¨$\ss1$.
890: \end{exam}
891:
892: The computation for $\nn \ge 3$ is more complicated, and we
893: postpone it. For the moment, we just point that, as the
894: numbers¨$\bbb\nn\dd\bs$ obey the linear
895: recurrence¨\eqref{E:Number}, standard arguments imply that
896: they can be expressed in terms of the eigenvalues
897: of¨$\Mat\nn$:
898:
899: \begin{prop} \label{P:Expansion}
900: Let $\ev_1$, \dots, $\ev_\ir$ be the non-zero
901: eigenvalues of¨$\Mat\nn$. Then, for
902: each simple $\nn$-braid¨$\bs$, there
903: exist polynomials $P_1, \dots, P_\ir$ with $\deg(P_\ind)$
904: at most the multiplicity of¨$\ev_\ind$ for¨$\Mat\nn$ such
905: that, for each¨$\dd \ge 0$, we have
906: \begin{equation} \label{E:General}
907: \bbb\nn\dd\bs = P_1(\dd) \ev_1^\dd + \cdots +
908: P_\ir(\dd) \ev_\ir^\dd.
909: \end{equation}
910: \end{prop}
911:
912: \begin{coro}
913: For all¨$\nn, \bs$, the generating function of the
914: numbers¨$\bbb\nn\dd\bs$'s with respect to¨$\dd$ is rational.
915: \end{coro}
916:
917: \subsection{Reducing the size}
918:
919: The size¨$\nn!$ of the incidence¨$\Mat\nn$ is uselessly
920: large, and we shall see now how to lower it. This will be
921: done in two steps. The first one relies on the fact,
922: pointed out in Lemma¨\ref{L:Stack}$(iii)$, that many columns
923: in¨$\Mat\nn$ are equal. For subsequent use, it will be
924: useful to introduce a new sequence of numbers:
925:
926: \begin{defi}
927: For $\II, \JJ \ince \Int{\nno}$, we denote
928: by $\aa\nn\II\JJ$ (\resp $\aas\nn\II\JJ$) the
929: number of simple $\nn$-braids satisfying $\DivL(\bs) =
930: \II$ (\resp $\DivL(\bs) \supseteq \II$) and
931: $\DivR(\bs) \supseteq \JJ$.
932: \end{defi}
933:
934: \begin{lemm} \label{L:Reduction}
935: For $\nn \ge 1$, let $\Matt\nn$ be the $2^\nno \times
936: 2^\nno$ matrix with entries indexed by subsets
937: of¨$\Int{\nno}$ defined by $(\Matt\nn)_{\II, \JJ} =
938: \aa\nn\II\JJ$. Then the characteristic polynomials
939: of¨$\Matt\nn$ and $\Mat\nn$ coincide up to a power of¨$\xx$,
940: and, for every simple¨$\bt$ with $\DivL(\bt) = \JJ$ and
941: every $\dd \ge 1$, we have
942: \begin{equation} \label{E:Numberr}
943: \bbb\nn\dd{\bt} = ((1, 1, \dots, 1) \,
944: \Matt\nn\!{}^{\dd-1})_\JJ.
945: \end{equation}
946: \end{lemm}
947:
948: \begin{proof}
949: Gathering the columns corresponding to simples with
950: the same¨$\DivL$¨set and summing the corresponding
951: lines amounts to replacing¨$\Mat\nn$ with a similar
952: matrix of the form
953: $\left(\begin{matrix}
954: \Matt\nn&0\\\dots&0 \end{matrix}\right)$, so the
955: result about the characteristic polynomial is clear.
956:
957: As for the value of¨$\bbb\nn\dd\bt$, the argument
958: is similar to that for Lemma¨\ref{L:Comput}. The
959: induction starts as $\aa\nn\II{\Int{\nno}} =
960: 1$ holds for each¨$\II$. For the general step, we
961: find
962: \begin{align*} \label{E:Comput}
963: \bbb\nn\dd{\bt}
964: &= \sum_{(\bs, \bt)\mbox{ \Small normal}}
965: \bbb\nn{\dd-1}{\bs}
966: = \sum_{\DivR(\bs) \supseteq\JJ}
967: \bbb\nn{\dd-1}{\bs}
968: = \sum_\II \sum_{
969: \underset{\scriptstyle \DivL(\bs) = \II}
970: {\scriptstyle \DivR(\bs) \supseteq\JJ}}
971: \bbb\nn{\dd-1}{\bs} \\
972: &= \sum_\II \bbbb\nn{\dd-1}{\II} \aa\nn\II\JJ
973: = \sum_\II ((1, \dots, 1)
974: \Matt\nn{}^{\dd-2})_{\II} (\Matt\nn)_{\II, \JJ}
975: = ((1, \dots, 1)
976: \Matt\nn{}^{\dd-1})_{\JJ},
977: \end{align*}
978: where $\bbbb\nn\dd\II$ denotes the common
979: value of $\bbb\nn\dd\bs$ for $\bs$ with¨$\DivL(\bs) =
980: \II$.
981: \end{proof}
982:
983: For $\nn = 3$, and using the enumeration $\emptyset$,
984: $\{1\}$, $\{2\}$, $\{1,2\}$ that is induced by our
985: enumeration of simple braids, we obtain
986: $\Matt3 = \left(\begin{matrix}
987: 1&0&0&0\\2&1&1&0\\
988: 2&1&1&0\\
989: 1&1&1&1
990: \end{matrix}\right)$. Observe that the second
991: and third columns in¨$\Matt3$ coincide, which suggests a
992: further reduction step.
993:
994: \section{Partitions associated with a simple braid}
995: \label{S:Part}
996:
997: We can indeed reduce the size of the matrices once more:
998: we can replace the
999: incidence matrix¨$\Matt\nn$ with a new matrix¨$\Mattt\nn$,
1000: whose size is¨$p(\nn)$, the number of partitions of¨$\nn$.
1001: Here the result is deduced from elementary remarks
1002: about simple braids (or, equivalently, about
1003: permutations); it can also be deduced from classical
1004: results about Solomon's algebra of descents---and
1005: therefore extends to all Artin--Tits groups of
1006: spherical type.
1007:
1008: \subsection{Computation of¨$\aas\nn\II\JJ$}
1009:
1010: We shall start from an explicit determination of the value
1011: of the numbers¨$\aas\nn\II\JJ$ in terms of the block
1012: compositions of¨$\II$ and¨$\JJ$. We first recall the notions
1013: of composition and partition.
1014:
1015: \begin{defi}
1016: For $\II \ince \Int{\nno}$, the {\it
1017: $\nn$-composition}¨$\comp\nn\II$ of~$\II$ is defined to be
1018: the sequence $(\ip_1, \ip_2-\ip_1,
1019: \dots,
1020: \ip_\ir-\ip_{\ir-1})$, where
1021: $\ip_1, \dots, \ip_\ir$ is the increasing
1022: enumeration of $\Int\nn \setminus \II$. The {\it
1023: $\nn$-partition}¨$\part\nn\II$ of~$\II$ is the
1024: non-increasing rearrangement of¨$\comp\nn\II$.
1025: \end{defi}
1026:
1027: \begin{exam}
1028: The composition of~$\II$ consists of the sizes of the blocks
1029: of adjacent elements in~$\II$, augmented by¨$1$: for
1030: instance, the $10$-composition of
1031: $\{1,2,4,5,6,9\}$ is¨$(3, 4, 1, 2)$, as the blocks are of
1032: size¨$2, 3, 0, 1$. The $10$-partition of~$\II$ is therefore
1033: $(4, 3, 2, 1)$. Note that the $\nn$-composition of~$\II$
1034: determines~$\II$, but its $\nn$-partition does not.
1035: \end{exam}
1036:
1037: The geometric observation is the following one:
1038:
1039: \begin{lemm} \label{L:Sequence}
1040: Assume $\JJ \ince \Int{\nno}$, and let
1041: $(\iq_1, \dots, \iq_\is)$ be the $\nn$-composition
1042: of¨$\JJ$. For $\bs$ a simple $\nn$-braid, define
1043: $\ff_\bs : \Int\nn \to \Int\is$ by $\ff_\bs(i) = j$
1044: if the $i$th strand of¨$\bs$ finishes in the
1045: $j$th block of¨$\JJ$. Then $\bs \mapsto \ff_\bs$
1046: establishes a bijection between the simple
1047: $\nn$-braids¨$\bs$ that satisfy
1048: $\DivR(\bs) \supseteq \JJ$ and the functions
1049: from¨$\Int\nn$ to¨$\Int\is$. Moreover, for
1050: $\DivR(\bs) \supseteq \JJ$, we have, for every¨$i$,
1051: \begin{equation} \label{E:Equiv}
1052: i \in \DivL(\bs)
1053: \quad \Longleftrightarrow \quad
1054: \ff_\bs(i) \ge \ff_\bs(i+1).
1055: \end{equation}
1056: \end{lemm}
1057:
1058: \begin{proof}
1059: Assume that $\bs$ is a simple $\nn$-braid
1060: satisfying $\DivR(\bs) \supseteq \JJ$, \ie, that is
1061: right divisible by¨$\ss j$ for each¨$j$ in¨$\JJ$.
1062: By hypothesis, the first block of consecutive
1063: elements of¨$\JJ$ is $1, \dots \iq_1-1$. Then $\bs$
1064: being right divisible by $\ss1$, \dots,
1065: $\ss{\iq_1-1}$ is equivalent to its being right
1066: divisible by the left lcm of these elements, which
1067: is¨$\D_{\iq_1}$. Similarly, the second block
1068: in¨$\JJ$ is $\ss{\iq_1+1}$, \dots, $\ss{\iq_1 +
1069: \iq_2-1}$, and being right divisible by these
1070: elements amounts to being right divisible by their
1071: left lcm, which is $\sh^{\iq_1}(\D_{\iq_2})$, where
1072: $\sh$ denotes the shift endomorphism of¨$B_\infty$
1073: that maps each¨$\ss i$ to¨$\ss{i+1}$. Now, by
1074: construction, the blocks $\{1, \dots, \iq_1-1\}$
1075: and $\{\iq_1+1 \dots, \iq_1 + \iq_2-1\}$ are
1076: separated by¨$\iq_1$, and, therefore, the
1077: corresponding¨$\s$'s commute. In particular,
1078: $\D_{\iq_1}$ and $\sh^{iq_1}(\D_{\iq_2})$ commute,
1079: and their left lcm is their product. Finally, a
1080: simple braid¨$\bs$ satisfies
1081: $\DivR(\bs) \supseteq \JJ$ if and only if it is a
1082: right multiple of the element
1083: $$\D_\JJ = \D_{\iq_1} \, \sh^{\iq_1}(\D_{\iq_2}) \,
1084: \dots \, \sh^{\iq_1 + \cdots +
1085: \iq_{\is-1}}(\D_{\iq_\is}),$$
1086: \ie, we have $\bs = \bs' \D_\JJ$ for some¨$\bs'$.
1087:
1088: We claim that $\ff_\bs$ determines¨$\bs'$,
1089: hence¨$\bs$. Indeed, in a simple braid, any two
1090: strands cross at most once. Now, in a $\D$-diagram,
1091: any two strands cross. So, if the $i$th and the
1092: $i'$th strands go to the same block of¨$\JJ$, \ie,
1093: if we have $\ff_\bs(i) = \ff_\bs(i')$, then these
1094: strands cross in the final¨$\D$-part, and therefore
1095: they cannot cross in (any diagram
1096: representing)¨$\bs'$. So, when
1097: $\ff_\bs$ is given, there is only one way to
1098: construct¨$\bs'$, namely taking the strands to the
1099: entrance of the specified $\D$-block in increasing
1100: order (Figure¨\ref{F:Blocks}).
1101:
1102: Consider now¨$i$, $1 \le i < \nn$. We wonder
1103: whether $\ss i$ is a left divisor of¨$\bs$, \ie, if
1104: the $i$th and the $i + 1$st strands cross in the
1105: diagram of¨$\bs$. If we have $\ff_\bs(i) =
1106: \ff_\bs(i+1)$, the $i$th and $i+1$st strand go to
1107: the same $\JJ$-block, where they certainly
1108: cross. If we have $\ff_\bs(i) >
1109: \ff_\bs(i+1)$, then the $i$th strand goes to a
1110: $\JJ$-block on the right of the $\JJ$-block to
1111: which the $i+1$st strand goes, so they must cross
1112: in the $\bs'$ part. On the contrary, for
1113: $\ff_\bs(i) < \ff_\bs(i+1)$, the strands cannot
1114: cross in the $\bs'$ part---if they crossed once,
1115: they would have to cross a second time before
1116: exiting, and this is forbidden---and they do not
1117: cross in the $\D$ part either. So \eqref{E:Equiv}
1118: holds.
1119: \end{proof}
1120:
1121: \begin{figure} [htb]
1122: \begin{picture}(92,44)(0, 0)
1123: \put(0,4){\includegraphics{Blocks.eps}}
1124: \put(75,32){the $\bs'$ part}
1125: \put(75,14){the $\D$ part}
1126: \put(1,15){$\D_{\iq_1}$}
1127: \put(46,15){$\D_{\iq_\is}$}
1128: \put(0,0){first block of¨$\JJ$}
1129: \put(30,0){$\dots$}
1130: \put(55,35){$\dots$}
1131: \put(45,0){last block of¨$\JJ$}
1132: \end{picture}
1133: \caption{\smaller A simple braid divisible by all
1134: $\ss j$'s with $j$ in¨$\JJ$ can be represented by a
1135: diagram finishing with $\D$'s corresponding to the
1136: blocks of¨$\JJ$; the strands going to the same block
1137: cannot cross in the $\bs'$ part, as they cross
1138: inside the block; so the strands starting from¨$i$
1139: and¨$i'$ with $i < i'$ cross if and only if they
1140: go to different blocks and $i$ goes to a block on
1141: the right of the block $i'$ goes to.}
1142: \label{F:Blocks}
1143: \end{figure}
1144:
1145: We deduce the following characterization
1146: of¨$\aas\nn\II\JJ$:
1147:
1148: \begin{prop} \label{P:EES}
1149: Assume that $\II, \JJ$ are subsets of~$\Int{\nno}$
1150: with respective $\nn$-composi\-tions $(\ip_1, \dots,
1151: \ip_\ir)$ and $(\iq_1, \dots, \iq_\is)$. Then $\aas\nn\II\JJ$
1152: is the number of $\ir \times \is$ matrices
1153: with nonnegative integer entries such that, for
1154: all~$i, j$, the $i$th row has sum~$\ip_i$ and the
1155: $j$th~column has sum~$\iq_j$. In particular, we have
1156: \begin{equation} \label{E:Value}
1157: \aas\nn\II\emptyset = \frac{\nn!}{\ip_1! \dots
1158: \ip_\ir!}
1159: \mbox{\quad and \quad}
1160: \aas\nn\emptyset\JJ = \frac{\nn!}{\iq_1! \dots
1161: \iq_\is!}.
1162: \end{equation}
1163: \end{prop}
1164:
1165: \begin{proof}
1166: Lemma¨\ref{L:Sequence} immediately implies
1167: \begin{gather}
1168: \label{E:Sequence}
1169: \aa\nn\II\JJ = \card
1170: \{\ff \!:\! \Int\nn \!\to\! \Int\is \,;\,
1171: (\forall j)(\card\ff\inv(j) = \iq_j)
1172: \,\&\,
1173: (i \in \II \Leftrightarrow \ff(i) \ge \ff(i+1))\},\\
1174: \label{E:SequenceBis}
1175: \aas\nn\II\JJ = \card\{\ff \!:\! \Int\nn \!\to\!
1176: \Int\is \,;\, (\forall j)(\card\ff\inv(j) = \iq_j)
1177: \,\&\,
1178: (i \in \II \Rightarrow \ff(i) \ge \ff(i+1))\}.
1179: \end{gather}
1180:
1181:
1182:
1183: Assume that $\ff$ is a function of~$\Int\nn$
1184: to~$\Int\is$ satisfying the constraints
1185: of~\eqref{E:SequenceBis}. Let $A_\ff$ be the $\ir
1186: \times \is$-matrix whose $(i, j)$-entry is the
1187: number of~$k$'s in the $i$th block of~$\II$
1188: satisfying~$\ff(k) = \iq_j$. By construction, the
1189: the sum of the $i$th row of~$A_\ff$ is the
1190: size~$\ip_i$ of the $i$th block of~$\II$, while
1191: the sum of the $j$th column is the number of~$k$'s
1192: satisfying $\ff(k) = j$, \ie, it is~$\iq_j$. We
1193: claim that $A_\ff$ determines~$\ff$. Indeed,
1194: \eqref{E:SequenceBis} requires that $\ff$ be
1195: non-increasing on each block of~$\II$, so there is
1196: only one possibility once the number of~$k$'s
1197: going to the various~$j$ is fixed.
1198:
1199: The first equality in¨\eqref{E:Value} follows:
1200: for¨$\is = \nn$, there is exactly one nonzero entry
1201: in each column, so choosing a convenient matrix
1202: amounts to choosing among $\nn$~elements the
1203: $\iq_1$~columns with a¨$1$ in the first row, the
1204: $\iq_2$~columns with a¨$1$ in the second row, \etc\
1205: The second equality is similar with rows and
1206: columns exchanged.
1207: \end{proof}
1208:
1209: \begin{coro} \label{C:Depends}
1210: $(i)$ The number¨$\aas\nn\II\JJ$ only depends on the
1211: partitions¨$\part\nn\II$ and¨$\part\nn\JJ$.
1212:
1213: $(ii)$ For each~$\nn$ and¨$\II$, the number¨$\aa\nn\II\JJ$
1214: only depends on the partition¨$\part\nn\JJ$.
1215: \end{coro}
1216:
1217: \begin{proof}
1218: Point¨$(i)$ directly follows from the characterization of
1219: Proposition¨\ref{P:EES}, as the latter clearly involves the
1220: sizes of the blocks of¨$\II$ and¨$\JJ$ only. As for¨$(ii)$,
1221: the usual inclusion-exclusion formula gives
1222: $$\aa\nn\II\JJ = \sum_{\KK \cap\II = \emptyset}
1223: (-1)^{\card\KK} \aas\nn{\II \cup \KK}\JJ.$$
1224: By¨$(i)$, each term in the sum only depends
1225: on¨$\part\nn\JJ$, and so does the sum.
1226: \end{proof}
1227:
1228: It is easy to check that the value of¨$\aa\nn\II\JJ$ does
1229: not only depend on¨$\part\nn\II$ in
1230: general: when we apply the inclusion-exclusion formula,
1231: the sizes of the blocks in¨$\II \cup \KK$ do not only depend
1232: on the sizes of the block in¨$\II$.
1233:
1234: \begin{rema} \label{R:Hohlweg}
1235: Corollary¨\ref{C:Depends} can also be deduced from the classical results by
1236: Solomon about the descent algebra---and, therefore, it extends to all
1237: Artin--Tits groups of spherical type. The argument is as follows. For¨$f$ a
1238: permutation, let
1239: $D(f)$ denote the sets of descents of¨$f$. In the group
1240: algebra¨$\QQ[\Sym\nn]$, let $d_I:= \sum\{f; D(f) = I\}$ and
1241: $e_\JJ= \sum\{f; D(f) \cap \JJ = \emptyset\}$. Using $w_0$
1242: for the flip permutation, we have $w_0 e_\JJ = \sum\{f; D(f) \supseteq
1243: \JJ\}$, and therefore $\aa\nn\II\JJ = \langle d_\II, w_0e_\JJ\rangle$,
1244: where $\langle.,.\rangle$ is the inner product defined by $\langle f,
1245: g\rangle = 1$ for $g = f\inv$, and¨$=0$ otherwise. Using the isometry result
1246: of¨\cite{Hoh}, we deduce $\aa\nn\II\JJ = \langle \theta(\dd_\II),
1247: \theta(w_0e_\JJ)\rangle$, where $\theta(e_\KK)$ denotes the character
1248: of¨$\Sym\nn$ induced by the trivial character of the standard parabolic
1249: subgroup generated by (the transpositions¨$s_i$ with¨$i$ in)¨$\KK$.
1250: By¨\cite{Sol}, the subspace of¨$\QQ[\Sym\nn]$ generated by the¨$e_\KK$ is
1251: a subalgebra, and the kernel of¨$\theta$ is generated by the elements
1252: $\dd_\II - \dd_{\JJ}$ with
1253: $\II, \JJ$ associated with the same partition, and it follows from the above
1254: expression that $\aa\nn\II\JJ$ only depends on the partition
1255: associated with¨$\JJ$.
1256: \end{rema}
1257:
1258:
1259: \subsection{The matrix¨$\Mattt\nn$}
1260:
1261: We can now come back to the matrix¨$\Matt\nn$, and
1262: replace it for the computation of the numbers¨$\bb\nn\dd$
1263: with a new matrix of smaller size. Indeed, a direct
1264: application of Corollary¨\ref{C:Depends} is
1265:
1266: \begin{lemm}
1267: Assume that $\JJ$, $\JJ'$ are subsets of¨$\Int\nno$ with
1268: the same $\nn$-partition. Then the $\JJ$th and $\JJ'$th
1269: columns of¨$\Matt\nn$ are equal.
1270: \end{lemm}
1271:
1272: Thus the process used to replace¨$\Mat\nn$ with¨$\Matt\nn$
1273: can be applied again, \ie, the form a new matrix by
1274: gathering the equal columns and summing the corresponding
1275: rows.
1276:
1277: \begin{defi}
1278: $(i)$ For¨$\l$ a partition (or a composition) of¨$\nn$, we
1279: denote by¨$\set\l$ the unique subset¨$\II$ of¨$\Int\nno$
1280: satisfying $\comp\nn\II = \l$.
1281:
1282: $(ii)$ For $\l, \m \vdash \nn$ (\ie, partitions of¨$\nn$),
1283: we put
1284: $$\aab\l\m
1285: = \sum_{\part\nn\II = \l} \aa\nn\II{\set\m}
1286: = \card\{\bs \,;\, \part\nn{\DivL(\bs)} = \l
1287: \,\&\, \DivR(\bs) \supseteq \set\m\},$$
1288: and we let $\Mattt\nn$ be the matrix with rows and columns
1289: indexed by partitions of¨$\nn$ and whose $(\l, \m)$-entry
1290: is¨$\aab\l\m$.
1291: \end{defi}
1292:
1293: In this way the size of the matrix has been reduced
1294: from¨$\nn!$ to¨$\nbpart\nn$, the number of partitions
1295: of¨$\nn$. For instance, enumerating partitions in the order
1296: induced by the previous order on $\Pw(\Int\nn)$, we obtain
1297: $$\Mattt3 =
1298: \left(\begin{matrix}
1299: 1&0&0\\4&2&0\\1&1&1
1300: \end{matrix}\right), \quad
1301: \Mattt4 =
1302: \left(\begin{matrix}1&0&0&0&0\\11&4&1&0&0\\5&3&2&1&0\\
1303: 6&4&2&2&0\\1&1&1&1&1
1304: \end{matrix}\right), \quad
1305: \Mattt5 =
1306: \left(\begin{matrix}
1307: 1&0&0&0&0&0&0\\
1308: 26&8&0&2&0&0&0\\
1309: 23&12&4&5&0&1&0\\
1310: 43&21&5&10&0&2&0\\
1311: 8&6&4&4&2&2&0\\
1312: 18&12&6&8&2&4&0\\
1313: 1&1&1&1&1&1&1
1314: \end{matrix}\right).$$
1315:
1316: Applying the same argument as for
1317: Lemma¨\ref{L:Reduction}, we obtain:
1318:
1319: \begin{prop} \label{P:NNNumber}
1320: For $\nn \ge 1$, the characteristic polynomials
1321: of¨$\Mattt\nn$ and $\Mat\nn$ coincide up to a power of¨$\xx$,
1322: and, for every simple $\nn$-braid¨$\bs$, we have for
1323: each¨$\dd$
1324: \begin{equation}
1325: \bbb\nn\dd{\bs} = ((1,1, \dots, 1) \,
1326: \Mattt\nn^{\dd-1})_\l,
1327: \end{equation}
1328: where $\l$ is the $\nn$-partition of¨$\DivL(\bs)$.
1329: In particular we have
1330: \begin{equation}
1331: \bb\nn\dd = ((1,1, \dots, 1) \,
1332: \Mattt\nn^{\dd})_{(1, 1, \dots, 1)}.
1333: \end{equation}
1334: \end{prop}
1335:
1336: Table¨\ref{T:Values} gives the first few values deduced
1337: from the above formulas.
1338:
1339: \begin{table}[htb]
1340: $$\begin{tabular}{c|r|r|r|r|r|r}
1341: \quad$\dd$
1342: &1&2&3&4&5&6\\
1343: \hline
1344: \vrule width0pt height12pt depth6pt
1345: $\bbb2\dd1$
1346: & 1
1347: & 2 & 3 & 4 & 5 & 6 \\
1348: \hline
1349: \vrule width0pt height12pt
1350: $\bbb3\dd1$
1351: & 1
1352: & 6 & 19 & 48 & 109 & 234 \\
1353: \vrule width0pt depth6pt
1354: $\bbb3\dd{\D_2}$
1355: & 1
1356: & 3 & 7 & 15 & 31 & 63 \\
1357: \hline
1358: \vrule width0pt height12pt
1359: $\bbb4\dd1$
1360: & 1
1361: &24&211&1\,380&8\,077&45\,252
1362: \\
1363: $\bbb4\dd{\D_2}$
1364: & 1
1365: &12&83&492&2\,765&15\,240\\
1366: \vrule width0pt depth6pt
1367: $\bbb4\dd{\D_3}$
1368: & 1
1369: &4&15&64&309&1\,600 \\
1370: \hline
1371: \vrule width0pt height12pt
1372: $\bbb5\dd1$
1373: & 1
1374: & 120 & 3\,651 & 79\,140
1375: & 1\,548\,701
1376: & 29\,375\,460
1377: \\
1378: $\bbb5\dd{\D_2}$
1379: & 1
1380: & 60 & 1\,501 & 30\,540
1381: & 585\,811
1382: & 11\,044\,080
1383: \\
1384: $\bbb5\dd{\D_3}$
1385: & 1
1386: & 20 & 311 & 5\,260
1387: & 94\,881
1388: & 1\,755\,360
1389: \\
1390: \vrule width0pt depth6pt
1391: $\bbb4\dd{\D_4}$
1392: & 1
1393: & 5 & 31 & 325
1394: & 4\,931
1395: & 86\,565
1396: \\
1397: \hline
1398: \vrule width0pt height12pt
1399: $\bbb6\dd1$
1400: & 1
1401: & 720
1402: & 90\,921
1403: & 7\,952\,040
1404: & 634\,472\,921
1405: & 49\,477\,263\,360
1406: \\
1407: $\bbb6\dd{\D_2}$
1408: & 1
1409: & 360
1410: & 38\,559
1411: & 3\,228\,300
1412: & 254\,718\,389
1413: & 19\,808\,530\,620\\
1414: $\bbb6\dd{\D_3}$
1415: & 1
1416: & 120 & 8\,727 & 649\,260
1417: & 49\,654\,757
1418: & 3\,831\,626\,580 \\
1419: $\bbb6\dd{\D_4}$
1420: & 1
1421: & 30 & 1\,075 & 61\,620
1422: & 4\,387\,195
1423: & 332\,578\,230 \\
1424: $\bbb6\dd{\D_5}$
1425: & 1
1426: & 6 & 63 & 1\,955
1427: & 116\,423
1428: & 8\,448\,606
1429: \end{tabular}$$
1430: \medskip
1431: \caption{\smaller First values of $\bbb\nn\dd{\D_\rr}$
1432: for $1 \le \rr < \nn$; the values of¨$\bb\nn\dd$ can be also
1433: read, as Proposition¨\ref{P:Total} gives $\bb\nn\dd =
1434: \bbb\nn{\dd+1}1$, so, for instance, we find $\bb33= 48$, and
1435: $\bb45 = 45\,252$.}
1436: \label{T:Values}
1437: \end{table}
1438:
1439: \subsection{Small values of¨$\nn$}
1440:
1441: For small values of¨$\nn$, it is easy to complete the
1442: computations and to obtain an explicit form for the
1443: expansion of¨$\bbb\nn\dd\bs$ announced in
1444: Proposition¨\ref{P:Expansion}.
1445:
1446: \begin{exam}
1447: Assume $\nn = 3$. The matrix¨$\Mattt3$ is invertible with
1448: eigenvalues¨$1$ (double) and¨$2$. By solving the
1449: recurrences, we find
1450: \begin{equation*} %\label{E:Number3}
1451: \bbb3\dd\bs =
1452: \begin{cases}
1453: 4 \cdot 2^\dd - 3\dd -4
1454: &\mbox{¨for $\bs$ with partition $(1, 1, 1)$, \ie,
1455: $\bs = 1$}\\
1456: 2^\dd - 1,
1457: &\mbox{¨for $\bs$ with partition $(2, 1)$,
1458: \ie, $\bs = \ss1, \ss2, \ss2\ss1$, or $\ss1\ss2$,}\\
1459: 1
1460: &\mbox{¨for $\bs$ with partition $(3)$, \ie, $\bs =
1461: \D_3$,}
1462: \end{cases}
1463: \end{equation*}
1464: and we deduce $\bb3\dd = 8 \cdot 2^\dd - 3\dd - 7$.
1465: \end{exam}
1466:
1467: \begin{exam}
1468: Assume now $\nn = 4$. The matrix¨$\Mattt4$ admits 4
1469: eigenvalues, namely those of¨$\Mattt3$, plus $\ev_1 = 3 +
1470: \sqrt6$ and $\ev_2 = 3 - \sqrt6$. Solving the recurrences
1471: yields for¨$\bbb\nn\dd\bs$ with associated partition as
1472: indicated
1473: $$\begin{array}{lccrccrccrc}
1474: (1,1,1,1):
1475: &\frac1{20}(18 + 7\sqrt6)
1476: &\hspace{-3mm}\ev_1^\dd
1477: &+
1478: &\frac1{20}(18-7\sqrt6)
1479: &\hspace{-3mm}\ev_2^\dd
1480: &-
1481: &\frac{256}5
1482: &\hspace{-3mm}\cdot\, 2^\dd
1483: &+
1484: &6k + 11,\\
1485: (2,1,1):
1486: &\frac1{60}(18 + 7\sqrt6)
1487: &\hspace{-3mm}\ev_1^\dd
1488: &+
1489: &\frac1{60}(18-7\sqrt6)
1490: &\hspace{-3mm}\ev_2^\dd
1491: &-
1492: &\frac85
1493: &\hspace{-3mm}\cdot\, 2^\dd
1494: &+
1495: &1,\\
1496: (2,2):
1497: &\frac1{12}(\sqrt6)
1498: &\hspace{-3mm}\ev_1^\dd
1499: &-
1500: &\frac1{12}(\sqrt6)
1501: &\hspace{-3mm}\ev_2^\dd,\\
1502: (3,1):
1503: &\frac1{60}(6 -\sqrt6)
1504: &\hspace{-3mm}\ev_1^\dd
1505: &+
1506: &\frac1{60}(6+\sqrt6)
1507: &\hspace{-3mm}\ev_2^\dd
1508: &+
1509: &\frac45
1510: &\hspace{-3mm}\cdot\, 2^\dd
1511: &-
1512: &1,\end{array}
1513: $$
1514: and¨$1$ for associated partition $(4)$, \ie, for $\bs =
1515: \D_4$. As the characteristic polynomial of¨$\Mattt4$ is
1516: $(x^2 - 6x + 3)(x - 2)(x-1)^2$,
1517: we can equivalently determine $\bbb4\dd\bs$ and $\bb4\dd$ by
1518: inductions on¨$\dd$ of the form
1519: \begin{equation} \label{E:Recur}
1520: \var_\dd = 6\var_{\dd-1} - 3\var_{\dd-2}
1521: + \alpha 2^\dd + \beta \dd + \gamma
1522: \end{equation}
1523: where $\alpha, \beta, \gamma$ are determined using special
1524: values of¨$\var_\dd$. For instance, $\bb4\dd$ is determined
1525: by¨\eqref{E:Recur} with $\alpha = 32$, $\beta = -12$,
1526: $\gamma = -34$ and the values $\var_{-1} = 0$, $\var_0= 1$.
1527: Generating functions can be deduced easily.
1528: \end{exam}
1529:
1530: \subsection{Eigenvalues of $\Mat\nn$}
1531:
1532: By Proposition¨\ref{P:Expansion}, the value of¨$\bb\nn\dd$
1533: and¨$\bbb\nn\dd\bs$, and in particular its asymptotic
1534: behaviour when $\dd$ grows to infinity, are connected with
1535: the non-zero eigenvalues of¨$\Mat\nn$, which, by
1536: Proposition¨\ref{P:NNNumber}, coincide with those
1537: of¨$\Mattt\nn$. The characteristic polynomial
1538: of¨$\Mattt\nn$---hence of¨$\Mat\nn$ up to an
1539: $x^\dd$¨factor---for small values of¨$\nn$ is displayed in
1540: Table¨\ref{T:Charpol}.
1541:
1542: \begin{table}[t]
1543: \begin{tabbing}
1544: \hspace{10mm}\=\hspace{85mm}\=\hspace{2cm}\kill
1545: \> $\charpol{\Mattt1}
1546: = x - 1$\\
1547: \> $\charpol{\Mattt2}
1548: = \charpol{\Mattt1} \cdot (x - 1)$\\
1549: \> $\charpol{\Mattt3}
1550: = \charpol{\Mattt2} \cdot (x - 2)$\\
1551: \> $\charpol{\Mattt4}
1552: = \charpol{\Mattt3} \cdot (x^2 - 6x +3)$\\
1553: \> $\charpol{\Mattt5}
1554: =\charpol{\Mattt4} \cdot (x^2 - 20 x +24)$\\
1555: \> $\charpol{\Mattt6}
1556: = \charpol{\Mattt5}
1557: \cdot (x^4 - 82 x^3 +359x^2 - 260x + 60)$\\
1558: \> $\charpol{\Mattt7}
1559: = \charpol{\Mattt6} \cdot
1560: (x^4 - 390 x^3 + 6024 x^2 - 13680 x + 8640)$\\
1561: \> $\charpol{\Mattt8}
1562: = \charpol{\Mattt7}
1563: \cdot (x^7 - 2134 x^6 + 139976 x^5 - 1321214 x^4
1564: + 3780975 x^3$\\
1565: \> \hspace{6cm}$- 3305160 x^2 + 1341900 x - 226800)$
1566: \end{tabbing}
1567: \begin{tabular}{c|c|c|c|c|c|c|c|c}
1568: $\nn$&1&2&3&4&5&6&7&8\\
1569: \hline
1570: \vline width0pt height10pt depth6pt
1571: $\ev_{max}(\Mat\nn)$&1&1&2&5.449 &18.717
1572: &77.405 &373.990 &2066.575 \\
1573: $\displaystyle\frac{\ev_{max}(\Mat\nn)}{\nn\cdot\ev_{max}(\Mat\nno)}$
1574: &-&0.5& 0.667 &0.681 &0.687 &0.689 &0.690 &0.691\\
1575: \end{tabular}
1576: \bigskip
1577: \caption{\smaller Characteristic polynomial
1578: of¨$\Mattt\nn$ for $\nn \le 8$, and the
1579: corresponding largest eigenvalue---which is to be
1580: compared with¨$\nn!$, the growth rate
1581: for the number of $\nn$-braids of degree at
1582: most¨$\dd$ if all sequences were normal.}
1583: \label{T:Charpol}
1584: \end{table}
1585:
1586: These values support the following
1587:
1588: \begin{conj} \label{C:Inclusion}
1589: For each¨$\nn$, the characteristic polynomial
1590: of¨$\Mat\nno$ divides that of¨$\Mat\nn$. More precisely,
1591: the sprectrum of¨$\Mattt\nn$ is the spectrum
1592: of¨$\Mattt\nno$, plus $p(\nn) - p(\nno)$ simple non-zero
1593: eigenvalues.
1594: \end{conj}
1595:
1596: It is not hard to check the above statement for $\nn \le
1597: 10$. Specifically, let¨$\widehat M_\nn$ be the size
1598: $p(\nn)-2$ matrix obtained from¨$\Mattt\nn$ by deleting the
1599: first and the last rows and columns. For each small value
1600: of¨$\nn$, one can directly check that $\widehat M_\nn$ is
1601: similar to a matrix of the form $\left(\begin{matrix}
1602: \widehat{M}_\nno&0\\\dots&\dots\end{matrix}\right)$, and
1603: deduce the properties asserted in
1604: Conjecture¨\ref{C:Inclusion}. But ne generic argument is
1605: known so far.
1606:
1607: The growth rate of the numbers¨$\bbb\nn\dd\bs$
1608: is connected with the largest eigenvalue
1609: $\ev_{max}(\Mat\nn)$ of¨$\Mat\nn$. For¨$\nn \le 6$,
1610: all¨$\bbb\nn\dd\bs$ except¨$\bbb\nn\dd{\D_\nn}$,
1611: which is¨$1$, and therefore all $\bb\nn\dd$ as
1612: well, grow like $\ev_{max}(\Mat\nn)^\dd$.
1613:
1614: \begin{ques}
1615: Do all¨$\bbb\nn\dd\bs$ except¨$\bbb\nn\dd{\D_\nn}$
1616: grow like $\ev_{max}(\Mat\nn)^\dd$?
1617: \end{ques}
1618:
1619: \begin{ques}
1620: What is the asymptotic behaviour of $\ev_{max}(\Mat\nn)$
1621: with¨$\nn$?
1622: \end{ques}
1623:
1624: The trivial upper bound of¨\eqref{E:Upper} suggests to
1625: compare $\ev_{max}(\Mat\nn)$ with¨$\nn!$, or, rather,
1626: $\ev_{max}(\Mat\nn)$ with¨$\nn \cdot \ev_{max}(\Mat\nno)$.
1627:
1628: \section{Letting the degree vary}
1629: \label{S:Deg}
1630:
1631: So far, we kept the braid index¨$\nn$ fixed, and studied
1632: how the numbers¨$\bb\nn\dd$ or¨$\bbb\nn\dd\bs$ vary
1633: with¨$\dd$, thus letting linear inductions appear. Quite
1634: different induction schemes appear when we fix the degree
1635: and let the braid index vary. No systematic method is
1636: known so far, and we only mention a few partial results
1637: motivated by the approach of¨\cite{Dhh}.
1638:
1639: \subsection{The numbers¨$\bb\nn2$}
1640:
1641: Very little is known about¨$\bb\nn\dd$ in general. The case
1642: $\dd = 1$ is trivial, as we already observed the equality
1643: $$\bb\nn1 = \nn!.$$
1644: For $\dd = 2$, the value can be deduced from earlier results
1645: of¨\cite{CSV, CSV2} about permutations. We shall use the
1646: following very general observation about duality in Garside
1647: groups:
1648:
1649: \begin{lemm} \label{L:Dual}
1650: For¨$\xx$ in¨$\DD\nn{}$, let $\dual\xx$
1651: and $\xx^*$ be defined by $\dual\xx\,\xx = \xx \,
1652: \xx^* = \D_\nn$. Then $\xx \mapsto \dual\xx$ and
1653: $\xx \mapsto \xx^*$ are permutations of¨$\DD\nn{}$,
1654: and, for each simple¨$\xx$, we have
1655: \begin{equation} \label{E:Dual}
1656: \DivR(\dual\xx) = \Int\nn \setminus \DivL(\xx)
1657: \mbox{\quad and \quad}
1658: \DivL(\xx^*) = \Int\nn \setminus \DivR(\xx).
1659: \end{equation}
1660: \end{lemm}
1661:
1662: \begin{proof}
1663: Assume $\xx \in \DD\nn{}$. Then, by hypothesis,
1664: $\xx$ is a left and a right divisor of¨$\D_\nn$,
1665: hence $\dual\xx$ and $\xx^*$ are positive
1666: braids, and they are divisors of¨$\D_\nn$
1667: in¨$B_\nn^+$, so they are simple. That the
1668: mappings $\xx \mapsto \dual\xx$ and
1669: $\xx \mapsto \xx^*$ are injective is clear,
1670: and the surjectivity follows from the
1671: finiteness of¨$\DD\nn{}$.
1672:
1673: Now, $\ss i$ being a right divisor of¨$\dual\xx$ is
1674: equivalent to $\dual\xx \ss i$ not being simple,
1675: hence to the non-existence of¨$\yy$ satisfying
1676: $\dual\xx \ss i \yy = \D_\nn$, and finally to the
1677: non-existence of¨$\yy$ satisfying $\xx = \ss i
1678: \yy$. This implies the first equality
1679: in¨\eqref{E:Dual}. The second equality follows from
1680: a symmetric argument.
1681: \end{proof}
1682:
1683: \begin{prop}
1684: The numbers¨$\bb\nn2$ are determined by the
1685: induction
1686: \begin{equation}
1687: \bb02=1, \qquad
1688: \bb\nn2 = \sum_{\ind=0}^\nno (-)^{\nn + \ind
1689: +1} {\nn \choose \ind}^2
1690: \bb\ind2.
1691: \end{equation}
1692: Their double exponential generating function is
1693: \begin{equation}
1694: \sum_{\nn = 0}^\infty \bb\nn2
1695: \frac{x^\nn}{\nn!^2} =
1696: \bigg( \sum_{\nn = 0}^\infty (-1)^\nn
1697: \frac{x^\nn}{\nn!^2}
1698: \bigg)\inv = \frac1{J_0(\sqrt\xx)},
1699: \end{equation}
1700: where $J_0(\xx)$ is the Bessel function.
1701: \end{prop}
1702:
1703: \begin{proof}
1704: By definition, $\bb\nn2$ is the number of pairs
1705: of simple¨$\nn$-braids $(\sx_1, \sx_2)$ satisfying
1706: $\DivR(\sx_1) \supseteq \DivL(\sx_2)$, \ie, by
1707: Lemma¨\ref{L:Dual},
1708: $\DivR(\sx_1) \cap \DivR(\dual\sx_2) = \emptyset$. By
1709: Lemma¨\ref{L:Exchange}, this number is also
1710: the number of pairs of permutations¨$(\ff, \gg)$
1711: in¨$\Sym\nn$ with no descent in common, \ie, such
1712: that there exists no¨$i$ satisfying both $\ff(i) >
1713: \ff(i+1)$ and $\gg(i) > \gg(i+1)$. Such pairs of
1714: permutations have been counted
1715: in¨\cite{CSV, CSV2} (see also¨\cite{Rio}),
1716: with the result indicated above.
1717: \end{proof}
1718:
1719: \subsection{The numbers $\bbb\nn2{\D_{\nn - \rrr}}$}
1720:
1721: Specific results appear when we consider the
1722: numbers¨$\bbb\nn\dd\bs$ with¨$\bs$ of the form¨$\D_{\nn
1723: - \rrr}$ with $1 \le \rrr \le \nn$. In particular, we can
1724: complete the computation when $\rrr$ is fixed and $\dd$ is
1725: small. We obviously have
1726: $\bbb\nn1{\D_{\nn - \rrr}} = 1$ for $1 \le \rrr \le \nn$, so
1727: the first case to consider is $\dd = 2$. The general
1728: principle that makes the computation of $\bbb\nn\dd{\D_{\nn
1729: - \rrr}}$ relatively easy is the following observation:
1730:
1731: \begin{lemm} \label{L:Prelim}
1732: For all¨$\nn, \dd, \rrr$, we have
1733: \begin{equation} \label{E:Prelim}
1734: \bbb\nn\dd{\D_{\nn - \rrr}} = \sum_{\mbox{\Small\rm $\bs$ right divisible
1735: by¨$\D_{\nn - \rrr}$}} \bbb\nn{\dd-1}\bs.
1736: \end{equation}
1737: \end{lemm}
1738:
1739: \begin{proof}
1740: The argument is similar to that for
1741: Proposition¨\ref{P:Total}. A sequence $(\bs_1,
1742: \dots, \bs_{\dd-1}, \D_{\nn-\rr})$ is normal if and only if
1743: both $(\bs_1, \dots, \bs_{\dd-1})$ and $(\bs_{\dd-1},
1744: \D_{\nn-\rr})$ are normal. Now $(\bs_{\dd-1}, \D_{\nn-\rr})$
1745: is normal if and only if every¨$\ss i$
1746: dividing¨$\D_{\nn-\rr}$ on the left divides¨$\bs_{\dd-1}$
1747: on the right. The $\ss i$'s dividing $\D_{\nn-\rr}$ on the
1748: left are $\ss1$, \dots, $\ss{\nn - \rr - 1}$. The simple
1749: braids that are right divisible by $\ss1$, \dots, $\ss{\nn -
1750: \rr - 1}$ are those right divisible by¨$\D_{\nn-\rr}$. Then
1751: \eqref{E:Prelim} follows.
1752: \end{proof}
1753:
1754: \begin{prop}
1755: For $1 \le \rrr \le \nn$, we have
1756: \begin{equation} \label{E:K1}
1757: \bbb\nn2{\D_{\nn - \rrr}} = \frac{\nn!}{(\nn - \rrr)!}.
1758: \end{equation}
1759: \end{prop}
1760:
1761: \begin{proof}
1762: By Lemma¨\ref{L:Prelim}, $\bbb\nn2{\D_{\nn - \rrr}}$ is
1763: the number of simple $\nn$-braids¨$\bs$ that are
1764: right divisible by¨$\D_{\nn - \rrr}$, \ie., that satisfy
1765: $\DivR(\bs) \supseteq \Int{{\nn - \rrr}}$. The block
1766: composition of¨$\Int{\nn - \rrr}$ in¨$\nn$ is $({\nn - \rrr}, 1,
1767: \dots, 1)$, so \eqref{E:Value} directly
1768: gives¨\eqref{E:K1}.
1769: \end{proof}
1770:
1771: \subsection{The numbers $\bbb\nn3{\D_{\nn -\rrr}}$}
1772:
1773: Things become more interesting for¨$\dd = 3$.
1774:
1775: \begin{prop} \label{P:K2}
1776: For $1 \le \rrr \le \nn$, there exist polynomials¨$P_1$, \dots, $P_{n-r}$
1777: with integer coefficients and $P_i$ of degree at most¨${\nn - \rrr} - i +
1778: 1$ such that, for every¨$\nn$, we have
1779: \begin{align}
1780: \label{E:K2}
1781: \bbb\nn3{\D_{\nn-\rrr}}
1782: &= (\nn - \rrr)!\,({\nn - \rrr}+1)^\nn + \sum_{i = 1}^{\nn -
1783: \rrr} P_i(\nn) \, i^{\rrr +i-1}.\\
1784: \intertext{The explicit values for $\rrr
1785: = 1, 2$ are}
1786: \label{E:K21}
1787: \bbb\nn3{\D_{\nno}} &= 2^\nno, \\
1788: \label{E:K22}
1789: \bbb\nn3{\D_{\nn-2}} &= 2 \cdot 3^\nn - (\nn + 6) \cdot
1790: 2^\nno + 1.
1791: \end{align}
1792: \end{prop}
1793:
1794: \begin{proof}
1795: We begin with¨\eqref{E:K21}. By
1796: Lemma¨\ref{L:Prelim}, $\bbb\nn3{\D_{\nno}}$ is the sum of
1797: all $\bbb\nn2\bs$ with $\bs$ right divisible by¨$\D_{\nno}$, \ie, it is the number of normal sequences
1798: $(\sx_1, \sx_2)$ such that $\sx_2$ is right divisible
1799: by¨$\D_{\nno}$. Let $\SS$ be the set of all such
1800: normal sequences. We partition¨$\SS$ according to the
1801: value of¨$\DivL(\sx_2)$, \ie, for each subset¨$\II$
1802: of¨$\Int{\nno}$, we count how many pairs¨$(\sx_1,
1803: \sx_2)$ satisfy¨$\DivL(\sx_2) = \II$. So assume
1804: that $\sx_2$ is right divisible by¨$\D_{\nno}$. Two
1805: cases are possible. Either $\sx_2$ is right
1806: divisible by (hence equal to)¨$\D_\nn$, and then we
1807: have
1808: $\DivL(\sx_2) = \Int{\nno}$. Or
1809: $\sx_2$ is not divisible by¨$\D_{\nno}$, and then
1810: Lemma¨\ref{L:Sequence} shows that
1811: $\sx_2$ must be $\ss{i, \nn}\D_{\nno}$ for some¨$i$
1812: with $2 \le i \le \nn$, so that the
1813: block composition of¨$\DivL(\sx_2)$ is $(i-1,
1814: \nn - i+1)$. So the possible compositions
1815: for the set¨$\DivL(\sx_2)$ are¨$(\nn)$, and $(\ip,
1816: \nn - \ip)$ with $1 \le \ip \le \nno$. Conversely,
1817: the previous analysis shows that, for each¨$\II$ of
1818: the previous form, there exists exactly one
1819: possible¨$\sx_2$. Now, Proposition¨\ref{P:EES} says
1820: that there is one choice for¨$\sx_1$ in the case
1821: of¨$(\nn)$---namely $\sx_1 = \D_\nn$---and $\nn
1822: \choose \ip$ choices for¨$\sx_1$ in the case
1823: of¨$(\ip, \nn - \ip)$. We deduce
1824: $$\bbb\nn3{\D_{\nno}} = 1 + \sum_{\ip= 1}^\nno {\nn
1825: \choose{\ip}} = 2^\nno.$$
1826:
1827: The method is similar for
1828: computing¨$\bbb\nn3{\D_{\nn-2}}$ in¨\eqref{E:K22}. Assume
1829: that
1830: $(\sx_1, \sx_2)$ is a normal sequence with¨$\sx_2$ right
1831: divisible by¨$\D_{\nn-2}$. The hypothesis is $\DivR(\sx_2) \supseteq
1832: \Int{\nn-3}$, so three cases may occur, namely $\DivR(\sx_2)
1833: \supseteq \Int{\nn-2}$, $\DivR(\sx_2) = \Int{\nn-3}$, and
1834: $\DivR(\sx_2) = \Int{\nn-3} \cup \{\nno\}$. The first case was analysed
1835: above. In the second case, $\DivL(\sx_2)$ has three blocks, and,
1836: conversely, each set¨$\II$ with three blocks gives
1837: exactly one eligible¨$\sx_2$. In the third case,
1838: $\DivL(\sx_2)$ has either two blocks, or it has
1839: three blocks with the middle one at least¨$2$;
1840: conversely, each set¨$\II$ of the previous form
1841: gives one eligible¨$\sx_2$. Using as above
1842: Proposition¨\ref{P:EES} to count the
1843: eligible¨$\sx_1$'s for each possible¨$\II$, we
1844: obtain that $\bbb\nn3{\D_{\nn-2}}$ is
1845: \begin{equation} \label{E:Sum}
1846: \bbb\nn3{\D_{\nno}}
1847: + \sum_{\overset{\scriptstyle\ip_1 + \ip_2 + \ip_3 =
1848: \nn} {\ip_1, \ip_2, \ip_3 \ge 1}}
1849: \frac{\nn!}{\ip_1! \ip_2! \ip_3!}
1850: + \sum_{\overset{\scriptstyle\ip_1 + \ip_2 =
1851: \nn} {\ip_1, \ip_2 \ge 1}}
1852: \frac{\nn!}{\ip_1! \ip_2!}
1853: + \sum_{\overset{\scriptstyle\ip_1 + \ip_2 + \ip_3 =
1854: \nn} {\ip_1, \ip_3 \ge 1,\ip_2 \ge 2}}
1855: \frac{\nn!}{\ip_1! \ip_2! \ip_3!}.
1856: \end{equation}
1857: Using the fact that $3^\nn$ is the sum of all
1858: $\frac{\nn!}{\ip_1! \ip_2! \ip_3!}$ with $\ip_1
1859: + \ip_2 + \ip_3 = \nn$, one deduces¨\eqref{E:K22}
1860: by bookkeeping.
1861:
1862: Applying the same method in the general case leads
1863: to¨\eqref{E:K2}. Indeed, always by
1864: Lemma¨\ref{L:Sequence}, specifying a simple
1865: $\nn$-braid¨$\sx_2$ satisfying¨$\DivR(\sx_2)
1866: \supseteq \Int{\rrr -1}$ amounts to choosing a
1867: permutation of the ${\nn - \rrr}$¨last strands and the
1868: ${\nn - \rrr}$¨positions $(i_1, \dots, i_{\nn - \rrr})$ where these
1869: strands start from. In the generic case, the
1870: resulting set¨$\DivL(\sx_2)$ is $\{i_1 - 1, \dots,
1871: i_{\nn - \rrr} - 1\}$, whose composition consists of
1872: ${\nn - \rrr}$¨blocks. The special cases are when at least
1873: two adjacent strands among the last ${\nn - \rrr}$¨ones
1874: start from adjacent positions; according to
1875: whether these strands cross or not in the final
1876: part, one then obtains either a composition with a
1877: block of size¨2 at least, or a composition with
1878: less than ${\nn - \rrr}$¨blocks. Conversely, for every
1879: subset¨$\II$ of¨$\Int{\nno}$ with ${\nn - \rrr}$¨blocks,
1880: there exists in general $(\nn - \rrr)!$¨eligible¨$\sx_2$'s,
1881: one for each choice of the final permutation of the
1882: last $\nn - \rrr$¨strands. There may be less than
1883: $(\nn - \rrr)!$¨choices for¨$\sx_2$ when $1$ occurs in the
1884: composition of¨$\II$. Also, subsets
1885: of¨$\Int{\nno}$ with fewer than¨${\nn - \rrr}$ blocks
1886: may lead to eligible¨$\sx_2$'s. Multiplying by the
1887: number of eligibles¨$\sx_1$'s for each¨$\II$ and
1888: summing up yields an expression similar
1889: to¨\eqref{E:Sum}, involving $(\nn - \rrr)!$¨sums of the form
1890: $\sum_{\ip_1 + \dots + \ip_{{\nn - \rrr}+1} = \nn}
1891: \frac{\nn!}{\ip_1! \dots \ip_{{\nn - \rrr}+1}!}$ with
1892: possible order constraints on¨$\ip_1$, \dots,
1893: $\ip_{{\nn - \rrr}+1}$. Each of them leads to a
1894: factor¨$({\nn - \rrr}+1)^\nn$, plus additional factors
1895: corresponding to specializing arguments to¨$0$
1896: or¨$1$ or to grouping them.
1897: \end{proof}
1898:
1899: \subsection{The numbers $\bbb\nn4{\D_{\nno}}$}
1900:
1901: For $\dd = 4$, it seems hopeless to complete the computation
1902: of¨$\bbb\nn\dd{\D_{\nn-\rr}}$. However, this can be done
1903: for¨$\rrr = 1$. The remarkable point is that still another
1904: induction scheme appears.
1905:
1906: \begin{prop}
1907: For¨$\nn \ge 1$, we have
1908: \begin{equation} \label{E:KKK3}
1909: \bbb\nn4{\DD{\nno}{}} = \sum_{\ind=0}^\nno
1910: \frac{\nn!}{\ind!}.
1911: \end{equation}
1912: \end{prop}
1913:
1914: \begin{proof}
1915: According to Lemma¨\ref{L:Prelim} again, we have
1916: now to count the normal sequences
1917: $(\sx_1, \sx_2, \sx_3)$ with¨$\sx_3$ of the
1918: form¨$\ss{i, \nn}\D_{\nno}$, $2 \le i \le \nn$.
1919: We partition the family according to the value¨$\II$
1920: of¨$\DivL(\sx_2)$, and count how many sequences may
1921: correspond to a given¨$\II$. Let $(\ip_1, \dots,
1922: \ip_\ir)$ denote the block composition of¨$\II$.
1923:
1924: Let us first consider the case $\II =
1925: \Int{\nno}$. Then we must have $\sx_2 = \D_\nn$,
1926: hence $\sx_1 = \D_\nn$ as well. There are
1927: $\nn$¨possible choices for¨$\sx_3$, and the total
1928: number of corresponding sequences¨$(\sx_1, \sx_2,
1929: \sx_3)$ is¨$\nn$.
1930:
1931: We assume now $\II \not= \Int{\nno}$, \ie, $\ir
1932: \ge 2$. As for¨$\sx_1$, Proposition¨\ref{P:EES} directly
1933: gives the number of choices, namely
1934: $\frac{\nn!}{\ip_1! \dots \ip_\ir!}$. So we are left
1935: with counting how many pairs¨$(\sx_2, \sx_3)$ are
1936: eligible. The case $\sx_3 = \D_n$ is excluded
1937: since it implies $\sx_2 = \D_\nn$ hence $\II =
1938: \Int{\nno}$. As in the case of¨$\bbb\nn3{\D_{\nno}}$,
1939: the hypothesis that $\sx_3$ is $\ss{i, \nn}
1940: \D_{\nno}$ for some¨$i$ with $2 \le i \le \nn$
1941: implies that the block composition
1942: of¨$\DivL(\sx_3)$ consists of two nonempty blocks,
1943: and, conversely, each partition of¨$\Int\nn$ into
1944: two nonempty blocks gives a unique¨$\sx_3$ of the
1945: convenient form. So the number of pairs¨$(\sx_2,
1946: \sx_3)$ associated with¨$\II$ is the number
1947: of¨$\sx_2$'s satisfying
1948: $\DivL(\sx_2) = \II$ and such that $\DivR(\sx_2)$
1949: has two blocks.
1950:
1951: By¨\eqref{E:Sequence}, this number is the
1952: number of functions¨$\ff$ of¨$\Int\nn$ to¨$\{1,
1953: 2\}$ such that $\ff(i) < \ff(i+1)$ holds exactly for
1954: $i \notin \II$. As only two values are possible,
1955: this condition means that we have $\ff(i) = 1$ and
1956: $\ff(i+1) = 2$ for¨$i \notin \II$, and
1957: $\ff(i+1) \le \ff(i)$ for¨$i \in \II$. Consider
1958: the blocks of¨$\II$. In each block,
1959: except possibly the first and the last ones, the
1960: value of¨$\ff$ has to be¨$2$ on the first element,
1961: and to be¨$1$ on the last element. Inbetween, $\ff$
1962: is non-increasing. So the values consist of a series
1963: of¨$2$'s, followed by a series of¨$1$'s. The
1964: only parameter to specify is the position
1965: where¨$\ff$ switches from¨$2$ to¨$1$, so, for a
1966: block of size¨$\ip$, there are $\ip-1$ possible
1967: choices (see Figure¨\ref{F:Function}). The cases of
1968: the first and the last blocks are special, because
1969: there is no constraint on the left for the first
1970: block, and on the right for the last block. So, in
1971: these special cases, there are
1972: $\ip$¨choices instead of¨$\ip-1$. The conclusion is
1973: that, for¨$\II$ of block composition $(\ip_1, \dots,
1974: \ip_\ir)$, there are
1975: $\ip_1(\ip_2-1) \dots (\ip_{\ir-1}-1) \ip_\ir$
1976: choices for the pairs¨$(\sx_2, \sx_3)$ associated
1977: with¨$\II$. Merging the result for¨$\sx_1$ and
1978: for¨$(\sx_2, \sx_3)$ and summing up over¨$\II$
1979: gives
1980: \begin{equation} \label{E:K3}
1981: %\sum_{\underset{\scriptstyle p_1, \dots, p_r \ge 1}{p_1 +
1982: %\cdots + p_r = n}}
1983: \bbb\nn4{\DD{\nno}{}} = \sum
1984: \frac{\nn!}{\ip_1! \dots \ip_\ir!} \,
1985: \ip_1(\ip_2-1)
1986: \dots (\ip_{\ir-1}-1)\ip_\ir.
1987: \end{equation}
1988: the sum being taken over all finite compositions
1989: $(\ip_1, \dots, \ip_\ir)$ of¨$\nn$: indeed, the
1990: value for¨$\II = \Int{\nno}$, namely¨$\nn$,
1991: corresponds to the missing term $\frac{\nn!}{\nn!}
1992: \nn$ of the sum.
1993:
1994: \begin{figure} [htb]
1995: \begin{picture}(82,20)(0, 0)
1996: \put(0,3){\includegraphics{Function.eps}}
1997: \put(15,0){$\ip_1$}
1998: \put(38,0){$\ip_2$}
1999: \put(57,0){$\ip_3$}
2000: \put(72,0){$\dots$}
2001: \end{picture}
2002: \caption{\smaller Proof of¨\eqref{E:K3}: the
2003: rises are fixed, so it just remains to choose the
2004: position of the fall in each block of¨$\II$,
2005: whence $\ip-1$ choices for a size¨$\ip$ block
2006: except the first and the last ones.}
2007: \label{F:Function}
2008: \end{figure}
2009:
2010: We can now simplify the right hand term
2011: in¨\eqref{E:K3}. To this end, we observe that
2012: \begin{equation} \label{E:KK3}
2013: \sum_{\underset{\scriptstyle \ip_1, \dots, \ip_\ir \ge 1}{\ip_1 +
2014: \cdots + \ip_\ir = \ind + 1}}
2015: \frac{\ip_1-1}{\ip_1!} \dots \frac{p_{r-1}-1}{p_{r-1}!} \,
2016: \frac{\ip_\ir}{\ip_\ir!} = 1
2017: \end{equation}
2018: holds for¨$\ind \ge 0$. Indeed, let $F(\ind)$ be the
2019: left hand side of¨\eqref{E:KK3}. We
2020: prove¨\eqref{E:KK3} using induction on¨$\ind$. For
2021: $\ind=0$, we get $1=1$. Assume $\ind \ge 1$ and
2022: consider the sequences
2023: $(\ip_1, \dots, \ip_\ir)$ satisfying $\ip_1 + \cdots +
2024: \ip_\ir = \ind + 1$. On the one hand, we have
2025: $(\ind + 1)$, whose contribution to¨$F(\ind)$ is
2026: $\frac{1}{\ind!}$. On the other hand, we have
2027: the sequences of length at least¨$2$. Now, for
2028: each¨$\ip$ with $1 \le \ip \le \ind$, the
2029: contribution of $(\ip, \ip_2, \dots, \ip_\ir)$
2030: to¨$F(\ind)$ is
2031: $\frac{\ip-1}{\ip!}$ times the contribution
2032: of $(\ip_2, \dots, \ip_\ir)$ to¨$F(\ind - \ip)$.
2033: Hence the total contribution of the sequences
2034: beginning with¨$\ip$ to¨$F(\ind)$ is
2035: $\frac{\ip-1}{\ip!}
2036: \, F(\ind - \ip)$, so, by induction hypothesis, it
2037: is $\frac{(\ip-1)}{\ip!}$. We deduce
2038: $F(\ind) = \frac{0}{1!} + \frac{1}{2!} + \cdots +
2039: \frac{\ind -1}{\ind!} + \frac1{\ind!}$,
2040: which is clearly¨$1$.
2041:
2042: Consider now the right hand side in¨\eqref{E:K3}.
2043: For
2044: $0 \le \ind < \nno$, the contribution of $(\ind+1,
2045: \ip_2, \dots, \ip_\ir)$ to the sum is
2046: $\frac{\nn!}{\ind!}$ times the quantity
2047: $\frac{\ip_1-1}{\ip_1!} \,
2048: \frac{\ip_{\ir_1}-1}{\ip_{\ir-1}!}
2049: \, \frac{\ip_\ir}{\ip_\ir!}$ involved
2050: in¨\eqref{E:KK3}. Using the latter equality, we
2051: deduce that the total contribution of the sequences
2052: beginning with¨$\ind + 1$ is
2053: $\frac{\nn!}{\ind!}$. As for¨$\ind = \nno$, the
2054: contribution of¨$(\nn)$ to the right
2055: hand side in¨\eqref{E:K3} is¨$\nn$, which is
2056: $\frac{\nn!}{(\nno)!}$, so the general formula
2057: remains valid. By summing over¨$\ind$, we
2058: obtain¨\eqref{E:KKK3}.
2059: \end{proof}
2060:
2061: \begin{coro}
2062: The numbers $\bbb\nn4{\DD{\nno}{}}$ are
2063: determined by the induction
2064: $$\var_1 = 1, \qquad
2065: \var_\nn = \nn \var_{\nno} +
2066: 2\nno.$$
2067: \end{coro}
2068:
2069: Another consequence of¨\eqref{E:KKK3} is the equality
2070: $$\bbb\nn4{\DD{\nno}{}} = \lfloor \nn! e \rfloor - 1,$$
2071: with $e = \exp(1)$.
2072:
2073: \begin{thebibliography}{99}
2074:
2075: \Reff Adj; S.I.\,Adyan; Fragments of the word Delta in a
2076: braid group; Mat. Zam. Acad. Sci. SSSR; 36-1; 1984;
2077: 25--34; translated Math. Notes of the Acad. Sci. USSR;
2078: 36-1 (1984) 505--510.
2079:
2080: \Ref Bes; D. Bessis; The dual braid monoid; An. Sci. Ec.
2081: Norm. Sup.; 36; 2003; 647--683.
2082:
2083: \Ref Bir; J.\,Birman; Braids, links, and mapping
2084: class groups; Annals of Math. Studies 82, Princeton
2085: Univ. Press (1975).
2086:
2087: \Reff BKL; J.~Birman, K.H. Ko \& S.J. Lee; A new approach
2088: to the word problem in the braid groups; Advances in
2089: Math.; 139-2; 1998; 322-353.
2090:
2091: \Ref Bro; A.\,Bronfman; Growth function of a class of
2092: monoids; Preprint (2001).
2093:
2094: \Reff CSV; L.\,Carlitz, R.\,Scoville \&
2095: T.\,Vaughan; Enumeration of pairs of permutations
2096: and sequences; Bull. Amer. Math. Soc.; 80;
2097: 1974; 881--884.
2098:
2099: \Reff CSV2; L.\,Carlitz, R.\,Scoville \&
2100: T.\,Vaughan; Enumeration of
2101: pairs of permutations; Discrete Math.; 14;
2102: 1976; 215--239.
2103:
2104: \Reff Chb; R.\,Charney; Geodesic automation and
2105: growth functions for Artin groups of finite type; Math.
2106: Ann.; 301-2; 1995; 307--324.
2107:
2108: \Ref CrH; C.R.\,Cromwell \& S.\,Humphries; Counting
2109: fundamental paths in $2$-generator Artin semigroups;
2110: Preprint (2004).
2111:
2112: \Reff Dgk; P.\,Dehornoy; Groupes de Garside;
2113: Ann. Scient. Ec. Norm. Sup.; 35; 2002; 267--306.
2114:
2115: \Ref Dgr; P.\,Dehornoy, I. Dynnikov, D. Rolfsen,
2116: B. Wiest; Why are braids orderable?; Panoramas \&
2117: Synth\`eses vol. 14, Soc. Math. France (2002).
2118:
2119: \Ref Dhh; P.\,Dehornoy; Still another approach to the
2120: braid ordering; arXiv: math.GR/0506495.
2121:
2122: \Reff Dfx; P.\,Dehornoy \& L.\,Paris; Gaussian
2123: groups and Garside groups, two generalizations of Artin
2124: groups; Proc. London Math. Soc.; 79-3; 1999; 569--604.
2125:
2126: \Reff Dlg; P. Deligne; Les immeubles des groupes de
2127: tresses g\'en\'eralis\'es; Invent. Math.; 17; 1972;
2128: 273--302.
2129:
2130: \Reff ElM; E.A.\,Elrifai \& H.R.\,Morton; Algorithms for
2131: positive braids; Quart. J. Math. Oxford; 45-2; 1994;
2132: 479--497.
2133:
2134: \Ref Eps; D.\,Epstein, J.\,Cannon, D.\,Holt, S.\,Levy,
2135: M.\,Paterson \& W.\,Thurston; Word Processing in Groups;
2136: Jones \& Bartlett Publ. (1992).
2137:
2138: \Reff Eps2; D.\,Epstein, A.\,Iano-Fletscher, \& U.\,Zwick;
2139: Growth functions and automatic groups; Experiment. Math; 5;
2140: 1996; 297--315.
2141:
2142: \Reff Gar; F.A.\,Garside; The braid group and
2143: other groups; Quart. J. Math. Oxford; 20-78; 1969;
2144: 235--254.
2145:
2146: \Ref Geb; V.\,Gebhardt; A new approach to the conjugacy
2147: problem in Garside groups; J. of Algebra, to appear;
2148: math.GT/0306199.
2149:
2150: \Ref Hoh; C.\,Hohlweg; Properties of the Solomon algebra
2151: homomorphism; arXiv: math.RT/0302309.
2152:
2153: \Ref KoL; K.H.\,Ko, S.\,Lee, J.H.\,Cheon, J.W.\,Han,
2154: J.\,Kang, C.\,Park; New public-key cryptosystem using
2155: braid groups; Crypto 2000, 166--184.
2156:
2157: \Ref McC; J.\,Mc Cammond; An introduction to
2158: Garside structures; Preprint (2005).
2159:
2160: \Reff Rio; J.\,Riordan; Inverse relations and
2161: combinatorial identities; Amer. Math. Monthly;
2162: 71; 1964; 485--498.
2163:
2164: \Reff Sol; L.\,Solomon; A Mackey formula in the
2165: group ring of a Coxeter group; J. Algebra;
2166: 41; 1976; 255--268.
2167:
2168: \Reff Xua; P.\,Xu; Growth of the positive braid
2169: groups; J. Pure Appl. Algebra; 80; 1992;
2170: 197--215.
2171:
2172: \end{thebibliography}
2173:
2174: \end{document}
2175:
2176: