math0506495/Dhh.tex
1: \documentclass{amsart}
2: \usepackage{amsmath, amssymb, amsthm, amscd, graphicx}
3: 
4: \textwidth=13.5cm
5: \textheight=20cm
6: 
7: % for 100% printing
8: \hoffset-1.2cm
9: \voffset+0.5cm
10: 
11: %for 110% printing
12: \hoffset-1.8cm
13: \voffset-0.6cm
14: 
15: \setlength{\unitlength}{1mm}
16: 
17: \theoremstyle{plain}
18: \newtheorem{prop}{Proposition}[section]
19: \newtheorem{coro}[prop]{Corollary}
20: \newtheorem{conj}[prop]{Conjecture}
21: \newtheorem{lemm}[prop]{Lemma}
22: \newtheorem{thrm}[prop]{Theorem}
23: 
24: \theoremstyle{definition}
25: \newtheorem{defi}[prop]{Definition}
26: \newtheorem*{propA}{Property A}
27: \newtheorem*{propC}{Property C}
28: \newtheorem{nota}[prop]{Notation}
29: \newtheorem{exam}[prop]{Example}
30: \newtheorem{rema}[prop]{Remark}
31: 
32: \numberwithin{equation}{section}
33: 
34: % bibliography
35: 
36: \def\Reff#1; #2; #3; #4; #5; #6; #7\par{%
37: \bibitem{#1} #2, {\it #3}, #4 {\bf #5} (#6) #7}
38: 
39: \def\Ref#1; #2; #3; #4\par{%
40: \bibitem{#1} #2, {\it #3}, #4}
41: 
42: % Specific macros of this text $$$$
43: 
44: \catcode`\¨=\active\def¨{~} 
45: 
46: \renewcommand{\a}{\mathtt{a}}
47: \newcommand{\A}{\mathtt{A}}
48: 
49: \renewcommand{\Bar}[1]{\overline{\vline width0pt
50: height5.5pt #1}}
51: %\renewcommand{\Bar}{\bar}
52: \renewcommand{\b}{\mathtt{b}}
53: \newcommand{\B}{\mathtt{B}}
54: \newcommand{\bbb}[3]{b_{#1,#2}(#3)}
55: \newcommand{\bbbb}[3]{b'_{#1,#2}(#3)}
56: \newcommand{\bx}{x}%_A_braid
57: \newcommand{\by}{y}
58: 
59: \renewcommand{\c}{\mathtt{c}}
60: \newcommand{\card}{\mathtt{\#}}
61: \newcommand{\cc}{c}
62: \newcommand{\ccc}[2]{\cc(\DD{#1}{#2})}
63: \newcommand{\CC}{C}
64: \newcommand{\CCC}{\mathbb{C}}
65: \newcommand{\CG}{\Gamma}
66: \newcommand{\cl}[1]{[#1]}
67: \newcommand{\Class}{C}
68: 
69: \newcommand{\D}{\Delta}
70: \newcommand{\Dd}[2]{\delta_{#1,#2}}
71: \newcommand{\Ddt}[2]{\widetilde\delta_{#1,#2}}
72: \newcommand{\DD}[2]{\D_{#1}^{#2}}
73: \newcommand{\dd}{d}%_Degree
74: \newcommand{\DDD}[2]{\Div(\D_{#1}^{#2})}
75: \newcommand{\DDDD}[2]{(\Div(\D_{#1}^{#2}),\nobreak\sm\nobreak)}
76: \renewcommand{\div}{\prec}
77: \newcommand{\Div}{\mathrm{Div}}
78: \newcommand{\DDiv}[1]{(\Div(#1),\nobreak\sm\nobreak)}
79: \newcommand{\dive}{\preccurlyeq}
80: \newcommand{\DivL}{D_L}
81: \newcommand{\DivR}{D_R}
82: \newcommand{\pz}{z}
83: 
84: \newcommand{\ea}{a}
85: \newcommand{\eb}{b}
86: \newcommand{\eg}{{\it e.g.}}
87: \newcommand{\etc}{{\it etc.}}
88: \newcommand{\ev}{\rho}%_Eigenvalue
89: \newcommand{\expo}{e}%_Number_of_occurrences
90: 
91: \newcommand{\ff}{f}
92: \newcommand{\flip}[1]{\phi_{#1}}
93: \newcommand{\first}[1]{(#1)_{\!_1}}
94: 
95: \newcommand{\g}{\gamma}
96: \let\ge=\geqslant
97: \newcommand{\G}{\Gamma}
98: \renewcommand{\gcd}{\mathrm{gcd}}
99: \renewcommand{\gg}{g}
100: \newcommand{\GG}[2]{\CG(\DD{#1}{#2})}
101: \newcommand{\gr}{>}
102: \newcommand{\gre}{\ge}
103: 
104: \newcommand{\hh}[1]{h_{#1}}
105: \newcommand{\hhh}[3]{\hh{#1}(\DD{#2}{#3})}
106: 
107: \newcommand{\ie}{{\it i.e.}}
108: \let\ince=\subseteq
109: \newcommand{\ind}{i}
110: \newcommand{\Ind}{\mathrm{ind}}
111: \newcommand{\indd}{j}
112: \newcommand{\Init}[1]{R_{#1}}
113: \newcommand{\Initt}[1]{R'_{#1}}
114: \newcommand{\Inittt}[1]{R''_{#1}}
115: \newcommand{\Int}[1]{[\![#1]\!]}
116: \newcommand{\inv}{^{-1}}
117: \newcommand{\Inv}{{}^{-1}}
118: \newcommand{\ik}{i}
119: \newcommand{\ip}{p}
120: \newcommand{\iq}{q}
121: \newcommand{\ir}{k}
122: \newcommand{\is}{\ell}
123: 
124: \newcommand{\kk}{k}
125: 
126: \renewcommand{\l}{\lambda}
127: \newcommand{\last}[1]{(#1)_{\!_\infty}}
128: \let\le=\leqslant
129: \newcommand{\lcm}{\mathrm{lcm}}
130: 
131: \newcommand{\m}{\mu}
132: \newcommand{\mm}{m}
133: 
134: \newcommand{\Na}[1]{\mathtt{\#}_{\!\mathstrut\ss1}#1}
135: \newcommand{\Nb}[1]{\mathtt{\#}_{\!\mathstrut\ss2}#1}
136: \newcommand{\nbpart}[1]{p(#1)}% number of partitions
137: \newcommand{\nn}{n}
138: \newcommand{\NN}{N}
139: \newcommand{\NNN}{N}
140: 
141: \renewcommand{\o}{\omega}
142: \newcommand{\op}{\cdot}
143: \newcommand{\opp}{\mathord{\cdot}}
144: 
145: \newcommand{\pp}{p}
146: \newcommand{\ppp}{p}
147: \newcommand{\pppp}{p}
148: 
149: \newcommand{\qq}{q}
150: 
151: \newcommand{\resp}{{\it resp.{~}}}
152: \newcommand{\rr}{r}
153: \newcommand{\RR}{\mathbb{R}}
154: 
155: \let\s=\sigma
156: \newcommand{\sh}{\mathrm{sh}}
157: \newcommand{\sm}{<}
158: \newcommand{\sme}{\le}
159: \renewcommand{\sp}{s}%_A_simple
160: \newcommand{\spp}{t}%_Another_simple
161: \renewcommand{\ss}[1]{\sigma_{#1}}
162: \newcommand{\sss}[2]{\sigma_#1^{#2}}
163: \renewcommand{\SS}[2]{S_{#1}^{#2}}
164: \newcommand{\SSS}[2]{\underline{S}_{#1}^{#2}}
165: \newcommand{\sx}{x}%_A_simple_factor_of_x
166: \newcommand{\Sym}[1]{\mathfrak{S}_{#1}}
167: 
168: \newcommand{\uu}{u}
169: 
170: \newcommand{\var}{u}
171: \newcommand{\vs}{\emph{vs.} }
172: \newcommand{\vv}{v}
173: 
174: \newcommand{\ww}{w}
175: \newcommand{\WW}[2]{\vec w_{#1}^{#2}}
176: 
177: \newcommand{\xx}{x}
178: \newcommand{\XX}{X}
179: 
180: \newcommand{\yy}{y}
181: 
182: \newcommand{\zz}{z}
183: 
184: \hyphenation{mon-oid mon-oids Lem-ma}
185: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
186: \begin{document}
187: 
188: \author{Patrick DEHORNOY}
189: \address{Laboratoire de Math\'ematiques Nicolas
190: Oresme UMR 6139\\ Universit\'e de Caen,
191: 14032~Caen, France}
192: \email{dehornoy@math.unicaen.fr}
193: \urladdr{//www.math.unicaen.fr/\textasciitilde dehornoy}
194: 
195: \title{Still another approach to the braid ordering}
196: 
197: \keywords{braid group; orderable group;
198: well-ordering; normal form; fundamental braid}
199: 
200: \subjclass{20F36, 05A05, 20F60}
201: 
202: \begin{abstract}
203: We develop a new approach to the linear ordering of the
204: braid group~$B_\nn$, based on investigating its
205: restriction to the set¨$\DDD\nn\dd$ of all divisors
206: of~$\DD\nn\dd$ in the monoid¨$B_\infty^+$, \ie, to
207: positive $n$-braids whose normal form has length at
208: most¨$\dd$. In the general case, we compute several
209: numerical parameters attached with the finite orders
210: $\DDDD\nn\dd$. In the case of $3$~strands, we moreover
211: give a complete description of the increasing
212: enumeration of~$\DDDD3\dd$. We deduce a new and
213: specially direct construction of the ordering
214: on~$B_3$, and a new proof of the result that its
215: restriction to~$B_3^+$ is a well-ordering of ordinal
216: type~$\omega^\omega$.
217: \end{abstract}
218: 
219: \maketitle
220: 
221: The general aim of this paper is to investigate the
222: connection between the Garside structure of Artin's
223: braid groups and their distinguished linear ordering
224: (sometimes called the Dehornoy ordering). This leads
225: to a new, alternative construction of the ordering.
226: 
227: Artin's braid groups¨$B_\nn$ are endowed with several
228: interesting combinatorial structures. One of them
229: stems from Garside's analysis
230: \cite{Gar} and is nowadays known as a Garside
231: structure¨\cite{Dgk, McC}. It describes $B_\nn$ as the group
232: of fractions of a monoid¨$B_\nn^+$ with a rich divisibility
233: theory. One of the outcomes of this theory is a unique
234: normal decomposition for every braid in¨$B_\nn$ in
235: terms of simple braids, which are the divisors of
236: Garside's fundamental braid~$\D_\nn$, a finite
237: family of¨$B_\nn^+$ in one-to-one correspondence
238: with the permutations of $\nn$~objects. One obtains a
239: natural graduation of the monoid¨$B_\nn^+$ by
240: considering the family~$\DDD\nn\dd$ of all divisors
241: of~$\DD\nn\dd$, which also are the elements
242: of~$B_n^+$ whose normal form has length at
243: most~$\dd$.
244: 
245: On the other hand, the braid groups are equipped with a
246: distinguished linear ordering, which is compatible with
247: multiplication on the left, and admits a simple
248: combinatorial characterization¨\cite{Dfb}: a
249: braid¨$\bx$ is smaller than another braid¨$\by$ if,
250: among all expressions of the quotient¨$\bx\inv\by$ in
251: terms of the standard generators¨$\ss i$, there exists
252: at least one expression in which the
253: generator¨$\ss\mm$ with maximal (or minimal)
254: index¨$\mm$ appears only positively, \ie,
255: $\ss\mm$ occurs, but $\ss\mm\inv$ does not. Several
256: deep results about that ordering are known, in
257: particular the fact that its restriction
258: to¨$B_\infty^+$ is a well-ordering, and a number of
259: equivalent constructions are known¨\cite{Dgr}.
260: 
261: Although both combinatorial in nature, the previous
262: structures remain mostly uncon\-nected---and connecting
263: them may appear as one of the most natural questions of
264: braid combinatorics. For degree¨$1$, \ie, for simple
265: braids, the linear ordering corresponds to a
266: lexicographical ordering of the associated
267: permutations¨\cite{Dgb}. But this connection does not
268: extend to higher degrees, and almost nothing is known
269: about the restriction of the linear ordering to positive
270: braids of a given degree. In particular, no connection is
271: known between the above mentioned Garside normal form and
272: the alternative normal form constructed by S.\,Burckel
273: in¨\cite{Bus, But, Buu}, one that makes comparison with
274: respect to the linear ordering easy: to give an example,
275: the Garside normal form of¨$\DD3{2\dd}$
276: is¨$(\ss1\ss2\ss1)^{2\dd}$, while its Burckel normal form
277: is $(\ss2\ss1^2\ss2)^\dd\ss1^{2\dd}$.
278: 
279: Our aim in this paper is to investigate the finite
280: linearly ordered sets $\DDDD\nn\dd$. A
281: nice way of thinking of this structure is to consider the
282: increasing enumeration of¨$\DDD\nn\dd$, and to view it as
283: a distinguished path from¨$1$ to¨$\DD\nn\dd$ in the Cayley
284: graph of¨$B_\nn$. A complete description of this path
285: would arguably be an optimal solution to the rather vague
286: question of connecting the Garside and the ordered
287: structures of braid groups. Such a description seems to
288: be extremely intricate from a combinatorial point of
289: view, and it remains out of reach for the moment, but we
290: prove partial results in this direction, namely 
291: 
292: - $(i)$ in the general case, a determination of some
293: numerical parameters attached with $\DDDD\nn\dd$ that in
294: some sense measure its size, with explicit values for
295: small values of¨$\nn$ and¨$\dd$, and
296: 
297: - $(ii)$ in the special case $\nn = 3$, a complete
298: description of the increasing enumeration
299: of¨$\DDDD\nn\dd$.
300: 
301: More specifically, the parameters we investigate are
302: the complexity and the heights.
303: The complexity¨$\ccc\nn\dd$ is defined as the
304: maximal number of occurrences of¨$\ss{\nn-1}$ in an
305: expresion of¨$\DD\nn\dd$ containing no¨$\ss{\nn-1}$.
306: It is connected with the termination of the handle
307: reduction algorithm of¨\cite{Dfo}, and its
308: determination was left as an open question in the
309: latter paper. The
310: $\rr$-height¨$\hhh\rr\nn\dd$ is defined to
311: be the number of¨$\rr$-jumps in the increasing
312: enumeration of $\DDDD\nn\dd$ (augmented by¨$1$), where
313: the term $\rr$-jump refers to some natural filtration of
314: the linear ordering¨$\sm$ by a sequence of partial
315: orderings¨$\sm_\rr$. When $\rr$ increases, $\rr$-jumps
316: are higher and higher, so $\hhh\rr\nn\dd$ counts how
317: many big jumps exist in $\DDDD\nn\dd$.  We prove that the
318: complexity $\ccc\nn\dd$ equals the height
319: $\hhh{\nn-1}\nn\dd$ (Proposition¨\ref{P:MainHeight}), and
320: that, for each¨$\rr$, the $\rr$-height $\hhh\rr\nn\dd$ is
321: the number of divisors of¨$\DD\nn\dd$ whose $\dd$th
322: factor of the normal form is right divisible by¨$\D_\rr$
323: (Proposition¨\ref{P:Main}). Together with the
324: combinatorial results of¨\cite{Dhi}, this allows for
325: computing the explicit values listed in
326: Table¨\ref{T:Values}, and for establishing
327: various inductive formulas (Propositions¨\ref{P:Values}
328: and¨\ref{P:Values34}, among others).
329: 
330: Besides the enumerative results, we also prove a
331: general structural result that connects the ordered set
332: $\DDDD\nn\dd$ with (subsets of) $\DDDD{\nn-1}\dd$
333: (Corollary¨\ref{C:Structure}). This result
334: suggests an inductive method for directly constructing
335: the increasing enumeration of $\DDDD\nn\dd$ starting
336: from those of $\DDDD{\nn-1}\dd$ and $\DDDD\nn{\dd-1}$.
337: This approach is completed here for $\nn = 3$
338: (Proposition¨\ref{P:Enum3}). In some sense,
339: $3$~strand braids are simple objects, and the result
340: may appear as of modest interest; however, the order
341: on~$B_3^+$ is a well-ordering of ordinal
342: type~$\omega^\omega$, hence not a so simple object. The
343: interesting point is that this approach leads to a
344: new, alternative construction of the braid ordering,
345: with in particular a new and simple proof for the
346: so-called Comparison Property which is the hard core
347: in the construction, namely the part that guarantees
348: the linearity of the ordering. In this way, one
349: obtains not only one more construction of an ordering
350: that already has many constructions
351: \cite{Dgr}, but arguably the optimal one, as it makes
352: all proofs simple once the initial inductive
353: definition is correctly stated, and as the connection
354: with the Garside structure is then explicit.
355: 
356: The paper is organized as follows. After a first
357: introductory section recalling basic properties and setting
358: the notation, we introduce the parameters $\ccc\nn\dd$
359: and $\hhh\rr\nn\dd$ in Section¨\ref{S:Meas}, and we
360: establish their connection. In Section¨\ref{S:Deter}, we
361: connect in turn $\hhh\rr\nn\dd$ with the number of
362: $\nn$-braids whose $\dd$th factor in the normal form
363: satisfy certain constraints, and deduce explicit
364: values. Finally, in Section¨\ref{S:N3}, we study the
365: specific case of¨$\DDDD3\dd$ and describe its
366: increasing enumeration, resulting in the new
367: construction of the braid ordering in this case.
368: 
369: \section{Background and preliminary results}
370: 
371: Our notation is standard, and we refer to textbooks
372: like¨\cite{Bir} or¨\cite{Eps} for basic results about
373: braid groups. We recall that the $\nn$¨strand braid group¨$B_\nn$
374: is defined for $\nn \ge 1$ by the presentation
375: \begin{equation} \label{E:Present}
376: B_\nn = \left<\ss1, \dots, \ss{\nn-1} \,;\,
377: \begin{array}{cl}
378: \ss i \ss j = \ss j \ss i 
379: &\text{\quad for $\vert
380: i - j\vert \ge 2$}\\
381: \   \ss i \ss j \ss i  = \ss j \ss i \ss j
382: &\text{\quad for $\vert i - j\vert = 1$}
383: \end{array} \right>.
384: \end{equation}
385: So, $B_1$ is the trivial one-element group, while
386: $B_2$ is the free group generated by¨$\ss1$. The
387: elements of¨$B_\nn$ are called $\nn$¨strand braids,
388: or simply {\it $\nn$-braids}. We use $B_\infty$
389: for the group generated by an infinite sequence
390: of¨$\ss i$'s subject to the relations
391: of¨\eqref{E:Present}, \ie, the direct limit of
392: all¨$B_\nn$'s with respect to the inclusion of¨$B_\nn$
393: into¨$B_{n+1}$.
394: 
395: By definition, every $\nn$-braid¨$\bx$ admits (infinitely
396: many) expressions in terms of the generators¨$\ss i$, $1
397: \le i < n$. Such an expression is called an
398: $\nn$¨strand {\it braid word}. Two braid words¨$\ww,
399: \ww'$ representing the same braid are said to be {\it
400: equivalent}; the braid represented by a braid
401: word¨$\ww$ is denoted¨$\cl\ww$. 
402: 
403: \subsection{Positive braids and the element $\D_\nn$}
404: 
405: We denote by¨$B_\nn^+$ the monoid
406: admitting the presentation¨\eqref{E:Present},
407: and by¨$B_\infty^+$ the union (direct limit) of
408: all¨$B_\nn^+$'s. The elements of¨$B_\nn^+$ are called {\it
409: positive} $\nn$-braids. In¨$B_\infty^+$, no element
410: except¨$1$ is invertible, and we have a natural notion of
411: divisibility: 
412: 
413: \begin{defi}
414: For $\bx, \by$ in¨$B_\nn^+$, we say that $\bx$ is a {\it
415: left divisor} of¨$\by$, denoted $\bx \dive \by$, or,
416: equivalently, that $\by$ is a {\it right multiple}
417: of¨$\bx$, if $\by = \bx \pz$ holds for some¨$\pz$
418: in¨$B_\nn^+$. We denote by¨$\Div(\by)$ the (finite) set of
419: all left divisors of¨$\by$ in¨$B_\nn^+$.
420: \end{defi}
421: 
422: The monoid~$B_\nn^+$ is not commutative for $\nn \ge
423: 3$, and therefore there are distinct symmetric
424: notions of a right divisor and a left multiple---but
425: we shall mostly use left divisors here. Note that
426: $\bx$ is a (left) divisor of¨$\by$ in the sense
427: of¨$B_\nn^+$ if and only if it is a (left) divisor in
428: the sense of¨$B_\infty^+$, so there is no need to
429: specify the index¨$\nn$.
430: 
431: According to Garside's theory \cite{Gar},
432: $B_\nn^+$ equipped with the left divisibility
433: relation is a lattice: any two
434: positive $\nn$-braids¨$\bx,
435: \by$ admit a greatest common left divisor, denoted
436: $\gcd(\bx, \by)$, and a least common right multiple,
437: denoted¨$\lcm(\bx, \by)$. A special role is played by the
438: lcm¨$\D_\nn$ of¨$\ss1$, \dots, $\ss{\nn-1}$, which can
439: be inductively defined by
440: \begin{equation} \label{E:Delta}
441: \D_1 = 1, \qquad
442: \D_\nn = \ss1 \ss2 \dots \ss{n-1} \, \D_{\nn-1}.
443: \end{equation}
444: It is well known that $\D_\nn^2$ belongs to the centre
445: of¨$B_\nn$ (and even generates it for $\nn \ge 3$),
446: and that the flip automorphism¨$\flip n$ of¨$B_\nn$
447: corresponding to conjugation by¨$\D_\nn$
448: exchanges¨$\ss i$ and¨$\ss{n-i}$ for
449: $1 \le i \le \nn-1$. It follows that $\flip n$
450: also exchanges¨$\ss{n, 1}$ and¨$\ss{1, n}$.
451: 
452: In¨$B_\nn^+$, the left and the right divisors of¨$\D_\nn$
453: coincide, and they make a finite sublattice of¨$(B_\nn^+,
454: \dive)$ with $\nn!$¨elements. These braids will be called
455: {\it simple} in the sequel. When braid
456: words are represented by diagrams as
457: mentioned in Figure~\ref{F:Diagram}, simple braids
458: are those positive braids that can be represented by a
459: diagram in which any two strands cross at most
460: once.
461: 
462: \begin{figure}[htb]
463: \begin{center}
464: \begin{picture}(83,21)(0,0)
465: \put(10,0){\includegraphics{Sigma.eps}}
466: \put(0,13){$\ss i$~:}
467: \put(0,3){$\ss i\inv$~:}
468: \put(9.5,19){$1$}
469: \put(19.5,19){$2$}
470: \put(50,19){$i$}
471: \put(58,19){$i{+}1$}
472: \put(27,3){\dots}
473: \put(77,3){\dots}
474: \put(27,13){\dots}
475: \put(77,13){\dots}
476: \end{picture}
477: \end{center}
478: \caption{\smaller One associates with
479: every $\nn$¨strand braid word¨$\ww$ an $\nn$¨strand
480: braid diagram by stacking elementary diagrams as
481: above; then two braid words are equivalent if and only
482: if the associated diagrams are the projections of
483: ambient isotopic figures in¨$\RR^3$, \ie, one can
484: deform one diagram into the other without allowing the
485: strands to cross or moving the endpoints.}
486: \label{F:Diagram}
487: \end{figure}
488: 
489: By mapping¨$\ss i$ to the transposition¨$(i, i+1)$, one
490: defines a surjective homomorphism¨$\pi$ of¨$B_\nn$ onto the
491: symmetric group¨$\Sym\nn$. The restriction of¨$\pi$ to
492: simple braids is a bijection: for every permutation¨$f$
493: of¨$\{1, \dots, \nn\}$, there exists exactly one simple
494: braid¨$\bx$ satisfying $\pi(\bx) = f$. It follows that
495: the number of simple $\nn$-braids is¨$\nn!$. 
496: 
497: \begin{exam} \label{X:Degree}
498: The set¨$\DDD3{}$ consists of six elements, namely $1$,
499: $\ss1$, $\ss2$, $\ss2\ss1$, $\ss1\ss2$, and¨$\D_3$. In
500: examples, we shall often use the shorter notation¨$\tt a$
501: for¨$\ss1$, $\tt b$ for¨$\ss2$, \etc\ Thus, the six
502: simple $3$-braids are $1$, $\tt a$, $\tt b$, $\tt ba$,
503: $\tt ab$, $\tt aba$.  
504: \end{exam}
505: 
506: \subsection{The normal form}
507: 
508: For each positive $\nn$-braid¨$\bx$ distinct of¨$1$,
509: the simple braid $\gcd(\bx, \D_\nn)$ is the maximal
510: simple left divisor of¨$\bx$, and we obtain a
511: distinguished expression $\bx = \sx_1 \bx'$
512: with¨$\sx_1$ simple. By decomposing¨$\bx'$ in the
513: same way and iterating,  we obtains the so-called
514: normal expression¨\cite{ElM, Eps}. 
515: 
516: \begin{defi}
517: A sequence $(\sx_1, \dots, \sx_\dd)$ of simple
518: $\nn$-braids is said to be {\it normal} if, for
519: each¨$\kk$, one has $\sx_\kk =
520: \gcd(\D_\nn, \sx_\kk \dots \sx_\dd)$.
521: \end{defi}
522: 
523: Clearly, each positive braid admits a unique normal
524: expression. It will be convenient here to consider the
525: normal expression as unbounded on the right by
526: completing it with as many trivial factors¨$1$ as needed.
527: In this way, we can speak of the {\it $\dd$th factor} (in
528: the normal form) of¨$\bx$ for each positive braid¨$\bx$.
529: We say that a positive braid has {\it degree¨$\dd$} if
530: $\dd$ is the largest integer such that the $\dd$th factor
531: of¨$\bx$ is not¨$1$. We shall use the following two
532: properties of the normal form: 
533: 
534: \begin{lemm} \cite{ElM} \label{L:Normal}
535: A sequence of simple $\nn$-braids $(\sx_1, \dots,
536: \sx_\dd)$ is normal if and only if, for each¨$\kk <
537: \dd$, each¨$\ss i$ that divides¨$\sx_{\kk+1}$
538: on the left divides¨$\sx_\kk$ on the right.
539: \end{lemm}
540: 
541: \begin{lemm} \cite{ElM} \label{L:Degree}
542: For $\bx$ a positive braid in¨$B_\nn^+$, the following
543: are equivalent:
544: 
545: $(i)$ The braid¨$\bx$ belongs to¨$\DDD\nn\dd$, \ie, is a
546: (left or right) divisor of¨$\DD\nn\dd$; 
547: 
548: $(ii)$ The degree of¨$\bx$ is at most¨$\dd$.
549: \end{lemm}
550: 
551: \begin{exam} \label{X:Enum}
552: There are 19 divisors of~$\DD32$, which also are the
553: $3$-braids of degree at most~$2$. Their enumeration in
554: normal form---in an ordering that may seem strange now,
555: but should become familiar soon---is: $1$,
556: $\a$, $\a\opp\a$, $\b$, $\b\a$,
557: $\b\a\opp\a$, $\b\opp\b$, $\b\opp\b\a$, $\a\b$,
558: $\a\b\a$, $\a\b\a\opp\a$, $\a\b\opp\b$,
559: $\a\b\opp\b\a$, $\a\opp\a\b$, $\a\b\a\opp\b$,
560: $\a\b\a\opp\b\a$, $\b\a\opp\a\b$, $\a\b\a\opp\a\b$,
561: $\a\b\a\opp\a\b\a$. 
562: \end{exam}
563: 
564: By Lemma~\ref{L:Degree}, every divisor
565: of¨$\DD\nn\dd$ can be expressed as the product of at
566: most¨$\dd$ divisors of¨$\D_\nn$, so we certainly have
567: $\card\DDD\nn\dd \le (\nn!)^\dd$ for all¨$\nn,
568: \dd$.
569: 
570: \subsection{The braid ordering}
571: 
572: The basic notion is the following one:
573: 
574: \begin{defi}
575: Let $\ww$ be a nonempty braid word. We say that
576: $\ss\mm$ is the {\it main} generator in¨$\ww$
577: if $\ss\mm$ or $\ss\mm\inv$ occurs in~$\ww$,
578: but no $\ss i^{\pm1}$ with $i > \mm$ does. We
579: say that $\ww$ is {\it
580: $\s$-positive} (\resp {\it $\s$-negative}) if the main
581: generator occurs only positively (\resp negatively)
582: in¨$\ww$.
583: \end{defi}
584: 
585: A positive nonempty braid word, \ie, one that contains
586: no¨$\ss i\inv$ at all, is $\s$-positive, but the inclusion
587: is strict: for instance, $\ss1\inv \ss2$ is not positive,
588: but it is $\s$-positive, as its main generator,
589: namely¨$\ss2$, occurs positively (one¨$\ss2$) but not
590: negatively (no¨$\ss2\inv$).
591: 
592: Then we have the following two properties, which have
593: received a number of independent proofs¨\cite{Dgr}:
594: 
595: \begin{propA}
596: {\it A $\s$-positive braid word does not represent¨$1$.}
597: \end{propA}
598: 
599: \begin{propC}
600: {\it Every braid except¨$1$ can be represented by a
601: $\s$-positive word or by a $\s$-negative word.}
602: \end{propC}
603: 
604: It is then straightforward to order the braids:
605: 
606: \begin{defi}
607: If $\bx, \by$ are braids, we say that $\bx \sm \by$ holds
608: if the braid¨$\bx\inv\by$ admits at least one $\s$-positive
609: representative.
610: \end{defi} 
611: 
612: It is clear that the relation¨$\sm$ is transitive and
613: compatible with multiplication on the left;  Property¨A
614: implies that $\sm$ has no cycle, hence is a
615: strict partial order, and Property¨C then implies that it is
616: actually a linear order.
617: 
618: As every nonempty positive braid word is
619: $\s$-positive, $\bx \dive \by$ implies $\bx \sme \by$
620: for all positive braids¨$\bx, \by$, but the converse is
621: not true: $\ss1$ is not a left divisor of¨$\ss2$, but $\ss1
622: \sm \ss2$ holds, since $\ss1\inv\ss2$ is a $\s$-positive
623: word. 
624: 
625: \begin{exam} \label{X:List}
626: The increasing enumeration of the set¨$\DDD3{}$ is
627: $$\tt 1 \sm a \sm b \sm ba \sm ab \sm aba.$$
628: For
629: instance, we have $\tt ba \sm ab$, \ie, $\ss2\ss1
630: \sm \ss1\ss2$, as the quotient, namely $\ss1\inv
631: \ss2\inv \ss1 \ss2$ (or $\tt ABab$), also admits the
632: expression $\ss2 \ss1\inv$, a $\s$-positive word.
633: Similarly, the reader can check that the increasing
634: enumeration of¨$\DDD32$ is the one given in
635: Example~\ref{X:Enum}.
636: \end{exam}
637: 
638: \begin{lemm} \label{L:Extend}
639: The linear ordering¨$\sm$ extends the left divisibility
640: ordering¨$\div$.
641: \end{lemm}
642: 
643: \begin{proof}
644: By definition, $1 \sm \ss i$ holds for every¨$i$. As the
645: ordering¨$\sm$ is compatible with multiplication on the
646: left, it follows that $\xx \sm \xx \ss i$ holds for
647: all¨$i, \xx$, and, therefore, $\xx \sm \xx \yy$ holds
648: whenever $\yy$ is a non-trivial positive braid.
649: \end{proof}
650: 
651: Lemma¨\ref{L:Extend} implies that $1$ is always the
652: first element of¨$\DDDD\nn\dd$, and
653: $\DD\nn\dd$ is always its last element.
654: It may be noted that a deep result by
655: Laver¨\cite{Lve} shows that, although $\sm$ is not
656: compatible with right multiplication in general,
657: nevertheless $\xx \sm \ss i
658: \xx$ always holds,
659: \ie, $\sm$ also extends the right divisibility ordering. 
660: 
661: By Property~C, every nontrivial braid admits
662: at least one $\s$-positive or $\s$-negative
663: expression. In general, such a $\s$-positive or
664: $\s$-negative expression is not unique, but the main
665: generator in such expressions is uniquely defined:
666: 
667: \begin{lemm} \label{L:Index}
668: If a braid¨$\bx$ admits a $\s$-positive 
669: expression, then the main generators in any
670: two $\s$-positive expressions of¨$\bx$
671: coincide.
672: \end{lemm}
673: 
674: \begin{proof}
675: Assume that $\ww$, $\ww'$ are
676: $\s$-positive  expressions of¨$\bx$, and let
677: $\ss\mm$, $\ss{\mm'}$ be their main
678: generators. Assume for instance $\mm <
679: \mm'$. Then $\ww\inv \ww'$ is a
680: $\s$-positive word, and it represents the
681: trivial braid¨$1$: this contradicts Property¨A.
682: \end{proof}
683: 
684: So, there will be no ambiguity in referring to
685: {\it the} main generator of some non-trivial braid¨$\bx$: 
686: this means the main generator in any $\s$-positive (or
687: $\s$-negative) expression of¨$\bx$.
688: 
689: \begin{rema}
690: Our current definition corresponds to the order¨$<^\phi$
691: of¨\cite{Dgr}. It differs from the one most usually
692: considered in literature in that we refer to the maximal
693: index rather than to the minimal one in the
694: definition of a $\s$-positive word. Switching from one
695: definition to the other amounts to conjugating
696: by¨$\D_\nn$, \ie, to applying the flip automorphism. Results
697: are entirely similar for both versions. However, it is much
698: more convenient to consider the ``max'' choice here, because
699: it guarantees that $B_\nn^+$ is an initial segment
700: of¨$B_{n+1}^+$. Using the ``min'' convention would make the
701: statements of the forthcoming sections less natural.
702: \end{rema}
703: 
704: \section{Measuring the ordered sets¨$\DDDD\nn\dd$}
705: \label{S:Meas}
706: 
707: As was said above, our aim is
708: to investigate the finite ordered sets¨$\DDDD\nn\dd$,
709: and, more generally, $(\Div(\pz), \sm)$ for¨$\pz$ a
710: positive braid. We shall do it by defining numerical
711: parameters that somehow measure their size. The first
712: parameter involves the length of the
713: $\s$-positive words that are, in some natural sense
714: defined below, drawn in the Cayley graph
715: of¨$\DD\nn\dd$. It will be called the {\it complexity}
716: of¨$\DD\nn\dd$, because it is directly connected with
717: the complexity analysis of the handle reduction
718: algorithm of¨\cite{Dfo}. The other parameters involve
719: a filtration of the linear ordering by
720: the¨$\ss i$'s, and they will be called the {\it
721: heights} of¨$\DD\nn\dd$ because they count the jumps
722: of a given height in¨$\DDDD\nn\dd$. 
723: 
724: \subsection{Sigma-positive paths in the Cayley graph}
725: 
726: The first parameter we attach to $(\Div(\pz),
727: \sm)$ involves the $\s$-positive paths in the Cayley
728: graph of¨$\pz$.
729: 
730: We recall that the Cayley graph of
731: the group¨$B_\nn$ with respect to the standard
732: generators $\ss i$ is the labeled graph  with vertex
733: set¨$B_\nn$ and such that there exists a $\ss
734: i$-labeled edge from¨$\bx$ to¨$\by$ if and only if
735: $\by = \bx \ss i$ holds. The Cayley graph of the
736: monoid¨$B_\nn^+$ is obtained by restricting the vertices
737: to¨$B_\nn^+$. Note that the Cayley graph of¨$B_\nn$
738: (and a fortiori of¨$B_\nn^+$) can be seen as a subgraph
739: of the Cayley graph of¨$B_\infty$.
740: 
741: \begin{defi}
742: (Figure¨\ref{F:Cayley})
743: For $\pz$ a positive braid, we denoted by¨$\CG(\pz)$
744: the subgraph of the Cayley graph of¨$B_\infty$
745: obtained by restricting the vertices to¨$\Div(\pz)$,
746: and only keeping those edges that connect two vertices
747: in¨$\Div(\pz)$.
748: \end{defi}
749: 
750: As every element of¨$B_\nn^+$ is a left divisor of¨$\DD\nn\dd$
751: for¨$\dd$ large enough, the Cayley graph of¨$B_\nn^+$ is 
752: the union of all graphs¨$\GG n k$ when $\dd$¨varies.
753: 
754: \begin{figure} [htb]
755: \begin{picture}(110,51)(0, -3)
756: \put(5,8){\includegraphics{Path31.eps}}
757: \put(50,0){\includegraphics{Path32.eps}}
758: \put(2,20){$1$}
759: \put(10,34){$\ss2$}
760: \put(27,34){$\ss2\ss1$}
761: \put(10,7){$\ss1$}
762: \put(27,7){$\ss1\ss2$}
763: \put(36,20){$\D_3$}
764: \put(47,20){$1$}
765: \put(55,33){$\ss2$}
766: \put(60,44){$\ss2^2$}
767: \put(70,44){$\ss2^2\ss1$}
768: \put(80,44){$\ss2\ss1^2$}
769: \put(90,44){$\ss2\ss1^2\ss2$}
770: \put(65,33){$\ss2\ss1$}
771: \put(65,33){$\ss2\ss1$}
772: \put(86,33){$\D_3\ss1$}
773: \put(99,33){$\D_3\ss1\ss2$}
774: \put(55,8){$\ss1$}
775: \put(65,8){$\ss1\ss2$}
776: \put(86,8){$\D_3\ss2$}
777: \put(99,8){$\D_3\ss2\ss1$}
778: \put(82,20){$\D_3$}
779: \put(108,20){$\D_3^2$}
780: \put(60,-3){$\ss1^2$}
781: \put(70,-3){$\ss1^2\ss2$}
782: \put(80,-3){$\ss1\ss2^2$}
783: \put(90,-3){$\ss1\ss2^2\ss1$}
784: \end{picture}
785: \caption{\smaller The graphs
786: of¨$\GG3{}$ and $\GG32$;
787: the dotted edges represent¨$\ss1$, the plain
788: ones¨$\ss2$; observe that the graph
789: of¨$\DD32$ is not planar; in grey: two
790: $\s$-positive words traced in the graphs, namely
791: $\tt aAbab$ and $\tt bbabAbab$ ({\it cf.}
792: Lemma¨\ref{L:Path})}
793: \label{F:Cayley}
794: \end{figure}
795: 
796: A path in the Cayley graph can be specified by its
797: initial vertex and the list of the labels of its
798: successive edges, \ie, by a braid word. For each¨$i <
799: n$ and each¨$\bx$ in¨$B_\nn$, there is exactly one $\ss
800: i$-labeled edge with target¨$\bx$, and one
801: $\ss i$-labeled edge with source¨$\bx$ in the Cayley
802: graph of¨$B_\nn$. Hence, in the complete
803: Cayley graph of¨$B_\nn$, for each initial vertex¨$\bx$ and
804: each $\nn$-braid word¨$\ww$, there is always one path
805: labeled¨$\ww$ starting from¨$\bx$. When we restrict to some
806: fragment¨$\G$, this need not be the case, but we have an
807: unambiguous notion of¨$\ww$ being drawn in¨$\G$
808: from¨$\bx$. Formally, this corresponds to
809: 
810: \begin{defi}
811: For $\G$ a subgraph of the Cayley graph
812: of¨$B_\infty$, and $\bx$ a vertex in¨$\G$, we say
813: that a braid word¨$\ww$ is {\it drawn} from¨$\bx$
814: in¨$\G$ if, for every prefix¨$\vv\ss i$ (\resp $\vv\ss
815: i\inv$) of¨$\ww$, there exists a $\ss i$-labeled edge
816: starting from (\resp finishing at)¨$\bx
817: \, \cl{\vv}$ in¨$\G$.
818: \end{defi}
819: 
820: For instance, we can check on Figure¨\ref{F:Cayley} that
821: the word¨$\ss1^2$ is drawn from¨$\ss2$
822: in¨$\GG32$, but not in¨$\GG3{}$. In algebraic terms,
823: we have the following characterization:
824: 
825: \begin{lemm} \label{L:Path}  
826: Assume that $\pz$ is a positive braid, and $\ww$ is a
827: braid word. Then $\ww$ is drawn from¨$\bx$ in¨$\CG(\pz)$
828: if and only if $\bx \cl{\vv} \dive \pz$ holds for each
829: prefix¨$\vv$ of¨$\ww$.
830: \end{lemm}               
831: 
832: \begin{proof}
833: The condition is sufficient. Indeed, assume it is
834: satisfied by¨$\ww$, and $\vv \ss i$ is a prefix of¨$\ww$.
835: Then, by hypothesis, $\bx \cl{\vv}$ and
836: $\bx \cl{\vv}\ss i$ are left divisors of¨$\pz$, hence
837: are vertices in¨$\CG(\pz)$, and, therefore,
838: there is a $\ss i$-labeled edge between¨$\bx \cl{\vv}$ and
839: $\bx \cl{\vv} \ss i$ in¨$\CG(\pz)$. The argument is
840: similar for a prefix of the form $\vv \ss i\inv$. Using
841: induction on the length of¨$\ww$, we deduce that $\ww$ is
842: drawn from¨$\bx$ in¨$\CG(\pz)$. 
843: 
844: Conversely, if there exists a $\ww$-labeled path
845: from¨$\bx$ in¨$\CG(\pz)$, then, for each prefix¨$\vv$
846: of¨$\ww$, the braid¨$\bx \cl{\vv}$ has to represent some
847: vertex in¨$\CG(\pz)$, hence it is a left divisor
848: of¨$\pz$. 
849: \end{proof}
850: 
851: For¨$\pz$ a positive braid, we shall investigate the
852: $\s$-positive words drawn in the graph¨$\CG(\pz)$. It is
853: clear that, even if $\Div(\pz)$ is a finite set,
854: arbitrary long words are drawn in¨$\CG(\pz)$ whenever the
855: latter contains at least 2¨vertices, \ie, $\pz$ is
856: not¨$1$. The example of Figure¨\ref{F:Cayley} shows that
857: restricting to $\s$-positive words does not change the
858: result: for instance, for each¨$\kk$, the
859: word $(\ss1\ss1\inv)^\kk\ss2\ss1\ss2$ is a
860: $\s$-positive expression of¨$\D_3$, and it is drawn
861: in¨$\GG3{}$. So we cannot hope for any finite upper
862: bound on the length of the $\s$-positive words drawn
863: in¨$\CG(\pz)$ in general. Now, the situation changes
864: if we concentrate on the main generators, \ie, we
865: forget about the generators with non-maximal index.
866: 
867: \begin{lemm} \label{L:Finite}
868: Assume that $\G$ is subgraph of the Cayley graph
869: of¨$B_\infty$, and $\ww$ is a $\s$-positive word drawn
870: in¨$\CG(\pz)$. Then the number of occurrences of the main
871: generator in¨$\ww$ is at most the number of
872: non-terminal vertices in¨$\G$.
873: \end{lemm}
874: 
875: \begin{proof}
876: Assume that $\ww$ is drawn from¨$\bx$ in¨$\G$. Let
877: $\ss\mm$ be the main generator in¨$\ww$. As there is at
878: most one $\ss\mm$-labeled edge starting from each vertex
879: of¨$\G$, it suffices to show that the number of¨$\ss\mm$'s
880: in¨$\ww$ is bounded above by the number of
881: $\ss\mm$-edges in¨$\G$. Hence, it suffices to show that
882: the path¨$\gamma$ associated with¨$\ww$ cannot cross the
883: same $\ss\mm$-edge twice. Now assume that some
884: $\ss\mm$-edge starts from the vertex¨$\by$, and that $\g$
885: crosses this edge twice. This means that $\g$ contains a
886: loop from¨$\by$ to¨$\by$. Let
887: $\vv$ be the subword of¨$\ww$ labeling that loop. By
888: construction, $\vv$ begins with¨$\ss\mm$, it contains
889: no¨$\ss\mm\inv$ and no $\ss i^{\pm1}$ with $i > m$ as
890: it is a subword of¨$\ww$, and it represents the
891: braid¨$1$ as it labels a loop in the Cayley graph
892: of¨$B_\infty$: this means that $\vv$ is a
893: $\s$-positive word representing¨$1$, which contradicts
894: Property¨A.
895: \end{proof}
896: 
897: Lemma¨\ref{L:Finite} applies in particular to every
898: graph¨$\CG(\pz)$ with¨$\pz$ a positive braid. So we can
899: introduce our first parameter measuring
900: the size of the ordered set¨$\DDiv\pz$:
901: 
902: \begin{defi}
903: (Figure~\ref{F:Cayley})
904: For¨$\pz$ a positive braid with main generator¨$\ss\mm$,
905: the {\it complexity}¨$\cc(\pz)$ of¨$\pz$ is defined to
906: be the maximal number of¨$\ss\mm$'s in a $\s$-positive
907: word drawn in¨$\CG(\pz)$. 
908: \end{defi}
909: 
910: \begin{exam}
911: The word¨$\ss2 \ss1 \ss2$ is a
912: $\s$-positive word drawn from¨$1$ in¨$\GG3{}$, and
913: it contains two¨$\ss2$'s, hence we have $\ccc3{}
914: \ge 2$. Actually, it is not hard to obtain the
915: exact value $\ccc3{} = 2$. Indeed, if a
916: $\s$-positive path¨$\g$ contains the two
917: $\ss2$-edges starting from¨$1$ and¨$\ss1\ss2$, it
918: cannot come back to¨$\ss2$ for possibly crossing the
919: third $\ss2$-edge; and if $\g$ contains the
920: $\ss2$-edge that starts from¨$\ss1$, it can never
921: come back to¨$1$ or to¨$\ss2\ss1$ and therefore
922: contains at most one $\ss2$-edge. 
923: As we have $\DD3\dd = (\ss2 \ss1 \ss2)^\dd$,
924: we deduce $\ccc3\dd \ge 2\dd$ for every¨$\dd$;
925: this value is certainly not optimal, since the
926: example displayed in Figure¨\ref{F:Cayley} contains
927: five¨$\ss2$'s, proving $\ccc32 \ge 5$---the exact
928: value is¨$6$, and, more generally, we have $\ccc\nn\dd
929: = 2^{\dd+1} - 2$, as will be seen in
930: Section¨\ref{S:Deter}.
931: \end{exam}
932: 
933: \begin{rema}
934: Restricting to $\s$-positive words drawn in¨$\CG(\pz)$
935: is essential: for instance, for each¨$\kk$, we have
936: \begin{equation} \label{E:Unbounded}
937: \D_3 = \ss2^{\kk+1} \ss1\ss2\ss1^{-\kk},
938: \end{equation}
939: a $\s$-positive word containing $\kk + 2$ letters¨$\ss2$.
940: Now, for¨$\kk \ge 1$, the word involved
941: in¨\eqref{E:Unbounded} is not drawn in¨$\GG3{1}$, as
942: its prefix¨$\ss2^2$ is not. Thus the
943: parameter¨$\cc(\pz)$ really involves the left
944: divisors of¨$\pz$.
945: \end{rema}
946: 
947: A direct application of Lemma¨\ref{L:Finite} gives:
948: 
949: \begin{prop}
950: Every positive braid has a finite complexity; more
951: precisely, for¨$\pz$ of length¨$\ell$ in¨$B_\nn^+$  with
952: $\nn \ge 3$, we have $\cc(\pz) \le (\nn-1)^\ell$.
953: \end{prop}
954: 
955: \begin{proof}
956: The number of non-terminal vertices in¨$\CG(\pz)$, \ie,
957: the number of proper left divisors of¨$\pz$, is at most
958: $1 + (\nn-1) + (\nn-1)^2 + \cdots + (\nn-1)^{\ell-1}$.
959: \end{proof}
960: 
961: As the length of any positive expression of¨$\D_\nn$ is
962: $\nn(\nn-1)/2$, we obtain in particular for all¨$\nn, \dd$
963: \begin{equation} \label{E:Coarse}
964: \ccc\nn\dd \le
965: (\nn-1)^{\dd\nn(\nn-1)/2}.
966: \end{equation}
967: 
968: Before going further, let us observe that, in the
969: definition of the complexity of¨$\pz$, we can restrict
970: to decompositions of¨$\pz$, \ie, instead of considering
971: paths starting from and finishing at arbitrary vertices,
972: we can restrict to paths starting from¨$1$ and finish
973: at¨$\pz$:
974: 
975: \begin{lemm} \label{L:Decomp}
976: Assume that $\pz$ is a positive braid with main
977: generator¨$\ss\mm$. Then $\cc(\pz)$ is the
978: maximal number of¨$\ss\mm$'s in any $\s$-positive
979: decomposition of¨$\pz$ drawn in¨$\CG(\pz)$.
980: \end{lemm}
981: 
982: \begin{proof}
983: Let $\cc'(\pz)$ be the number involved in the above
984: statement. Clearly we have $\cc'(\pz) \le \cc(\pz)$.
985: Conversely, assume that $\ww$ is drawn in¨$\CG(\pz)$
986: from¨$\bx$, and that the $\ww$-labeled path starting
987: from¨$\bx$ finishes at¨$\by$. Let $\uu$ be a positive
988: expression of¨$\bx$, and $\vv$ be a positive expression
989: of¨$\by\inv \pz$. The latter exists as, by hypothesis,
990: $\by$ is a left divisor of¨$\pz$. Then $\uu\ww\vv$ is a
991: $\s$-positive decomposition of¨$\pz$ drawn in¨$\CG(\pz)$.
992: Hence we have $\cc'(\pz) \ge \cc(\pz)$.
993: \end{proof}
994: 
995: 
996: \begin{rema}
997: Let us call Property~A$^*$ the fact that all
998: numbers¨$\ccc\nn\dd$ are finite. Above we derived 
999: Property~A$^*$ from Property¨A. Actually, the
1000: implication is an equivalence, \ie, we can also
1001: deduce Property¨A from Property~A$^*$. Indeed,
1002: assume that some $\s$-positive braid word¨$\ww$
1003: represents¨$1$. The word¨$\ww$ may involve negative
1004: letters and the problem is to find a vertex¨$\bx$
1005: such that there exists a path labeled¨$\ww$
1006: from¨$\bx$ in some¨$\GG\nn\dd$. Let $\ss m$ be the
1007: main generator in¨$\ww$. The
1008: word¨$\ww$ has finitely many prefixes, say
1009: $\ww_0$, \dots, $\ww_\ell$. By Garside's theory, each
1010: word¨$\ww_i$ is equivalent to a word of the form¨$u_i\inv
1011: \vv_i$ with $\uu_i, \vv_i$ positive. Let $\bx$
1012: be the least common left multiple of the
1013: positive braids¨$\cl{\uu_0}$,
1014: \dots, $\cl{\uu_\ell}$. Then, for each¨$i$, the
1015: braid¨$\bx\cl{\ww_i}$ is positive. Moreover, there
1016: exist¨$\nn$ and¨$\dd$ such that $\bx\cl{\ww_0}$, \dots,
1017: $\bx\cl{\ww_\ell}$ all are divisors of¨$\DD\nn\dd$. This
1018: means that the word¨$\ww$ is drawn from¨$\bx$ 
1019: in¨$\GG\nn\dd$, and the associated path is a loop
1020: around¨$\bx$. It follows that $\ww^\kk$ is drawn
1021: in¨$\GG\nn\dd$ from¨$\bx$ for each¨$\kk$. By
1022: construction, $\ww^\kk$ contains at least
1023: $\kk$¨generators¨$\ss m$, hence $\ccc\nn\dd$ cannot
1024: be finite.
1025: \end{rema}
1026: 
1027: \subsection{Connection with handle reduction}
1028: 
1029: Handle reduction¨\cite{Dfo} is an algorithmic solution to
1030: the word problem of braids that relies on the braid
1031: ordering---actually the most efficient method
1032: available to-date in practice. It is proved to be
1033: convergent, but the complexity upper bound resulting
1034: from the argument of¨\cite{Dfo} is exponential with
1035: respect to the length of the input word, seemingly
1036: very far from sharp. 
1037: 
1038: Each step of handle reduction involves a specific
1039: generator¨$\ss i$, and, for an induction, the point is to
1040: obtain an upper bound on the number of reduction steps
1041: involving the main generator. The latter will naturally
1042: be called the {\it main} reduction steps. The
1043: connection between handle reduction and the complexity
1044: as defined above relies on the following
1045: technical result: 
1046: 
1047: \begin{lemm} \cite{Dfo}
1048: Assume that $\pz$ is a positive braid with main
1049: generator¨$\ss\mm$, and $\ww$ is drawn in¨$\CG(\pz)$.
1050: Then, for each sequence of handle reductions from¨$\ww$,
1051: \ie, each sequence¨$\vec\ww$ with $\ww_0= \ww$ such that
1052: $\ww_\kk$ is obtained by reducing one handle
1053: from¨$\ww_{\kk-1}$ for each¨$\kk$, there exists a
1054: witness-word¨$\uu$ that is  $\s$-positive, drawn
1055: in¨$\Div(\pz)$, and such that the number of¨$\ss\mm$'s
1056: in¨$\uu$ is the number of main reductions in¨$\vec\ww$.
1057: \end{lemm}
1058: 
1059: It follows that the number of main reduction steps in any
1060: sequence of handle reductions starting with a word drawn
1061: in¨$\CG(\pz)$ is bounded above by¨$\cc(\pz)$. In
1062: particular, if we start with an $n$¨strand braid
1063: word¨$\ww$ of length¨$\ell$, then it is easy to show
1064: that $\ww$ is drawn in¨$\GG\nn\ell$, and, applying the
1065: upper bound of¨\eqref{E:Coarse}, we deduce an
1066: upper bound for the number of possible main reductions
1067: from¨$\ww$, one that is exponential with respect
1068: to¨$\ell$.
1069: 
1070: A natural way of improving this coarse upper bound would
1071: be to determine the value of¨$\ccc\nn\dd$ more precisely.
1072: This will be done in Section¨\ref{S:Deter} below.
1073: However, the explicit formulas show that, for $\nn \ge
1074: 3$, the growth rate with respect to¨$\dd$ is exponential,
1075: thus discarding any hope of proving the expected
1076: polynomial upper bound for the number of reduction steps
1077: by this approach.
1078: 
1079: \subsection{A filtration of the braid ordering}
1080: 
1081: We now introduce new numerical parameters for the ordered
1082: sets¨$\DDiv\pz$. These numbers appear in connection with a
1083: natural filtration of the ordering¨$\sm$, using an
1084: increasing sequence of partial orderings that we
1085: introduce now.
1086: 
1087: By Lemma¨\ref{L:Index}, the index of the main generator
1088: of a non-trivial braid is well defined. We can use this
1089: index to  measure the height of the jump between two
1090: braids¨$\bx, \by$ satisfying $\bx \sm \by$:
1091: 
1092: \begin{defi}
1093: For $\bx, \by$ in~$B_\infty$ and $\rr
1094: \ge 1$, we say that $\bx \sm_\rr \by$ holds
1095: or, equivalently, that $(\bx, \by)$ is a
1096: {\it $\rr$-jump}, if $\bx\inv\by$ admits a
1097: $\s$-positive expression in which the main
1098: generator is~$\ss\mm$ with $\mm \ge \rr$.
1099: \end{defi}
1100: 
1101: \begin{lemm}
1102: For each $\rr \ge 1$, the
1103: relation~$\sm_\rr$ is a strict partial
1104: order that refines~$\sm$; the
1105: relation~$\sm_1$ coincides with~$\sm$,
1106: and $\rr \le \qq$ implies that $\sm_\qq$
1107: refines~$\sm_\rr$. 
1108: \end{lemm}
1109: 
1110: \begin{proof}
1111: That $\sm_\rr$ is transitive follows from
1112: the fact that the concatenation of a
1113: $\s$-positive word with main
1114: generator~$\ss{\mm}$ and a $\s$-positive word 
1115: with main generator~$\ss{\mm'}$ is a
1116: $\s$-positive word with main
1117: generator~$\ss{\max(\mm, \mm')}$. 
1118: \end{proof}
1119: 
1120: In the sequel, we consider the $\sm_\rr$-chains
1121: included in~$\Div(\pz)$, and their length:
1122: 
1123: \begin{defi}
1124: For $\pz$ a positive braid and $\rr \ge 1$,
1125: we define the {\it $\rr$-height}¨$\hh\rr(\pz)$ of¨$\pz$
1126: to be the maximal length of a $\sm_\rr$-chain included
1127: in~$\Div(\pz)$. 
1128: \end{defi}
1129: 
1130: Before giving examples, we observe the
1131: connection between¨$\hh\rr(\pz)$ and
1132: the increasing enumeration of the
1133: set~$\Div(\pz)$:
1134: 
1135: \begin{lemm} \label{L:Enum}
1136: Let $\pz$ be a positive braid and $\rr \ge
1137: 1$. Then $\hh\rr(\pz)-1$ is the
1138: number of $\rr$-jumps in the increasing
1139: enumeration of~$\DDiv\pz$.
1140: \end{lemm}
1141: 
1142: \begin{proof}
1143: If the number of $\rr$-jumps in the
1144: increasing enumeration of~$\Div(\pz)$
1145: is~$\NN_\rr-1$, we can extract from~$\Div(\pz)$ a
1146: $\sm_\rr$-chain of length~$\NN_\rr$. Conversely,
1147: assume that $(\by_0, \dots \by_{\NN_\rr})$ is a
1148: $\sm_\rr$-chain in~$\Div(\pz)$. 
1149: Let $\pz_0 \sm \ldots \sm \pz_\NNN$ be the
1150: increasing enumeration of~$\Div(\pz)$. As
1151: $\sm_\rr$ refines~$\sm$, there exists an
1152: increasing function~$f$ of $\{0, \ldots,
1153: \NN_\rr\}$ into $\{0, \ldots, \NNN\}$ such that
1154: $\by_i = \pz_{f(i)}$ holds for every~$i$. Now
1155: the hypothesis $\pz_{f(i)} \sm_\rr
1156: \pz_{f(i+1)}$ implies that there exists at
1157: least one $\rr$-jump between $\pz_{f(i)}$
1158: and~$\pz_{f(i+1)}$. Indeed, by
1159: Lemma~\ref{L:Index}, it is impossible that a
1160: concatenation of $\mm$-jumps with $\mm < \rr$
1161: results in a $\rr$-jump. So the number
1162: of $\rr$-jumps in $(\pz_0, \dots, \pz_\NNN)$ is
1163: at least~$\NN_\rr$.
1164: \end{proof}
1165: 
1166: In other words, in order to determine¨$\hh\rr(\pz)$,
1167: there is no need to consider arbitrary chains: it
1168: is enough to consider the maximal chain
1169: obtained by enumerating $\Div(\pz)$ exhaustively. 
1170: 
1171: \begin{exam} \label{X:Height}
1172: Refining the increasing enumeration of¨$\DDD3{}$
1173: given in Example¨\ref{X:List} by indicating for
1174: each step the height of the corresponding jump, we
1175: obtain:
1176: \begin{equation} \label{E:List31}
1177: \tt 1 \sm_1 a \sm_2 b
1178: \sm_1 ba
1179: \sm_2 ab \sm_1 \D_3,
1180: \end{equation}
1181: where we recall $\tt a, b, \dots$ stand for
1182: $\ss1, \ss2, \dots$. For instance,
1183: $\tt (ba, ab)$ is a $2$-jump, as we have $\tt
1184: (ba)\inv(ab) = ABab = AabA = bA$, a
1185: $\s$-positive decomposition with main
1186: generator~$\ss2$. The number of $1$-jumps
1187: in~\eqref{E:List31},
1188: \ie, the number of symbols~$\sm_\rr$ with
1189: $\rr \ge 1$, is~$5$, while
1190: the number of~$2$-jumps is~$2$, so, by
1191: Lemma¨\ref{L:Enum}, we deduce
1192: $\hhh13{} = 6$ and $\hhh23{} = 3$.
1193: Similarly, we obtain for¨$\DD32$ 
1194: \begin{multline*}
1195: \tt 1 \sm_1 a \sm_1 aa \sm_2 b \sm_1 ba 
1196: \sm_1 baa \sm_2 bb \sm_1 bba 
1197: \sm_2 ab \sm_1
1198: aba  \sm_1 abaa \sm_2 abb 
1199: \\ \tt 
1200: \sm_1 abba \sm_2
1201: aab
1202: \sm_1  aaba \sm_1 aabaa \sm_2 baab \sm_1 baaba
1203: \sm_1 baabaa,
1204: \end{multline*}
1205: leading to $\hhh132 = 19$ and $\hhh232 = 7$.
1206: \end{exam}
1207: 
1208: \begin{prop} \label{P:Decreasing}
1209: $(i)$ For every braid~$\pz$ in~$B_\nn^+$, we
1210: have
1211: \begin{equation}
1212: \hh1(\pz) = \card\Div(\pz)  \ge
1213: \hh2(\pz)
1214: \ge \dots \ge \hh\nn(\pz) = 1.
1215: \end{equation}
1216: 
1217: $(ii)$ For all positive braids~$\pz, \pz'$
1218: and $\rr \ge 1$, we have
1219: \begin{equation} \label{E:Addition}
1220: \hh\rr(\pz\pz') \ge \hh\rr(\pz) + \hh\rr(\pz').
1221: \end{equation}
1222: \end{prop}
1223: 
1224: \begin{proof}
1225: $(i)$ A $\sm_1$-chain is simply a $\sm$-chain,
1226: hence every subset of~$\Div(\pz)$ gives such
1227: a chain. So the maximal $\sm_1$-chain in
1228: $\Div(\pz)$ is $\Div(\pz)$ itself, and
1229: $\hh1(\pz)$ is the cardinality of~$\Div(\pz)$.
1230: 
1231: On the other hand, no $\sm_\nn$-chain
1232: in~$B_\nn^+$ has length more than~$1$, as the
1233: main generator of a $\s$-positive $\nn$~strand
1234: braid word cannot be $\ss\nn$ or above. So
1235: $\hh\nn(\pz)$ is~$1$.
1236: 
1237: Then, for $\qq \le \rr$, every
1238: $\sm_\rr$-chain is a $\sm_\qq$-chain, which
1239: implies $\hh\rr(\pz) \ge \hh\qq(\pz)$.
1240: 
1241: Point~$(ii)$ is obvious, as the
1242: concatenation of two $\sm_\rr$-chains is a
1243: $\sm_\rr$-chain.
1244: \end{proof}
1245: 
1246: From~\eqref{E:Addition} we deduce
1247: $\hh\rr(\pz^\dd) \ge \dd \cdot \hh\rr(\pz)$
1248: for all~$\rr, \pz$. By Lemma¨\ref{L:Degree}, every
1249: divisor of~$\DD\nn\dd$ can be decomposed as the
1250: product of at most~$\dd$ divisors
1251: of~$\D_\nn$, and the latter are
1252: $\nn!$~in number, so we obtain the (coarse) bounds
1253: \begin{equation}
1254: \dd \cdot \hhh\rr\nn{} \le \hhh\rr\nn\dd \le
1255: (\nn!)^\dd
1256: \end{equation}
1257: for all~$\rr, \nn, \dd$. Better estimates
1258: will be given below.
1259: 
1260: \begin{rema}
1261: Instead of restricting to subsets of~$B_\infty$ of the
1262: form~$\Div(\pz)$, we can define the complexity and the
1263: $\rr$-height for every (finite) set of braids~$\XX$. Most
1264: general results extend, but, when $\XX$ is not closed
1265: under left division, nothing can be said about the
1266: number of~$\ss\rr$'s involved in a
1267: $\rr$-jump. Considering such an extension is not useful
1268: here. 
1269: \end{rema}
1270: 
1271: \subsection{Connection with the complexity}
1272: 
1273: We shall now connect the complexity¨$\cc(\pz)$ with
1274: the numbers¨$\hh\rr(\pz)$ just defined. The result
1275: is simple:
1276: 
1277: \begin{prop} \label{P:MainHeight}
1278: For $\pz$ a positive braid with main
1279: generator¨$\ss\mm$, we have
1280: \begin{equation}
1281: \cc(\pz) = \hh\mm(\pz) - 1.
1282: \end{equation}
1283: In particular, for $\nn \ge 2$ and $\dd
1284: \ge 0$, we have 
1285: \begin{equation} \label{E:Complexity}
1286: \ccc\nn\dd = \hhh{\nn-1}\nn\dd - 1.
1287: \end{equation}
1288: \end{prop}
1289: 
1290: One inequality is easy:
1291: 
1292: \begin{lemm} \label{L:MainHeight}
1293: For $\pz$ a positive braid with main
1294: generator¨$\ss\mm$, we have $\cc(\pz) \le
1295: \hh\mm(\pz) - 1$.
1296: \end{lemm}
1297: 
1298: \begin{proof} 
1299: The argument is reminiscent of that used for
1300: Lemma~\ref{L:Enum}, but requires a little more care.
1301: Assume that $\ww$ is a $\s$-positive word drawn
1302: in¨$\CG(\pz)$ from¨$\bx$ containing
1303: $\NN_\mm$¨occurrences of¨$\ss\mm$. By
1304: Lemma¨\ref{L:Decomp}, we can assume $\bx = 1$ without
1305: loss of generality. Let $\pz_0
1306: \sm \pz_1 \sm \dots \sm \pz_\NNN$ be the increasing
1307: enumeration of¨$\Div(\pz)$. By definition, all prefixes
1308: of¨$\ww$ represent divisors of¨$\pz$, so, letting¨$\ell$
1309: be the length of¨$\ww$, there exists a mapping $f : \{0,
1310: \dots, \ell\} \to  \{0,
1311: \dots, \NNN\}$ such that, for each¨$\kk$, the
1312: length¨$\kk$ prefix of¨$\ww$
1313: represents¨$\pz_{f(\kk)}$. By construction, we have
1314: $f(0) = 0$ and $f(\ell) = \NNN$.
1315: 
1316: The difference with Lemma~\ref{L:Enum} is
1317: that $f$ need not be increasing. Now, let
1318: $p_1$, \dots, $p_{\NN_\mm}$ be the
1319: $\NN_\mm$~positions in¨$\ww$ where the
1320: generator~$\ss m$ occurs, completed with
1321: $p_0 = 0$. Then, in the prefix of¨$\ww$ of
1322: length¨$p_1$, \ie, in the subword of¨$\ww$
1323: corresponding to positions from¨$p_0+1$
1324: to¨$p_1$, there is one¨$\ss m$, plus
1325: letters¨$\ss i^{\pm1}$ with $i <
1326: m$ (Figure¨\ref{F:Jump}). This subword is
1327: therefore $\s$-positive, hence we
1328: must have $\pz_{f(p_0)} \sm \pz_{f(p_1)}$,
1329: which requires
1330: $f(p_0) < f(p_1)$. Moreover, the quotient
1331: $\pz_{f(p_0)}\inv
1332: \pz_{f(p_1)}$ is a braid that admits at least
1333: one $\s$-positive expression containing¨$\ss m$,
1334: hence $\pz_{f(p_0)}
1335: \sm_\mm \pz_{f(p_1)}$ holds. Now the same
1336: is true between¨$f(p_1)$ and¨$f(p_2)$,
1337: \etc\ Hence the number of $\mm$-jumps in the
1338: increasing enumeration of~$\Div(\pz)$ is at
1339: least~$\NN_\mm$, \ie, we have $\hh\mm(\pz) \ge
1340: \NN_\mm + 1$.
1341: \end{proof}
1342: 
1343: \begin{figure} [htb]
1344: \begin{picture}(120,27)(0, -0)
1345: \put(0,2){\includegraphics{Jump.eps}}
1346: \put(87,23){almost-positive}
1347: \put(87,19){expression}
1348: \put(102,15){of¨$\pz$}
1349: \put(87,5){increasing }
1350: \put(87,1){enumeration of $\Div(\pz)$}
1351: \put(0,0){$\pz_0$}
1352: \put(10,0){$\pz_1$}
1353: \put(20,0){$\pz_2$}
1354: \put(24,0){$\sm_\mm$}
1355: \put(30,0){$\pz_3$}
1356: \put(40,0){$\pz_4$}
1357: \put(50,0){$\pz_5$}
1358: \put(54,0){$\sm_\mm$}
1359: \put(60,0){$\pz_6$}
1360: \put(70,0){$\pz_7$}
1361: \put(80,0){$\pz_8$}
1362: \put(2,9){$\scriptstyle f(0)$}
1363: \put(12,9){$\scriptstyle f(2)$}
1364: \put(22,9){$\scriptstyle f(1)$}
1365: \put(32,9){$\scriptstyle f(4)$}
1366: \put(42,9){$\scriptstyle f(3)$}
1367: \put(62,9){$\scriptstyle f(5)$}
1368: \put(62,12){$\scriptstyle f(7)$}
1369: \put(72,9){$\scriptstyle f(6)$}
1370: \end{picture}
1371: \caption{\smaller Proof of
1372: Lemma~\ref{L:MainHeight}: the main
1373: generator~$\ss\mm$ corresponds to the bold
1374: arrow: the function¨$f$ need not be
1375: increasing, but the projection of a bold
1376: arrow upstairs must include at least one bold
1377: arrow downstairs, \ie, at least one
1378: $\mm$-jump.}
1379: \label{F:Jump}
1380: \end{figure}
1381: 
1382: It remains to prove the second inequality in
1383: Proposition~\ref{P:MainHeight}, \ie, to prove
1384: that, if $\pz$ is a positive $\nn$-braid
1385: satisfying $\hh{\mm}(\pz) = \NN+1$, then
1386: $\pz$ admits an almost-positive expression
1387: containing $\NN$~generators~$\ss\mm$. 
1388: The problem is as follows: if $\pz$ is a
1389: positive braid and $\bx, \by$ are left
1390: divisors of~$\pz$ satisfying $\bx \sm \by$,
1391: then, by definition, the quotient $\bx\inv
1392: \by$ admits some $\s$-positive
1393: expression~$\ww$, but nothing {\it
1394: a priori } guarantees that $\ww$ be drawn
1395: in~$\CG(\pz)$. In other words, we might have
1396: $\bx \sm \by$ but no $\s$-positive witness
1397: for this inequality inside~$\Div(\pz)$.
1398: This however cannot happen, but the proof
1399: requires a rather delicate argument. 
1400: 
1401: \begin{prop} \label{P:Handle}
1402: Let $\pz$ be a positive braid. Then, 
1403: for all $\bx, \by$ in¨$\Div(\pz)$,
1404: the following are equivalent:
1405: 
1406: $(i)$ The relation $\bx \sm \by$ holds,
1407: \ie, there exists a $\s$-positive path
1408: from~$\bx$ to~$\by$ in the Cayley graph
1409: of~$B_\infty$;
1410: 
1411: $(ii)$ There exists a $\s$-positive path 
1412: from¨$\bx$ to¨$\by$ in the Cayley graph
1413: of~$B_\nn$;
1414: 
1415: $(iii)$ There exists a $\s$-positive path 
1416: from¨$\bx$ to¨$\by$ in¨$\CG(\pz)$.
1417: \end{prop}
1418: 
1419: \begin{proof}
1420: Clearly $(iii)$ implies¨$(ii)$, which 
1421: implies¨$(i)$. We shall prove that $(i)$
1422: implies~$(iii)$---and thus reprove that
1423: $(i)$ implies~$(ii)$, which was first
1424: proved in¨\cite{Lrb}---by
1425: using the handle reduction method of¨\cite{Dfo, Dgr}. The
1426: problem is to prove that, among all $\s$-positive paths
1427: connecting¨$\bx$ to¨$\by$ in the Cayley graph
1428: of¨$B_\infty$, at least one is drawn in~$\CG(\pz)$.
1429: 
1430: Now, let $\uu, \vv$ be positive words representing
1431: $\bx$ and¨$\by$. Then the word¨$\uu\inv\vv$
1432: represents¨$\bx\inv\by$, and, by hypothesis, it is drawn
1433:  in~$\CG(\pz)$ from¨$\bx$. Handle
1434: reduction is an operation that transforms a braid word into
1435: equivalent words and eventually produces a $\s$-positive
1436: word if it exists. It is proved in¨\cite{Dfo} that, for
1437: every $\nn$¨strand braid word¨$\ww$, there exists a
1438: finite fragment¨$\G_\ww$ of the Cayley
1439: graph of~$B_\nn^+$ and a vertex¨$\bx_\ww$
1440: of¨$\G_\ww$ such that
1441: $\ww$ and all words obtained from¨$\ww$ by handle reduction
1442: are drawn from¨$\bx_\ww$ in¨$\G_\ww$. Moreover,
1443: when $\ww$ has the form¨$\uu\inv\vv$ with $\uu, \vv$
1444: positive, then all vertices in¨$\G_\ww$ are the
1445: left divisors of the least common right multiple of
1446: the braids represented by¨$\uu$ and¨$\vv$, here
1447: ¨$\bx$ and¨$\by$,  while $\bx_\ww$ is the braid
1448: represented by¨$\uu$, \ie,¨$\bx$. As
1449: $\bx$ and¨$\by$ are divisors of¨$\pz$, so is their
1450: least common right multiple, and the
1451: graph¨$\G_\ww$ is included in~¨$\CG(\pz)$.
1452: It follows that every word obtained
1453: from¨$\uu\inv\vv$ using handle reduction is
1454: drawn from¨$\bx$ in¨$\CG(\pz)$. The
1455: termination of handle reduction guarantees
1456: that, among these words, at least one is
1457: $\s$-positive, so $(iii)$ follows.
1458: \end{proof}
1459: 
1460: A direct application of Proposition~\ref{P:Handle} is the
1461: existence of $\s$-positive quotient-sequences drawn in
1462: the Cayley graph. The definition is as follows:
1463: 
1464: \begin{defi} \label{D:Witness}
1465: Assume that $\pz$ is a positive braid and
1466: $\XX$ is a subset of~$\Div(\pz)$. Let
1467: $\bx_0 \sm \dots \sm \bx_\NN$ be the
1468: increasing enumeration of~$\XX$. We say that
1469: a sequence of words $\vec\ww = (\ww_1, \dots,
1470: \ww_\NNN)$ is a {\it quotient-sequence} for~$\XX$ if,
1471: for each~$\kk$, the word~$\ww_\kk$ is an  expression
1472: of~$\bx_{\kk-1}\inv \bx_\kk$ for each¨$\kk$. We
1473: say that $\vec\ww$ is {\it $\s$-positive} if every
1474: entry in¨$\vec\ww$ is $\s$-positive, and that
1475: $\vec\ww$ is {\it drawn in¨$\CG(\pz)$} (from¨$\bx_0$)
1476: if $\ww_\kk$ is drawn from~$\bx_{\kk-1}$ in~$\CG(\pz)$
1477: for each¨$\kk$.
1478: \end{defi}
1479: 
1480: \begin{coro} \label{C:Witness}
1481: Assume that $\pz$ is a positive braid. Then
1482: every subset of~$\Div(\pz)$ admits a
1483: $\s$-positive quotient-sequence drawn in¨$\CG(\pz)$.
1484: \end{coro}
1485: 
1486: \begin{exam} \label{X:Jump}
1487:  (Figure¨\ref{F:MaxPath})
1488: By computing the successive quotients in the
1489: increasing enumeration of¨$\DDD32$ given in
1490: Example¨\ref{X:List}, we easily find that
1491: $$\tt (a, a, AAb, a, a, AAb, a, AAb, a, a, bAA, a,
1492: bAA, a, a, bAA, a, a)$$
1493: is a $\s$-positive quotient-sequence for $\Div(\DD32)$
1494: drawn in¨$\GG32$. This sequence turns out to be the
1495: unique sequence with the above properties, but this
1496: uniqueness is specific to the case of $3$-braids ({\it
1497: cf.} Figure~\ref{F:Decomp41} below). 
1498: \end{exam}
1499: 
1500: \begin{figure} [htb]
1501: \begin{picture}(60,45)(0, -0)
1502: \put(0,0){\includegraphics{MaxPath32.eps}}
1503: \put(1,26){$1$}
1504: \put(57,18){$\D_3^2$}
1505: \end{picture}
1506: \caption{\smaller The increasing enumeration of the
1507: divisors of¨$\DD32$, and a $\s$-positive
1508: quotient-sequence drawn in¨$\GG32$: the associated path
1509: visits every vertex, and is labeled
1510: $\tt aaAAbaaAAbabAAaaAAbabAAaabAAaa$; it crosses
1511: $6$ $\ss2$-edges (and no $\ss2\inv$)}
1512: \label{F:MaxPath}
1513: \end{figure}
1514: 
1515: We can now easily complete the proof of
1516: Proposition~\ref{P:MainHeight}:
1517: 
1518: \begin{proof}[Proof of
1519: Proposition~\ref{P:MainHeight}]
1520: Let $(\pz_0, \dots, \pz_\NN)$
1521: be the $\sm$-increasing enumeration\break
1522: of $\Div(\pz)$. By Corollary¨\ref{C:Witness},
1523: there exists a $\s$-positive quotient-sequence¨$\vec\ww$
1524: for¨$\Div(\pz)$ that is drawn in¨$\CG(\pz)$. Let $\ww =
1525: \ww_1 \ldots \ww_\NN$. By construction, $\ww$ is a
1526: $\s$-positive word drawn in¨$\CG(\pz)$, and the number of
1527: occurrences of the main generator¨$\ss\mm$ in¨$\ww$ is
1528: (at least) the number of $\mm$-jumps in¨$(\pz_0, \dots,
1529: \pz_\NN)$. So we have $\cc(\pz) \ge \hh\mm(\pz)
1530: -1$. Owing to Lemma¨\ref{L:MainHeight}, this
1531: completes the proof.
1532: \end{proof}
1533: 
1534: \begin{rema}
1535: Assume that $\vec\ww$ is a $\s$-positive
1536: quotient-sequence  for~$\Div(\pz)$, and
1537: $\ss\mm$ is the main generator
1538: occurring in~$\vec\ww$. Then
1539: each word~$\ww_i$ contains zero
1540: or one letter~$\ss\mm$. Indeed, if $\ww_i$
1541: contained two~$\ss\mm$'s or more, then the
1542: vertex reached after the first~$\ss\mm$ ought
1543: to lie in the open $\sm$-interval determined by two
1544: successive entries of~$\vec\pz$, and the
1545: latter is empty by construction since all
1546: elements of~$\Div(\pz)$ occur in~$\vec\pz$.
1547: \end{rema}
1548: 
1549: \section{A decomposition result for~$\DDiv\pz$}
1550: \label{S:Deter}
1551: 
1552: In this section, we establish a structural result
1553: describing $(\DDD\nn\dd, <)$ as the concatenation of
1554: $\ccc\nn\dd+1$ intervals isomorphic to subsets
1555: of¨$(\DDD{\nn-1}\dd, <)$. We deduce an explicit formula
1556: connecting¨$\hhh\rr\nn\dd$ with the number of braids
1557: in¨$\DDD\nn\dd$ whose $\dd$th factor is right divisible
1558: by¨$\D_\rr$, which in turn enables us to complete the
1559: computation of¨$\ccc\nn\dd$ and¨$\hhh\rr\nn\dd$ for
1560: small values of¨$\rr$, $\nn$ and¨$\dd$.
1561: 
1562: \subsection{$B_\rr$-classes}
1563: 
1564: In order to analyse the linearly ordered
1565: sets¨$\DDDD\nn\dd$, and, more generally, 
1566: $(\Div(\pz), \sm\nobreak)$ for¨$\pz$ a positive braid,
1567: we introduce convenient partitions. As $B_\rr$ is a
1568: group for each~$\rr$, it is clear that the relation
1569: $\bx\inv \by \in B_\rr$ defines an equivalence
1570: relation on (positive) braids, so we may put:
1571: 
1572: \begin{defi}
1573: For $\rr \ge 1$ and $\bx, \by$
1574: in¨$B_\infty^+$, we say that $\bx$ and
1575: $\by$ are {\it $B_\rr$-equivalent} if
1576: $\bx\inv \by$ belongs to¨$B_\rr$.
1577: \end{defi}
1578: 
1579: By construction, $B_\rr$-equivalence is
1580: compatible with multiplication on the left.
1581: In the sequel, we consider the restriction
1582: of $B_\rr$-equivalence to finite subsets
1583: of~$B_\infty^+$ of the form~$\Div(\pz)$,
1584: \ie, we use $B_\rr$-equivalence to partition
1585: $\Div(\pz)$ into subsets, naturally called
1586: $B_\rr$-classes.
1587: 
1588: \begin{exam}
1589: As $B_1$ is trivial, $B_1$-equivalence is
1590: equality, and, therefore, the $B_1$-classes are
1591: singletons. On the other hand, any two elements
1592: of¨$B_\nn$ are
1593: $B_\rr$-equivalent for each¨$\rr \ge n$, so, for $\pz$
1594: in¨$B_\nn^+$, there is only one $B_\rr$-class for $\rr \ge
1595: n$, and the only interesting relations arise for
1596: $1 < \rr < n$. For instance, $\DDD3{}$
1597: contains three $B_2$-classes, while $\DDD32$
1598: contains seven of them
1599: (Figure¨\ref{F:Classes}).
1600: \end{exam}
1601: 
1602: \begin{figure} [htb]
1603: \begin{picture}(100,43)(0, -0)
1604: \put(0,0){\includegraphics{Classes.eps}}
1605: \end{picture}
1606: \caption{\smaller The $B_2$-classes
1607: in¨$\DDD3{}$ and¨$\DDD32$}
1608: \label{F:Classes}
1609: \end{figure}
1610: 
1611: Saying that there is an $\rr$-jump
1612: between two braids¨$\bx$ and¨$\by$ means that
1613: $\bx\inv \by$ is $\s$-positive and does not
1614: belong to¨$B_\rr$, so, for $\bx \sm \by$, we
1615: have the equivalence 
1616: \begin{equation} \label{E:Jump}
1617: \big(\mbox{$\bx, \by$ are not
1618: $B_\rr$-equivalent}\big)
1619: \Longleftrightarrow
1620: \left(
1621: \begin{matrix}
1622: \mbox{there is a $\rr$-jump between} \\
1623: \mbox{between¨$\bx$ and¨$\by$}
1624: \end{matrix}
1625: \right).
1626: \end{equation} 
1627: 
1628: \begin{lemm} \label{L:Interval}
1629: Assume that $\pz$ is a positive braid. Then,
1630: each $B_\rr$-class in¨$\Div(\pz)$ is an
1631: interval for¨$\sm$ and there is an $\rr$-jump
1632: between each $B_\rr$-class and the next one.
1633: \end{lemm}
1634: 
1635: \begin{proof}
1636: Assume $\bx \sm \by \in \Div(\pz)$. By
1637: \eqref{E:Jump}, if $\bx$ and $\by$
1638: are not $B_\rr$-equivalent, there is an
1639: $\rr$-jump between¨$\bx$ and¨$\by$, hence
1640: between¨$\bx$ and any element of¨$\Div(\pz)$
1641: above¨$\by$, so no such element may be
1642: $B_\rr$-equivalent to¨$\bx$. This implies
1643: that each $B_\rr$-class is an
1644: $\sm$-interval.
1645: \end{proof}
1646: 
1647: \begin{coro} \label{C:NbClasses}
1648: For each $\rr \ge 1$, 
1649: the number of $B_\rr$-classes in¨$\Div(\pz)$
1650: is¨$\hh\rr(\pz)$.
1651: \end{coro}
1652: 
1653: \begin{proof}
1654: By¨\eqref{E:Jump}, there is no
1655: $\rr$-jump between two elements of the same
1656: $B_\rr$-class, and there is one between two
1657: elements not in the same $B_\rr$-class. Thus
1658: the number of $B_\rr$-classes is the number
1659: of $\rr$-jumps in the $\sm$-increasing
1660: enumeration of¨$\Div(\pz)$ augmented by¨$1$,
1661: hence, by Lemma~\ref{L:Enum}, it is¨$\hh\rr(\pz)$.
1662: \end{proof}
1663: 
1664: $B_\rr$-equivalence provides a partition
1665: of¨$(\Div(\pz), \sm)$ into finitely many
1666: subintervals. The interest of this partition
1667: is that we can describe $B_\rr$-classes rather
1668: precisely and, typically, connect them with
1669: subsets of¨$B_\rr$. In particular, this will
1670: allow for connecting the ordered
1671: sets¨$\DDDD\nn\dd$ with the
1672: sets¨$(\DDD{\nn-1}\dd, \sm)$.
1673: 
1674: \begin{prop} \label{P:Lattice}
1675: (Figure¨\ref{F:Decomp})
1676: Assume $\pz \in B_\infty^+$ and $\rr \ge 1$. Let
1677: $\Class$ be a
1678: $B_\rr$-class in¨$\Div(\pz)$, and let
1679: $\ea, \eb$ be its $\sm$-extremal elements. Then $\ea$
1680: divides every element of¨$\Class$ on the left, and the left
1681: translation by¨$\ea$ defines an isomorphism of
1682: $(\Div(\ea\inv\eb), \dive, \sm)$ onto¨$(\Class, \dive,
1683: \sm)$. In particular, $(\Class, \dive)$ is a
1684: lattice.
1685: \end{prop}
1686: 
1687: \begin{proof}
1688: By Lemma¨\ref{L:Interval}, $\Class$ is the
1689: $\sm$-interval determined by¨$\ea$ and¨$\eb$,
1690: \ie, we have $$\Class = \{\bx
1691: \in \Div(\pz) ; \ea \sm \bx \sm \eb\}.$$
1692: 
1693: We know that $\Div(\pz)$ is a
1694: lattice with respect to left divisibility:
1695: any two elements¨$\bx, \by$ of¨$\Div(\pz)$
1696: admit a greatest left common divisor, here
1697: denoted $\gcd(\bx, \by)$, and a least common
1698: right multiple, denoted¨$\lcm(\bx, \by)$.
1699: Firstly, we claim that $\Class$ is a lattice with
1700: respect to left divisibility, \ie, the left gcd 
1701: and the right lcm of two elements of¨$\Class$
1702: lie in¨$\Class$. So assume $\bx, \by \in \Class$. Let $\bx_0,
1703: \by_0$ be defined by $\bx = \gcd(\bx, \by)
1704: \bx_0$ and $\by = \gcd(\bx,
1705: \by) \by_0$.  The hypothesis that $\bx\inv\by$ belongs
1706: to¨$B_\rr$ implies that there exist $\bx_1, \by_1$ in¨$B_\rr^+$
1707: satisfying $\bx\inv \by = \bx_1\inv \by_1$. By definition of
1708: the gcd, there must exist a positive
1709: braid¨$\pz_1$ satisfying $\bx_1 = \pz_1 \bx_0$
1710: and $\by_1 = \pz_1 \by_0$. Because
1711: $\pz_1$ is positive, $\bx_1 \in B_\rr^+$
1712: implies $\bx_0 \in B_\rr^+$, hence
1713: $\gcd(\bx, \by) \in \Class$. As for the lcm,
1714: the conjunction of $\bx = \gcd(\bx,
1715: \by)
1716: \bx_0$ and $\by = \gcd(\bx, \by) \by_0$
1717: implies $$\lcm(\bx, \by) = \gcd(\bx,
1718: \by)\lcm(\bx_0, \by_0).$$ As $\bx_0, \by_0
1719: \in B_\rr^+$ implies $\lcm(\bx_0, \by_0) \in
1720: B_\rr^+$, we deduce $\lcm(\bx, \by) \in \Class$.
1721: 
1722: As $\Class$ is finite, it follows that $\Class$
1723: admits a global gcd. Because the linear
1724: ordering¨$\sme$ extends the partial
1725: divisibility ordering¨$\dive$, this global
1726: gcd must be the $\sm$-minimum¨$\ea$ of¨$\Class$.
1727: Symmetrically, $\Class$ admits a global lcm,
1728: which must be the $\sm$-maximum¨$\eb$. So, at
1729: this point, we know that $\ea$ is a left
1730: divisor of every element in¨$\Class$, and $\eb$
1731: is a right multiple of each such element,
1732: \ie, we have 
1733: \begin{equation} \label{E:Inclusion}
1734: \Class \ince \{\bx \in B_\infty^+ ; \ea \dive
1735: \bx \dive \eb\}.
1736: \end{equation}
1737: Moreover, $\ea \dive \bx \dive \eb$
1738: implies  $\ea \sme \bx \sme \eb$, hence $\bx
1739: \in \Class$, so the inclusion
1740: in¨\eqref{E:Inclusion} is an
1741: equality.
1742: 
1743: Now, put $F(\bx) = \ea\bx$ for¨$\bx$
1744: in¨$\Div(\ea\inv \eb)$. As $B_\infty^+$ is
1745: left cancellative, $F$ is injective.
1746: Moreover, for $\bx$ a positive braid, $\bx
1747: \dive
1748: \ea\inv
1749: \eb$ is equivalent to $\ea\bx \dive \eb$, so
1750: the image of¨$F$ is $\{\bx \in B_\infty^+ ;
1751: \ea \dive
1752: \bx \dive \eb\}$, hence is¨$\Class$. Finally, by
1753: construction, $F$ preserves both¨$\dive$
1754: and¨$\sm$.
1755: \end{proof}
1756: 
1757: \begin{figure} [htb]
1758: \begin{picture}(86,44)(0, 0)
1759: \put(2,0){\includegraphics{Decomp.eps}}
1760: \put(25,32){$C$}
1761: \put(25,7){$a$}
1762: \put(39,29){$b$}
1763: \put(0,0){$1$}
1764: \put(21,19){$\ss\rr$}
1765: \put(41,23){$\ss\rr$}
1766: \put(62,27){$\ss\rr$}
1767: \put(86,40){$\pz$}
1768: \end{picture}
1769: \caption{\smaller Decomposition of $(\Div(\pz),
1770: \sm)$ into $B_\rr$-classes: each class¨$C$ is a
1771: lattice with respect to divisibility; the increasing
1772: enumeration of¨$\Div(\pz)$ exhausts the first class,
1773: then jumps to the next one by an $\rr$-jump, \etc\ The
1774: number of classes is¨$\hh\rr(\pz)$.}
1775: \label{F:Decomp}
1776: \end{figure}
1777: 
1778: For¨$\rr = 1$, each $B_\rr$-class is a singleton, and
1779: Proposition¨\ref{P:Lattice} says nothing; similarly, if
1780: the main generator of¨$\pz$ is¨$\ss\mm$,
1781: there is only one $B_\rr$-class for $\rr >
1782: \mm$, and we gain no information. But, for
1783: $1 < \rr \le \mm$, and specially for $\rr =
1784: \mm$, Proposition¨\ref{P:Lattice} states that
1785: the chain¨$\Div(\pz)$ is obtained by
1786: concatenating $\hh\rr(\pz)$ copies of sets of
1787: the form¨$\Div(\pz')$ with¨$\pz'$ of index at
1788: most¨$\rr$. In particular, for $\pz = \DD
1789: n\dd$, we have
1790: 
1791: \begin{coro} \label{C:Structure}
1792: For each¨$\nn$ and each¨$\rr$ with $\rr < \nn$, the
1793: chain $\DDDD\nn\dd$ is obtained by  concatenating
1794: $\hhh\rr\nn\dd$ in\-tervals, each of which,
1795: when equipped with¨$\dive$, is a translated
1796: copy of some initial sublattice
1797: of¨$(\DDD\rr\dd, \dive)$.
1798: \end{coro}
1799: 
1800: The case of¨$\DD32$ and¨$\DD4{}$ are illustrated in
1801: Figures¨\ref{F:Decomp32} and¨\ref{F:Decomp41}.
1802: 
1803: \begin{figure} [htb]
1804: \begin{picture}(64,43)(0, 0)
1805: \put(2,0){\includegraphics{Decomp32.eps}}
1806: \put(1,25){$1$}
1807: \put(63,20){$\DD32$}
1808: \end{picture}
1809: \caption{\smaller Decomposition of
1810: $\DDDD32$ into $B_2$-classes: the increasing
1811: enumeration of¨$\DDDD32$ is the concatenation
1812: of the increasing enumeration of the successive
1813: classes, separated by $2$-jumps (compare with
1814: Figure¨\ref{F:MaxPath}); in this case, $B_2$-classes
1815: are simply chains with respect to divisibility}
1816: \label{F:Decomp32}
1817: \end{figure}
1818: 
1819: \begin{figure} [htb]
1820: \begin{picture}(104,48)(0, 0)
1821: \put(2,0){\includegraphics{Decomp41.eps}}
1822: \put(7,0){$1$}
1823: \put(97,46){$\DD4{}$}
1824: \end{picture}
1825: \caption{\smaller Decomposition of
1826: $\DDDD4{}$ into $B_3$-classes; note that the
1827: $\ss3$-arrows (thick) corresponding to 3-jumps
1828: are not unique; in this case, all $B_3$-classes are
1829: isomorphic to the lattice $(\DDD3{}, \sm, \dive)$, \ie,
1830: to the Cayley graph of¨$\D_3$}
1831: \label{F:Decomp41}
1832: \end{figure}
1833: 
1834: 
1835: \subsection{Extremal elements}
1836: 
1837: The next step is to observe that extremal
1838: points in $B_\rr$-classes admit a simple
1839: characterization in terms of divisibility.
1840: 
1841: \begin{prop} \label{P:Extremal}
1842: Assume that $\pz$ is a positive braid. 
1843: 
1844: $(i)$ An element¨$\bx$ of¨$\Div(\pz)$ is the maximum of its
1845: $B_\rr$-class if and only if the relation $\bx\ss i
1846: \dive\pz$ fails for $1 \le i < \rr$.
1847: 
1848: $(ii)$ An element¨$\bx$ of¨$\Div(\pz)$ is the minimum of its
1849: $B_\rr$-class if and only if no¨$\ss i$ with $1 \le i <
1850: \rr$ divides¨$\bx$ on the right.
1851: \end{prop}
1852: 
1853: \begin{proof}
1854: $(i)$ The condition is necessary: if $\bx \ss i$ lies
1855: in¨$\Div(\pz)$ for some¨$i$ with $i < \rr$, then $\bx \ss
1856: i$ lies in the same $B_\rr$-class as¨$\bx$, and it is larger
1857: both for¨$\dive$ and¨$\sm$, so $\bx$ cannot be maximal in
1858: its $B_\rr$-class. Conversely, assume that $\bx$ is
1859: not maximal in its $B_\rr$-class. Then there
1860: exists¨$\by$ satisfying $\bx \sm \by$ and $\by$ is
1861: $B_\rr$-equivalent to¨$\bx$. Now, by 
1862: Proposition¨\ref{P:Lattice},
1863: the lcm of¨$\bx$ and¨$\by$ is also $B_\rr$-equivalent
1864: to¨$\bx$, which means that there exists¨$\by_1$ in¨$B_\rr^+$
1865: satisfying $\lcm(\bx, \by) = \bx\by_1$. Now $\bx \sm \by$
1866: implies $\by_1 \not= 1$, so there must exist¨$i < m$ such
1867: that $\ss i$ is a left divisor of¨$\by_1$. Then we
1868: have $\bx\ss i \dive \bx\by_1 \dive \pz$, hence $\bx
1869: \ss i \dive \pz$.
1870: 
1871: $(ii)$ The argument is symmetric. If we have $\bx = 
1872: \by \ss i$ for some positive braid¨$\by$ and $i < \rr$, then
1873: $\by$ belongs to the $B_\rr$-class of¨$\bx$, and $\bx$
1874: cannot be minimal in its $B_\rr$-class. Conversely, assume
1875: that $\bx$ is not minimal in its $B_\rr$-class. Then there
1876: exists¨$\by$ satisfying $\by \sm \bx$ and $\by$ is
1877: $B_\rr$-equivalent to¨$\bx$. By
1878: Proposition¨\ref{P:Lattice} again, the gcd
1879: of¨$\bx$ and¨$\by$ is also $B_\rr$-equivalent
1880: to¨$\bx$, which means that there
1881: exists¨$\by_0$ in¨$B_\rr^+$ satisfying
1882: $\gcd(\bx, \by) \by_0 = \bx$. As $\by \sm
1883: \bx$ implies $\by_0 \not= 1$, there must exist¨$i < m$
1884: such that $\ss i$ is a right divisor of¨$\by_0$, hence
1885: of¨$\bx$.
1886: \end{proof}
1887: 
1888: When we apply the previous criterion to the
1889: braids¨$\DD\nn\dd$, we obtain:
1890: 
1891: \begin{prop} \label{P:MaxClass}
1892: For $\bx$ in¨$\DDD\nn\dd$ and $1 \le
1893: \rr \le \nn$, the
1894: following are equivalent:
1895: 
1896: $(i)$ The element¨$\bx$ is $\sm$-maximal in its
1897: $B_\rr$-class;
1898: 
1899: $(ii)$ The element¨$\bx \ss i$ belongs to¨$\Div(\DD\nn\dd)$
1900: for no¨$i < \rr$;
1901: 
1902: $(iii)$ The $\dd$th factor of¨$\bx$ is
1903: right divisible by¨$\D_\rr$.
1904: 
1905: $(iv)$ The $\dd+1$st factor of¨$\bx \D_\rr$
1906: is¨$\D_\rr$.
1907: \end{prop}
1908: 
1909: \begin{proof}
1910: The equivalence of¨$(i)$ and¨$(ii)$ is given
1911: by Proposition¨\ref{P:Extremal}$(i)$. It
1912: remains to establish the equivalence
1913: of¨$(ii)$--$(iv)$. For $\rr = 1$, $(ii)$ is
1914: vacuously true, while $(iii)$ and
1915: $(iv)$ always hold. So the expected equivalences are true.
1916: We henceforth assume $\rr \ge 2$.
1917: 
1918: Let $\bx$ belong to¨$\DDD\nn\dd$, and let $\sx_\dd$ be the $\dd$th
1919: factor in the normal form of¨$\bx$. For $i < n$, saying that
1920: $\bx \ss i$ does not belong to¨$\Div(\DD\nn\dd)$ means that
1921: the normal form of¨$\bx \ss i$ has length $\dd+1$, hence,
1922: equivalently, that the normal form of¨$\sx_\dd \ss i$ has
1923: length¨$2$. This occurs if and only if
1924: $\ss i$ is a right divisor of¨$\sx_\dd$. So, for
1925: $\rr \le n$,
1926: $(ii)$ is equivalent to $\sx_\dd$ being right divisible by
1927: all¨$\ss i$'s with $1 \le i < \rr$, hence to
1928: $\sx_\dd$ being right divisible by the (left)
1929: lcm of these elements, which is¨$\D_\rr$.
1930: 
1931: Finally, $(iii)$ and $(iv)$ are equivalent. Indeed, if
1932: the $\dd$th factor¨$\sx_\dd$ in the normal form of¨$\bx$ is
1933: divisible by¨$\D_\rr$ on the right, then $(\sx_\dd, \D_\rr)$ is a
1934: normal sequence as no $\ss i$ with $i < \rr$
1935: from¨$\D_\rr$ may pass to¨$\sx_\dd$. Hence
1936: $(\sx_1, \ldots, \sx_\dd, \D_\rr)$ is a normal
1937: sequence, necessarily the normal form
1938: of¨$\bx\D_\rr$.  Conversely, assume that the
1939: normal form of¨$\bx\D_\rr$ is
1940: $(\sx_1, \ldots, \sx_\dd, \D_\rr)$. The hypothesis that
1941: $(\sx_\dd, \D_\rr)$ is normal implies that $\sx_\dd$ is divisible
1942: on the right by each¨$\ss i$ with $i < \rr$,
1943: hence is divisible on the right by¨$\D_\rr$.
1944: Now $(\sx_1, \ldots,
1945: \sx_\dd)$ is the normal form of¨$\bx$.
1946: \end{proof}
1947: 
1948: Observe that, for $\rr \ge 2$, an element
1949: of¨$\DDD\nn\dd$ that is $\sm$-maximal in its
1950: $B_\rr$-class cannot
1951: belong to¨$\DDD\nn{\dd-1}$, \ie, cannot have
1952: degree $\dd-1$ or less, since the
1953: $\dd$th factor of its normal form cannot be¨$1$.
1954: 
1955: Similar conditions characterize the minimal elements of the
1956: $B_\rr$-classes. Because the normal form has a priviledged
1957: orientation, the results are not entirely symmetric of
1958: those of Proposition¨\ref{P:MaxClass}
1959: 
1960: \begin{prop} \label{P:MinClass}
1961: For $\bx$ in¨$\DDD\nn\dd$ and $1 \le
1962: \rr \le \nn$, the following are equivalent:
1963: 
1964: $(i)$ The element¨$\bx$ is $\sm$-minimal of its
1965: $B_\rr$-class;
1966: 
1967: $(ii)$ No¨$\ss i$ with $i < \rr$ is a right divisor
1968: of¨$\bx$;
1969: 
1970: $(iii)$ The degrees of¨$\bx$ and¨$\bx \D_\rr$ are equal.
1971: \end{prop}
1972: 
1973: \begin{proof}
1974: The equivalence of¨$(i)$ and¨$(ii)$ is given by
1975: Proposition¨\ref{P:Extremal}$(ii)$. On the other hand,
1976: everything is obvious for¨$\rr = 1$. So it remains to
1977: establish the equivalence of¨$(ii)$ and¨$(iii)$ in the case
1978: $\rr \ge 2$. Now, assume that¨$(ii)$ holds an $\bx$ has
1979: degree¨$\dd$. The hypothesis that $\ss i$ is not a right
1980: divisor of¨$\bx$ implies that $\bx \ss i$ is a divisor
1981: of¨$\DD\nn\dd$. As this holds for each $i <
1982: \rr$, the lcm of $\bx \ss1$, \dots, $\bx \ss{\rr-1}$, which
1983: is $\bx \D_\rr$, also divides¨$\DD\nn\dd$, which means that
1984: $\bx \D_\rr$ has degree (at most)¨$\dd$. So $(ii)$
1985: implies¨$¨(iii)$.
1986: 
1987: Conversely, assume that $\ss i$ divides¨$\bx$ on the right.
1988: Then the degree of¨$\bx \ss i$ is strictly larger than that
1989: of¨$\bx$, and, {\it a fortiori}, the same is true for¨$\bx
1990: \D_\rr$.
1991: \end{proof}
1992: 
1993: \subsection{Determination of¨$\hhh\rr\nn\dd$}
1994: 
1995: A direct application of the previous results is a formula
1996: connecting the number of $B_\rr$-classes
1997: in¨$\Div(\DD\nn\dd)$, \ie, the
1998: numbers¨$\hhh\rr\nn\dd$, with the number of
1999: braids whose normal form ends with some
2000: specific factor.
2001: 
2002: \begin{defi}
2003: For $\nn, \dd \ge 1$ and for¨$\sp$ a simple
2004: $\nn$-braid, we denote by¨$\bbb\nn\dd\sp$ the number
2005: of positive braids of degree at most¨$\dd$, \ie, of
2006: divisors of¨$\DD\nn\dd$, whose $\dd$th factor is¨$\sp$.
2007: \end{defi}
2008: 
2009: \begin{prop} \label{P:Main}
2010: For $1 \le \rr \le \nn$, we have
2011: \begin{equation} \label{E:Main}
2012: \hhh\rr\nn\dd = \sum_{\mbox{\Small $s$ right divisible by
2013: $\D_\rr$}} \bbb\nn\dd s  = \bbb \nn{\dd+1}{\D_\rr} .
2014: \end{equation} 
2015: \end{prop}
2016: 
2017: In words: The number of $\rr$-jumps in $\DDDD\nn\dd$ is
2018: the number of
2019: $\nn$-braids of degree at most¨$\dd$ whose $\dd$th factor
2020: is right divisible by¨$\D_\rr$.
2021: 
2022: \begin{proof}
2023: By Corollary¨\ref{C:NbClasses},
2024: $\hhh\rr\nn\dd$ is the number of
2025: $B_\rr$-classes in¨$\Div(\DD\nn\dd)$. Each
2026: class contains exactly one maximum element, and, by
2027: Proposition¨\ref{P:MaxClass}, the latter are
2028: characterized by the property that their
2029: $\dd$th factor is right
2030: divisible by¨$\D_\rr$. The first equality
2031: in¨\eqref{E:Main} follows. The second one
2032: follows from the equivalence of¨$(iii)$
2033: and¨$(iv)$ in Proposition¨\ref{P:MaxClass}.
2034: \end{proof}
2035: 
2036: For $\rr = 1$, as every simple braid is divisible by¨$1$ on
2037: the right, Relation¨\eqref{E:Main} reduces to
2038: \begin{equation} \label{E:MainOne}
2039: \hhh1\nn\dd = \sum_{\sp} \bbb\nn\dd\sp = \bbb
2040: \nn{\dd+1}{1},
2041: \end{equation}
2042: a special case of the relation $\hh1(\pz) =
2043: \card\Div(\pz)$ of
2044: Proposition¨\ref{P:Decreasing}. 
2045: For $\rr = \nn$, as the only normal sequence of
2046: length¨$\dd$  that finishes with¨$\D_\nn$ is  $(\D_\nn,
2047: \dots, \D_\nn)$, Relation¨\eqref{E:Main}
2048: reduces to
2049: \begin{equation} \label{E:Mainn}
2050: \hhh\nn\nn\dd = 1,
2051: \end{equation}
2052: also noted in Proposition¨\ref{P:Decreasing}.
2053: Finally, for $\rr = \nn-1$, we obtain using
2054: Proposition¨\ref{P:MainHeight}:
2055: 
2056: \begin{coro} \label{C:Main}
2057: For $\nn \ge 2$, we have
2058: \begin{equation} \label{E:MainNO}
2059: \ccc\nn\dd = \hhh{\nn - 1}\nn\dd -1 = \sum_{i =
2060: 2}^{\nn}
2061: \bbb\nn\dd{\ss i \ss{i+1} \dots \ss{\nn-1}\D_{\nn - 1}} =
2062: \bbb\nn{\dd+1}{\D_{\nn - 1}} - 1.
2063: \end{equation}
2064: \end{coro}
2065: 
2066: \begin{proof}
2067: The simple $\nn$-braids that are right divisible
2068: by¨$\D_{\nn-1}$ are the braids of the form
2069: $\ss i \ss{i+1} \dots \ss{\nn-1}$ with $1 \le i\le
2070: \nn$. Indeed,
2071: it is clear that every such braid is simple
2072: and right divisible by¨$\D_{\nn-1}$.
2073: Conversely, the only possibility for $\pz
2074: \D_{\nn-1}$ to be simple is that $\pz$ moves the $\nn$th
2075: strand to some position between¨$1$ and¨$\nn$, but
2076: introduces no crossing between the remaining strands.
2077: Finally, $\ss1 \ss2 \dots \ss{\nn-1}\D_{\nn-1}$
2078: is¨$\D_\nn$, and we already observed that
2079: $\bbb\nn\dd{\D_\nn}$ is¨$1$, so we obtain the first
2080: equality in¨\eqref{E:MainNO}.
2081: \end{proof}
2082: 
2083: \subsection{Computation of¨$\bbb\nn\dd\sp$}
2084: 
2085: By Lemma¨\ref{L:Normal}, normal sequences are
2086: characterized by a local condition involving only pairs
2087: of consecutive elements. It follows that the set of all
2088: normal sequences is a rational set, \ie, it can be
2089: recognized by a finite state automaton. Standard
2090: arguments then show that the numbers¨$\bbb\nn\dd\sp$ obey
2091: a linear recurrence. Building on this observation,
2092: seemingly first used in the case of braids in¨\cite{Chb},
2093: we can obtain explicit formulas for the
2094: parameters¨$\ccc\nn\dd$ and¨$\hhh\rr\nn\dd$ for small
2095: values of¨$\rr$, $\nn$, and/or¨$\dd$. We shall not go
2096: into details in the current paper, and refer
2097: to¨\cite{Dhi} where all formulas are established---and
2098: where, more generally, the rich combinatorics underlying
2099: the normal form of braids is investigated.
2100: 
2101: In the sequel, we write $(M)_{\xx, \yy}$ for the $(\xx,
2102: \yy)$-entry of a matrix¨$M$. The general principle for
2103: computing the numbers¨$\bbb\nn\dd\sp$ for some
2104: fixed¨$\nn$ is as follows:
2105: 
2106: \begin{lemm} \label{L:Comput}
2107: For $\nn \ge 1$, let $M_\nn$ be the square matrix with
2108: entries indexed by simple $\nn$-braids defined by
2109: $$
2110: (M_\nn)_{\sp, \spp} = 
2111: \begin{cases}
2112: 1 & \mbox{if $(\sp, \spp)$ is normal,}\\
2113: 0 & \mbox{otherwise.}
2114: \end{cases}$$
2115: Then, for every simple¨$\spp$ and every
2116: $\dd \ge 1$, we have $\bbb\nn\dd{\spp} = ((1, 1, \dots, 1)
2117: \, M_\nn\!\!\!{}^{\dd-1})_\spp$.
2118: \end{lemm}
2119: 
2120: The proof is an easy induction on¨$\dd$.
2121: 
2122: \begin{exam}
2123: The matrix¨$M_1$ is
2124: $(1)$, corresponding to
2125: $\bbb1\dd1 = 1$. For $\nn=2$, using the enumeration $(1,
2126: \ss1)$ of simple $2$-braids, we find $M_2 = 
2127: \left(\begin{matrix}1&0\\1&1\end{matrix}\right)$,
2128: leading to
2129: $\bbb2\dd1 = \dd$, $\bbb2\dd{\ss1} = 1$, as could
2130: be expected: there are $\dd+1$ braids of degree
2131: at most¨$\dd$, namely the braids¨$\ss1^\expo$ with
2132: $\expo < \dd$, whose $\dd$th factor is¨$1$, and
2133: $\ss1^\dd$, whose $\dd$th factor is¨$\D_2$, \ie,
2134: ¨$\ss1$. For $\nn = 3$, using the enumeration $(1, \ss1,
2135: \ss2,\ss2\ss1, \ss1\ss2, \D_3)$ of simple $3$-braids,  we
2136: obtain
2137: $$M_3 = \left(
2138: \begin{matrix}
2139: 1&0&0&0&0&0\\
2140: 1&1&0&0&1&0\\
2141: 1&0&1&1&0&0\\
2142: 1&1&0&0&1&0\\
2143: 1&0&1&1&0&0\\
2144: 1&1&1&1&1&1
2145: \end{matrix}
2146: \right),$$
2147: from which we can deduce for instance $\bbb331= 19$ or
2148: $\bbb34{\ss1} = 15$ using Lemma¨\ref{L:Comput}.
2149: \end{exam}
2150: 
2151: Using Proposition¨\ref{P:Main}, we deduce
2152: 
2153: \begin{prop} \label{P:Values}
2154: With $M_\nn$ as in Lemma¨\ref{L:Comput}, we have
2155: for $\nn \ge \rr \ge 1$ and $\dd \ge 1$
2156: \begin{gather*}
2157: \ccc\nn\dd = ((1, 1, \dots, 1) \, M_\nn^\dd)_{\D_{\nn-1}}
2158: - 1,\\
2159: \hhh\rr\nn\dd = ((1, 1, \dots, 1) \,
2160: M_\nn^\dd)_{\D_{\rr}}.
2161: \end{gather*}
2162: \end{prop}
2163: 
2164: \begin{coro}
2165: $(i)$ For fixed¨$\nn, \rr$, the generating functions for
2166: the sequences $\ccc\nn\dd$ and $\hhh\rr\nn\dd$ are
2167: rational.
2168: 
2169: $(ii)$ For fixed¨$\nn, \rr$, the numbers $\ccc\nn\dd$ and
2170: $\hhh\rr\nn\dd$ admit expressions of the form
2171: \begin{equation} \label{E:General}
2172: P_1(\dd) \ev_1^\dd + \cdots +
2173: P_\ir(\dd) \ev_\ir^\dd.
2174: \end{equation} 
2175: where $\ev_1$, \dots, $\ev_\ir$ are the non-zero
2176: eigenvalues of¨$M_\nn$ and $P_1, \dots,
2177: P_\ir$ are polynomials with $\deg(P_\ind)$ at
2178: most the multiplicity of¨$\ev_\ind$ for¨$M_\nn$.
2179: \end{coro}
2180: 
2181: As the matrix¨$M_\nn$ is an $\nn! \times \nn!$ matrix,
2182: completing the computation is not so easy, even for small
2183: values of¨$\nn$. Actually, it is shown in¨\cite{Dhi} how
2184: to replace¨$M_\nn$ with a smaller matrix¨$\overline
2185: M_\nn$ of size $p(\nn) \times p(\nn)$, where $p(\nn)$ is
2186: the number of partitions of¨$\nn$. The property is
2187: connected with classical results by Solomon about the
2188: descents of permutations¨\cite{Sol}. With such methods,
2189: one easily obtains the values listed in
2190: Table¨\ref{T:Values}.
2191: 
2192: \begin{table}[htb]
2193: $$\begin{tabular}{c|r|r|r|r|r|r|r}
2194: \quad$\dd$
2195: & 0 
2196: &1&2&3&4&5&6\\
2197: \hline
2198: \vrule width0pt height12pt depth6pt
2199: $\hhh12\dd$
2200: & 1
2201: & 2 & 3 & 4 & 5 & 6 & 7 \\
2202: \hline
2203: \vrule width0pt height12pt 
2204: $\hhh13\dd$
2205: & 1
2206: & 6 & 19 & 48 & 109 & 234 & 487\\
2207: \vrule width0pt depth6pt
2208: $\hhh23\dd$
2209: & 1
2210: & 3 & 7 & 15 & 31 & 63 & 127\\
2211: \hline
2212: \vrule width0pt height12pt 
2213: $\hhh14\dd$
2214: & 1
2215: &24&211&1\,380&8\,077&45\,252&249\,223
2216: \\
2217: $\hhh24\dd$
2218: & 1
2219: &12&83&492&2\,765&15\,240
2220: & 83\,399 \\
2221: \vrule width0pt depth6pt
2222: $\hhh34\dd$
2223: & 1
2224: &4&15&64&309&1\,600 
2225: & 8\,547
2226: \\
2227: \hline
2228: \vrule width0pt height12pt 
2229: $\hhh15\dd$
2230: & 1
2231:  & 120 & 3\,651 & 79\,140 
2232: & 1\,548\,701 
2233: & 29\,375\,460 
2234: & 551\,997\,751 
2235: \\
2236: $\hhh25\dd$
2237: & 1
2238:  & 60 & 1\,501 & 30\,540 
2239: & 585\,811
2240: & 11\,044\,080 
2241: & 207\,154\,921 
2242: \\
2243: $\hhh35\dd$
2244: & 1
2245:  & 20 & 311 & 5\,260 
2246: & 94\,881
2247: & 1\,755\,360 
2248: & 32\,741\,851 
2249: \\
2250: \vrule width0pt depth6pt
2251: $\hhh45\dd$
2252: & 1
2253:  & 5 & 31 & 325 
2254: & 4\,931 
2255: & 86\,565 
2256: & 1\,590\,231 
2257: \\
2258: \hline
2259: \vrule width0pt height12pt 
2260: $\hhh16\dd$
2261: & 1
2262: & 720 
2263: &\!90\,921 
2264: &\!7\,952\,040 
2265: &\!634\,472\,921
2266: &\!49\,477\,263\,360
2267: &\!3\,836\,712\,177\,121
2268: \\
2269: $\hhh26\dd$
2270: & 1
2271: & 360 
2272: &\!38\,559 
2273: &\!3\,228\,300 
2274: &\!254\,718\,389
2275: &\!19\,808\,530\,620
2276: &\!1\,535\,016\,069\,499 \\
2277: $\hhh36\dd$
2278: & 1
2279:  & 120 & 8\,727 & 649\,260 
2280: & 49\,654\,757
2281: & 3\,831\,626\,580 
2282: & 296\,361\,570\,667\\
2283: $\hhh46\dd$
2284: & 1
2285:  & 30 & 1\,075 & 61\,620 
2286: & 4\,387\,195
2287: & 332\,578\,230 
2288: & 25\,612\,893\,355\\
2289: $\hhh56\dd$
2290: & 1
2291:  & 6 & 63 & 1\,956
2292: & 116\,423
2293: & 8\,448\,606 
2294: & 643\,888\,543
2295: \end{tabular}$$
2296: \bigskip
2297: \caption{\smaller First values of $\hhh\rr\nn\dd$
2298: for $1 \le \rr < \nn$---the value is¨$1$ for $\rr
2299: \ge \nn$. For instance, we read that the number of
2300: $3$-strand braids of degree at most¨$2$, \ie,
2301: $\hhh132$, is¨$19$---as was seen in
2302: Example¨\ref{X:Height}---while the maximal number
2303: of¨$\ss3$'s in a $\s$-positive word drawn in¨$\GG44$,
2304: \ie, $\ccc44$, which is $\hhh344-1$
2305: according to Proposition¨\ref{P:MainHeight}, is¨$308$.}
2306: \label{T:Values}
2307: \end{table}
2308: 
2309: Using the reduced matrices 
2310: $\overline M_3 = \left(\begin{matrix}1&0&0\\4&2&0\\1&1&1
2311: \end{matrix}\right)$ and 
2312: $\overline M_4 =
2313: \left(\begin{matrix}1&0&0&0&0\\11&4&1&0&0\\5&3&2&1&0\\
2314: 6&4&2&2&0\\1&1&1&1&1
2315: \end{matrix}\right)$, we obtain the
2316: following explicit form for¨\eqref{E:General} involving
2317: the non-zero eigenvalues of¨$M_3$, namely $1$ (double),
2318: $2$ and of¨$M_4$, namely $1$ (double), $2$, and $3 \pm
2319: \sqrt6$:
2320: 
2321: \begin{prop} \label{P:Values34}
2322: Let $\ev_1 = 3 + \sqrt{6}$ and $\ev_2 = 3 - \sqrt{6}$.
2323: Then, for $\dd \ge 1$, we have
2324: \begin{align*}
2325: \hhh13\dd &= 8\cdot 2^\dd - 3 \dd - 7,\\
2326: \hhh23\dd &= \ccc3\dd+1 = 2 \cdot 2^\dd - 1,\\
2327: \hhh14\dd &= 
2328: \frac3{20}(32 + 13\sqrt6) \ev_1^\dd  
2329: + \frac3{20}(32 - 13\sqrt6) \ev_2^\dd 
2330: - \frac{128}5 \cdot\, 2^\dd
2331: +  6\dd + 17,\\
2332: \hhh24\dd &= 
2333: \frac1{20}(32 + 13\sqrt6) \ev_1^\dd 
2334: + \frac1{20}(32 - 13\sqrt6) \ev_2^\dd
2335: - \frac{16}5 \cdot\, 2^\dd 
2336: + 1,\\
2337: \hhh34\dd &= \ccc34 +1 = 
2338: \frac1{20}(4 + \sqrt6) \ev_1^\dd
2339: + \frac1{20}(4 - \sqrt6) \ev_2^\dd 
2340: + \frac85 \cdot\, 2^\dd - 1.
2341: \end{align*}
2342: \end{prop}
2343: 
2344: The main interest of the above formulas are to show that
2345: each of the involved parameters has an exponential growth
2346: with respect to¨$\dd$, in $O(2^\dd)$ for $\nn = 3$, and in
2347: $O((3+\sqrt{6})^\dd)$ for $\nn = 4$. For practical
2348: purposes, it may be more convenient to resort to inductive
2349: formulas, for instance
2350: \begin{gather}
2351: \label{E:Recur3}
2352: \hhh13\dd = 2 \hhh13{\dd-1} + 3 \dd + 1,\\
2353: \label{E:Recur4}
2354: \hhh14\dd = 6\hhh14{\dd-1} - 3\hhh14{\dd-2}
2355: + 32 \cdot 2^\dd - 12 \dd - 34,
2356: \end{gather}
2357: together with initial values $\hhh130 = \hhh140 = 1$,
2358: $\hhh141 = 24$ (or $\hhh14{-1}= 0$).
2359: 
2360: \subsection{Small values of¨$\dd$}
2361: 
2362: Another approach is to keep¨$\dd$ fixed and let¨$\nn$ vary.
2363: Once again, we only mention a few results, and refer the
2364: reader to¨\cite{Dhi} for the proofs and additional
2365: comments. For $\dd = 1$, it is easy to determine all
2366: values:
2367: 
2368: \begin{prop} [\cite{Dhi}]
2369: For $\nn \ge \rr \ge 1$, we have
2370: \begin{equation*} \label{E:K1}
2371: \hhh\rr\nn{} = \frac{\nn!}{\rr!}.
2372: \end{equation*}
2373: \end{prop}
2374: 
2375: For $\dd = 2$, it is easier to complete the
2376: computation for $\nn - \rr$ rather than¨$\rr$ fixed. 
2377: 
2378: \begin{prop} [\cite{Dhi}] \label{P:K2}
2379: For $\nn \ge \rr \ge 1$, we have
2380: \begin{align*} 
2381: \hhh{\nn-\rr}\nn2 
2382: &= \rr!\,(\rr+1)^\nn + \sum_{i = 1}^\rr P_i(\nn) \,
2383: i^{\nn-\rr+i-1},\\
2384: \intertext{for some polynomial $P_i$ of degree at
2385: most¨$\rr - i + 1$. The values for $\rr =
2386: 1, 2$ are}
2387: \hhh{\nn-1}\nn2 &= 2^\nn - 1, \\
2388: \hhh{\nn-2}\nn2 &= 2 \cdot 3^\nn - (\nn + 6) \cdot
2389: 2^{\nn-1} + 1.
2390: \end{align*}
2391: \end{prop}
2392: 
2393: For $\rr$ fixed, no general formula is known. Let us
2394: mention the case of¨$\hhh1\nn2$, which follows from
2395: results of¨\cite{CSV}:
2396: 
2397: \begin{prop} [\cite{Dhi}]
2398: The numbers¨$\hhh1\nn2$ are determined by the
2399: induction
2400: \begin{equation*}
2401: \hhh102=1, \qquad 
2402: \hhh1\nn2 = \sum_{\ind=0}^{\nn-1} (-1)^{\nn + \ind
2403: +1} {\nn \choose \ind}^2 
2404: \hhh1\ind2.
2405: \end{equation*}
2406: Their double exponential generating function is, with 
2407: $J_0(\xx)$ is the Bessel function,
2408: \begin{equation*}
2409: \sum_{\nn = 0}^\infty \hhh1\nn2
2410: \frac{z^\nn}{\nn!^2} = 
2411: \bigg( \sum_{\nn = 0}^\infty (-1)^\nn
2412: \frac{z^\nn}{\nn!^2}
2413: \bigg)\inv = \frac1{J_0(\sqrt{z})}.
2414: \end{equation*}
2415: \end{prop}
2416: 
2417: Finally, for $\dd = 3$, the computation can be completed
2418: at least in the case $\nn - \rr = 1$:  
2419: 
2420: \begin{prop} [\cite{Dhi}]
2421: For¨$\nn \ge 1$, we have, with $e = \exp(1)$,
2422: \begin{equation*} \label{E:KKK3}
2423: \hhh{\nn-1}\nn3 = \sum_{\ind=0}^{\nn-1}
2424: \frac{\nn!}{\ind!} = \lfloor\nn! e\rfloor - 1.
2425: \end{equation*}
2426: \end{prop}
2427: 
2428: Using Proposition¨\ref{P:MainHeight}, we deduce the
2429: following explicit values for $\ccc\nn\dd$, \ie, for the
2430: maximal number of occurrences of¨$\ss{\nn-1}$ is a
2431: $\s$-positive word drawn in the Cayley graph
2432: of¨$\DD\nn\dd$:
2433: $$\ccc\nn{} = \nn - 1, \qquad
2434: \ccc\nn2 = 2^\nn - 2, \qquad
2435: \ccc\nn3 = \sum_{\ind=0}^{\nn-1}
2436: \frac{\nn!}{\ind!} - 1 = \lfloor\nn! e\rfloor - 2.
2437: $$
2438: The formulas listed above show that a number
2439: of different induction schemes appear, suggesting that the
2440: combinatorics of normal sequences of braids is very rich. 
2441: 
2442: \section{A complete description of $\DDDD3\dd$}
2443: \label{S:N3}
2444: 
2445: Our ultimate goal would be a complete description
2446: of each chain $\DDDD\nn\dd$, this typically meaning
2447: that we are able to explicitly specify the increasing
2448: enumeration of its elements. This goal remains out of
2449: reach in the general case, but we shall show now how the
2450: process can be completed in the case $\nn = 3$. The
2451: counting formulas of Section¨\ref{S:Deter} play a key
2452: role in the construction, and, in particular, the Pascal
2453: triangle of Table¨\ref{T:Triangle} below is directly
2454: connected with the $2^\dd$¨factor in the inductive
2455: formulas of Proposition¨\ref{P:Values34}.  As an
2456: application, we deduce a new proof of Property¨C and of
2457: the well-ordering property, hence a complete
2458: re-construction of the braid ordering in the case $\nn
2459: = 3$.
2460: 
2461: The general principle is to make the decomposition of
2462: Corollary¨\ref{C:Structure} explicit. The latter shows
2463: that, for all¨$\nn$ and¨$\dd$, the chain $\DDDD\nn\dd$
2464: can be decomposed into $\ccc\nn\dd$¨subintervals
2465: each of which is a copy of some fragment
2466: of¨$\DDDD{\nn-1}\dd$. Moreover, the approach of
2467: Section¨\ref{S:Deter} suggests an induction on¨$\dd$
2468: as well, so, finally, we are led to looking for a
2469: description of¨$\DDDD\nn\dd$ in terms
2470: of¨$\DDDD{\nn-1}\dd$ and¨$\DDDD\nn{\dd-1}$---\ie, in
2471: the current case, a description of $\DDDD3\dd$ in terms
2472: of
2473: $\DDDD2\dd$ and $\DDDD3{\dd-1}$.
2474: 
2475: \subsection{The braids¨$\Dd\nn\pp$}
2476: 
2477: The subsequent construction will appeal to a double
2478: series of braid called¨$\Dd\nn\pp$, and we begin with a
2479: few preliminary properties of these braids.
2480: 
2481: \begin{defi}
2482: For $\nn \ge 2$, let $\ss{\nn, 1}$ and $\ss{1, \nn}$
2483: respectively denote the braid words
2484: $\ss{\nn-1} \ss{\nn-2} \dots \ss1$ and $\ss1 \ss2
2485: \dots \ss{\nn-1}$. For $\pp \ge 0$, we define 
2486: $\Ddt\nn\pp$ to  be (the braid represented by) the
2487: length¨$\pp$ prefix of the right infinite
2488: word¨$(\ss{\nn, 1}\ss{1, \nn})^\infty$, and $\Dd\nn\pp$ to
2489: be (the braid represented by) the length¨$\pp$ suffix of
2490: the left infinite word¨$^\infty(\ss{\nn, 1}\ss{1, \nn})$. 
2491: \end{defi}
2492: 
2493: For instance, we find
2494: $\Dd30 = 1$, $\Dd31 = \b$, $\Dd32 = \a\b$, \dots,
2495: $\Dd34 = \b\a\a\b$, \dots, $\Dd37 = \a\a\b\b\a\a\b$, \etc\
2496: Similarly, we have $\Dd46 = \c\b\a\a\b\c$, and, more
2497: generally, $\Dd\nn{2\nn-2} = \Ddt\nn{2\nn-2}= \ss{\nn,
2498: 1}\ss{1, \nn}$. Note that, as a word,
2499: $\Dd\nn\pp$ is obtained by reversing the order of the
2500: letters in¨$\Ddt\nn\pp$.
2501: 
2502: \begin{lemm}
2503: For $\nn \ge 2$ and $\pp, \qq \ge 0$ satisfying $\pp
2504: + \qq = \dd(\nn-1)$, we have
2505: \begin{equation} \label{E:Deltadelta}
2506: \Dd\nn\pp \, \DD{\nn-1}\dd \, \Ddt\nn\qq = \DD\nn\dd.
2507: \end{equation}
2508: \end{lemm}
2509: 
2510: \begin{proof}
2511: We first prove using induction on¨$\dd$ the relation
2512: \begin{equation} \label{E:Deltak0}
2513: \Dd\nn{\dd(\nn-1)} \, \DD{\nn-1}\dd = \DD\nn\dd,
2514: \end{equation}
2515: \ie, \eqref{E:Deltadelta} with $\qq =
2516: 0$. For $\dd = 0$, \eqref{E:Deltak0} reduces to $1 =
2517: 1$. Assume $\dd \ge 1$.  By definition,
2518: $\Dd\nn{\dd(\nn-1)}$ is $\ss{\nn,1} \, 
2519: \Dd\nn{(\dd-1)(\nn-1)}$ for
2520: $\dd$¨odd, and is $\ss{1, \nn} \,
2521: \Dd\nn{(\dd-1)(\nn-1)}$ for
2522: $\dd$¨even, so, in any case, we can write
2523: $$\Dd\nn{\dd(\nn-1)} = \flip\nn^{\dd-1}(\ss{1, \nn}) \,
2524: \Dd\nn{(\dd-1)(\nn-1)},$$ 
2525: where we recall $\flip\nn$ denotes the flip
2526: automorphism of¨$B_\nn$ that exchanges¨$\ss i$ and
2527: $\ss{\nn - i}$. Using the induction hypothesis
2528: and¨\eqref{E:Delta}, we find
2529: \begin{align*}
2530: \Dd\nn{\dd(\nn-1)} \, \DD{\nn-1}\dd 
2531: &= \flip\nn^{\dd-1}(\ss{1, \nn}) \, \Dd\nn{(\dd-1)(\nn-1)} \,
2532: \DD{\nn-1}{\dd-1} \, \D_{\nn-1}\\
2533: &= \flip\nn^{\dd-1}(\ss{1, \nn}) \, \DD n{\dd-1} \, \D_{\nn-1}
2534: = \DD n{\dd-1} \, \ss{1, \nn} \, \D_{\nn-1}
2535: = \DD n{\dd-1} \, \D_\nn 
2536: = \DD\nn\dd.
2537: \end{align*} 
2538: 
2539: We return to the general case
2540: of¨\eqref{E:Deltadelta}. For $\dd$ even, we have
2541: $\Dd\nn{\dd(\nn-1)} = \Ddt n{\dd(\nn-1)}$, hence
2542: $\Ddt\nn\qq \, \Dd\nn\pp = \Dd\nn{\dd(\nn-1)}$. If
2543: $\dd$ is odd, we have $\Dd\nn{\dd(\nn-1)} =
2544: \flip\nn(\Ddt n{\dd(\nn-1)})$, which implies
2545: $\flip\nn(\Ddt\nn\qq) \, \Dd\nn\pp = \Dd\nn{\dd(\nn-1)}$. So $\flip\nn^\dd(\Ddt\nn\qq) \, \Dd\nn\pp = \Dd\nn{\dd(\nn-1)}$ holds in both
2546: cases. Now, using¨\eqref{E:Deltak0}, we find
2547: \begin{align*}
2548: \flip\nn(\Ddt\nn\qq) \, \Dd\nn\pp \, \DD{\nn-1}\dd \, \Ddt\nn\qq
2549: &= \Dd\nn{\dd(\nn-1)} \, \DD{\nn-1}\dd \, \Ddt\nn\qq
2550: = \DD\nn\dd \, \Ddt\nn\qq
2551: = \flip\nn(\Ddt\nn\qq) \, \DD\nn\dd,
2552: \end{align*} 
2553: from which we deduce¨\eqref{E:Deltadelta} by cancelling 
2554: $\flip\nn(\Ddt\nn\qq)$ on the left.
2555: \end{proof}
2556: 
2557: \begin{lemm}
2558: For $1 \le i \le \nn-2$ we have
2559: \begin{equation}
2560: \label{E:Commut}
2561: \Dd\nn{\dd(\nn-1)} \, \ss i = \ss{i+e} \,
2562: \Dd\nn{\dd(\nn-1)}
2563: \end{equation}
2564: with $e = 0$ if $\dd$ is even, and $e = 1$ if $\dd$ is odd.
2565: \end{lemm}
2566: 
2567: \begin{proof}
2568: For $1 \le i \le \nn-2$, we have
2569: \begin{equation}
2570: \ss{1,n} \, \ss i = \ss{i+1} \, \ss{1,n},
2571: \mbox{\quad and \quad}
2572: \ss{\nn, 1} \, \ss{i+1} = \ss i \, \ss{\nn, 1}, 
2573: \end{equation}
2574: as an easy induction shows. This implies
2575: $\ss{\nn,1} \, \ss{1,n} \, \ss i = \ss i \,
2576: \ss{\nn,1} \, \ss{1,n}$, and therefore $(\ss{\nn,1}
2577: \, \ss{1,n})^\dd \, \ss i = \ss i \,
2578: (\ss{\nn,1} \, \ss{1,n})^\dd$, \ie, 
2579: $\Dd\nn{2\dd(\nn-1)} \, \ss i = \ss i \,
2580: \Dd\nn{2\dd(\nn-1)}$, for every¨$\dd$. On the other
2581: hand, we have 
2582: $\Dd\nn{(2\dd+1)(\nn-1)} = \ss{1,n} \, \Dd\nn{2\dd(\nn-1)}$, hence
2583: $$\Dd\nn{(2\dd+1)(\nn-1)} \, \ss i 
2584: = \ss{1,n} \, \ss i \, \Dd\nn{2\dd(\nn-1)} 
2585: = \ss{i+1} \, \ss{1,n} \, \Dd\nn{2\dd(\nn-1)} =
2586: \ss{i+1} \, \Dd\nn{(2\dd+1)(\nn-1)},$$
2587: as was expected.
2588: \end{proof}
2589: 
2590: \subsection{A Pascal triangle}
2591: 
2592: We shall now construct for every¨$\dd$ a sequence of
2593: positive braids¨$\SS3\dd$ that will turn out to be
2594: the increasing enumeration of¨$\DDDD3\dd$.
2595: The construction relies on an induction similar to
2596: a Pascal triangle. In order to make it easily
2597: understandable, it is convenient to start with a
2598: construction in the (trivial) cases¨$\nn =
2599: 1$ and¨$\nn = 2$.
2600: 
2601: As $B_1$ is the trivial group, then for every¨$\dd$
2602: there is exactly one element of degree at
2603: most¨$\dd$, namely¨$1$, and we can state:
2604: 
2605: \begin{prop}
2606: Let $\SS1\dd$ be defined for
2607: $\dd \ge 0$ by
2608: \begin{equation} \label{E:Def1}
2609: \SS1\dd = (1).
2610: \end{equation}
2611: Then $\SS1\dd$ is the increasing enumeration
2612: of¨$\DDD1\dd$.
2613: \end{prop}
2614: 
2615: The group¨$B_2$ is the rank¨$1$ free group
2616: generated by¨$\ss1$. The fundamental braid¨$\D_2$
2617: is just¨$\ss1$, and the braids of degree at
2618: most¨$\dd$, \ie, the divisors of¨$\DD2\dd$, consist
2619: of the $\dd+1$ braids¨$1, \ss1, \ldots, \sss1\dd$.
2620: On the other hand, we have $\ss{1, 2} = \ss{2, 1} =
2621: \ss1$, and $\Dd1i = \sss1i$ for every¨$i$. 
2622: 
2623: \begin{nota}
2624: If $S_1, S_2$ are sequences (of braids), we denote
2625: by $S_1 + S_2$ the concatenation of¨$S_1$ and¨$S_2$,
2626: \ie, the sequence obtained by appending¨$S_2$ after¨$S_1$.
2627: If $S$ is a sequence of braids, and $\xx$ is a
2628: braid, we denote by¨$\xx S$ the translated sequence
2629: obtained by multiplying each entry in¨$S$ by¨$\xx$
2630: on the left. 
2631: \end{nota}
2632: 
2633: With these conventions, the sequence $(1, \ss1,
2634: \dots, \ss1^\dd)$ can be expressed as $\Dd20(1)
2635: + \Dd21(1) + \cdots + \Dd2\dd(1)$, and we can
2636: state:
2637: 
2638: \begin{prop} \label{P:Enum3}
2639: Let $\SS2\dd$ be defined for $\dd \ge 0$ by
2640: \begin{equation} \label{E:Def2}
2641: \SS2\dd = \Dd20\SS1\dd + \Dd21\SS1\dd + \cdots + \Dd2\dd\SS1\dd.
2642: \end{equation}
2643: Then $\SS2\dd$ is the increasing enumeration
2644: of¨$\DDD2\dd$.
2645: \end{prop}
2646: 
2647: We repeat the process for¨$\nn = 3$, introducing
2648: a sequence¨$\SS3\dd$ by a definition similar
2649: to¨\eqref{E:Def2} that involves $\SS2\dd$
2650: and¨$\SS3{\dd-1}$. The result we shall prove is:
2651: 
2652: \begin{prop} \label{P:Main3}
2653: Let $\SS3\dd$ be defined for $\dd \ge 0$ by
2654: \begin{align} \label{E:Def3}
2655: \SS3\dd =  \Dd30\SS2 \dd + \SS3{\dd, 1}
2656: + \Dd31\SS2 \dd + \cdots 
2657: + \Dd3{2\dd-1}\SS2 \dd + \SS3{\dd, 2\dd} + \Dd3{2\dd} \SS2 \dd,
2658: \end{align}
2659: where $\SS3{\dd, 1}, \cdots, \SS3{\dd, 2\dd}$ are 
2660: defined by $\SS3{\dd, 1} =
2661: \SS3{\dd, 2\dd} =
2662: \emptyset$ and, for $2 \le \ppp \le 2\dd-1$, 
2663: \begin{equation*}
2664: \SS3{\dd, \ppp} = 
2665: \begin{cases}
2666: \phantom{\ss2}\ss1(\SS3{\dd-1, \ppp-1} +
2667: \Dd3{\ppp-1}\SS2{\dd-1} +
2668: \SS3{\dd-1, \ppp}) & \mbox{for $\ppp = 0 \pmod 4$,}\\
2669: \ss2\ss1(\SS3{\dd-1, \ppp-2} + \Dd3{\ppp-1}\SS2{\dd-1} +
2670: \SS3{\dd-1, \ppp-1}) & \mbox{for $\ppp = 1 \pmod 4$,}\\
2671: \phantom{\ss2}\ss2(\SS3{\dd-1, \ppp-1} +
2672: \Dd3{\ppp-1}\SS2{\dd-1} +
2673: \SS3{\dd-1, \ppp}) & \mbox{for $\ppp = 2 \pmod 4$,}\\
2674: \ss1\ss2(\SS3{\dd-1, \ppp-2} + \Dd3{\ppp-1}\SS2{\dd-1} +
2675: \SS3{\dd-1, \ppp-1}) & \mbox{for $\ppp = 3 \pmod 4$.}
2676: \end{cases}
2677: \end{equation*}
2678: Then $\SS3\dd$ is the increasing enumeration
2679: of¨$\DDD3\dd$.
2680: \end{prop}
2681: 
2682: The general scheme is illustrated in 
2683: Table¨\ref{T:Triangle}: the sequence¨$\SS3\dd$ is
2684: constructed by starting with $2\dd+1$¨copies
2685: of¨$\SS2\dd$ translated by¨$\Dd30$, \dots,
2686: $\Dd3{2\dd}$ and inserting (translated copies
2687: of) fragments of the previous sequence¨$\SS3{\dd-1}$.  
2688: 
2689: \begin{table}[t]
2690: \begin{picture}(130,47)(1, 0)
2691: \put(0,10){$\Dd30 \SS23$}
2692: \put(10,10){$(\SS3{3,1})$}
2693: \put(20,10){$\Dd31 \SS23$}
2694: \put(32,10){$\SS3{3,2}$}
2695: \put(40,10){$\Dd32 \SS23$}
2696: \put(52,10){$\SS3{3,3}$}
2697: \put(60,10){$\Dd33 \SS23$}
2698: \put(72,10){$\SS3{3,4}$}
2699: \put(80,10){$\Dd34 \SS23$}
2700: \put(92,10){$\SS3{3,5}$}
2701: \put(100,10){$\Dd34 \SS23$}
2702: \put(110,10){$(\SS3{3,6})$}
2703: \put(120,10){$\Dd36 \SS23$}
2704: \put(10,9){$\underbrace{\hbox to 28mm{\hfill}}$}
2705: \put(14,4){$\ss2\cdot$}
2706: \put(29,4){$\ss1\ss2\cdot$}
2707: \put(17.5,3){$\swarrow$}
2708: \put(26.5,3){$\searrow$}
2709: \put(52,9){$\underbrace{\hbox to 26mm{\hfill}}$}
2710: \put(56,4){$\ss1\cdot$}
2711: \put(70,4){$\ss2\ss1\cdot$}
2712: \put(59,3){$\swarrow$}
2713: \put(68,3){$\searrow$}
2714: \put(92,9){$\underbrace{\hbox to 27mm{\hfill}}$}
2715: \put(96,4){$\ss2\cdot$}
2716: \put(111,4){$\ss1\ss2\cdot$}
2717: \put(99,3){$\swarrow$}
2718: \put(108,3){$\searrow$}
2719: \put(14,0){$\cdots$}
2720: \put(31,0){$\cdots$}
2721: \put(54,0){$\cdots$}
2722: \put(71,0){$\cdots$}
2723: \put(95,0){$\cdots$}
2724: \put(112,0){$\cdots$}
2725: 
2726: \put(20,22){$\Dd30 \SS22$}
2727: \put(30,22){$(\SS3{2,1})$}
2728: \put(40,22){$\Dd31 \SS22$}
2729: \put(52,22){$\SS3{2,2}$}
2730: \put(60,22){$\Dd32 \SS22$}
2731: \put(72,22){$\SS3{2,3}$}
2732: \put(80,22){$\Dd33 \SS22$}
2733: \put(90,22){$(\SS3{2,4})$}
2734: \put(100,22){$\Dd34 \SS22$}
2735: \put(30,21){$\underbrace{\hbox to
2736: 28mm{\hfill}}$}
2737: \put(36,16){$\ss2\cdot$}
2738: \put(50,16){$\ss1\ss2\cdot$}
2739: \put(38,14.5){$\swarrow$}
2740: \put(47,14.5){$\searrow$}
2741: \put(72,21){$\underbrace{\hbox to
2742: 27mm{\hfill}}$}
2743: \put(77,16){$\ss1\cdot$}
2744: \put(91,16){$\ss2\ss1\cdot$}
2745: \put(79,14.5){$\swarrow$}
2746: \put(88,14.5){$\searrow$}
2747: 
2748: \put(40,34){$\Dd30 \SS21$}
2749: \put(50,34){$(\SS3{1,1})$}
2750: \put(60,34){$\Dd31 \SS21$}
2751: \put(70,34){$(\SS3{1,2})$}
2752: \put(80,34){$\Dd32 \SS21$}
2753: \put(50,33){$\underbrace{\hbox to
2754: 28mm{\hfill}}$}
2755: \put(56,28){$\ss2\cdot$}
2756: \put(70,28){$\ss1\ss2\cdot$}
2757: \put(58,26.5){$\swarrow$}
2758: \put(67,26.5){$\searrow$}
2759: 
2760: \put(60,44){$\Dd30 \SS20$}
2761: \end{picture}
2762: \smallskip
2763: \caption{\smaller The inductive construction
2764: of¨$\SS3\dd$ as a Pascal triangle: the subsequence¨$
2765: \SS3{\dd,\ppp}$ is obtained by (translating and)
2766: concatenating the previous subsequences¨$
2767: \SS3{\dd-1,\ppp-1}$ and¨$
2768: \SS3{\dd-1,\ppp}$, or¨$\SS3{\dd-1,\ppp-2}$ and¨$
2769: \SS3{\dd-1,\ppp-1}$, depending on the parity of¨$\ppp$;
2770: the parenthesized sequences are empty; if we forget about
2771: the subsequences¨$\Dd3\qq\SS2\dd$, we have the
2772: Pascal triangle.}
2773: \label{T:Triangle}
2774: \end{table}
2775: 
2776: \begin{exam} \label{X:SS3}
2777: The difference between the definition of¨$\SS3\dd$
2778: in¨\eqref{E:Def3} and that of¨$\SS2\dd$ in¨\eqref{E:Def2} is
2779: the insertion of the additional factors $ \SS3{\dd,\ppp}$
2780: between the consecutive terms¨$\Dd3\qq\SS2\dd$. Because
2781: $\SS3{\dd, 1}$ and $\SS3{\dd, 2\dd}$ are empty, the difference
2782: occurs for $\dd \ge 2$ only. The first values are:
2783: 
2784: $\SS30 = \Dd30 \SS20 = (1)$,
2785: 
2786: $\SS31 = \Dd30 \SS21 + \SS3{1,1} + \Dd31 \SS21 +
2787: \SS3{1,2} + \Dd32 \SS21$
2788: 
2789: \hspace{1cm} $= (1, \a) + \emptyset
2790: + \b(1, \a) + \emptyset 
2791: + \a\b(1, \a)$
2792: 
2793: \hspace{1cm} $= (1, \a, \b, \b\a, \a\b, \a\b\a)$,
2794: 
2795: $\SS32  = \Dd30 \SS22 + \SS3{2,1} +
2796: \Dd31 \SS22 + \SS3{2,2} +
2797: \Dd32 \SS22 + \SS3{2,3} +
2798: \Dd33 \SS22 + \SS3{2,4} +
2799: \Dd34 \SS22$
2800: 
2801: \hspace{1cm} $ = (1, \a, \a\a) + \emptyset
2802: + \b(1, \a, \a\a) + \b(\b, \b\a)
2803: + \a\b(1, \a, \a\a)$
2804: 
2805: \hspace{4cm} $ + \a\b(\b, \b\a)
2806: + \a\a\b(1, \a, \a\a) + \emptyset 
2807: + \b\a\a\b(1, \a, \a\a)$
2808: 
2809: \hspace{1cm} $ = (1, \a, \a\a, \b, \b\a, \b\a\a,
2810: \b\b, \b\b\a,
2811: \a\b, \a\b\a, \a\b\a\a,
2812: \a\b\b, \a\b\b\a, \a\a\b$, 
2813: 
2814:  \hspace{6cm} $\a\a\b\a, \a\a\b\a\a, \b\a\a\b,
2815: \b\a\a\b\a, \b\a\a\b\a\a)$.\\
2816: It is easy to check directly that the sequence
2817: $\SS3\dd$ provides the increasing enumeration
2818: of¨$\DDD3\dd$ for $\dd = 0, 1, 2$. 
2819: \end{exam}
2820: 
2821: The proof of Proposition¨\ref{P:Main3} will be
2822: split into several pieces, each of which is
2823: established using an induction on the degree¨$\dd$.
2824: 
2825: \begin{lemm} \label{L:Inclusion3}
2826: All entries in¨$\SS3\dd$ are divisors of¨$\DD3\dd$.
2827: \end{lemm}
2828: 
2829: \begin{proof}
2830: The result is true for $\dd = 0$. Assume $\dd \ge 1$. By
2831: construction, each entry in¨$\SS3\dd$ either is of the form
2832: $\Dd3\qq \ss1^\expo$ with $0 \le \qq \le 2\dd$ and $0 \le
2833: \expo \le \dd$, or belongs to some
2834: subsequence¨$\SS3{\dd,\ppp}$ with $2 \le \ppp \le 2\dd -
2835: 1$. In the first case,
2836: $\Dd3\qq\ss1^\expo$ is a right divisor of $\Dd3{2\dd}
2837: \ss1^\expo$, which itself is a left divisor of
2838: $\Dd3{2\dd} \ss1^\dd$. By¨\eqref{E:Deltadelta}, the latter
2839: is¨$\DD3\dd$. Hence each $\Dd3\qq
2840: \ss1^\expo$ is a divisor of¨$\DD3\dd$. As for the
2841: entries coming from some subsequence¨$\SS3{\dd,\ppp}$,
2842: by definition they are of the form¨$\xx\yy$ with $\xx$
2843: one of $\ss2, \ss1\ss2, \ss1,
2844: \ss2\ss1$ and $\yy$ an entry in¨$\SS3{\dd-1}$. Then
2845: $\xx$ is a divisor of¨$\D_3$, while, by induction
2846: hypothesis, $\yy$ is a divisor of¨$\DD3{\dd-1}$, so
2847: $\xx\yy$ is a divisor of¨$\DD3\dd$.
2848: \end{proof}
2849: 
2850: \begin{lemm} \label{L:Card3}
2851: The length of the sequence¨$\SS3\dd$ equals the
2852: cardinality of¨$\DDD3\dd$.
2853: \end{lemm}
2854: 
2855: \begin{proof}
2856: Let $\ell_\dd$ denote the length of¨$\SS3\dd$.
2857: Computing¨$\ell_\dd$ is not very difficult. However,
2858: there is need to do it, which amounts to solve a
2859: recursive formula unnecessarily. Indeed, we saw in
2860: Section¨\ref{S:Deter} that the cardinality¨$\hhh13\dd$
2861: of¨$\DDD3\dd$ obeys the inductive rule¨\eqref{E:Recur3}.
2862: So it will be enough to check that
2863: $\ell_\dd$ satisfies the relation
2864: \begin{equation} \label{E:Double}
2865: \ell_\dd = 2 \ell_{\dd-1} + 3\dd + 1,
2866: \end{equation}
2867: and starts from the initial¨$\ell_1 = 6$ (or $\ell_0 =
2868: 1$). The latter point was checked in Example¨\ref{X:SS3}.
2869: 
2870: Now Table¨\ref{T:Triangle} shows that most entries
2871: in¨$\SS3{\dd-1}$ give rise to two entries in¨$\SS3\dd$.
2872: More precisely, each entry of¨$\SS3{\dd-1}$ not
2873: belonging to a factor of the form¨$\Dd3{2\qq}
2874: \SS2{\dd-1}$ gives rise to two entries in¨$\SS3\dd$,
2875: and, conversely, each entry in¨$\SS3\dd$ not belonging
2876: to a factor¨$\Dd3{\qq}
2877: \SS2{\dd}$ comes from such an entry in¨$\SS3{\dd-1}$.
2878: There are $\dd$¨factors¨$\Dd3{2\qq} \SS2{\dd-1}$
2879: in¨$\SS3{\dd-1}$, each of length¨$\dd$, and
2880: $2\dd+1$¨factors¨$\Dd3{2\qq} \SS2{\dd}$ in¨$\SS3\dd$,
2881: each of length¨$\dd+1$. So we obtain
2882: $$\ell_\dd - (2\dd+1)(\dd+1) = 2(\ell_{\dd-1} - \dd^2),$$
2883: which gives¨\eqref{E:Double}.
2884: \end{proof}
2885: 
2886: At this point, we cannot (yet) conclude that each
2887: divisor of¨$\DD3\dd$ occurs exactly once in¨$\SS3\dd$,
2888: as there could be some repetitions.
2889: 
2890: \subsection{A quotient-sequence for¨$\SS3\dd$}
2891: 
2892: Our next aim is to show that $\SS3\dd$ is
2893: $\sm$-increasing. To this end, we shall explicitly
2894: determine the quotient of adjacent entries
2895: in¨$\SS3\dd$, \ie, we shall specify a quotient-sequence
2896: for¨$\SS3\dd$ in the sense of Definition¨\ref{D:Witness}.
2897: 
2898: A preliminary step consists in determining the first and
2899: the last entries of the sequence¨$ \SS3{\dd,\ppp}$.
2900: For¨$S$ a nonempty sequence, we denote by $\first S$
2901: (\resp $\last S$) the first (\resp last) entry in¨$S$.
2902: 
2903: \begin{lemm} \label{L:FirstLast}
2904: For $1 < i < 2\dd$, we have
2905: \begin{equation}
2906: \first{ \SS3{\dd,\ppp}} = \Dd3{\ppp-1} \, \ss2, 
2907: \mbox{\quad and \quad}
2908: \last{ \SS3{\dd,\ppp}} \, \ss2 = \Dd3\ppp \, \sss1\dd. 
2909: \end{equation}
2910: \end{lemm}
2911: 
2912: \begin{proof}
2913: The result is vacuously true for $\dd = 0, 1$. Assume $\dd \ge
2914: 2$ with $\ppp  = 0 \pmod 4$. Using the definition, the
2915: induction hypothesis, and¨\eqref{E:Commut}, we find
2916: \begin{gather*}
2917: \first{ \SS3{\dd,\ppp}} 
2918: = \ss1 \, \first{ \SS3{\dd-1,\ppp-1}} = 
2919: \ss1 \, \Dd3{\ppp-2} \, \ss2
2920: = \Dd3{\ppp-1} \, \ss2, \\
2921: \last{ \SS3{\dd,\ppp}} \, \ss2 
2922: = \ss1 \, \last{ \SS3{\dd-1,\ppp}} \, \ss2 
2923: =  \ss1 \, \Dd3\ppp \, \sss1{\dd-1} 
2924: = \Dd3\ppp \, \sss1\dd.
2925: \end{gather*}
2926: Similarly, for $\ppp  = 1 \pmod 4$, we have
2927: \begin{gather*}
2928: \first{ \SS3{\dd,\ppp}} 
2929: = \ss2\ss1 \, \first{ \SS3{\dd-1,\ppp-2}} = 
2930: \ss2\ss1 \, \Dd3{\ppp-3} \, \ss2
2931: = \Dd3{\ppp-1} \, \ss2, \\
2932: \last{ \SS3{\dd,\ppp}} \, \ss2 
2933: = \ss2\ss1 \, \last{ \SS3{\dd-1,\ppp-1}} \, \ss2 
2934: =  \ss2\ss1 \, \Dd3{\ppp-1} \, \sss1{\dd-1} 
2935: = \ss2\, \Dd3{\ppp-1} \, \sss1\dd
2936: = \Dd3\ppp \, \sss1\dd.
2937: \end{gather*}
2938: Then, for $\ppp  = 2 \pmod 4$, we have
2939: \begin{gather*}
2940: \first{ \SS3{\dd,\ppp}} 
2941: = \ss2 \, \first{ \SS3{\dd-1,\ppp-1}} = 
2942: \ss2 \, \Dd3{\ppp-2} \, \ss2
2943: = \Dd3{\ppp-1} \, \ss2, \\
2944: \last{ \SS3{\dd,\ppp}} \, \ss2 
2945: = \ss2 \, \last{ \SS3{\dd-1,\ppp}} \, \ss2 
2946: =  \ss2 \, \Dd3\ppp \, \sss1{\dd-1} 
2947: = \Dd3\ppp \, \sss1\dd.
2948: \end{gather*}
2949: Finally, for $\ppp  = 3 \pmod 4$, we find
2950: \begin{gather*}
2951: \first{ \SS3{\dd,\ppp}} 
2952: = \ss1\ss2 \, \first{ \SS3{\dd-1,\ppp-2}} = 
2953: \ss1\ss2 \, \Dd3{\ppp-3} \, \ss2
2954: = \Dd3{\ppp-1} \, \ss2, \\
2955: \last{ \SS3{\dd,\ppp}} \, \ss2 
2956: = \ss1\ss2 \, \last{ \SS3{\dd-1,\ppp-1}} \, \ss2 
2957: =  \ss1\ss2 \, \Dd3{\ppp-1} \, \sss1{\dd-1} 
2958: =  \ss1\ss2\ss1\ss2 \, \Dd3{\ppp-3} \, \sss1{\dd-1}\\
2959: \hspace{4cm}
2960: =  \ss1\ss1\ss2\ss1 \, \Dd3{\ppp-3} \, \sss1{\dd-1} 
2961: =  \ss1\ss1\ss2 \, \Dd3{\ppp-3} \, \sss1\dd
2962: = \Dd3\ppp \, \sss1\dd.
2963: \end{gather*}
2964: This completes the argument.
2965: \end{proof}
2966: 
2967: We shall now construct an explicit quotient-sequence
2968: for¨$\SS3\dd$, \ie, a sequence of braid words
2969: representing the quotients of the consecutive entries
2970: of¨$\SS3\dd$. Before doing it for¨$\SS3\dd$, let us
2971: consider the (trivial) cases of¨$\SS1\dd$ and¨$\SS2\dd$.
2972: As $\SS1\dd$ consists of one single entry, it vacuously
2973: admits the empty sequence as a quotient-sequence. As
2974: for¨$\SS2\dd$, we can state:
2975: 
2976: \begin{lemm}
2977: For $\dd \ge 0$, let $\WW1\dd$ be the empty sequence, and
2978: let¨$\WW2\dd$ be defined by
2979: \begin{equation} \label{E:DDef2}
2980: \WW2\dd = \WW1\dd + (\ss1) + \WW1\dd + \cdots + \WW1\dd
2981: + (\ss1) + \WW1\dd,
2982: \end{equation}
2983: $\dd$¨times¨$(\ss1)$. Then $\WW2\dd$ is a
2984: quotient-sequence for¨$\SS2\dd$.
2985: \end{lemm}
2986: 
2987: On a similar way, we shall prove:
2988: 
2989: \begin{prop} \label{P:Quotient3}
2990: Let $\WW3\dd$ be the sequence defined by
2991: $\WW30 = \emptyset$ and 
2992: \begin{align} \label{E:DDef3}
2993: \WW3\dd =  \WW2\dd 
2994: \!+\! (\ss1^{-\dd}\ss2) 
2995: &+ \WW2\dd 
2996: \!+\! (\ss1^{-\dd}\ss2) 
2997: \!+\! \WW3{\dd, 2} \!+\!
2998: (\ss2\ss1^{-\dd}) \\ 
2999: \notag &+ \WW2\dd 
3000: \!+\! (\ss1^{-\dd}\ss2) 
3001: \!+\! \WW3{\dd, 3} 
3002: \!+\!
3003: (\ss2\ss1^{-\dd}) 
3004: \!+\! \cdots \\  
3005: \notag &+ \WW2\dd 
3006: \!+\! (\ss1^{-\dd}\ss2) 
3007: \!+\! \WW3{\dd, 2\dd-1} \!+\!
3008: (\ss2\ss1^{-\dd}) 
3009: \!+\! \WW2\dd 
3010: \!+\! (\ss2\ss1^{-\dd})
3011: \!+\! \WW2\dd,
3012: \end{align}
3013: \begin{tabbing} 
3014: with \= $\WW3{\dd, 2} = \WW3{\dd, 3} 
3015: = \WW2{\dd-1} 
3016: \!+\! (\ss2\sss1{-\dd+1}) 
3017: \!+\! \WW3{\dd-1, 2},$\\ 
3018: \> \vline width0pt height15pt
3019: $\WW3{\dd, 2\dd-2} =\WW3{\dd, 2\dd-1}
3020: = \WW3{\dd-1, 2\dd-3} 
3021: \!+\! (\sss1{-\dd+1}\ss2) 
3022: \!+\! \WW2{\dd-1}$,\\
3023: and
3024: \> \vline width0pt height15pt
3025: $\WW3{\dd, 2\pppp}  = \WW3{\dd, 2\pppp+1} 
3026: = \WW3{\dd-1,2\pppp-1} 
3027: \!+\! (\sss1{-\dd+1}\ss2) 
3028: \!+\! \WW2{\dd-1} 
3029: \!+\! (\ss2\sss1{-\dd+1}) 
3030: \!+\! \WW3{\dd-1, 2\pppp}$
3031: \end{tabbing}
3032: for $4 \le 2\pppp \le 2\dd-4$. Then
3033: $\WW3\dd$ is a quotient-sequence for¨$\SS3\dd$.
3034: \end{prop}
3035: 
3036: \begin{exam} \label{X:Quotient}
3037: We find
3038: $\WW31 
3039:  = \WW21 + (\A\b) + \WW21 + (\b\A) + \WW21
3040:  = (\a, \A\b, \a, \b\A, \a)$, and
3041: \begin{align*}
3042: \WW32 
3043: = \WW22 + (\A\A\b) 
3044: &+ \WW22 + (\A\A\b) 
3045: + \WW3{2, 2} + (\b\A\A)\\
3046: &+ \WW22 +
3047: (\A\A\b) + \WW3{2, 3} + (\b\A\A) + \WW22 +
3048: (\b\A\A) + \WW22
3049: \end{align*}
3050:  with $\WW3{2, 2} = \WW3{2, 3} = \WW21 = (\a)$,
3051: whence
3052: $$\WW32 
3053:  = (\a, \a, \A\A\b, \a, \a, \A\A\b, \a, \b\A\A, \a, \a,
3054: \A\A\b, \a, \b\A\A, \a, \a, \b\A\A, \a, \a).$$
3055: \end{exam}
3056: 
3057: \begin{proof} [Proof of Proposition¨\ref{P:Quotient3}]
3058: We prove using induction on¨$\dd$ that $\WW3\dd$ is a
3059: quotient-sequence for¨$\SS3\dd$ with the $4\dd-2$¨terms
3060: in¨\eqref{E:DDef3} corresponding to the $4\dd-1$¨nonempty
3061: terms in¨\eqref{E:Def3}---so, in particular, for $2
3062: \le \ppp \le 2\dd-1$, the subsequence¨$\WW3{\dd,\ppp}$ is
3063: a quotient-sequence for ¨$ \SS3{\dd,\ppp}$. The result is
3064: vacuously true for
3065: $\dd=0$.  
3066: Assume $\dd \ge 1$. By definition, the sequence¨$\SS3\dd$
3067: consists of the concatenation of the $2\dd+1$ sequences
3068: $\Dd30 \SS2 \dd, \cdots, \Dd3{2\dd} \SS2 \dd$, in which the $2\dd-2$
3069: sequences $\SS3{\dd, 2}, \dots, \SS3{\dd, 2\dd-1}$ are inserted.
3070: We shall consider these subsequences separately, and then
3071: consider the transitions between consecutive subsequences.
3072: 
3073: First, $\WW2\dd$ is a quotient-sequence for¨$\SS2\dd$,
3074: hence it is a quotient-sequence for every sequence
3075: $\Dd3\qq
3076: \SS2\dd$ as well, since, by definition, the quotients we
3077: consider are invariant under left translation. Then, by
3078: construction, each subsequence
3079: $\SS3{\dd, 2\pppp}$ or $\SS3{\dd, 2\pppp+1}$ appearing in¨$\SS3\dd$ is
3080: obtained by translating some subsequence¨$S$
3081: of¨$\SS3{\dd-1}$, namely
3082: $$S =  \SS3{\dd-1, 2\pppp-1} + \Dd3{\qq-1} \SS2{\dd-1} +
3083: \SS3{\dd-1, 2\pppp}.$$
3084: By induction hypothesis, the sequence
3085: $$\WW3{\dd-1, 2\pppp-1} +
3086: (\sss1{-\dd+1}\ss2) + \WW2{\dd-1} +
3087: (\ss2\sss1{-\dd+1}) +
3088: \WW3{\dd-1, 2\pppp},$$
3089: which is precisely $\WW3{\dd, 2\pppp}$ and $\WW3{\dd, 2\pppp+1}$
3090: by definition, is a quotient-sequence for¨$S$. The
3091: property remains true in the special cases $\pppp = 1$
3092: and $\pppp = \dd$, which correspond to removing the
3093: initial term¨$\SS3{\dd-1,2\pppp-1}$, and/or the final
3094: term¨$\SS3{\dd-1, 2\pppp}$, respectively. Then
3095: $\WW3{\dd, 2\pppp}$ and $\WW3{\dd, 2\pppp+1}$ are also
3096: quotient-sequences for any sequence obtained from¨$S$ by
3097: a left translation, in particular for¨$\SS3{\dd,
3098: 2\pppp}$ and¨$\SS3{\dd, 2\pppp+1}$.
3099: 
3100: So it only remains to study the transitions between the
3101: consecutive terms in the expression¨\eqref{E:Def3}
3102: of¨$\SS3\dd$, \ie, to compare the last entry in each term
3103: with the first entry in the next term. Four cases are to be
3104: considered, namely the special case of the first two terms
3105: and of the final two terms, and the generic cases of the
3106: transitions from¨$\Dd3\qq \SS2\dd$ to¨$\SS3{\dd,\ppp+1}$
3107: and from¨$\SS3{\dd,\ppp}$ to¨$\Dd3\qq \SS2\dd$.
3108: 
3109: As for the first two terms, namely $\Dd30 \SS2\dd$ and
3110: $\Dd31 \SS2\dd$, \ie, $\SS2\dd$ and $\ss2\SS2\dd$, the last
3111: entry in¨$\SS2\dd$ is¨$\sss1\dd$, while the first entry
3112: in¨$\ss2\SS2\dd$ is¨$\ss2$, so $\sss1{-\dd}\ss2$ is a quotient.
3113: As for the last two terms, namely $\Dd3{2\dd-1} \SS2\dd$ and
3114: $\Dd3{2\dd} \SS2\dd$, the last entry in¨$\Dd3{2\dd-1}\SS2\dd$
3115: is¨$\Dd3{2\dd-1}\,\sss1\dd$, while the first entry
3116: in¨$\Dd3{2\dd}\SS2\dd$ is¨$\Dd3{2\dd}$. Now,
3117: by¨\eqref{E:Deltadelta}, we have 
3118: $\Dd3{2\dd-1}\,\sss1\dd\ss2 =
3119: \Dd3{2\dd}
3120: \,\sss1\dd$, so $\ss2\sss1{-\dd}$ is an expression of the
3121: quotient.
3122: 
3123: Consider now the transition from¨$\Dd3\qq \SS2\dd$
3124: to¨$\SS3{\dd,\qq+1}$. The last entry in¨$\Dd3\qq \SS2\dd$
3125: is¨$\Dd3\qq\, \sss1\dd$, while, by Lemma¨\ref{L:FirstLast}, the
3126: first entry in¨$\SS3{\dd,\qq+1}$ is¨$\Dd3\qq \,\ss2$. Hence
3127: $\sss1{-\dd}\ss2$ represents the quotient. Finally,
3128: consider the transition from¨$\SS3{\dd,\ppp}$ to¨$\Dd3\qq
3129: \SS2\dd$. By Lemma¨\ref{L:FirstLast} again, the last
3130: entry¨$\xx$ in¨$\Dd3\qq \SS2\dd$ satisfies¨$\xx\, \ss2
3131: = \Dd3\qq \, \sss1\dd$, while the first entry in¨$\Dd3\qq
3132: \SS2\dd$ is¨$\Dd3\qq$. Hence $\ss2\sss1{-\dd}$
3133: represents the quotient.
3134: \end{proof}
3135: 
3136: \begin{coro} \label{C:Distinct3}
3137: For each¨$\dd$ the sequence¨$\SS3\dd$
3138: is $\sm$-increasing; so, in particular, it consists of
3139: pairwise distinct braids.
3140: \end{coro}
3141: 
3142: \begin{proof}
3143: By definition, every word in¨$\WW3\dd$ is
3144: $\s$-positive, hence, by Property¨A, it does not
3145: represent¨$1$.
3146: \end{proof}
3147: 
3148: As $\SS3\dd$ consists of pairwise distinct divisors
3149: of¨$\DD3\dd$, Lemma¨\ref{L:Card3} implies that every
3150: divisor of¨$\DD3\dd$ occurs exactly once in¨$\SS3\dd$.
3151: Then, as $\SS3\dd$ is $\sm$-increasing, it must be the
3152: increasing enumeration of¨$\DDD3\dd$, and the proof of
3153: Proposition¨\ref{P:Main3} is complete.
3154: 
3155: \begin{rema}
3156: Once we know that $\SS3\dd$ is the increasing
3157: enumeration of¨$\DDD3\dd$ and that $\WW3\dd$ is a
3158: $\s$-positive quotient-sequence for¨$\SS3\dd$, we can
3159: count the $2$-jumps in¨$\SS3\dd$ and obtain the value
3160: of¨$\hhh23\dd$ directly: this amounts to forgetting
3161: about all¨$\ss1^{\pm1}$ in the construction
3162: of¨$\WW3\dd$, and it is then fairly obvious that there
3163: only remains $2^\dd-2$ times¨$\ss2$.
3164: \end{rema}
3165: 
3166: \subsection{Larger values of¨$\nn$}
3167: 
3168: The same construction can be developped for¨$\nn = 4$
3169: and further. The general scheme is clear, namely to
3170: define¨$\SS4\dd$ using an inductive rule
3171: \begin{align} \label{E:Def4}
3172: \SS4\dd =  \Dd40 \SS3\dd + \SS4{\dd, 1}
3173: + \Dd41 \SS3\dd + \cdots 
3174: + \Dd4{3\dd-1} \SS3\dd + \SS4{\dd, 3\dd} + \Dd4{3\dd}
3175: \SS3\dd,
3176: \end{align}
3177: where the intermediate factor¨$\SS4{\dd,\ppp}$ is
3178: constructed by concatenating and translating convenient
3179: fragments of¨$\SS4{\dd-1}$. Owing to the inductive
3180: rule¨\eqref{E:Recur4} satisfied by the number of
3181: elements¨$\hhh14\dd$ of¨$\DDD4\dd$, we can expect that the
3182: generic entry of¨$\SS4{\dd-1}$ has to be repeated 6¨times
3183: in¨$\SS4\dd$, but with some entries from¨$\SS4{\dd-2}$
3184: repeated 3¨times only. Once the inductive definition
3185: of¨$\SS4\dd$ is made complete, showing that the sequence
3186: is $\sm$-increasing and counting its entries should be
3187: easy. As we have no complete description so far, we
3188: leave the question open here.
3189: 
3190: \subsection{A new construction for the linear ordering
3191: of¨$B_3$}
3192: 
3193: The main interest of the approach described above is not
3194: only to connect the Garside structure of¨$B_\nn$ with
3195: its linear ordering, but also to provide a new
3196: independent construction of the braid ordering, at least
3197: in the case of¨$B_3$ as the latter is the only one for
3198: which the construction was completed so far.
3199: 
3200: As was recalled in the introduction, the existence of
3201: the linear ordering of braids relies on two properties
3202: of braids, namely Property¨A and Property¨C. These
3203: properties have received a number of independent
3204: proofs¨\cite{Dgr}. In particular, Property¨A has now
3205: a very short proof based on Dynnikov's coordinization
3206: for triangulations of a punctured disk (\cite{Dgr},
3207: Chapter¨9). As for Property¨C, no really simple proof
3208: exists so far: not to mention the initial argument
3209: involving self-distributive algebra, the combinatorial
3210: proofs based on handle reduction or on Burckel's
3211: uniform tree approach, as well as the geometric proofs
3212: based on standardization of curve diagrams require some
3213: care. So, at the moment, one can estimate that the
3214: optimal proof of Property¨C is still missing.
3215: 
3216: A direct application of our construction of the
3217: sequence¨$\SS3\dd$ is
3218: 
3219: \begin{prop}
3220: Property¨C holds for¨$B_3$, \ie, every non-trivial
3221: $3$-braid admits a $\s$-positive or a $\s$-negative
3222: expression.
3223: \end{prop}
3224: 
3225: \begin{proof}[New proof]
3226: We take as an hypothesis that Property¨A is true, so
3227: that the relation¨$\sm$ is a partial ordering, but we do
3228: not assume that $\sm$ is linear. As every braid in¨$B_3$
3229: is the quotient of two positive braids in¨$B_3^+$,
3230: proving Property¨C for¨$B_3$ amounts to proving that, if
3231: $\bx, \by$ are arbitrary elements of¨$B_3^+$, then the
3232: quotient $\bx\inv \by$ admits a $\s$-positive or a
3233: $\s$-negative expression.
3234: 
3235: Now the construction of¨$\SS3\dd$ is self-contained, and
3236: so is that of¨$\WW3\dd$. Then, by construction, every
3237: word in¨$\WW3\dd$ is $\s$-positive. As any concatenation
3238: of $\s$-positive words is $\s$-positive, it follows
3239: that, if $\bx, \by$ are any braids occurring
3240: in¨$\bigcup_\dd\SS3\dd$, then the quotient¨$\bx\inv \by$
3241: admits a $\s$-positive or a $\s$-negative expression,
3242: according to whether $\bx$ occurs before or after¨$\by$
3243: in¨$\SS3\dd$. So, in order to conclude that Property¨C is
3244: true, it just remains to check that each positive
3245: $3$-braid occurs in¨$\bigcup_\dd\SS3\dd$. As
3246: every entry of¨$\SS3\dd$ belongs to¨$\DDD3\dd$, this
3247: is equivalent to proving that each divisor of¨$\DD3\dd$
3248: occurs in¨$\SS3\dd$. Property¨A guarantees that the
3249: entries of¨$\SS3\dd$ are pairwise distinct
3250: (Corollary¨\ref{C:Distinct3}), so it suffices to compare
3251: the length of¨$\SS3\dd$ with the cardinality
3252: of¨$\DDD3\dd$, and this is what we
3253: made in Lemma¨\ref{L:Card3}.
3254: \end{proof}
3255: 
3256: Actually the construction of¨$\SS3\dd$ gives more.
3257: The approach developped by S.\,Burc\-kel in¨\cite{Bus}
3258: consists in introducing a convenient notion of normal
3259: braid words such that every positive braid admits
3260: exactly one normal expression. In the case of $3$¨
3261: strand braids, the definition is as follows. Every
3262: positive $3$¨strand braid word¨$\ww$ can be written as
3263: an alternated product of blocks¨$\ss1^{\expo}$
3264: and¨$\ss2^{\expo}$. Then we define the {\it code}
3265: of¨$\ww$ to be the sequence made by the sizes of these
3266: blocks. To avoid ambiguity, we decide that the last
3267: block is considered to be a block of¨$\ss1$'s, \ie, we
3268: decide that the code of¨$\ss1$ is¨$(1)$, while the code
3269: of¨$\ss2$ is¨$(1, 0)$. For instance, the code
3270: of¨$\ss2^2
3271: \ss1^3\ss2^5$ is $(2, 3, 5, 0)$.
3272: 
3273: \begin{defi}
3274: A positive $3$¨strand braid word¨$\ww$ is said to be
3275: {\it normal in the sense of Burckel} if its code has the
3276: form $(e_1, \dots, e_\ell)$ with $e_\kk \ge 2$ for $2
3277: \le \kk\le \ell-2$.
3278: \end{defi}
3279: 
3280: Burckel shows in¨\cite{Bus} that every positive
3281: $3$-braid admits a unique normal expression and,
3282: moreover, that $\bx \sm \by$ holds if and only if the
3283: normal form of¨$\bx$ is $\mathtt{ShortLex}$-smaller than
3284: the normal form of¨$\by$, where $\mathtt{ShortLex}$
3285: refers to the variant of the lexicographic ordering of
3286: sequences in which the length is given priority: $(e_1,
3287: \dots, e_\ell) <_{\mathtt{ShortLex}} (e'_1, \dots,
3288: e'_{\ell'})$ holds for $\ell < \ell'$, and for $\ell =
3289: \ell'$ and $(e_1, \dots, e_\ell)$ lexicographically
3290: smaller than¨$(e'_1, \dots, e'_{\ell'})$. Burckel's
3291: method consists in defining an iterative reduction
3292: process on non-normal braid words. Our current
3293: approach allows for an alternative, simpler method.
3294: First, a direct inspection shows:
3295: 
3296: \begin{lemm}
3297: Let $\SSS3\dd$ be the sequence of braid words defined by
3298: the inductive rule¨\eqref{E:Def3}. Then $\SSS3\dd$
3299: consists of words that are normal in the sense of
3300: Burckel.
3301: \end{lemm}
3302: 
3303: Then, by construction, every braid in¨$\SS3\dd$ is
3304: represented by a word of¨$\SSS3\dd$, so, as every
3305: positive $3$-braid occurs in¨$\bigcup\SS3\dd$, we
3306: immediately deduce:
3307: 
3308: \begin{prop}
3309: Every positive $3$-braid admits an expression that is
3310: normal in the sense of Burckel.
3311: \end{prop}
3312: 
3313: This in turn enables us to obtain a simple proof for
3314: the following deep, and so far not very well understood
3315: result originally due to Laver¨\cite{Lve} (and
3316: to Burckel \cite{Bus} for the ordinal type):
3317: 
3318: \begin{coro}
3319: The restriction of¨$\sm$ to¨$B_3^+$ is a well-ordering
3320: of ordinal type¨$\omega^\omega$.
3321: \end{coro}
3322: 
3323: \begin{proof}
3324: The $\mathtt{ShortLex}$ ordering of sequences of
3325: nonnegative integers is a well-ordering of ordinal
3326: type¨$\omega^\omega$, so its restriction to codes of
3327: normal words in the sense of Burckel is a well-ordering
3328: as well. The type of the latter cannot be less
3329: than¨$\omega^\omega$ as one can easily exhibit an
3330: increasing sequence of length¨$\omega^\omega$.
3331: \end{proof}
3332: 
3333: Burckel's approach extends to all braid
3334: monoids¨$B_\nn^+$, at the expense of introducing a
3335: convenient notion of normal word and defining an
3336: associated reduction process which is very intricate.
3337: Completing the construction of the sequences¨$\SS4\dd$
3338: and, more generally, $\SS\nn\dd$ along the lines
3339: described above would hopefully allow for an alternative
3340: much simpler approach. In particular, once the correct
3341: definition is given, all subsequent proofs should reduce
3342: to easy inductive verifications.
3343: 
3344: \begin{thebibliography}{99}
3345: 
3346: \Ref Bir; J.\,Birman; Braids, links, and mapping 
3347: class groups; Annals of Math. Studies 82, Princeton
3348: Univ.  Press  (1975). 
3349: 
3350: \Reff Bus; S.\,Burckel; The wellordering on
3351: positive braids; J. Pure Appl. Algebra; 120-1; 1997; 1--17.
3352: 
3353: \Reff But; S.\,Burckel; Computation of the
3354: ordinal of braids; Order; 16; 1999; 291--304.
3355: 
3356: \Reff Buu; S.\,Burckel; Syntactical methods for
3357: braids of three strands; J. Symb. Comp.; 31;
3358: 2001; 557--564.
3359: 
3360: \Reff CSV; L.\,Carlitz, R.\,Scoville \&
3361: T.\,Vaughan; Enumeration of
3362: pairs of permutations; Discrete Math.; 14;
3363: 1976; 215--239.
3364: 
3365: \Reff  Chb; R. Charney; Geodesic automation and
3366: growth functions for Artin groups of finite type; Math.
3367: Ann.; 301-2; 1995; 307--324.
3368: 
3369: \Reff Dfb; P. Dehornoy; Braid groups and left
3370: distributive  operations; Trans. Amer. Math. Soc.; 345-1;
3371: 1994; 115--151. 
3372: 
3373: \Reff Dfo; P.\,Dehornoy; A fast method for comparing
3374: braids; Advances in Math.; 125; 1997; 200--235.
3375: 
3376: \Reff Dgb; P.\,Dehornoy; Strange questions about braids;
3377: J. Knot Th. and its Ramifications; 8-5; 1999; 589--620.
3378: 
3379: \Reff Dgk; P.\,Dehornoy; Groupes de Garside; 
3380: Ann. Scient. Ec. Norm. Sup.; 35; 2002; 267--306.
3381: 
3382: \Ref Dgr; P.\,Dehornoy, I. Dynnikov, D. Rolfsen, B.
3383: Wiest; Why are braids orderable?; Panoramas \&
3384: Synth\`eses vol. 14, Soc. Math. France (2002).
3385: 
3386: \Ref Dhi; P.\,Dehornoy; Combinatorics of normal sequences
3387: of braids; Preprint.
3388: 
3389: \Reff ElM; E.A.\,Elrifai \& H.R.\,Morton; Algorithms for
3390: positive braids; Quart. J. Math. Oxford; 45-2; 1994;
3391: 479--497.
3392: 
3393: \Ref Eps; D.\,Epstein, J.\,Cannon, D.\,Holt, S.\,Levy,
3394: M.\,Paterson \& W.\,Thurston; Word Processing in Groups;
3395: Jones \& Bartlett Publ. (1992).
3396: 
3397: \Reff Gar; F.A.\,Garside; The braid group and
3398: other groups; Quart. J. Math. Oxford; 20-78; 1969;
3399: 235--254.
3400: 
3401: \Ref Lrb; D.\,Larue; Left-distributive and
3402: left-distributive idempotent algebras; Ph D Thesis,
3403: University of Colorado, Boulder (1994).
3404: 
3405: \Reff Lve; R.\,Laver; Braid group actions on left
3406: distributive structures and well-orderings in the braid
3407: group; J. Pure Appl. Algebra; 108-1; 1996; 81--98.
3408: 
3409: \Ref McC; J.\,Mc Cammond; An introduction to Garside
3410: structures; Preprint (2005).
3411: 
3412: \Reff Sol; L.\,Solomon; A Mackey formula in the group ring
3413: of a Coxeter group; J. Algebra; 41; 1976; 255-268.
3414: 
3415: \end{thebibliography}
3416: 
3417: \end{document}
3418: