math0305156/cb.tex
1: % This is LATEX
2: % using the style file of G&T: http://www.maths.warwick.ac.uk/gt
3: %
4: \documentclass{gtart}
5: \usepackage{amssymb, enumerate, epsfig}
6: %\input epsf
7: %\usepackage{pictex}
8: %
9: \def\<{\langle}
10: \def\>{\rangle}
11: \def\implies{\Longrightarrow}
12: \def\inv{^{-1}}
13: \def\basl{\backslash}
14: \def\Q{{\mathbb Q}}
15: \def\C{{\mathbb C}}
16: \def\N{{\mathbb N}}
17: \def\Z{{\mathbb Z}}
18: \def\R{{\mathbb R}}
19: \def\bH{{\mathbb H}}
20: \def\MCG{\mathcal{MCG}}
21: \def\marg#1{{\marginpar{\tiny #1}}}
22: \def\calG{\mathcal{G}}
23: \def\pA{pseudo-Anosov}
24: \def\BnD{B_n / \langle \Delta^2 \rangle}
25: %\def\Bp{B_\bigcirc}
26: %\def\Bpp{B_\odot}
27: %\def\Dp{D_\bigcirc}
28: %\def\Dpp{D_\odot}
29: \def\tld{\widetilde}
30: \renewcommand{\phi}{\varphi}
31: %
32: \newtheorem{lemma}{Lemma}[section]
33: \newtheorem{theorem}[lemma]{Theorem}
34: \newtheorem{prop}[lemma]{Proposition}
35: \newtheorem{cor}[lemma]{Corollary}
36: \newtheorem{definition}[lemma]{Definition}
37: \newtheorem{lem}[lemma]{Lemma}
38: \newtheorem{conj}[lemma]{Conjecture}
39: \newtheorem{remark}[lemma]{Remark}
40: \newtheorem{example}[lemma]{Example}
41: \newtheorem{question}[lemma]{Question}
42: %
43: 
44: \begin{document}
45: \title{On the structure of the centralizer of a braid}
46: \author{Juan Gonz\'alez-Meneses\footnote{Partially supported by MCYT,
47: BFM2001-3207 and FEDER.} and Bert Wiest}
48: \address{Dpto.~de Matem\'atica Aplicada I, E.T.S.~Arquitectura, Universidad
49: de Sevilla, Avda.~Reina Mercedes, 2. 41012 Sevilla, Spain;
50: {\tt meneses@us.es}\\
51: and\\
52: IRMAR (UMR 6625 du CNRS), Universit\'e de Rennes 1, Campus de Beaulieu,
53: \\ 35042 Rennes cedex, France; {\tt bertold.wiest@math.univ-rennes1.fr}}
54: 
55: \begin{abstract}
56: The mixed braid groups are the subgroups of Artin braid groups whose elements preserve a
57: given partition of the base points. We prove that the centralizer of any braid can be
58: expressed in terms of semidirect and direct products of mixed braid groups. Then we
59: construct a generating set of the centralizer of any braid on $n$ strands, which has at most
60: $\frac{k(k+1)}{2}$ elements if $n=2k$, and at most $\frac{k(k+3)}{2}$ elements if $n=2k+1$.
61: These bounds are shown to be sharp, due to work of N.V.Ivanov and of S.J.Lee. Finally, we
62: describe how one can explicitly compute this generating set.
63: \end{abstract}
64: \keywords{braid, centralizer, Nielsen-Thurston theory.}
65: \primaryclass{20F36}\secondaryclass{20E07, 20F65.}
66: %\maketitlepage
67: \makeshorttitle
68: 
69: % ------------------------------------------------------------------
70: 
71: \section{Introduction and statement of the results}\label{S:IN}
72: 
73: In 1971, Makanin \cite{Makanin} gave an algorithm for computing a
74: generating set of the centralizer $Z(\beta)$ of any given element $\beta$
75: of the $n$-string braid group $B_n$. His method, however, tends to yield
76: very large, and highly redundant generating sets.
77: %
78: One hint that much smaller generating sets could be found came from the experimental results
79: of Gonz\'alez-Meneses and Franco, which were obtained with a radically improved version of
80: Makanin's algorithm, based on new theoretical work \cite{GMFr}. Also, it has probably been
81: clear to specialists for a long time that Nielsen-Thurston theory could be used to improve
82: upon Makanin's results. However, there seems to be no such result in the literature, and the
83: aim of the present paper is to fill this gap.
84: 
85: Although our main interest was to compute, for any given $\beta\in B_n$,
86: a small generating set of $Z(\beta)$, we succeed in describing this
87: centralizer in terms of semidirect and direct products of
88: {\em mixed braid groups} (see~\cite{Manfredini,Orevkov}). These groups
89: are defined as follows: let $X=\{P_1,\ldots,P_n\}$ be the base points of
90: the braids in $B_n$. Given a partition ${\cal P}$ of $X$, the mixed braid
91: group $B_{\cal P}$ consists of those braids whose associated permutation
92: preserves each coset of ${\cal P}$.
93: 
94: The well known classification of mapping classes of a punctured surface
95: into periodic, reducible and pseudo-Anosov ones, yields an analogous
96: classification for braids. If $\beta$ is reducible, then one can decompose
97: it, in a certain sense, into a {\em tubular braid} $\widehat{\beta}$, and
98: some {\em interior braids} $\beta_{[1]},\ldots, \beta_{[t]}$, all of
99: them having less than $n$ stands. The main result of this paper is the
100: following:
101: 
102: \begin{theorem}\label{T:main1}
103: Let $\beta\in B_n$. One has:
104: \begin{enumerate}
105: \item If $\beta$ is pseudo-Anosov , then $Z(\beta)\simeq \mathbb{Z}^2$.
106: 
107: \item If $\beta$ is periodic, then $Z(\beta)$ is either $B_n$ or isomorphic
108: to a braid group on an annulus.
109: 
110: \item If $\beta$ is reducible, then there exists a split exact sequence:
111: $$
112: 1 \longrightarrow Z(\beta_{[1]})\times \cdots \times Z(\beta_{[t]})
113: \longrightarrow Z(\beta) \longrightarrow Z_0(\widehat{\beta})\longrightarrow 1,
114: $$
115: where $Z_0(\widehat{\beta})$ is a subgroup of $Z(\widehat{\beta})$,
116: isomorphic either to $\Z^2$ or to a mixed braid group.
117: \end{enumerate}
118: \end{theorem}
119: 
120: Notice that $\mathbb{Z}\simeq B_2=B_{\{\{1,2\}\}}$, also
121: $B_n=B_{\{\{1,\ldots,n\}\}}$, and finally the braid group over an annulus
122: on $k$ strands is isomorphic to $B_{\{\{1,\ldots,k\},\{k+1\}\}}\subset
123: B_{k+1}$. Hence all these groups can be seen as mixed braid groups. Then,
124: by recurrence on the number of strands we deduce the following:
125: 
126: \begin{cor}
127:  For every $\beta\in B_n$, the centralizer $Z(\beta)$ can be expressed in
128: terms of semidirect and direct products of mixed braid groups.
129: \end{cor}
130: 
131: Using the above structure we shall construct, for any braid $\beta\in B_n$,
132: a generating set of $Z(\beta)$ having very few elements. More precisely,
133: we obtain:
134: 
135: \begin{theorem}\label{T:main}
136: If $\beta\in B_n$, then the centralizer $Z(\beta)$ can
137: be generated by at most $\frac{k(k+1)}{2}$ elements if $n=2k$,
138: and at most $\frac{k(k+3)}{2}$ elements if $n=2k+1$.
139: \end{theorem}
140: 
141: We will present an example, communicated to us by S. J. Lee, showing that
142: the above bound is sharp.
143: That is, we will define, for every positive
144: integer $n$, a braid in $B_n$ whose centralizer cannot be generated by
145: less than $\frac{k(k+1)}{2}$ elements if $n=2k$, or less than
146: $\frac{k(k+3)}{2}$ elements if $n=2k+1$.
147: (The first to observe that the number of generators of the centralizer
148: may grow quadratically with the number of strands was N.V.Ivanov
149: \cite{IvanovTalk}.)
150: 
151: However, the above bound refers to the worst case, and one could be interested
152: in the minimal number of generators of a particular braid. We shall give
153: a generating set which is in some sense the smallest ``natural'' generating
154: set for the centralizer of a braid. However, we shall also give an example
155: that illustrates the difficulty of finding the absolutely minimum possible
156: number of generators.
157: 
158: Let us mention that, for the special case of reducible braids conjugated
159: to a generator $\sigma_i$, its centralizer has already been described
160: in~\cite{FRZ}.
161: 
162: The plan of the paper is as follows: in section \ref{S:NiTh} we set up
163: notation and some standard machinery, and give the mentioned example by
164: S.\ J.\ Lee. In section~\ref{S:per} we study $Z(\beta)$ in the case where
165: $\beta$ is periodic, section~\ref{S:pA} deals with the {\pA} case, and
166: section~\ref{S:red} the reducible one, which is the most involved. In
167: section~\ref{S:upperbound} we define a generating set which is no larger
168: than the stated upper bound. In section~\ref{S:minimal} we describe a
169: generating set which is as small as possible while still reflecting the
170: geometric structure of the Nielsen-Thurston decomposition. We also give an
171: example to show that by algebraic trickery, even smaller sets can be obtained.
172: Finally in section~\ref{S:alg} we discuss how the generating set that
173: we defined can be found algorithmically.
174: 
175: % ------------------------------------------------------------------
176: 
177: \section{Prerequisites from Nielsen-Thurston theory}\label{S:NiTh}
178: 
179: We denote by $D$ the closed disk of radius $2$ centered at $0$ in the
180: complex plane. For any $n\in \N$, the disk $D$, together with any choice
181: of $n$ distinct points in its interior, is denoted $D_n$, and the
182: distinguished points are called the {\sl punctures}. We shall use
183: different choices for the exact position of the punctures at different
184: times - they may be lined up on the real axis, or regularly distributed on
185: a circle of radius $1$, or again one of them may be in the centre while
186: the remaining $n-1$ are distributed over the circle of radius $1$. In most
187: instances, the position of the punctures is irrelevant, and we shall leave
188: it unspecified.
189: 
190: We recall that the braid group $B_n$ is the group of isotopy classes of
191: homeomorphisms fixing (pointwise) the boundary and permuting the punctures
192: of $D_n$. Here the isotopies must fix pointwise the boundary and the punctures.
193: Alternatively, $B_n$ could be defined as the group of isotopy
194: classes of disjoint movements of the punctures, starting and ending with
195: the configuration of $D_n$. Yet another definition of $B_n$ is as the set
196: of isotopy classes of braids with $n$ strings in the cylinder
197: $D\times [0,1]$, where the start and end points of the strings are exactly
198: the puncture points in $D_n\times \{0\}$ and $D_n\times \{1\}$. We shall
199: use all three points of view.
200: 
201: We shall often work with a certain quotient of the group $B_n$, rather
202: than with $B_n$ itself.
203: We recall that the center of $B_n$ is isomorphic to the integers,
204: and generated by the full twist $\Delta^2$ (where $\Delta$ is Garside's
205: half twist). Geometrically, the group projection
206: $B_n \to B_n/\langle \Delta^2 \rangle$ is given by smashing the boundary
207: curve of $D_n$ to a puncture, so that $\BnD$ is naturally a subgroup
208: of the mapping class group of the sphere with $n+1$ punctures.
209: In order to keep notation manageable, we shall use the same letters for
210: elements of the braid group $B_n$ and for their image in the quotient $\BnD$.
211: This abuse of notation should not cause confusion.
212: 
213: We say that an element $\beta\in B_n$
214: is {\sl periodic} if the element of $\BnD$ represented by
215: $\beta$ is of finite order. Equivalently, $\beta$ is periodic if there
216: exists a $k\in \N$  such that in $B_n$ we have that $\beta^k$ is equal
217: to some power of $\Delta^2$.
218: 
219: We say an element $\beta$ of $B_n$ is {\sl reducible} if there exists a
220: nonempty multicurve $C$ in $D_n$ (i.e.~a system of disjoint simple closed
221: curves in $D_n$, none of them isotopic to the boundary or enclosing a
222: single puncture)
223: which is stabilized by $\beta$, i.e.~such that $\beta(C)$ is isotopic to $C$.
224: Note that $\beta$ may permute different components of the multicurve $C$.
225: 
226: The following definition is taken from \cite{BLM} (see also \cite{Ivanov}).
227: To every reducible braid $\beta\in B_n$ one can associate a canonical
228: invariant multicurve: its {\sl canonical  reduction system}, which by
229: definition is the collection of all isotopy classes $c$ of simple closed
230: curves which have the following two properties:
231: firstly, $c$ must be stabilized by some power of $\beta$, and secondly
232: any simple closed curve which has non-zero geometric intersection number
233: with $c$ must {\it not} be stabilized by any power of $\beta$. For instance,
234: let us
235: consider the punctured disk $D_6$, where the 6 punctures are arranged
236: uniformly on the circle of radius $1$ around $0$. Then the
237: rotation of the punctures around the circle by an angle of $\frac{2\pi}{3}$
238: is a periodic element of $B_6$ (of period $3$), it is
239: also reducible (e.g.~the three simple closed curves encircling punctures
240: 1 and 2, 3 and 4, and 5 and 6 respectively form an invariant multicurve),
241: but its canonical  reduction system is empty. This example, however, is
242: somewhat untypical: if a {\it non-periodic} braid is reducible, then
243: its canonical reduction system is nonempty (see \cite{Ivanov}).
244: 
245: If $C$ is an invariant multicurve of a reducible braid $\beta$,
246: then we define the {\sl tubular braid} induced by $\beta$ and $C$ to be the
247: braid on fewer strings obtained from $\beta$ by removing from $D_n$ all the
248: disks bounded by outermost curves of $C$, and collapsing each outermost
249: curve of $C$ to a puncture point. It should be stressed that this braid
250: is only defined up to conjugacy.
251: 
252: An alternative way to look at the same defintion is the following:
253: let us consider again $\beta$ as an isotopy class of $n$ disjoint strings
254: in $D\times [0,1]$ with extremal points at the puncture points of
255: $D_n\times\{0\}$ and $D_n\times \{1\}$, such that each disk
256: $D\times \{t\}$ intersects each string exactly once.
257: Now our picture can be completed by embedded cylinders in
258: $D\times [0,1]$ which are disjoint from each other and from the strings
259: of the braid, each of which intersects each disk $D\times \{t\}$ in exactly
260: one circle, and whose boundary components are exactly the outermost curves
261: of $C$ in $D\times \{0\}$ and $D\times \{1\}$. We can interpret the solid
262: cylinders bounded by these cylinders as ``fat strings'', and the resulting
263: braid with some fat strings is exactly the tubular braid defined above.
264: 
265: The {\em interior} braids induced by $\beta$ and $C$ are the braids on
266: fewer strings induced by $\beta$ at the interior of the discs bounded by
267: the outermost curves of $C$. They can be thought of as the braids `inside'
268: the tubes of the tubular braid. Therefore, for every reducible braid
269: $\beta$, and every invariant multicurve $C$, we can decompose $\beta$ into
270: one tubular braid and some interior braids -- as many as the number of
271: outermost curves in $C$.
272: 
273: Finally, we have the notion of a {\sl \pA} element of $B_n$, for which we
274: refer to \cite{FLP} or \cite{Ivanov}. Roughly speaking, $\beta\in B_n$
275: is {\pA} if it is represented by a homeomorphism of $D_n$ which preserves
276: two transverse measured foliations on $D_n$ (called the ``stable'' and the
277: ``unstable'' foliation), while scaling the measure of the unstable one by
278: some factor $\lambda$ which is greater than 1, and the measure of the stable
279: one by $\frac{1}{\lambda}$.
280: 
281: Thurston's theorem \cite{Thclass,FLP} states that every irreducible
282: element of $B_n$ is either periodic or pseudo-Anosov.
283: 
284: We end this section with the promised example, due to S. J. Lee, that should
285: be helpful for understanding the relationship between the Nielsen-Thurston
286: decomposition and the centralizer subgroup of a braid $\beta\in B_n$.
287: This example was also found independently by N. V. Ivanov and
288: H. Hamidi-Tehrani~\cite{IvanovRecent}.
289: 
290: \begin{example}\label{E:purebraid} \rm
291:  Suppose that $n=2m$, and denote by $\sigma_i$ the standard generator of
292: $B_n$, in which the $i$th and the $(i+1)$st punctures permute their
293: positions in a clockwise sense. We define $\beta=\sigma_1 \sigma_3^2
294: \sigma_5^3 \cdots \sigma_{2m-1}^{m}$.
295: 
296: 
297: The canonical reduction system of $\beta$ consists of $m$ circles, the
298: $i$th one enclosing the punctures $2i-1$ and $2i$. The corresponding
299: tubular braid is the trivial braid of $B_m$, and the interior braids are,
300: respectively, $\sigma_1$, $\sigma_1^2$, \ldots, $\sigma_1^m$ (notice that
301: all of them are non-conjugate, since conjugate braids have the same
302: exponent sum).
303: 
304: Let $D_{(1)}, \ldots, D_{(m)}$ be the disks bounded by the above circles.
305: As we shall see, any braid that commutes with $\beta$ has to send each
306: disk $D_{(i)}$ to itself (since the interior braids are non-conjugate).
307: A generating set of the centralizer subgroup of $\beta$ is given by
308: \begin{enumerate}
309: \item[(i)] for each $i\in \{1,\ldots,m\}$, the braid $\sigma_{2i-1}$, whose
310:  support is contained in $D_{(i)}$,
311: \item[(ii)] any generating set for the pure braid group on $m$ strings $P_m$--
312: all the generators here act as the identity on
313: $D_{(1)}\cup\ldots\cup D_{(m)}$, and can be seen as a pure tubular braid on $m$
314: strings (tubes), where the $i$th tube starts and ends at $D_{(i)}$.
315: \end{enumerate}
316: It can be easily shown that, in this case, $Z(\beta)\simeq\Z^m\times P_m$.
317: The essential observation now is the following: it can be deduced by the
318: presentation given in~\cite{Bir}, that the abelianization of $P_m$ is
319: isomorphic to $\Z^{m(m-1)/2}$  (see also~\cite{Arnold}). Hence, the
320: abelianization of $Z(\beta)$ is isomorphic to $\Z^m\times \Z^{m(m-1)/2}$.
321: Therefore, at least $m+\frac{m(m-1)}{2}=\frac{m(m+1)}{2}$ generators are
322: needed for the centralizer of the braid $\beta$.
323: 
324:  The case when $n=2m+1$ is analogous. The braid proposed by S. J. Lee is:
325: $\beta=\sigma_2 \sigma_4^2 \sigma_6^3 \cdots \sigma_{2m}^m$. This time the
326: first strand is not enclosed by any curve of the canonical reduction
327: system of $\beta$, and one has: $Z(\beta)\simeq \Z^m \times P_{m+1}$.
328: Hence, in this case the minimal possible number of generators is
329: $m+\frac{m(m+1)}{2}=\frac{m(m+3)}{2}$.
330: \end{example}
331: 
332: By proving theorem~\ref{T:main}, we will show that the above examples are
333: the worst one can find.
334: 
335: % ------------------------------------------------------------------
336: 
337: \section{The periodic case}\label{S:per}
338: 
339: We have to start by describing the periodic elements of $B_n$.
340: In order to state this classification result, which is classical, we
341: need to define two braids.
342: 
343: If $D_n$ is the disk with $n$ punctures arranged regularly on the circle
344: of radius $1$, then the braid which we shall call $\delta_{(n)}$
345: is represented by a clockwise
346: movement of all punctures on this circle by an angle $\frac{2\pi}{n}$.
347: If no confusion is possible, we shall simply write $\delta$, without
348: indicating the number of strands (note that this braid is the Garside
349: element of the Birman-Ko-Lee structure of $B_n$~\cite{BKL}).
350: 
351: 
352: Similarly, if we think of $D_{n}$ as having one puncture in the centre, and
353: $n-1$ punctures arranged circularly around it, then we define
354: $\gamma_{(n)}\in B_{n}$ to
355: be the braid given by a circular movement of the $n-1$ punctures by an
356: angle of $\frac{2\pi}{n-1}$, while leaving the central puncture fixed.
357: Again, for simplicity we shall often only write $\gamma$ instead of
358: $\gamma_{(n)}$.
359: 
360: 
361: The result that classifies periodic braids, which is due to
362: Eilenberg \cite{Eilenberg} and de~Ker\'ekj\'art\'o \cite{deK}
363: (see~\cite{coko} for a modern exposition) is:
364: 
365: \begin{lemma}\label{L:per_gamma_delta}
366: Every periodic braid in $B_n$ is conjugate to a power of $\delta_{(n)}$ or
367: $\gamma_{(n)}$.
368: \end{lemma}
369: 
370: Thus we only need to consider the centralizer subgroups of $\delta_{(n)}^k$
371: and $\gamma_{(n)}^k$ for all $n, k\in \Z$, since the centralizers of
372: conjugate elements are isomorphic by an inner automorphism of $B_n$. This
373: problem has been solved by Bessis, Digne and Michel~\cite{BDM}, on the
374: wider context of complex reflexion groups. We shall explain their result in
375: the particular case of braid groups:
376: 
377: We suppose first that $\beta=\delta_{(n)}^k$ where, without loss of
378: generality, $k\geqslant 0$. Let $d=\gcd (n,k)$. For $u=1,\ldots,n$,
379: we will denote $P_u=e^{i 2\pi u/n}$ the punctures of $D_n$,
380: so $\beta=\delta_{(n)}^k$ sends $P_u$ to $P_{u+k}$ for every $u$
381: (the indices are taken modulo $n$). Hence the permutation induced
382: by $\beta$ has $d$ orbits (cycles) of length $r=\frac{n}{d}$, that we
383: denote by ${\cal C}_1,\ldots, {\cal C}_d$. See in figure~\ref{F:delta_orbits}
384: an example where $n=12$, $k=9$, $d=3$ and $r=4$: the braid $\delta_{(12)}$
385: and the three orbits of $\delta_{(12)}^9$.
386: %
387: \begin{figure}[htb]
388: \centerline{
389: \epsfbox{delta_orbits.eps}
390: }
391: \caption{The braid $\delta\in B_{12}$, and the three orbits of
392: $\delta^9$ (in black, white and grey).}
393: \label{F:delta_orbits}
394: \end{figure}
395: 
396: If $r>1$ (that is if $d<n$), consider the once punctured disc
397: $D^*=D\backslash \{0\}$, and the $r-$sheeted covering $\theta=\theta_r\co
398: D^* \rightarrow D^*$ defined by $\theta(ae^{it})=ae^{itr}=ae^{itn/d}$. The
399: orbits ${\cal C}_1,\ldots,{\cal C}_d$ are sent by $\theta$ to the points
400: $Q_1,\ldots,Q_d$, where $Q_u=e^{i2\pi u/d}$. If we consider the half-line
401: $L=\{ae^{i\pi/d},\; a\in ]0,2]\}$
402: (notice that $L$ passes between $Q_d$ and $Q_1$), then $D^*\backslash L$
403: is a fundamental region for $\theta$ (see figure~\ref{F:theta}).
404: %
405: \begin{figure}[htb]
406: \centerline{ \epsfbox{theta.eps} } \caption{The covering map $\theta=\theta_4$ associated to
407: $\delta_{(12)}^9$. } \label{F:theta}
408: \end{figure}
409: 
410: Now notice that every braid in $B_d(D^*)$ can be lifted, by $\theta^{-1}$,
411: to a braid in $B_n$ in a natural way. The resulting braid is a $\frac{2\pi
412: d}{n}$-symmetric braid, that is, it is invariant under a rotation by an
413: angle of $\frac{2\pi d}{n}$. But then it is also invariant under a
414: rotation of angle $\frac{2\pi k}{n}$; in other words, the resulting braid
415: commutes with $\beta$. Hence we have a natural homomorphism: $\theta^*\co
416: B_d(D^*) \rightarrow B_n$ whose image is contained in $Z(\beta)$. Then one
417: has
418: 
419: \begin{theorem}[\cite{BDM}]\label{T:centrdelta}
420: The natural homomorphism
421: $\theta^*\co B_d(D^*)\rightarrow Z(\delta_{(n)}^k)$ is an isomorphism.
422: \end{theorem}
423: 
424: In other words, every element in the centralizer of $\beta=\delta_{(n)}^k$
425: can be seen (via $\theta$) as a braid on a once punctured disc, that is, a
426: braid on an annulus. Notice that if $r=1$ (that is, if $k$ is a multiple
427: of $n$), then $\beta$ is a power of $\delta_{(n)}^n=\Delta_{(n)}^2$. In this
428: case $\theta$ is the identity map, and the fundamental region is the whole
429: $D_n$. Hence the centralizer of $\beta$ is the whole $B_n$, as one should
430: expect.
431: 
432: 
433: Since we are interested in minimising the set of generators, we observe
434: that if $d=n$ (thus $r=1$), then $Z(\beta)=B_n$ is generated by two
435: elements, namely Artin's $\sigma_1$ and Birman-Ko-Lee's $\delta$. In a
436: similar way, if $1<d<n$, then the braid group $B_d(D^*)$ is generated by
437: just two elements, namely $\delta_{(n)}=\theta^*(\delta_{(d)})$ and the
438: braid $\theta^*(\sigma_1)$ shown in figure \ref{F:gendelgam}(a). Notice
439: that this case contains the above one, where $\theta^*$ is the identity.
440: Finally, if $d=1$ then $B_1(D^*)$ is cyclic. Thus we have:
441: 
442: \begin{prop}\label{P:centrdelta'}
443: If $k$ and $n$ are coprime, then $Z(\delta_{(n)}^k)$ is generated by a single
444: element, namely $\delta_{(n)}$. If, by contrast, $gcd(k,n)\geqslant 2$, then
445: $Z(\delta_{(n)}^k)$ is generated by two elements: $\delta_{(n)}$ and the braid
446: $\theta^*(\sigma_1)$.
447: \end{prop}
448: 
449: \begin{figure}[htb]
450: \centerline{
451: \epsfbox{gendelgam.eps} % exported at 35pc
452: }
453: \caption{Generators $\theta_3^*(\sigma_1)$ and $\bar{\theta}_3^*(\sigma_1)$
454: of the centralisers of $\delta_{(12)}^4$ and $\gamma_{(13)}^4$.}
455: \label{F:gendelgam}
456: \end{figure}
457: 
458: It is clear that the generating set given by proposition~\ref{P:centrdelta'}
459: is indeed minimal.
460: %Following with the work in~\cite{BDM}, we will study the centralizer of
461: %$\beta=\gamma_{(n)}^k$.
462: Next we study the centralizer of $\beta=\gamma_{(n)}^k$, still following the
463: work in \cite{BDM}.  This time we call $d=\gcd (n-1,k)$, and $r=(n-1)/d$.
464: If $d<n-1$, the above map $\theta$ induces a natural homomorphism
465: $\bar{\theta}^*=\bar{\theta}_r^*\co B_d(D^*)\rightarrow B_n$, where this
466: time the central point of $D$ is considered as a puncture. Hence, the central
467: strand of every braid coming from $B_d(D^*)$ is trivial. We observe that the
468: image of this homomorphism is contained in $Z(\beta)$, and in fact one has:
469: 
470: \begin{theorem}[\cite{BDM}]\label{T:centrdelta2}
471: The natural homomorphism
472: $\bar{\theta}_r^*\co B_d(D^*)\rightarrow Z(\gamma_{(n)}^k)$ is an isomorphism.
473: \end{theorem}
474: 
475: By contrast, if $d=n-1$, then $\beta$ is a power of $\gamma^{n-1}=\Delta^2$,
476: so $\overline{\theta}_r^*=1$, $Z(\beta)=B_n$ and everything works as above.
477: Hence we have
478: 
479: \begin{prop}\label{P:centrgamma}
480: If $k$ and $n-1$ are coprime, then $Z(\gamma_{(n)}^k)$ is generated by a
481: single element, namely $\gamma_{(n)}$. If, by contrast, $gcd(k,n-1)=d
482: \geqslant 2$, then $Z(\gamma_{(n)}^k)$ is generated by two elements:
483: $\gamma_{(n)}=\bar{\theta}^*(\delta_{(d)})$ and the braid
484: $\bar{\theta}^*(\sigma_1)$.
485: \end{prop}
486: 
487: See figure \ref{F:gendelgam}(b) for an illustration of the braid
488: $\bar{\theta}^*(\sigma_1)$. We summarize all the results in this
489: section as follows:
490: 
491: \begin{cor}
492:   The centralizer of any periodic braid in $B_n$ either equals $B_n$ or is
493:   isomorphic to $B_d(D^*)$, for some $d<n$. In particular, it can be
494:   generated by at most two elements.
495: \end{cor}
496: 
497: We end with a result that will be helpful later:
498: 
499: \begin{cor}\label{C:addstring}
500: If $k$ is not a multiple of $n$, then
501: $Z(\delta_{(n)}^k)\cong Z(\gamma_{(n+1)}^k)$.
502: \end{cor}
503: 
504: \begin{proof}
505:  Both groups are isomorphic to $B_d(D^*)$, where $d=\gcd(n,k)$.
506: An actual isomorphism can be defined as follows: take any element
507: $\alpha\in Z(\delta_{(n)}^k)$, isotope it to make it
508: $\frac{2\pi k}{n}$-symmetric, and then add a
509: trivial strand based at the central point of $D_n$.
510: \end{proof}
511: 
512: 
513: % ------------------------------------------------------------------
514: 
515: \section{The pseudo-Anosov case}\label{S:pA}
516: 
517: \begin{prop}\label{P:pA}
518: If $\beta\in B_n$ is pseudo-Anosov, then the centralizer of $B_n$ is free
519: abelian and generated by two elements: some {\pA} $\alpha$ which has the
520: same stable and unstable projective measured foliation as $\beta$ (possibly
521: $\beta$ itself), and one periodic braid $\rho$ (a root of $\Delta^2$,
522: possibly $\Delta^2$ itself).
523: \end{prop}
524: 
525: We stress that the generating set promised by proposition \ref{P:pA} is
526: obviously minimal.
527: For proving this result, it is more convenient to think about the quotient
528: group $\BnD$. Since $\langle \Delta^2 \rangle$ is the center of $B_n$,
529: it is contained in the centralizer of any element. Hence the centralizer of
530: an element in $B_n$ is just the preimage of the centralizer of its
531: corresponding mapping class in $\BnD$. Thus, for the rest of this section,
532: we shall work in this quotient $\BnD$; we shall prove the following
533: result, from which proposition \ref{P:pA} will then be deduced:
534: 
535: \begin{prop}\label{P:pA'} If $\beta\in \BnD$ is \pA, then the centralizer
536: of $\beta$ is abelian, and is generated by some {\pA} $\alpha$ which has the
537: same stable and unstable projective measured foliation as $\beta$, and
538: possibly one element $\rho$ of finite order.
539: \end{prop}
540: 
541: \begin{proof}[Proof of proposition \ref{P:pA'}]
542: We start by observing that the {\pA} element $\beta$ cannot commute with
543: any reducible element $a\in \BnD$, except possibly with periodic ones --
544: thus all elements of $Z(\beta)\subset \BnD$ are either {\pA} or periodic.
545: To see this, let us assume that the canonical reduction system $C$ of $a$
546: is non-empty. Then the canonical reduction system of
547: $\beta^{-1} a\beta$ is $\beta(C)$. If it were true that $\beta^{-1} a\beta=a$,
548: then we would have $\beta(C)=C$, which is impossible since it is well known
549: that {\pA} homeomorphisms do not stabilise any curves or multicurves.
550: (This result is also a special case of corollary 7.13 of \cite{Ivanov}.)
551: 
552: Our next claim is that all pseudo-Anosov elements in $Z(\beta)$ have the
553: same stable and unstable projective measured foliations.
554: In order to prove this, we can apply Corollaries 7.15 and 8.4 of
555: \cite{Ivanov}: since the centralizer subgroup of $\beta$
556: is infinite and irreducible,
557: it follows that $Z(\beta)$ contains an infinte cyclic group as a subgroup
558: of finite index. It follows that if $a$ is any {\pA} element in the
559: centralizer of $\beta$, then there exist $k, k'\in \N$ such that
560: $a^k=\beta^{k'}$. Since all powers of a {\pA} element have the same stable
561: and unstable projective measured foliation, it follows that $a$ has
562: the same stable and unstable projective measured foliations as $\beta$, and
563: so do all {\pA} elements of $Z(\beta)\subseteq \BnD$.
564: 
565: Next we make an essential observation which only works for braid groups,
566: and does not generalize to mapping class groups of surfaces with no boundary,
567: or with more than two boundary components: all elements of
568: $\BnD$, regarded as a subgroup of the mapping class group of the $n+1$ times
569: punctured sphere, fix the puncture which came from collapsing the boundary
570: of $D_n$. Moreover, there are singular leaves of the stable and unstable
571: foliation of $\beta$ emanating from this puncture, at least one of each
572: (like for every other puncture). In the cyclic ordering around the puncture,
573: singular leaves of the stable and unstable foliation alternate. If an
574: element $a$ of $\BnD$ commutes with $\beta$, then the action of $a$ has to
575: preserve the projective stable
576: and unstable foliations. Thus in the cyclic ordering around our preferred
577: puncture, the action of $a$ can only induce a cyclic (possibly trivial)
578: permutation of the singular leaves (sending stable to stable, and unstable
579: to unstable leaves, nevertheless).
580: 
581: Now we see that an element $a$ of $Z(\beta)\subseteq \BnD$ is uniquely
582: determined by just two data: firstly the stretch factor $\lambda$ by which
583: its action on the unstable measured foliation of $\beta$ multiplies the
584: measure of that foliation. (This factor $\lambda$ equals $1$ if $a$ is
585: periodic, and belongs to the set $\R_+\setminus \{1\}$ if $a$ is \pA).
586: And secondly by the cyclic permutation of the leaves of the stable projective
587: foliation emanating from the distinguished puncture of the $n+1$ times
588: punctured sphere. Indeed, if $a$ and $b$ share both data, then $ab^{-1}$ has
589: stretch factor 1 (so it is periodic), and preserves the singular leaves.
590: Hence it is the identity in $\BnD$, so $a=b$.
591: 
592: This implies that the set of periodic elements of
593: $Z(\beta)$ forms a subgroup of $Z(\beta)$ which is either trivial or
594: isomorphic to $\Z / k\Z$, where $k$ is a divisor of the number of singular
595: leaves of the stable foliation emanating from the preferred puncture.
596: Any generator of this subgroup can play the r\^ole of our desired generator
597: $\rho$ of $Z(\beta)\subseteq \BnD$.
598: 
599: Now $\rho$ commutes with any other element in $Z(\beta)$, because their
600: commutator has stretch factor 1 and induces the trivial permutation of the
601: prongs around the preferred singularity.
602: 
603: Now notice that the stretch factor yields a multiplicative map
604: from $Z(\beta)$ to $\mathbb{R}^+$. But it is known that the set of
605: possible stretch factors for a given foliation is discrete (see~\cite{Ivanov}),
606: so the image of $Z(\beta)$ under this map must be a cyclic subgroup of $\R^+$.
607: Take an element $\alpha$ whose stretch factor $\lambda$ generates this group.
608: Then $\alpha$ is {\pA} and the stretch factor of any element in
609: $Z(\beta)$ must be a power of $\lambda$.
610: 
611: 
612: We now have that $\alpha$ and $\rho$ generate $Z(\beta)\in \BnD$, because any
613: element in $Z(\beta)$ can be multiplied by some power of $\alpha$ so as
614: to obtain an element with stretch factor $1$, i.e.~a power of $\rho$.
615: 
616: It follows that $Z(\beta)\subset \BnD$ is isomorphic to $\Z \times \Z/ k\Z$,
617: with generators $\alpha$ and $\rho$. This completes the proof of
618: proposition \ref{P:pA'}.
619: \end{proof}
620: 
621: 
622: \begin{proof}[Proof of proposition~\ref{P:pA}]
623:  By proposition~\ref{P:pA'}, $Z(\beta)\subset \BnD$ is isomorphic to $\Z
624: \times \Z/ k\Z$, with generators $\alpha$ and $\rho$. But then
625: $Z(\beta)\subset B_n$ is just the preimage of $Z(\beta)\subset \BnD$ under
626: the natural projection. Consider the subgroup
627: $\langle \rho \rangle \subset Z(\beta) \subset \BnD$. Its preimage is an
628: infinte cyclic group in $B_n$ that contains $\langle \Delta^2 \rangle$.
629: We can suppose (up to choosing an appropriate $\rho$), that the generator
630: of this cyclic group projects to $\rho$, so we call it $\rho$ as well.
631: Notice that $\rho$ is a root of $\Delta^2$, since $\Delta^2$ belongs to
632: $\langle \rho \rangle$. Then we choose an element in $B_n$ that projects to
633: $\alpha$, and we also call it $\alpha$. We must prove that in $B_n$ we still
634: have $Z(\beta) = \langle \alpha \rangle \times \langle \rho \rangle$.
635: 
636: But every element in $Z(\beta)\subset B_n$ can be written as $\alpha^k
637: \rho^l \Delta^{2m}$. Since $\Delta^2$ is a power of $\rho$, then
638: $\{\alpha,\rho\}$ is a set of generators of $Z(\beta)$. On the other hand,
639: the commutator of $\alpha$ and $\rho$ projects to the trivial mapping
640: class, hence it equals $\Delta^{2k}$ for some $k$. But the algebraic
641: number of crossings of the braid $\Delta^{2k}$ is $kn(n-1)$, while for the
642: commutator of any two elements this number is zero. Hence $k=0$, so
643: $\alpha$ and $\rho$ commute. Finally, it is well known that $B_n$ is
644: torsion-free, so $Z(\beta)$ is isomorphic to $\Z \times \Z$, as we wanted
645: to show.
646: \end{proof}
647: 
648: 
649: % ------------------------------------------------------------------
650: 
651: \section{The reducible case}\label{S:red}
652: 
653: 
654: It remains to study the centralizer of a non-periodic reducible braid
655: $\beta$. Recall that for every braid $\gamma$ one has
656: $Z(\gamma^{-1}\beta\gamma)= \gamma^{-1} Z(\beta) \gamma$.
657: Hence, in general we will not study $Z(\beta)$, but the centralizer of a
658: suitable conjugate of $\beta$, which will be easier to describe.
659: Throughout this section we shall think of the punctures of the disk $D_n$
660: as being lined up on the real axis.
661: 
662: % ............................................................
663: 
664: \subsection{Reducible braids in regular form}\label{S:reg_form}
665: 
666: As we saw in section~\ref{S:NiTh}, if $\beta$ is a non-periodic reducible
667: element, then its canonical reduction system is nonempty. We denote by
668: $R'(\beta)$ the set of outermost curves in the canonical reduction system
669: of $\beta$. It is determined by $\beta$ up to isotopy fixing the
670: punctures. Since we can study any conjugate of $\beta$, we can suppose
671: that $R'(\beta)$ is a family of disjoint circles centered at the real axis,
672: with disjoint interiors, each one enclosing more than one and less than
673: $n$ punctures.
674: 
675: Notice that there could be punctures in $D_n$ not enclosed by any circle
676: in $R'(\beta)$. In order to simplify the notations below, we define the
677: system of curves $R(\beta)$ to contain exactly the curves of $R'(\beta)$,
678: plus one circle around each such puncture of $D_n$. These new circles are
679: called the degenerate circles of $R(\beta)$. We now have that every puncture
680: in $D_n$ is enclosed by exactly one circle in $R(\beta)$.
681: 
682:  Notice that $\beta$ preserves $R(\beta)$, but it could permute the circles.
683: We will suppose that this permutation has $t$ orbits (or cycles)
684: ${\cal C}_1, \ldots, {\cal C}_t$. That is, ${\cal C}_i$ is a family
685: of circles $\{C_{i,1},\ldots,C_{i,r_i}\}\subset R(\beta)$ such that $\beta$
686: sends $C_{i,k}$ to $C_{i,k+1}$ (here the second index is taken modulo
687: $r_i$). Then one has $R(\beta)={\cal C}_1\cup \cdots \cup {\cal C}_t =$
688: $\{C_{1,1},\ldots,C_{1,r_1}\}\cup \cdots \cup \{C_{t,1},\ldots,C_{t,r_t}\}$.
689: If $m_i$ is the number of punctures inside $C_{i,k}$, for any $k$, then
690: $1\leqslant m_i <n$ and $m_1r_1+\cdots +m_tr_t=n$.
691: 
692: 
693: Let $\widehat{\beta}$ be the tubular braid induced by $\beta$ and
694: $R(\beta)$. Then $\widehat{\beta}\in B_m$, where
695: $m=r_1+\cdots+r_t$. For $i=1,\ldots, t$ and $k=1,\ldots, r_i$, let
696: $\beta_{i,k}$ be the braid induced by $\beta$ in the interior of
697: $C_{i,k}$. In other words, $\beta_{i,k}$ is the braid inside the
698: tube of $\widehat{\beta}$ which starts at $C_{i,k}$ and ends at
699: $C_{i,k+1}$. We will call the braids $\beta_{i,k}$ the {\em interior braids}
700: of $\beta$. Notice that the interior braids of each degenerate circle is
701: just a trivial braid on one string.
702: %Notice that the degenerate circles of $R(\beta)$ are
703: %irrelevant, since they do not modify the corresponding tubular braid, and
704: %their corresponding interior braids are always trivial.
705: 
706: In figure~\ref{F:tubular} we can see an example of a reducible braid
707: $\beta\in B_{13}$, and its corresponding tubular braid $\widehat{\beta}\in
708: B_6$. In this example we have three orbits, and the following data:
709: $r_1=3$, $r_2=2$, $r_3=1$; $m_1=2$, $m_2=3$, $m_3=1$,
710: $\beta_{1,1}=\sigma_1^2$, $\beta_{1,2}=\sigma_1^{-1}$, $\beta_{1,3}=1$,
711: $\beta_{2,1}=\sigma_1 \sigma_2$, $\beta_{2,2}=\sigma_1^{-1}\sigma_2$,
712: $\beta_{3,1}=1$ and $\widehat{\beta}=\sigma_3^2\sigma_2\sigma_1
713: \sigma_5^2\sigma_4$.
714: 
715: 
716: \begin{figure}[htb]
717: \centerline{
718:  \epsfbox{tubular.eps} %
719: }
720: \caption{Example of a reducible braid $\beta$, and its corresponding tubular
721: braid $\widehat{\beta}$.}\label{F:tubular}
722: \end{figure}
723: 
724: 
725: It would be desirable for $\beta$ to have its interior braids as simple
726: as possible, in order to study its centralizer. We propose the following:
727: 
728: \begin{definition}
729: Let $\beta\in B_n$ be a non-periodic reducible braid. Then $\beta$ will be
730: said to be in {\em regular form} if (using the notation introduced above) it
731: satisfies the following conditions:
732: \begin{enumerate}
733: 
734: % \item $R(\beta)$ is a family of circles, whose orbits under $\beta$ will
735: %be denoted by ${\cal C}_1=\{C_{1,1},\ldots,C_{1,r_1}\},\ldots, {\cal C}_t=
736: %\{C_{t,1},\ldots,C_{t,r_t}\}$. * Remarque de Bert: ce n'est pas une condition!
737: 
738:  \item The only non-trivial interior braids in $\beta$ are
739: $\beta_{1,r_1}, \beta_{2,r_2},\ldots, \beta_{t,r_t}$ -- we shall denote these
740: braids by $\beta_{[1]}, \beta_{[2]}, \ldots, \beta_{[t]}$.
741: 
742:  \item For $i,j\in \{1,\ldots,t\}$, if $\beta_{[i]}$ and $\beta_{[j]}$ are
743: conjugate, then $\beta_{[i]}=\beta_{[j]}$.
744: 
745: \end{enumerate}
746: \end{definition}
747: 
748: 
749:  Hence, if $\beta$ is in regular form, there is at most one non-trivial
750: interior braid for each orbit, and any two interior braids are either equal or
751: non-conjugate. Fortunately, one can conjugate every non-periodic reducible
752: braid $\beta$ to another one in regular form, as we are going to see.
753: 
754: First, consider the subgroup $B_{R(\beta)}\subset B_n$ consisting of those
755: braids preserving $R(\beta)$. For $\alpha\in B_{R(\beta)}$, we can consider
756: the tubular braid $\widehat{\alpha}$ induced by $\alpha$ and $R(\beta)$.
757: Every $\alpha\in B_{R(\beta)}$ is completely determined by $\widehat{\alpha}$
758: and its interior braids $\alpha_{i,k}$, for
759: $i=1,\ldots t$ and $k=1,\ldots, r_i$.
760: 
761:  Now consider, in $\beta$, an orbit ${\cal C}_i =
762: \{C_{i,1},\ldots,C_{i,r_i}\}$ and the interior braids
763: $\beta_{i,1},\ldots, \beta_{i,r_i}\in B_{m_i}$. We define $\alpha\in
764: B_{R(\beta)}$ as follows: $\widehat{\alpha}$ is trivial, $\alpha_{j,k}=1$
765: if $j\neq i$, and $\alpha_{i,k}=\beta_{i,k}\beta_{i,k+1}\cdots
766: \beta_{i,r_i}$. If we conjugate $\beta$ by $\alpha$, we obtain
767: $\beta'=\alpha^{-1}\beta \alpha$, which has the following properties:
768: \begin{itemize} \item $\widehat{\beta}'=\widehat{\beta}$.
769: 
770: \item $\beta'_{j,k}=\beta_{j,k}$, for $j\neq i$.
771: 
772: \item $\beta'_{i,k}= (\alpha_{i,k})^{-1} \beta_{i,k} \alpha_{i,k+1} =
773: (\beta_{i,r_i}^{-1}\cdots \beta_{i,k}^{-1})(\beta_{i,k} \cdots
774: \beta_{i,r_i})=1$, for $k\neq r_i$.
775: 
776: \item $\beta'_{i,r_i}=(\alpha_{i,r_i})^{-1} \beta_{i,r_i} \alpha_{i,1} =
777: \beta_{i,r_i}^{-1} \beta_{i,r_i}(\beta_{i,1} \cdots \beta_{i,r_i}) =
778: \beta_{i,1} \cdots \beta_{i,r_i}$.
779: 
780: \end{itemize}
781: 
782: In other words, if we conjugate $\beta$ by $\alpha$ we `transfuse' all the
783: interior braids in ${\cal C}_i$ to the last tube $C_{i,r_i}$, so
784: $\beta'_{i,r_i}$ becomes the only nontrivial interior braid in ${\cal
785: C}_i$. In figure~\ref{F:transfuse} we can see an example of such a
786: conjugation, where $\beta_{[i]}$ denotes the product $\beta_{i,1}\cdots
787: \beta_{i,r_i}$. We can now do the same for every $i=1,\ldots,t$.
788: Therefore, since we are interested in $\beta$ up to conjugacy, we can
789: suppose that $\beta_{i,k}=1$ if $k\neq r_i$ and denote
790: $\beta_{[i]}=\beta_{i,r_i}$, for every $i=1,\ldots,t$.
791: 
792: 
793: \begin{figure}[htb]
794: \centerline{
795:  \epsfbox{transfuse.eps} %
796: }
797: \caption{How to conjugate $\beta$ to simplify interior braids.}
798: \label{F:transfuse}
799: \end{figure}
800: 
801: 
802: Now suppose that some $\beta_{[i]}$ is conjugate to some $\beta_{[j]}$,
803: and let $h_{i,j}$ be a conjugating braid, that is, $h_{i,j}^{-1}
804: \beta_{[i]} h_{i,j}= \beta_{[j]}$. Consider the braid $\alpha\in
805: B_{R(\beta)}$ such that $\widehat{\alpha}=1$, $\;\alpha_{j,k}=1$ for
806: $j\neq i$ and $\alpha_{i,k}=h_{i,j}$ for every $k$. As we can see in
807: figure~\ref{F:betai_betaj}, if we conjugate $\beta$ by $\alpha$, then
808: $\beta_{[i]}$ is replaced by $\beta_{[j]}$. Therefore, we can assume that
809: for $i,j\in \{1,\ldots,t\}$, either $\beta_{[i]}=\beta_{[j]}$ or
810: $\beta_{[i]}$ and $\beta_{[j]}$ are not conjugate, and therefore we can
811: suppose that $\beta$ is in regular form.
812: 
813: 
814: \begin{figure}[htb]
815: \centerline{
816:  \epsfbox{betai_betaj.eps} %
817: }
818: \caption{How to replace $\beta_{[i]}$ by $\beta_{[j]}$ if they are conjugate.}
819: \label{F:betai_betaj}
820: \end{figure}
821: 
822: 
823:  Notice that we have chosen to put $\beta_{[i]}$ into the tube starting at
824: $C_{i,r_i}$. But we can move it to any other tube of ${\cal C}_i$ if we
825: wish, by a suitable conjugation, and later on we will need to use this.
826: Hence we define, for $i\in \{1,\ldots,t\}$ and $k\in\{1,\ldots, r_i-1\}$,
827: a braid $\mu=\mu(i,k)$ that will `move' the interior braid $\beta_{[i]}$
828: to the tube $C_{i,k}$. This braid is defined as follows: the tubular braid
829: $\widehat{\mu}$ is trivial, and the interior braids are all trivial except
830: $\mu_{i,k+1}=\mu_{i,k+2}=\cdots =\mu_{i,r_i}=\beta_{[i]}$. We can see in
831: figure~\ref{F:mu_ik} how this works.
832: 
833: \begin{figure}[htb]
834: \centerline{
835:  \epsfbox{mu_ik.eps} %
836: }
837: \caption{How to move $\beta_{[i]}$  from $C_{i,4}$ to $C_{i,2}$,
838: when $r_i=4$.}\label{F:mu_ik}
839: \end{figure}
840: 
841: % ............................................................
842: 
843: \subsection{Centralizer of a braid in regular form}
844: 
845:  We will now study the centralizer of $\beta$, assuming that $\beta$ is in
846: regular form. Recall that the only non-trivial interior braids of $\beta$
847: are denoted $\beta_{[1]},\ldots,\beta_{[t]}$, and that $\widehat{\beta}$
848: is the tubular braid associated to $\beta$ and $R(\beta)$. In this section
849: we will show that there is an exact sequence:
850: $$
851:   1 \rightarrow Z(\beta_{[1]})\times \cdots \times Z(\beta_{[t]})
852: \stackrel{g}{\longrightarrow}
853: Z(\beta) \stackrel{p}{\longrightarrow} Z_0(\widehat{\beta}) \rightarrow 1,
854: $$
855: where $Z_0(\widehat{\beta})$ is a subgroup of $Z(\widehat{\beta})$. Later
856: on we will see that this sequence splits.
857: 
858: 
859: For $i\in \{1,\ldots,t\}$, consider the centralizer $Z(\beta_{[i]})$ in
860: $B_{m_i}$. We define a map
861: $g_i\co Z(\beta_{[i]}) \rightarrow B_{R(\beta)}$ as follows:
862: given $\gamma\in Z(\beta_{[i]})$, $g_i(\gamma)$
863: is the braid $\alpha\in B_{R(\beta)}$ satisfying $\widehat{\alpha}=1$,
864: $\;\alpha_{j,k}=1$ for $j\neq i$, and $\alpha_{i,k}=\gamma$ for
865: $k=1,\ldots,r_i$. We need to show the following:
866: 
867: 
868: \begin{prop}\label{P:gi_injective}
869:  The map $g_i$ defined above is an injective homomorphism, and its image
870: is contained in $Z(\beta)$.
871: \end{prop}
872: 
873: \begin{proof}
874: % Let $\gamma,\delta\in Z(\beta_{[i]})$. Since the tubular braids
875: %associated to $g_i(\gamma)$ and $g_i(\delta)$ are trivial, then the
876: %interior braid
877: %$(g_i(\gamma)g_i(\delta))_{j,k}=(g_i(\gamma))_{j,k}(g_i(\delta))_{j,k}$,
878: %for every possible $j$ and $k$. Hence, one has
879: %$(g_i(\gamma)g_i(\delta))_{j,k}=1$ if $j\neq i$, and one also has
880: %$(g_i(\gamma)g_i(\delta))_{i,k}=\gamma\delta$. So the interior braids of
881: %$g_i(\gamma)g_i(\delta)$ coincide with the interior braids of
882: %$g_i(\gamma\delta)$. Since the tubular braids are both trivial, this
883: %implies that $g_i(\gamma)g_i(\delta)=g_i(\gamma\delta)$, so $g_i$ is a
884: %homomorphism.
885: %
886: %Moreover, since $g_i(\gamma)_{i,1}=\gamma$ for every
887: %$\gamma\in Z(\beta_{[i]})$, then $g_i(\gamma)=1$ implies $\gamma=1$,
888: %so $g_i$ is one to one.
889: The map $g_i$ is given by the diagonal homomorphism
890: $Z(\beta_{[i]})\to Z(\beta_{[i]}) \times \ldots \times Z(\beta_{[i]})$
891: ($r_i$ factors), followed by the homomorphism induced by an inclusion of
892: $r_i$ copies of an $m_i$-times punctured disk into $r_i$ disjoint subdisks
893: (each containing $m_i$ punctures) of $D_n$. By the results of
894: \cite{ParisRolfsen} we can deduce that $g_i$ is indeed an injective
895: homomorphism.
896: 
897: 
898: It remains to show that for every $\gamma\in Z(\beta_{[i]})$ one has
899: $\alpha=g_i(\gamma)\in Z(\beta)$. Since $\widehat{\alpha}$ is trivial,
900: $\widehat{\alpha^{-1}\beta \alpha}= \widehat{\alpha}^{-1}\widehat{\beta}
901: \widehat{\alpha}=\widehat{\beta}$. So we just need to show that the
902: interior braids of $\alpha^{-1} \beta \alpha$ and $\beta$ coincide.
903:   For $j\neq i$, the braids $\alpha_{j,k}$ are trivial for every $k$, so
904: $\left(\alpha^{-1} \beta \alpha\right)_{j,k}=\beta_{j,k}$.
905:  Now, for $k\neq r_i$, one has $\left(\alpha^{-1} \beta \alpha\right)_{i,k} =
906: \alpha_{i,k}^{-1} \; \beta_{i,k}\;  \alpha_{i,k+1} =
907: \gamma^{-1} 1 \gamma = 1= \beta_{i,k}$.
908:  Finally, since $\gamma$ commutes with $\beta_{[i]}$, one has
909: $\left(\alpha^{-1} \beta \alpha\right)_{i,r_i} =
910: \alpha_{i,r_i}^{-1} \; \beta_{i,r_i}\;  \alpha_{i,1} =
911: \gamma^{-1} \beta_{[i]} \gamma = \beta_{[i]}= \beta_{i,r_i}$.
912: Therefore $\alpha^{-1}\beta \alpha = \beta$, so the image of $g_i$ is
913: contained in $Z(\beta)$.
914: \end{proof}
915: 
916: 
917: \begin{prop}\label{P:g_injective}
918:  The map  $\; g\co Z(\beta_{[1]})\times \cdots \times Z(\beta_{[t]})
919: \longrightarrow Z(\beta)$ defined by
920: $g(\gamma_1,\ldots,\gamma_t)=g_1(\gamma_1)\cdots g_t(\gamma_t)$
921: is an injective homomorphism.
922: \end{prop}
923: 
924: \begin{proof} Given $\gamma\in Z(\beta_{[i]})$, the only nontrivial
925: strands in $g_i(\gamma)$ are those inside the tubes
926: $C_{i,1},\ldots, C_{i,r_i}$.
927: Hence if $i\neq j$, $\gamma\in Z(\beta_{[i]})$ and
928: $\delta\in Z(\beta_{[j]})$, then $g_i(\gamma)$ and $g_j(\delta)$ commute.
929: Since every $g_i$ is a homomorphism, this shows that $g$ is also a
930: homomorphism. But we know by the previous proposition that $g_i$ is
931: injective for $i=1,\ldots,t$. Using an argument similar to the proof of
932: proposition \ref{P:gi_injective}, one can deduce that $g$ is also injective.
933: \end{proof}
934: 
935: Now we will relate $Z(\beta)$ and $Z(\widehat{\beta})$.
936: Every braid in $Z(\beta)$ preserves the canonical reduction system of
937: $\beta$ (see~\cite{Ivanov}), so it must preserve $R(\beta)$. That is,
938: $Z(\beta)\subset B_{R(\beta)}$. Let $p\co B_{R(\beta)}\rightarrow B_m$ be
939: the homomorphism which sends $\alpha$ to $\widehat{\alpha}$, the tubular
940: braid induced by $\alpha$ and $R(\beta)$. If we take $\alpha\in Z(\beta)$
941: then $\beta=\alpha^{-1}\beta \alpha$, so
942: $p(\beta)=p(\alpha^{-1}\beta \alpha)=p(\alpha)^{-1}p(\beta) p(\alpha)$.
943: Hence $p(\alpha)$ commutes with $p(\beta)=\widehat{\beta}$. Therefore, if
944: we restrict $p$ to $Z(\beta)$ we get
945: $\;p\co Z(\beta)\rightarrow Z(\widehat{\beta})$.
946: 
947: 
948: Unfortunately, neither $p\co B_{R(\beta)}\rightarrow B_m$ nor its restriction
949: $\;p\co Z(\beta)\rightarrow Z(\widehat{\beta})$ are surjective, but we shall
950: see that the
951: elements in the image of $p$ in either case can be easily characterised by
952: the permutation they induce. Notice that $p$ induces a bijection
953: $\widetilde{p}$ from $R(\beta)$ to $\{P_1,\ldots,P_m\}$, the punctures of
954: $D_m$. We denote by $\tau$ the inverse of $\widetilde{p}$.
955: 
956: \begin{definition}
957:  Let $\eta\in B_m$, and let $\pi_{\eta}$ be the permutation induced by
958: $\eta$ on the punctures of $D_m$. We say that $\pi_{\eta}$ is
959: {\em consistent with} $R(\beta)$ if, for $i=1,\ldots,m$,
960: $\;\tau(P_i)$ and $\tau(\pi_{\eta}(P_i))$ enclose the same number of punctures.
961: \end{definition}
962: 
963: \begin{prop}\label{P:cons_Rbeta}
964: An element $\eta\in B_m$ is in the image of
965: $\;p\co B_{R(\beta)}\rightarrow B_m$ if and only if
966: $\pi_{\eta}$ is consistent with $R(\beta)$.
967: \end{prop}
968: 
969: \begin{proof}
970:  If $\eta$ is in the image of $p$, let $\alpha\in B_{R(\beta)}$ with
971: $p(\alpha)=\eta$. Then, for every $i=1,\ldots,m$, $\;\tau(P_i)$ and
972: $\tau(\pi_{\eta}(P_i))$ are the top and bottom circles of a tube
973: determined by $\alpha$. Hence they must enclose the same number of
974: punctures (the number of strands inside the tube).
975: 
976: Conversely, suppose that $\pi_{\eta}$ is consistent with $R(\beta)$. Take
977: $i\in \{1,\ldots,m\}$ and suppose that $\tau(P_i)=C_{j,k}$. Then take the
978: $i$th strand of $\eta$ and consider it as a tube, enclosing the trivial
979: braid on $m_j$ strands. Do this for every $i=1,\ldots,m$. The resulting
980: braid, $\psi(\eta)$, is well defined since $\pi_{\eta}$ is consistent with
981: $R(\beta)$, and it belongs to $B_{R(\beta)}$. Moreover,
982: $p(\psi(\eta))=\eta$ by construction.
983: \end{proof}
984: 
985: The homomorphism $\psi$ introduced in this proof will play a prominent
986: r\^ole in what follows: if $\eta\in B_m$, then $\psi(\eta)$ is the braid
987: in $B_{R(\beta)}$ whose tubular braid equals $\eta$, and whose interior
988: braids are all trivial.
989: 
990: All the elements in $B_m$ that shall be considered from now on will have
991: permutations consistent with $R(\beta)$. Hence, by abuse of notation, we
992: will identify $C_{i,k}=\widetilde{p}(C_{i,k})$ and
993: ${\cal C}_i=\widetilde{p}({\cal C}_i)$ if it does not lead to confusion.
994: 
995: 
996: We still need to characterise the elements in the image of
997: $\;p\co Z(\beta)\rightarrow Z(\widehat{\beta})$. We just know that their
998: permutations must be consistent with $R(\beta)$, but this is not
999: sufficient. Recall that the permutation induced by $\beta$ on the
1000: components of $R(\beta)$ has $t$ orbits, ${\cal C}_1, \ldots {\cal C}_t$.
1001: The key observation now is that every element $\alpha\in Z(\beta)$
1002: preserves these orbits setwise, though it could permute them. Therefore,
1003: for $i=1,\ldots,t$, one has $\alpha({\cal C}_i)={\cal C}_j$ for some $j$.
1004: In the same way, for any $\eta\in Z(\widehat{\beta})$ one has
1005: $\alpha({\cal C}_i)={\cal C}_j$ for some $j$.
1006: 
1007: \begin{lemma}\label{L:perm-conj}
1008: Let $\alpha\in Z(\beta)$. If $\alpha({\cal C}_i)={\cal C}_j$ for some
1009: $i,j\in \{1,\ldots, t\}$, then $\beta_{[i]}=\beta_{[j]}$.
1010: \end{lemma}
1011: 
1012: 
1013: \begin{proof}
1014: Since $\alpha({\cal C}_i)={\cal C}_j$, the two orbits have the same length,
1015: which we shall denote $r$; thus $r=r_i=r_j$.
1016: Now $\beta^r$ is a braid that preserves $C_{i,k}$ and $C_{j,k}$ for every $k$,
1017: and is such that $(\beta^r)_{i,k}=\beta_{[i]}$ and
1018: $(\beta^r)_{j,k}=\beta_{[j]}$.
1019: Now since $\alpha$ commutes with $\beta$, then it also commutes with
1020: $\beta^{r}$. Suppose that $\alpha$ sends $C_{i,1}$ to $C_{j,k}$. Then
1021: $\beta_{[j]}=(\beta^{r})_{j,k} = (\alpha^{-1}\beta^{r}\alpha)_{j,k} =
1022: (\alpha_{i,1})^{-1} (\beta^{r})_{i,1} \alpha_{i,1} =
1023: (\alpha_{i,1})^{-1} \beta_{[i]}\alpha_{i,1}.$
1024: Therefore $\beta_{[i]}$ and $\beta_{[j]}$ are conjugate, and since $\beta$
1025: is in regular form, $\beta_{[i]}=\beta_{[j]}$, as we wanted to prove.
1026: \end{proof}
1027: 
1028: 
1029: Lemma~\ref{L:perm-conj} imposes another condition for a braid in
1030: $Z(\widehat{\beta})$ to be in $p(Z(\beta))$:
1031: 
1032: 
1033: \begin{definition}
1034:  Let $\eta\in Z(\widehat{\beta})$. We say that $\pi_{\eta}$ is
1035: {\em consistent with} $\beta$ if it is consistent with $R(\beta)$ and,
1036: furthermore, for every $i,j\in\{1,\ldots, t\}$ such that
1037: $\eta({\cal C}_i)={\cal C}_j$, one has $\beta_{[i]}=\beta_{[j]}$.
1038: \end{definition}
1039: 
1040: 
1041: \begin{definition}
1042: $Z_0(\widehat{\beta})$ is the subgroup of $Z(\widehat{\beta})$ consisting of
1043: those elements whose permutation is consistent with $\beta$.
1044: \end{definition}
1045: 
1046: Then lemma~\ref{L:perm-conj} can be restated as follows: If
1047: $\alpha\in Z(\beta)$ then $p(\alpha)\in Z_0(\widehat{\beta})$. Moreover,
1048: we can prove the following:
1049: 
1050: 
1051: \begin{prop}\label{P:p_surjective}
1052:  The homomorphism $p\co Z(\beta)\longrightarrow Z_0(\widehat{\beta})$ is
1053: surjective.
1054: \end{prop}
1055: 
1056: 
1057: \begin{proof}
1058:   Let $\eta\in Z_0(\widehat{\beta})$. We shall construct a preimage of $\eta$
1059: under $p$ in two steps. Since $\pi_{\eta}$ is consistent with $\beta$
1060: (thus with $R(\beta)$), we can, as a first step, consider the braid
1061: $\psi(\eta)\in B_n$.
1062: We then have $p(\psi(\eta))=\eta$; but $\psi(\eta)$ does not necessarily
1063: commute with
1064: $\beta$, since the interior braids of $\psi(\eta)^{-1}\beta \psi(\eta)$
1065: could differ from those of $\beta$. Actually, since the interior braids of
1066: $\psi(\eta)$ are all trivial, conjugating $\beta$ by $\psi(\eta)$ just
1067: permutes the interior braids of $\beta$. More precisely, the braid
1068: $\psi(\eta)^{-1} \beta \psi(\eta)$ equals $\beta$, except that, for each
1069: $i\in\{1,\ldots,t\}$, it may not be the tube $C_{i,r_i}$ which contains the
1070: nontrivial interior braid $\beta_{[i]}$, but some other tube from the
1071: family $\mathcal{C}_i$. Our aim in the second step is thus to
1072: fill the tubes of $\psi(\eta)$ with more suitable interior braids, in order
1073: to obtain a braid that commutes with $\beta$.
1074: %so that it will commute with $\beta$.
1075: 
1076:  For every $i\in\{1,\ldots,t\}$, we know that $\psi(\eta)$ sends ${\cal
1077: C}_i$ to some ${\cal C}_j$. Let $k_i\in \{1,\ldots,r_i\}$ be such that
1078: $\psi(\eta)$ sends $C_{i,k_i}$ to $C_{j,r_j}$, and consider the braid
1079: $\mu(i,k_i)$ defined at the end of Subsection~\ref{S:reg_form}. If we
1080: conjugate $\beta$ by $\mu(i,k_i)$ we move $\beta_{[i]}$ from $C_{i,r_i}$
1081: to $C_{i,k_i}$. If we further conjugate by $\psi(\eta)$, then
1082: $\beta_{[i]}$ goes to $C_{j,r_j}$. But $\eta$ is consistent with $\beta$,
1083: so $\beta_{[i]}=\beta_{[j]}$. Hence, the interior braids in ${\cal C}_j$
1084: are preserved. We can do this for $i=1,\ldots,t$, so we obtain that the
1085: braid
1086: $$
1087: \left(\prod_{i=1}^t{\mu(i,k_i)}\right)\psi(\eta)
1088: $$
1089: commutes with $\beta$ and its tubular braid is $\eta$, so it is in
1090: $p^{-1}(\eta)\cap Z(\beta)$. This shows the result.
1091: \end{proof}
1092: 
1093: 
1094: We can finally bring together all the results in this section to state the
1095: following:
1096: 
1097: 
1098: \begin{theorem}\label{T:exactseq}
1099: Let $\beta\in B_n$ be a non-periodic reducible braid in regular form.
1100: Then the sequence
1101: $$
1102: 1 \rightarrow Z(\beta_{[1]})\times \cdots \times Z(\beta_{[t]})
1103: \stackrel{g}{\longrightarrow}
1104: Z(\beta) \stackrel{p}{\longrightarrow}  Z_0(\widehat{\beta}) \rightarrow 1
1105: $$
1106: is exact.
1107: \end{theorem}
1108: 
1109: 
1110: \begin{proof}
1111:   By proposition~\ref{P:g_injective} $g$ is injective, and by
1112: proposition~\ref{P:p_surjective} $p$ is surjective. It just remains to
1113: show that $\mbox{im}(g) =\ker(p)$.
1114: 
1115: By construction, every element in the image of $g$ induces a trivial
1116: tubular braid, so $\mbox{im}(g) \subset \ker(p)$. Let then
1117: $\alpha\in \ker(p)$, that is, $\widehat{\alpha}=1$. Since
1118: $\alpha\in Z(\beta)$, we have $\alpha^{-1}\beta\alpha=\beta$, and since
1119: $\beta_{i,k}=1$ for $k\neq r_i$, we must have
1120: $\alpha_{i,k}^{-1} 1 \alpha_{i,k+1}=1$, so $\alpha_{i,k}=\alpha_{i,k+1}$ for
1121: $k=1,\ldots,r_i-1$. Hence $\alpha_{i,1}=\alpha_{i,2}=\cdots = \alpha_{i,r_i}$
1122: for every $i$. Moreover,  we have
1123: $\beta_{[i]}=\beta_{i,r_i}=\alpha_{i,r_i}^{-1} \beta_{i,r_i} \alpha_{i,1}=
1124: \alpha_{i,1}^{-1} \beta_{[i]} \alpha_{i,1}$, so
1125: $\alpha_{i,1}\in Z(\beta_{[i]})$. Therefore,
1126: $\alpha=g_1(\alpha_{1,1})g_2(\alpha_{2,1})\cdots g_t(\alpha_{t,1})=
1127: g(\alpha_{1,1},\alpha_{2,1},\ldots,\alpha_{t,1})$. That is,
1128: $\ker(p)\subset \mbox{im}(g)$.
1129: \end{proof}
1130: 
1131: % ............................................................
1132: 
1133: \subsection{Finding a section for $\mathbf{p}$}
1134: 
1135:  In this subsection we will prove that the exact sequence of
1136: theorem~\ref{T:exactseq} splits. We recall that $\widehat{\beta}$ is
1137: obtained from $\beta$ by collapsing the disks bounded by outermost curves
1138: in the canonical reduction system of $\beta$ to single punctures. In
1139: particular, the canonical reduction system of $\widehat{\beta}$ must be
1140: empty. Hence, $\widehat{\beta}$ is either periodic or pseudo-Anosov. We will
1141: distinguish these two cases, to define a multiplicative section for
1142: $p$, but first we will show an easy particular case. Recall that a braid
1143: is pure if it induces the trivial permutation of its base points.
1144: 
1145: \begin{prop}\label{P:section_trivial}
1146:  If $\widehat{\beta}$ is pure, there is a homomorphism
1147: $\:h\co Z_0(\widehat{\beta}) \rightarrow Z(\beta)$ such that $p\circ h=1$.
1148: \end{prop}
1149: 
1150: \begin{proof}
1151:   We shall prove that in this case, the homomorphism $\psi$ constructed
1152: in the proof of proposition \ref{P:cons_Rbeta} is such a section.
1153:   Let $\eta\in Z_0(\widehat{\beta})$. Since $\widehat{\beta}$ is pure,
1154: ${\cal C}_i=\{C_{i,1}\}$ for all $i$. Hence, if $\eta$ sends ${\cal C}_i$ to
1155: ${\cal C}_j$ then it sends the tube $C_{i,1}$ (containing $\beta_{[i]}$)
1156: to the tube $C_{j,1}$ (containing $\beta_{[j]}=\beta_{[i]}$, since $\beta$
1157: is in regular form). Therefore, filling every tube in $\eta$ with the
1158: trivial braid, that is, defining $h(\eta)=\psi(\eta)$, yields indeed an
1159: element of $Z(\beta)$.
1160: \end{proof}
1161: 
1162: Next we study the general case, depending whether $\widehat{\beta}$ is
1163: periodic or pseudo-Anosov.
1164: 
1165: \begin{prop}\label{P:section_periodic}
1166: If $\widehat{\beta}$ is periodic, there is a homomorphism
1167: $\:h\co Z_0(\widehat{\beta}) \rightarrow Z(\beta)$ such that $p\circ h=1$.
1168: \end{prop}
1169: 
1170: \begin{proof}
1171: Recall that we are studying $\beta$ up to conjugacy. This implies that we
1172: can also study $\widehat{\beta}$ up to conjugacy since, for every
1173: $\xi\in B_m$, if we conjugate $\beta$ by $\psi(\xi)$ we are conjugating
1174: $\widehat{\beta}$ by $\xi$. Moreover, after conjugating by $\psi(\xi)$,
1175: $\beta$ continues to be in regular form (up to renaming the circles in
1176: $R(\beta)$). Therefore we can suppose, up to conjugacy, that
1177: $\widehat{\beta}$ is a rigid rotation of the disc, that is, a power of
1178: $\delta_{(m)}$ or $\gamma_{(m)}$.
1179: 
1180: Suppose first that $\widehat{\beta}=\delta^k_{(m)}$ for some $k$. We can
1181: suppose that $k$ is not a multiple of $m$, since in that case
1182: $\widehat\beta$ would be a power of $\Delta^2_{(m)}$, thus it would be pure,
1183: and this case has already been studied in proposition~\ref{P:section_trivial}.
1184: Recall the analysis of periodic braids in section~\ref{S:per}: the base
1185: points $Q_1,\ldots, Q_m$ of $\widehat{\beta}$ will be evenly distributed
1186: along a circle of radius 1 around 0. Let $d=\gcd(m,k)<m$ and $r=m/d$.
1187: Then $\widehat\beta $ sends $Q_i$ to $Q_{i+k}$, and there are $d$ orbits
1188: ${\cal C}_1,\ldots,{\cal C}_d$ of length $r$. The orbit ${\cal C}_i$ will
1189: contain the points $Q_u$ where $u\equiv i$ (mod $d$). Since we can choose
1190: which tubes of $\beta $ contain the interior braids, we will suppose that
1191: these are the tubes starting at $Q_{m-d+1}, Q_{m-d+2},\ldots, Q_{m}$, that
1192: is, the last $d$ points of $D_m$.
1193: 
1194: We will consider now some line segments in $D$ which separate the points
1195: $Q_1,\ldots,Q_m$ into $r$ sets of $d$ points. Let $L$ be the line segment
1196: joining the origin with the border of $D$, passing between the points
1197: $Q_{m-d}$ and $Q_{m-d+1}$, and let $L'$ be the segment passing between
1198: $Q_m$ and $Q_1$. Notice that $L$ and $L'$ determine a sector which contains
1199: the points $Q_{m-d+1},\ldots, Q_m$, corresponding to the tubes of $\beta$
1200: with nontrivial interior braids. Let $\phi: \C \rightarrow \C $ be the
1201: rotation around the origin by an angle of $2\pi k/m$ (the angle induced
1202: by $\widehat\beta$), and denote $L_i=\phi^i(L)$. Since $\gcd(m,k)=d$, the
1203: segments $L_0,\ldots,L_{r-1}$ divide $D$ into $m/d=r$ sectors, each
1204: one of angle $2\pi/r$ and containing the points $Q_{id+1},\ldots, Q_{id+d}$
1205: for some $i$. Take the smallest integer $e>0$ such that $\phi^e (L)=L'$.
1206: Then one has $L_0=L$ and $L_e=L'$. We are interested in the union of segments
1207: ${\cal L}=L_1 \cup L_2\cup \cdots \cup L_e$ (see figure~\ref{F:L0_Le} for
1208: an example).
1209: %
1210: \begin{figure}[htb]
1211: \centerline{ \epsfbox{L0_Le.eps} }
1212: \caption{The segments $L$, $L'$, and the union of
1213: segments ${\cal L}$, for $\widehat\beta=\delta^6\in B_{15}$.} \label{F:L0_Le}
1214: \end{figure}
1215: 
1216: 
1217: Let then  $\eta\in Z_0(\widehat{\beta})$.  In order to define $h(\eta)$, it
1218: suffices to define its interior braids. This is done as follows: recall that,
1219: since $\eta$ commutes with $\widehat{\beta}$, it can be isotoped to a
1220: symmetric braid (with respect to the rotation $\phi $), so we take a
1221: symmetric representative of $\eta$. For every base point $Q_i$ of
1222: $\widehat{\beta}$ (corresponding to a circle $C_{j,u}$), consider the strand
1223: of $\eta$ starting at $Q_i$ (the $i$th strand of $\eta$). Then we define the
1224: interior braid $h(\eta)_{j,u}=(\beta_{[j]})^{L(\eta,i)}$, where
1225: $L(\eta,i)\in \Z$ is the algebraic number of times that the $i$th strand of
1226: $\eta$ crosses ${\cal L}$. This is well defined by theorem~\ref{T:centrdelta}
1227: (if you take two distinct representatives of $\eta$ as a symmetric braid,
1228: they are isotopic through symmetric braids, so the strands never touch the
1229: origin and the intersection number $L(\eta,i)$ is preserved).
1230: 
1231: In other words, we define $h(\eta)$ as follows: we start with trivial
1232: interior braids, and we follow the movement of the strands of $\eta$.
1233: Each time a strand crosses a segment of ${\cal L}$ in the positive sense,
1234: we multiply its interior braid by $\beta_{[j]}$ (where $j$ is the index
1235: of the orbit ${\cal C}_j$ of that strand). And every time a strand crosses
1236: ${\cal L}$ in the negative sense, we multiply its interior braid by
1237: $\beta_{[j]}^{-1}$.
1238: 
1239: We have thus defined a map
1240: $h\co Z_0(\widehat{\beta})\rightarrow B_{R(\beta)}$. To show that
1241: $h$ is a homomorphism, it suffices to see that the interior braids of
1242: $\eta \xi$ are the product of those of $\eta$ and $\xi$, for
1243: $\eta,\xi\in Z_0(\widehat{\beta})$. Suppose that the $i$th strand of
1244: $\eta $ goes from $Q_i$ (corresponding to $C_{j,u}$) to $Q_{i'}$
1245: (corresponding to $C_{j',u'}$). Hence $\eta $ sends ${\cal C}_j$ to
1246: ${\cal C}_{j'}$, and since $\eta\in Z_0(\widehat{\beta})$, it follows
1247: that $\beta_{[j]}=\beta_{[j']}$. One also has, by definition,
1248: $L(\eta\xi ,i)=L(\eta ,i)+ L(\xi ,i')$. Therefore $(\eta \xi)_{j,u} =
1249: (\beta_{[j]})^{L(\eta\xi,i)} =
1250: (\beta_{[j]})^{L(\eta,i)} (\beta_{[j]})^{L(\xi,i')} =
1251: \eta_{j,u}\xi_{j',u'}$, so  $h$ is a homomorphism.
1252: 
1253: We must finally show that, with this definition, $h(\eta)\in Z(\beta )$,
1254: for every $\eta \in Z_0(\widehat\beta )$. We will define first some special
1255: braids. For every $i,j\in \{1,\ldots,d\}$ such that  $i<j$ and
1256: $\beta_{[i]}=\beta_{[j]}$, define the symmetric braid $S_{i,j} =
1257: S_{j,i}=\theta_r^*(\sigma_i \cdots \sigma_{j-2} \sigma_{j-1} \sigma _{j-2}
1258: \cdots \sigma_i)$  (see figure~\ref{F:gendelgam} in section~\ref{S:per} to
1259: recall the definition of $\theta_r^*$, and figure~\ref{F:S_13} here for an
1260: example). The braid $S_{i,j}$ commutes with $\widehat\beta $ (since it is
1261: symmetric), and it permutes the orbits ${\cal C}_i$ and ${\cal C}_j$,
1262: preserving the others. Hence $S_{i,j}\in Z_0(\widehat\beta)$. Moreover, its
1263: strands do not cross ${\cal L}$, so by definition of $h$ one has
1264: $h(S_{i,j})=\psi(S_{i,j})$ (the interior braids are trivial).
1265: %
1266: \begin{figure}[htb]
1267: \centerline{ \epsfbox{S_13.eps} }
1268: \caption{The braid $S_{1,3}$, for
1269: $\widehat\beta=\delta^6\in B_{15}$ (assuming that $\beta_{[1]}=\beta_{[3]}$).}
1270: \label{F:S_13}
1271: \end{figure}
1272: 
1273: But $h(S_{i,j})$ commutes with $\beta$, since the only tubes it permutes
1274: are those of the orbits ${\cal C}_i$ and ${\cal C}_j$; among these tubes,
1275: the only two with non-trivial interior braids are exchanged, and their
1276: corresponding interior braids are equal ($\beta_{[i]}=\beta_{[j]}$). Hence
1277: the interior braids of $\beta$ are preserved by $\psi(S_{i,j})=h(S_{i,j})$,
1278: so $h(S_{i,j})\in Z(\beta)$.
1279: 
1280: Take then an arbitrary $\eta\in Z_0(\widehat\beta )$. We must show that
1281: $h(\eta) \in Z(\beta)$. Suppose that $\eta$ sends ${\cal C}_i$ to
1282: ${\cal C}_j$ for some $i,j$. Then $\beta_{[i]}=\beta_{[j]}$, so $S_{i,j}$
1283: is defined, and the braid $\eta S_{i,j}$ preserves the orbit ${\cal C}_i$.
1284: We can continue this way, until we obtain a braid
1285: $\eta S_{i_1,j_1} \cdots S_{i_k,j_k}$ that commutes with $\widehat\beta $
1286: and preserves every orbit ${\cal C}_i$, for $i=1,\ldots,d$.  Since
1287: $h(S_{i,j})\in Z(\beta )$ for every $i,j$, and $h$ is a homomorphism, in
1288: order to show that $h(\eta)\in Z(\beta)$ it suffices to show that $h(\eta
1289: S_{i_1,j_1}\cdots S_{i_k,j_k})\in Z(\beta)$. Therefore, we can suppose that
1290: $\eta$ preserves every orbit ${\cal C}_i$.
1291: 
1292: Denote $\alpha =h(\eta)$. We need to show that the interior braids of $\alpha^{-1} \beta
1293: \alpha $ coincide with those of $\beta$. Since $\eta$ preserves all orbits, we will consider
1294: just the tubes of ${\cal C}_1$, the other ones being analogous. Suppose that $\alpha$ sends
1295: the circle $C_{1,u}$ to $C_{1,r}$. Then it must send $C_{1,v}$ to $C_{1,v-u}$ for every $v$
1296: (the indices are taken modulo $r$).
1297: 
1298: We will identify the points $Q_1,\ldots,Q_m$ with their corresponding circles $C_{i,v}$. For
1299: every $v=1,\ldots,r$, let $b_v$ be the strand of $\eta $ starting at $C_{1,v}$. Since $\eta$
1300: is symmetric, we have $\phi (b_v)=b_{v+1}$. Suppose that $b_v$ crosses $t$ times the segment
1301: $L_i$, where $i\in\{0,\ldots,r-1\}$. Then $b_{v+1}$ will cross $t$ times the segment
1302: $\phi(L_i)=L_{i+1}$. Therefore, if $b_v$ crosses $l$ times ${\cal L}$, and if it crosses
1303: $l_0$ times $L_0$ and $l_e$ times $L_e$, then $b_{v+1}$ crosses $l-l_e+l_0$ times ${\cal
1304: L}$.
1305: 
1306: If $v\neq r$ and $v\neq u$, then $b_v$ neither starts nor ends at $C_{1,r}$. Then it crosses
1307: $L_0$ and $L_e$ the same number of times. Hence, $b_v$ and $b_{v+1}$ cross ${\cal L}$ the
1308: same number of times, say $l$. Therefore, if $v\neq r,u$, one has $(\alpha^{-1} \beta
1309: \alpha)_{1,v-u}=(\alpha_{1,v})^{-1} \beta_{1,v} \alpha_{1,v+1} = \beta_{[1]}^{-l} \: 1 \:
1310: \beta_{[1]}^l= 1 = \beta_{1,v-u}$.
1311: 
1312: If $u=v=r$, then $b_v$ starts at ends at $C_{1,r}$. Hence, as above, it crosses $L_0$ and
1313: $L_e$ the same number of times, so $b_v=b_r$ and $b_{v+1}=b_1$ cross ${\cal L}$ the same
1314: number of times, say $l$. We then have $(\alpha^{-1} \beta \alpha)_{1,v-u}=(\alpha^{-1}
1315: \beta \alpha)_{1,r} = (\alpha_{1,r})^{-1} \beta_{1,r} \alpha_{1,1}= \beta_{[1]}^{-l}
1316: \beta_{[1]} \beta_{[1]}^l = \beta_{[1]} = \beta_{1,r} = \beta_{1,v-u}$. Hence, if $u=r$, we
1317: have already seen all the possible cases. We will then suppose that $u\neq r$.
1318: 
1319: If $v=r$, then $b_v$ starts (but does not end) at $C_{1,r}$. Hence, it crosses $L_e$ one
1320: more time (in the positive sense) than it crosses $L_0$. Therefore, if $b_v=b_r$ crosses $l$
1321: times ${\cal L}$, then $b_{v+1}=b_1$ crosses it $l-1$ times. One has: $(\alpha^{-1} \beta
1322: \alpha)_{1,v-u}=(\alpha^{-1} \beta \alpha)_{1,r-u} = (\alpha_{1,r})^{-1} \beta_{1,r}
1323: \alpha_{1,1}= \beta_{[1]}^{-l} \beta_{[1]} \beta_{[1]}^{l-1} = 1 =
1324: \beta_{1,r-u}=\beta_{1,v-u}$.
1325: 
1326: Finally, if $v=u$ then $b_v$ ends (but does not start) at $C_{1,r}$. In this case, it
1327: crosses $L_e$ one less time (in the positive sense) than it crosses $L_0$. Hence, if
1328: $b_v=b_u$ crosses $l$ times ${\cal L}$, then $b_{v+1}=b_{u+1}$ crosses it $l+1$ times. One
1329: then has: $(\alpha^{-1} \beta \alpha)_{1,v-u}=(\alpha^{-1} \beta \alpha)_{1,r} =
1330: (\alpha_{1,u})^{-1} \beta_{1,u} \alpha_{1,u+1}= \beta_{[1]}^{-l} \: 1 \: \beta_{[1]}^{l+1} =
1331: \beta_{[1]} = \beta_{1,r} = \beta_{1,v-u}$.
1332: 
1333: Therefore, in every possible case we have  $(\alpha^{-1} \beta \alpha)_{1,v-u}=
1334: \beta_{1,v-u}$, for every $v$. This means that the interior braids of $(\alpha^{-1} \beta
1335: \alpha)$ and of $\beta $ coincide, that is, $\alpha =h(\eta)$ commutes with $\beta $, as we
1336: wanted to show.
1337: 
1338: This completes the proof of proposition \ref{P:section_periodic} in the case
1339: $\widehat{\beta}=\delta^k_{(m)}$, and it only remains to deal with the case when
1340: $\widehat\beta=\gamma_{(m)}^k$. As above, we can suppose that $k$ is not a multiple of
1341: $m-1$, since in that case $\widehat\beta$ would be pure, and this case has already been
1342: treated in proposition~\ref{P:section_trivial}. Hence, the only fixed point in the
1343: permutation induced by $\widehat\beta $ is the origin. Therefore, every $\eta$ commuting
1344: with $\widehat\beta $ must fix the origin. This means that, for every $\eta\in
1345: Z_0(\widehat\beta )$, we can fill its central tube with the trivial braid, and the other
1346: tubes in the same way as above (defining ${\cal L}$, and counting the number of times each
1347: strand crosses ${\cal L}$). This defines a homomorphism $h: Z_0(\widehat\beta)\rightarrow
1348: Z(\beta)$ which is a section of $p$. The proof is the same as above.
1349: \end{proof}
1350: 
1351: 
1352: It remains to study the case when $\widehat{\beta}$ is \pA.
1353: 
1354: \begin{prop}\label{P:section_pA}
1355: If $\widehat{\beta}$ is pseudo-Anosov, then there is a homomorphism
1356: $h\co Z_0(\widehat{\beta})\rightarrow Z(\beta)$ such that $p\circ h=1$.
1357: \end{prop}
1358: 
1359: \begin{proof}
1360: In this case, we know that $Z(\widehat{\beta})$ is a free abelian group of rank 2, generated
1361: by a pseudo-Anosov and a periodic braid. Hence, $Z_0(\widehat{\beta})$ is an abelian group
1362: of rank one or two. Notice that $\Delta_{(m)}^2\in Z_0(\widehat{\beta})$, because this braid
1363: commutes with $\widehat{\beta}$ and because $\pi_{\Delta^2}$ is trivial, and thus consistent
1364: with $\beta$. Hence $Z_0(\widehat{\beta})$ contains at least one periodic element. On the
1365: other hand, $\widehat{\beta}$ belongs itself to $Z_0(\widehat{\beta})$, since
1366: $\pi_{\widehat{\beta}}$ is clearly consistent with $\beta$. Hence in $Z_0(\widehat{\beta})$
1367: there are also pseudo-Anosov braids. Since all powers of a periodic braid are periodic, and
1368: all powers of a pseudo-Anosov braid are pseudo-Anosov, it follows that
1369: $Z_0(\widehat{\beta})$ has in fact rank two.  More precisely, $Z_0(\widehat{\beta}) =
1370: \langle \eta\rangle \times \langle \rho \rangle$, where $\eta$ is pseudo-Anosov and $\rho$
1371: is periodic. In particular, we have
1372: $\widehat\beta \in \langle \eta\rangle \times \langle \rho \rangle$, and the three braids
1373: $\widehat{\beta}$, $\eta$ and $\rho$ are mutually commuting.
1374: 
1375: Our aim is to define two commuting braids $h(\rho)$ and $h(\eta)$ in
1376: $Z(\beta)$ which are preimages of $\rho$ respectively $\eta$ under $p$.
1377: The definition of $h(\rho)$ is very simple: we take an arbitrary preimage
1378: of $\rho$ under $p$ -- this is possible since $p$ is surjective by
1379: proposition \ref{P:p_surjective}. It remains to construct $h(\eta)$.
1380: 
1381: \begin{lemma}\label{L:pAlem1}
1382: Suppose $\alpha\in B_{R(\beta)}$, that is, the braid $\alpha$ preserves the set of outermost
1383: curves in the canonical reduction system of $\beta$. Suppose also that $\mu, \nu \in
1384: Z(\widehat{\alpha})$. Suppose that $\iota_\mu\in B_{R(\beta)}$ is a braid with trivial tubes
1385: (i.e.\ $\widehat{\iota}_\mu=1$) such that $\psi(\mu)\cdot\iota_\mu \in Z(\alpha)$. Finally,
1386: suppose that $\mu$ and $\nu$ induce the same permutation. Then we have as well that
1387: $\psi(\nu)\cdot\iota_\mu \in Z(\alpha)$.
1388: \end{lemma}
1389: 
1390: In other words, if two tubular braids commute with $\widehat{\alpha}$, if they induce the
1391: same permutation, and if some ``filling'' of one of them commutes even with $\alpha$, then
1392: the same filling of the other will also commute with $\alpha$.
1393: 
1394: \begin{proof}[Proof of lemma \ref{L:pAlem1}] Conjugating $\alpha$ by
1395: $\psi(\nu)\cdot\iota_\mu \in Z(\alpha)$ yields a certain braid $\alpha'$; we have to check
1396: that $\alpha'=\alpha$. Firstly, we have an equality of tubular braids
1397: $\widehat{\alpha'}=\widehat{\alpha}$, because $\nu$, the tubular braid of
1398: $\psi(\nu)\cdot\iota_\mu$, commutes with $\widehat{\alpha}$. Moreover, since $\mu$ and $\nu$
1399: induce the same permutations, we have for $i=1,\ldots, m$ that the $i$th tube of $\alpha'$
1400: contains the same braid as the $i$th tube of
1401: $(\psi(\mu)\cdot\iota_\mu)^{-1}\cdot\alpha\cdot(\psi(\mu)\cdot\iota_\mu)$. Since
1402: $\psi(\mu)\cdot\iota_\mu$ commutes with $\alpha$, this is in turn the same as the $i$th tube
1403: of $\alpha$. In summary, $\alpha$ and $\alpha'$ have the same tubular braids, and
1404: corresponding tubes contain the same interior braids, which implies that $\alpha=\alpha'$.
1405: \end{proof}
1406: 
1407: Next we have to think in detail about the orbit structure of
1408: $\widehat{\beta}$. Let us choose arbitrarily a puncture $P$ of the disk
1409: $D_m$ (on which $\widehat{\beta}$ acts), and let $O(\widehat{\beta},\rho)$ be
1410: the orbit of that puncture under the action of the subgroup
1411: $\langle \widehat{\beta}\rangle \times \langle \rho \rangle$ of
1412: $Z_0(\widehat{\beta})$. Let $O(\widehat{\beta},\rho,\eta)$ be the orbit
1413: of $P$ under the action of the group
1414: $\langle\rho\rangle \times \langle \eta \rangle$ (note that this
1415: group is also isomorphic to $\Z^2$, and contains $\widehat{\beta}$).
1416: 
1417: We are going to suppose without loss of generality that
1418: $O(\widehat{\beta},\rho,\eta)$ contains {\it all} punctures of $D_m$,
1419: and we shall specify how the tubes of $\eta$ corresponding to this orbit
1420: shall be filled -- indeed, if there are other orbits, then these can be
1421: treated in same way, independently.
1422: 
1423: {\bf Special case: } Let us start by considering the simpler special case that
1424: $O(\widehat{\beta},\rho,\eta) = O(\widehat{\beta},\rho)$, i.e.\ that the
1425: action of $\eta$ preserves the $(\widehat{\beta},\rho)$-orbit. In this case
1426: we have
1427: %
1428: \begin{lemma}\label{L:pAlem2}
1429: There exist integers $k$ and $l$ such that $\eta$ and
1430: $\widehat{\beta}^k\cdot \rho^l$ induce the same permutations on
1431: $O(\widehat{\beta},\rho)$.
1432: \end{lemma}
1433: %
1434: \begin{proof}[Proof of lemma \ref{L:pAlem2}] One can choose $k$ and $l$ such
1435: that $\widehat{\beta}^k \rho^l(P) = \eta(P)$, simply because $\eta(P)$ is in
1436: the orbit of $P$ under the action of $\widehat{\beta}$ and $\rho$. Now if
1437: $P'$ is another point in the orbit, then
1438: $P'=\widehat{\beta}^\kappa \rho^\lambda(P)$ for some $\kappa, \lambda \in \Z$.
1439: Since $\widehat{\beta}, \rho$, and $\eta$ are mutually commuting, we get
1440: $\widehat{\beta}^k \rho^l(P') =
1441: \widehat{\beta}^k \rho^l(\widehat{\beta}^\kappa \rho^\lambda(P)) =
1442: \widehat{\beta}^\kappa\rho^\lambda(\widehat{\beta}^k \rho^l(P))=
1443: \widehat{\beta}^\kappa\rho^\lambda(\eta(P))=
1444: \eta(\widehat{\beta}^\kappa \rho^\lambda(P)) = \eta(P')$.
1445: \end{proof}
1446: 
1447: We already know a nice preimage of $\widehat{\beta}^k\cdot \rho^l$ under $p$:
1448: the braid $\beta^k\cdot h(\rho)^l$ belongs to $Z(\beta)$, because both
1449: $\beta$ and $h(\rho)$ do. This braid can be reexpressed as
1450: $\psi(\widehat{\beta}^k \rho^l)\cdot\iota$, where $\iota$ is
1451: some braid in $B_{R(\beta)}$ with $\widehat{\iota}=1$. (That is, we define
1452: $\iota$ to consist of the interior braids of the tubes of
1453: $\beta^k\cdot h(\rho)^l$).
1454: 
1455: Now we define our filling of $\eta$ by $h(\eta):=\psi(\eta)\cdot\iota$. By
1456: lemma \ref{L:pAlem1} we have that indeed $\psi(\eta)\cdot\iota\in Z(\beta)$.
1457: In order to see that $\psi(\eta)\cdot\iota$ lies also in the centralizer of
1458: $h(\rho)$ one can use a very similar argument. Explicitly, both $\eta$
1459: and $\widehat{\beta}^k\rho^l$ lie in the centralizer of $\rho$, and they
1460: induce the same permutation of the punctures. Moreover,
1461: $\psi(\widehat{\beta}^k \rho^l) \cdot\iota = \beta^k h(\rho)^l\in Z(h(\rho))$.
1462: By lemma \ref{L:pAlem1} we conclude again that
1463: $\psi(\eta)\cdot\iota \in Z(h(\rho))$, also.
1464: 
1465: {\bf General case: } In the case where $\eta$ does not preserve
1466: $O(\widehat{\beta},\rho)$, the strategy is to work not with $\beta$ itself
1467: but with a certain conjugate of $\beta$. The details are as follows.
1468: We have a finite number of disjoint $(\widehat{\beta},\rho)$-orbits in
1469: $O(\widehat{\beta},\rho,\eta)$, and since $\eta$ commutes with
1470: $\widehat{\beta}$ and $\rho$, the action of $\eta$ permutes these
1471: orbits cyclically:
1472: $$
1473: O(\widehat{\beta},\rho)\stackrel{\eta-{\rm action}}{\longrightarrow}
1474: \eta(O(\widehat{\beta},\rho)) \stackrel{\eta-{\rm action}}{\longrightarrow}
1475: \ldots \stackrel{\eta-{\rm action}}{\longrightarrow}
1476: \eta^s(O(\widehat{\beta},\rho))=O(\widehat{\beta},\rho).
1477: $$
1478: Let us denote $\widehat{\beta}_*, \rho_*$ and $\eta^s_*$ the braids
1479: which are obtained from $\widehat{\beta}, \rho$ and $\eta^s$ by retaining
1480: only the strands corresponding to $O(\widehat{\beta},\rho)$, and forgetting
1481: the strands corresponding to all other $(\widehat{\beta},\rho)$-orbits.
1482: Similarly, let $\beta_*$ be the corresponding restriction of $\beta$.
1483: Our first aim is to fill the tubes of $\rho_*$ and $\eta^s_*$ so as to obtain
1484: commuting braids in $Z(\beta_*)$. This can be done as in the ``special case'':
1485: for $\rho_*$ we choose any filling in $Z(\beta_*)$, and for $\eta^s_*$ there
1486: exists a braid $\iota_*$ with trivial tubes such that
1487: $\psi(\eta_*^s) \cdot \iota_*$ commutes with $\beta_*$ and the filling of
1488: $\rho_*$.
1489: 
1490: We have succeeded in finding a filling of certain tubes of $\eta^s$, but
1491: not yet of $\eta$ itself. Also, we have so far only filled the tubes
1492: of $\rho$ which correspond to $O(\widehat{\beta},\rho)$, but not yet
1493: those in the $\eta$-translates of this orbit. We first notice that the
1494: $\eta$-action sends $\widehat{\beta}$-orbits to $\widehat{\beta}$-orbits,
1495: and that in each $\widehat{\beta}$-orbit there is exactly one tube whose
1496: preimage in $\beta$ contains a nontrivial braid (the same for all
1497: $\widehat{\beta}$-orbits), and all other tubes are filled with
1498: a trivial braid. Thus, up to cyclically changing the numbering of the
1499: orbits of each tube of $\widehat{\beta}$, we may assume that the
1500: $\eta$-action sends each tube of $\widehat{\beta}$ in
1501: $O(\widehat{\beta},\rho)$ to a tube of $\widehat{\beta}$ in
1502: $\eta(O(\widehat{\beta},\rho))$ which is filled with the nontrivial braid
1503: if and only if the tube of $O(\widehat{\beta},\rho)$ is. Similarly, for
1504: $i=1,\ldots,s-1$ we may assume that $\eta^i$ sends each $\widehat{\beta}$-tube
1505: in $O(\widehat{\beta},\rho)$ to a $\widehat{\beta}$-tube in
1506: $\eta^i(O(\widehat{\beta},\rho))$ which has the same filling in $\beta$.
1507: 
1508: Now we can use the same property as a construction recipe for $h(\rho)$:
1509: a tube of $\rho$ in $\eta^i(O(\widehat{\beta},\rho))$ (where $i=1,\ldots,s$)
1510: is filled in the same way as its preimage under $\eta^i$. With this
1511: definition, $h(\rho)$ commutes with $\beta$. Finally we are ready to
1512: define $h(\eta)$: we take the braid $\psi(\eta)$, but modify the braids
1513: in the tubes that terminate at positions corresponding to
1514: $O(\widehat{\beta},\rho)$ by multiplying them on the right by $\iota_*$. In
1515: other words, the braid $h(\eta)$ is obtained from $\eta$ as follows: we fill
1516: those tubes of $\eta$ which connect points in $\eta^i(O(\widehat{\beta},\rho))$
1517: to points in $\eta^{i+1}(O(\widehat{\beta},\rho))$ (with $i=0,\ldots,s-2$)
1518: with the trivial braid, and we fill the tubes that start in
1519: $\eta^{s-1}(O(\widehat{\beta},\rho))$ and terminate in
1520: $O(\widehat{\beta},\rho)$ with the interior braids of $\iota_*$.
1521: By construction, this braid $h(\eta)$ commutes with both $\beta$ and $h(\rho)$.
1522: This concludes the proof of proposition \ref{P:section_pA}.
1523: \end{proof}
1524: 
1525: 
1526: \begin{cor}
1527: Suppose that $\beta$ is a non-periodic reducible braid in regular form. Then
1528: the exact sequence
1529: $$
1530: 1\longrightarrow Z(\beta_{[1]})\times \cdots \times Z(\beta_{[t]})
1531: \stackrel{g}{\longrightarrow} Z(\beta)  \stackrel{p}{\longrightarrow}
1532: Z_0(\widehat{\beta}) \longrightarrow 1
1533: $$
1534: splits. That is,
1535: $Z(\beta)\cong\left(Z(\beta_{[1]})\times \cdots
1536: \times Z(\beta_{[t]})\right) \rtimes Z_0(\widehat{\beta})$.
1537: \end{cor}
1538: 
1539: \begin{proof}
1540:  Since $\widehat{\beta}$ cannot be reducible, the result is a direct
1541: consequence of propositions \ref{P:section_trivial}, \ref{P:section_periodic}
1542: and \ref{P:section_pA}.
1543: \end{proof}
1544: 
1545: % ............................................................
1546: 
1547: \subsection{Structure of $Z_0(\widehat{\beta})$}\label{S:C_0beta}
1548: 
1549: %The only part of theorem~\ref{T:main1} that remains to be shown
1550: %says that, if $\beta$ is a non-periodic
1551: %reducible braid, then $Z_0(\widehat{\beta})$ is isomorphic either to
1552: %$\Z^2$ or to a mixed braid group. This will be shown in this subsection.
1553: The proof of theorem~\ref{T:main1} is now completed by the following result.
1554: \begin{prop}\label{P:C_0beta}
1555: Suppose that $\beta$ is a non-periodic reducible braid, and that its tubular
1556: braid $\widehat{\beta}$ has $m$ strands. Then $Z_0(\widehat{\beta})$ is
1557: isomorphic either to $\Z^2$ or to a mixed braid group on $k$ strands,
1558: where $k\leqslant m$.
1559: \end{prop}
1560: 
1561: \begin{proof}
1562: As usual, there are three subcases, depending wether $\widehat{\beta}$ is
1563: trivial, periodic or \pA.
1564: Recall that we are assuming that $\beta$ is in regular form.
1565: 
1566: Suppose first that $\widehat{\beta}=1$. In this case, $Z(\widehat{\beta})=B_m$.
1567: Hence $Z_0(\widehat{\beta})$ contains any braid whose permutation is
1568: consistent with $\beta$. Denote by ${\cal P}$ the following partition of
1569: $\{P_1,\ldots,P_m\}=\{C_{1,1},\ldots,C_{m,1}\}$: we say that $P_i$ and $P_j$
1570: belong to the same coset of ${\cal P}$ if and only if
1571: $\beta_{[i]}=\beta_{[j]}$. By definition, a braid's permutation is
1572: consistent with $\beta$ if and only if it preserves ${\cal P}$. Therefore,
1573: $Z_0(\widehat{\beta})=B_{\cal P}$, and we are done. (In this case, we have
1574: $k=m$).
1575: 
1576: 
1577: If $\widehat{\beta}$ is {\pA}, it is shown in proposition~\ref{P:section_pA}
1578: that $Z_0(\widehat{\beta})\simeq \Z^2$, so this case is already known.
1579: 
1580: Finally, suppose that $\widehat{\beta}$ is periodic. If it is a power of
1581: $\Delta^2$, then its centralizer is the whole $B_m$, and its corresponding
1582: permutation is trivial, so this case is equivalent to the first one.
1583: 
1584: If $\widehat{\beta}$ is periodic but not a power of $\Delta^2$, then we know
1585: by theorems~\ref{T:centrdelta} and \ref{T:centrdelta2} that
1586: $Z(\widehat{\beta})\simeq B_d(D_*)$, for some $d\geqslant 1$, where $D^*$
1587: is the once punctured disk. But every base point $Q_i$ in $D^*$ corresponds
1588: to an orbit ${\cal C}_i$ of $\widehat{\beta}$ (see figure~\ref{F:theta} in
1589: section~\ref{S:per}), so we can define the following partition ${\cal P}'$
1590: of $\{Q_1,\ldots,Q_d\}$: $Q_i$ and $Q_j$ belong to the same coset if and
1591: only if $\beta_{[i]}=\beta_{[j]}$. This partition lifts by $\theta^{-1}$ to
1592: a partition of $\{P_1,\ldots,P_m\}$, in such a way that any braid in
1593: $B_d(D_*)$ preserves ${\cal P}'$ if and only if its corresponding permutation
1594: in $Z(\widehat{\beta})$ is consistent with $\beta$. Therefore,
1595: $Z_0(\widehat{\beta})\simeq B_{{\cal P}'}(D^*)$. Now it suffices to consider
1596: the central puncture of $D^*$ as another base point, $Q_{d+1}$, and to
1597: notice that $B_{{\cal P}'}(D^*)\cong B_{\cal P}$, where
1598: ${\cal P}={\cal P}'\cup \{\{Q_{d+1}\}\}$. To summarize, in this case we have
1599: $Z_0(\widehat{\beta})\cong B_{{\cal P}'}(D^*)\cong B_{\cal P}$, and the
1600: partition $\cal P$ has $k=d+1$ cosets. Since $d$ must be a proper divisor
1601: of $m$, we get that $k=d+1<m$, and the result follows.
1602: \end{proof}
1603: 
1604: In particular, $Z_0(\widehat{\beta})$ is isomorphic either to $\Z^2$ or to
1605: a mixed braid group. Theorem~\ref{T:main1} is thus proven.
1606: 
1607: % ------------------------------------------------------------------
1608: 
1609: \section{An upper bound for the number of generators}\label{S:upperbound}
1610: 
1611: Once decomposed $Z(\beta)$, if $\beta$ is reducible, as a semi-direct product
1612: of $(Z(\beta_{[1]})\times \cdots \times Z(\beta_{[t]}))$ and
1613: $Z_0(\widehat{\beta})\subset Z(\widehat{\beta})$, we will define a small set
1614: of generators for $Z(\beta)$. We will proceed by induction on the number of
1615: strings, but we need to define first a generating set for
1616: $Z_0(\widehat{\beta})$. We do it as follows:
1617: 
1618: \begin{prop}\label{P:genC_0}
1619:  Let $\beta\in B_n$ be a non-periodic reducible braid, and let
1620: $\widehat{\beta}\in B_m$ be its corresponding tubular braid. Then
1621: $Z_0(\widehat{\beta})$ can be generated by at most $\frac{m(m-1)}{2}$ elements.
1622: \end{prop}
1623: 
1624: \begin{proof}
1625:  If $m=2$ then $Z_0(\widehat{\beta})$ is cyclic, so let us assume that
1626: $m\geqslant 3$. We know by subsection~\ref{S:C_0beta} that
1627: $Z_0(\widehat{\beta})$ is either isomorphic to $\Z^2$ or to a mixed braid
1628: group. The case $\Z^2$ satisfies our result, so we will assume that
1629: $Z_0(\widehat{\beta})$ is isomorphic to a mixed braid group on $k$ strings.
1630: 
1631: Mixed braid groups have been studied in~\cite{Manfredini}, where a
1632: presentation in terms of generators and relations is given. Since we are
1633: mainly interested in the generators, we will extract from those
1634: in~\cite{Manfredini} a small generating set: Let ${\cal P}$ be a partition
1635: of the set $\{1,\ldots,k\}$, having $d$ cosets of length $m_i$
1636: (for $i=1,\ldots,d$). A generating set for $B_{\cal P}$ is given by the
1637: following:
1638: \begin{enumerate}
1639: 
1640:  \item For $i=1,\ldots, d$, a generating set for $B_{m_i}$ (if $m_i>1$).
1641: 
1642:  \item A generating set for the pure braid group $P_d$.
1643: 
1644: \end{enumerate}
1645: It is clear that the first kind of generators corresponds to the movements
1646: of the points inside a coset, while the second one corresponds to the
1647: movement of the points of a coset with respect to those of the others.
1648: For instance, if $k=6$ and ${\cal P}=\{\{1\},\{2,3\},\{4,5,6\}\}$, then one
1649: possible generating set would be:
1650: $$
1651:    \{ \sigma_2\}\cup \{\sigma_4,\;\sigma_5\} \cup \{\sigma_1^2, \;
1652:    \sigma_1\sigma_2\sigma_3^2\sigma_2^{-1} \sigma_1^{-1},\; \sigma_3^2\}.
1653: $$
1654: In order to minimise these generators we recall that $B_2$ is cyclic and,
1655: if $m>2$, then $B_m$ can be generated by two elements. Hence, if we denote
1656: $e_i=m_i-1$ if $m_i< 3$ and $e_i=2$ otherwise, then $e_i$ is a minimal number
1657: of generators for $B_{m_i}$. On the other hand, a minimal number of generators
1658: for $P_d$ is $\frac{d(d-1)}{2}$. Therefore, the minimal number of generators
1659: for $B_{\cal P}$ is:
1660: $$
1661:    g_{\cal P}=\left(\sum_{i=1}^d{e_i}\right)+
1662:    \frac{d(d-1)}{2}\leqslant  \left(\sum_{i=1}^d(m_i-1)\right)+\frac{d(d-1)}{2}
1663: $$
1664: $$
1665:    = k-d+ \frac{d(d-1)}{2}=k+\frac{d(d-3)}{2} \leqslant  k+\frac{k(k-3)}{2}
1666:    =\frac{k(k-1)}{2}.
1667: $$
1668: Notice that if ${\cal P}=\{\{1\},\{2\},\ldots,\{k\}\}$ (so $d=k$), then
1669: $g_{\cal P}=\frac{k(k-1)}{2}$, and this is the worst possibility by the above
1670: formula.
1671: 
1672: Finally we recall from proposition \ref{P:C_0beta} that $k\leqslant m$, so
1673: that $g_{\cal P}\leqslant \frac{m(m-1)}{2}$.
1674: %Now we turn to $Z_0(\widehat{\beta})$. Recall that we are assuming that it
1675: %is isomorphic to a mixed braid groups on $k$ strings, so the result will
1676: %follow if we prove that $k\leqslant m$. We have two cases, depending
1677: %whether $\widehat{\beta}$ is trivial or periodic.
1678: %If $\widehat{\beta}=1$, then we saw in section~\ref{S:C_0beta} that $k=m$
1679: %(this proves that the bound is sharp). If $\widehat{\beta}$ is periodic,
1680: %then either it is a power of $\Delta^2$ -- and in this case $k=m$ again --
1681: %or $Z(\widehat{\beta})\simeq B_d(D^*)$, so $k=d+1$. But in this case,
1682: %$d$ must be a proper divisor of $m$, hence $k=d+1<m$, and the result follows.
1683: \end{proof}
1684: 
1685: 
1686: The first generating set $G'$ of $Z(\beta)$ that we will present
1687: is the following: if $\beta$ is periodic or {\pA}, we have already defined
1688: in sections~\ref{S:per} and \ref{S:pA} a minimal generating set of $Z(\beta)$,
1689: having one or two elements. So suppose that $\beta$ is reducible. Then, by
1690: induction on the number of strings, and by proposition~\ref{P:genC_0}, we
1691: can suppose that we have defined $G_1,\ldots, G_t$ and $G_0$, generating sets
1692: for $Z(\beta_{[1]}), \ldots, Z(\beta_{[t]})$ and $Z_0(\widehat{\beta})$
1693: respectively (if some $\beta_{[i]}$ has one string, then $G_i=\emptyset$).
1694: Then we define $G'=g_1(G_1)\cup \cdots \cup g_t(G_t) \cup h(G_0)$, which is
1695: clearly a generating set for $Z(\beta)$.
1696: 
1697: 
1698: \begin{proof}[Proof of theorem~\ref{T:main}.]
1699: Denote $p(n)$ the upper bound proposed in theorem~\ref{T:main}, that is,
1700: $p(n)=\frac{k(k+1)}{2}$ if $n=2k$ or $p(n)=\frac{k(k+3)}{2}$ if $n=2k+1$.
1701: We will show that the generating set $G'$ defined above has at most
1702: $p(n)$ elements. The case $n=2$ is trivial, so we can suppose that $n>2$ and
1703: that the result is true for any smaller number of strings. We can also assume
1704: that $\beta$ is non-periodic and reducible.
1705: 
1706: The strategy now is to successively replace $\beta$ by different braids,
1707: in such a way that during each replacement step the number of generators
1708: of its centralizer, as given by the above construction, increases.
1709: 
1710: The first modification of $\beta$ will be to replace the tubular braid
1711: $\widehat{\beta}$ by the trivial braid. At the same time, we shall modify the
1712: interior braids, with the aim of rendering them pairwise non-conjugate.
1713: More precisely, we notice that, for any braid $\alpha$ with at least two
1714: strings, the number of generators of $Z(\alpha)$ and $Z(\Delta^{2p}\alpha)$
1715: is the same, while $\Delta^{2p}\alpha$ and $\Delta^{2q}\alpha$ are conjugate
1716: if and only if $p=q$. Thus after multyplying each interior braid $\beta_{[i]}$
1717: by a suitable power of twists $\Delta_{(m_i)}^2$, we can assume that all the
1718: interior braids with at least two strings are pairwise non-conjugate, so that
1719: $t=m$. As seen in the proof of proposition~\ref{P:genC_0}, this first
1720: replacement has increased (or left unchanged) the number of generators of
1721: $G_0$, according to our construction.
1722: 
1723: % Recall that $\widehat{\beta}$ has $m$ strands, while $m_i$ is the number of strands of $\beta_{[i]}$.
1724: % By proposition~\ref{P:genC_0}, the number of elements of $G_0$ is maximised when $\widehat{\beta}$
1725: %is trivial, and all the interior braids are pairwise non-conjugate. Notice also that, for any braid $\alpha$
1726: %with at least two strings, the number of generators of $Z(\alpha)$ and $Z(\Delta^{2p}\alpha)$ is the same, while
1727: %$\Delta^{2p}\alpha$ and $\Delta^{2q}\alpha$ are conjugate if and only if $p=q$. Therefore, we can assume that
1728: %$\widehat{\beta}=1$ and (up to multiplying by some powers of $\Delta^2_{(m_i)}$) that all the interior braids
1729: %with at least two strings are non-conjugate, since this would not decrease the number of elements in $G'$.
1730: %We are thus assuming  that $t=m$.
1731: 
1732: Suppose, without loss of generality, that $m_1=m_2=\cdots=m_d=1$,  that $m_i=2s_i$ for $i=d+1,\ldots,d+u$,
1733: and that $m_i=2s_i+1$, for $i=d+u+1,\ldots,d+u+v$, where $d+u+v=m$. Hence $u$ is the number of interior braids
1734: with an even number of strings, and $v$ is the number of interior braids with an odd (but greater than one) number
1735: of strings. If $d\geqslant 2$, then we shall make further modifications to the braid $\beta$, with the aim
1736: of lowering $d$. More precisely, if $d\leqslant 2$, then we can decrease $d$ by multiplying $\beta$ by
1737: $\sigma_1^p$ for some $p$, where $p$ is chosen in such a way that no other interior braid of $\beta$ equals
1738: $\sigma_1^p$. This replacement increases $u$ by one,
1739: and decreases $d$ by two. Thus the number of generators in $G_0$ decreases by one (if $d=2$) or increases
1740: (if $d>2$). But we would have a new interior braid, $\sigma_1^p$, yielding one new generator. Hence, the total
1741: number of elements in $|G'|$ will not decrease. In other words, without decreasing the number of elements
1742: of $|G'|$ we can replace $\beta$ by a braid with $d\leqslant 1$.
1743: 
1744: %we can decrease $d$ if we multiply $\beta$ by $\sigma_1^p$ for some $p$: if no other
1745: %interior braid of $\beta$ equals $\sigma_1^p$, then we would increase $u$ by one and decrease $d$ by two.
1746: %The number of generators in $G_0$ would decrease by one (if $d=2$) or would increase (if $d>2$). But we would have
1747: %a new interior braid, $\sigma_1^p$, yielding one new generator. Hence, the total number of elements in $|G'|$ would
1748: %not decrease. In other words, we can suppose that $d\leqslant 1$.
1749: 
1750: Denote $a=s_{d+1}+\cdots +s_{d+u}$, $\; b=s_{d+u+1}+\cdots+s_{m}$ and $S=a+b$.
1751: Then one has $n=d+2S+v$. By induction on the number of strings, we have the following bound on the number of
1752: elements in $G'$:
1753: \begin{eqnarray*}
1754: |G'| & \leqslant & \sum_{i=d+1}^{m}{p(m_i)}+ \frac{m(m-1)}{2}  \\
1755:    & = & \sum_{i=d+1}^{d+u}{\frac{s_i(s_i+1)}{2}}+\sum_{i=d+u+1}^m{\frac{s_i(s_i+3)}{2}}+
1756: \left( \begin{array}{c} m \\2 \end{array}\right) \\
1757:   & = &  \sum_{i=d+1}^m{\left( \begin{array}{c} s_i+1 \\2 \end{array}\right)} + \sum_{i=d+u+1}^m{s_i} +
1758: \left( \begin{array}{c} m \\2 \end{array}\right) \\
1759:   & = &  \sum_{i=d+1}^m{\left( \begin{array}{c} s_i+1 \\2 \end{array}\right)} + b +
1760: \left( \begin{array}{c} m \\2 \end{array}\right) \\
1761: \end{eqnarray*}
1762: where $s_i\geqslant 1$ for $i=d+1,\ldots,m$.
1763: 
1764: 
1765:  Given two positive integers $x$ and $y$, one has:
1766: $$
1767:   \left( \begin{array}{c} x+1 \\ 2 \end{array}\right) + \left( \begin{array}{c} y+1 \\ 2 \end{array}\right) =
1768:  \left( \begin{array}{c} x+y+1 \\ 2 \end{array}\right) - xy.
1769: $$
1770: This yields:
1771: \begin{eqnarray*}
1772:  |G'| & \leqslant & \left( \begin{array}{c} S+1 \\ 2 \end{array}\right)
1773:  -\left(\sum_{d+1\leqslant i<j \leqslant m}{s_is_j}\right) + b +
1774: \left( \begin{array}{c} m \\ 2 \end{array}\right).
1775: \end{eqnarray*}
1776: 
1777:  Now we distinguish two cases. If $d=0$, then $m=u+v$ and $n=2S+v$. Also,
1778: \begin{eqnarray*}
1779:  |G'| & \leqslant & \left( \begin{array}{c} S+1 \\ 2 \end{array}\right)
1780:  -\left(\sum_{1\leqslant i<j\leqslant m}{s_is_j}\right) + b +
1781: \left( \begin{array}{c} m \\ 2 \end{array}\right) \\   & \leqslant & \left(
1782: \begin{array}{c} S+1 \\ 2 \end{array}\right)
1783:  -  \left( \begin{array}{c} m \\ 2 \end{array}\right) + b +
1784: \left( \begin{array}{c} m \\ 2 \end{array}\right)
1785: \quad = \quad \frac{S(S+1)}{2} + b.
1786: \end{eqnarray*}
1787: 
1788: 
1789: If $v=0$ one has $b=0$, so $S=a$ and
1790: $|G'|\leqslant \frac{S(S+1)}{2}=\frac{a(a+1)}{2}$; but also $n=2k=2a$,
1791: so $p(n)=\frac{a(a+1)}{2}$ and we are done.
1792: 
1793: If $v=1$ then $n=2S+1$, hence $k=S$ and $p(n)=\frac{S(S+3)}{2}$. But in this case
1794: $|G'|\leqslant \frac{S(S+1)}{2}+b\leqslant \frac{S(S+1)}{2}+S = \frac{S(S+3)}{2}=p(n)$.
1795: 
1796: If $v\geqslant 2$, since $n=2S+v$ one has $k\geqslant S+1$. Then $|G'|\leqslant \frac{S(S+1)}{2} + b
1797: < \frac{S(S+1)}{2} + (S+1) =\frac{(S+2)(S+1)}{2} \leqslant \frac{k(k+1)}{2} \leqslant p(n)$.
1798: 
1799: Therefore, the result is true if $d=0$. Suppose now that $d=1$. In this case $m=u+v+1$ and
1800: $n=2S+v+1$. Then one has:
1801: \begin{eqnarray*}
1802:  |G'| & \leqslant & \left( \begin{array}{c} S+1 \\ 2 \end{array}\right)
1803:  -\left(\sum_{2\leqslant i<j\leqslant m}{s_is_j}\right) + b +
1804: \left( \begin{array}{c} m \\ 2 \end{array}\right) \\
1805: & \leqslant & \left( \begin{array}{c} S+1 \\ 2 \end{array}\right)
1806:  -  \left( \begin{array}{c} m-1 \\ 2 \end{array}\right) + b +
1807: \left( \begin{array}{c} m \\ 2 \end{array}\right) \\
1808:  & = & \frac{S(S+1)}{2} + b + m-1 \quad = \quad \frac{S(S+1)}{2} + b+u+v \\
1809:  & \leqslant & \frac{S(S+1)}{2} + S+ v  \quad = \quad  \frac{S(S+3)}{2} + v.
1810: \end{eqnarray*}
1811: 
1812: 
1813: 
1814: If $v=0$ then $b=0$ and $\;k=S$, so $|G'|\leqslant \frac{S(S+3)}{2} = p(n)$.
1815: 
1816: If $v=1$ then $n=2S+2$ and $k=S+1$. Then $|G'|\leqslant \frac{S(S+3)}{2}+1=
1817: \frac{(S+1)(S+2)}{2}=p(n)$.
1818: 
1819: If $v=2$ then $n=2S+3$ and $k=S+1$. Then $|G'|\leqslant \frac{S(S+3)}{2}+2=
1820: \frac{S^2+3S+4}{2} <\frac{(S+1)(S+4)}{2}=p(n)$.
1821: 
1822: 
1823: Finally, if $v\geqslant 3$ then $n=2S+v+1$ so $k\geqslant S+v/2$. Hence
1824: $$
1825:   p(n)\; \geqslant \; \frac{(S+v/2)(S+v/2+1)}{2} \;=
1826:   \;\frac{S^2+(v+1)S+v(v+2)/4}{2}
1827: $$
1828: $$
1829:   \geqslant \; \frac{S(S+3)}{2} + S/2 + v/2
1830:   \; > \; \frac{S(S+3)}{2}+ v \; \geqslant \; |G'|.
1831: $$
1832: Therefore, in every case $|G'|\leqslant p(n)$, and theorem~\ref{T:main} is
1833: proved.
1834: \end{proof}
1835: 
1836: 
1837: Recall that, in example~\ref{E:purebraid}, %(which is due to S. J. Lee),
1838: we defined braids of any number of strands whose centralizer could not be
1839: generated by less than $p(n)$ elements. Therefore, the bound given by
1840: theorem~\ref{T:main} is the best possible one.
1841: 
1842: % ------------------------------------------------------------------
1843: 
1844: \section{Small generating sets}\label{S:minimal}
1845: 
1846: We saw in the previous section an upper bound for the number of generators of the
1847: centralizer of a braid $\beta$, in terms of its number of strings. But one could obtain a
1848: better bound if more information about $\beta$ is given. In this section we will define a
1849: new generating set $G$ for $Z(\beta)$, which is in most cases smaller than the set $G'$
1850: defined before. It is also the smallest possible ``natural'' generating set, in the sense
1851: that each generator belongs to one of the $t+1$ factors in the semidirect product
1852: decomposition in theorem \ref{T:main1}(c). Thus in a philosophical sense, $G$ is the
1853: ``right'' generating set, even though it is not in general the smallest possible one, as we
1854: shall see at the end of this section.
1855: 
1856: If $\beta$ is periodic or {\pA}, we already know a minimal generating set, with at most two
1857: elements. We also know a minimal generating set for any mixed braid group (see the proof of
1858: proposition~\ref{P:genC_0}). Hence we can define $G$ by induction on the number of strands,
1859: when $\beta$ is a reducible, non-periodic braid. We can also suppose that $\beta$ is in
1860: regular form. We recall that the interior braids are denoted $\beta_{[1]},\ldots,
1861: \beta_{[t]}$, and the tubular braid $\widehat{\beta}$.
1862: 
1863: \begin{definition}
1864: We will say that $i,j\in \{1,\ldots, t\}$ are {\em permutable}
1865: if there exists some $\eta\in Z_0(\widehat{\beta})$ such that
1866: $\eta({\cal C}_i)={\cal C}_j$.
1867: \end{definition}
1868: 
1869: Remark that permutability is an equivalence relation,
1870: and the definition of $Z_0(\widehat\beta )$ says that if $i$ and $j$
1871: are permutable then $\beta_{[i]}=\beta_{[j]}$.
1872: 
1873: Let then $\{i_1,\ldots,i_r\}\subset\{1,\ldots,t\}$ be coset representatives
1874: for permutability. Let $G_{i_k}$ be a minimal set of generators for
1875: $Z(\beta_{[i_k]})$, and $G_0$ be a minimal set of generators for
1876: $Z_0(\widehat{\beta})$. Then we define
1877: $G=g_{i_1}(G_{i_1})\cup \cdots \cup g_{i_r}(G_{i_r}) \cup h(G_H)$.
1878: Notice that $G\subset G'$, and they coincide if and only if there is no
1879: pair of permutable indices.
1880: 
1881: \begin{prop}
1882: $G$ is a generating set of $Z(\beta)$.
1883: \end{prop}
1884: 
1885: \begin{proof}
1886: From the exact sequence of theorem~\ref{T:exactseq} it follows that,
1887: if $G_i$ is a set of generators for $Z(\beta_{[i]})$, then a set of
1888: generators for $Z(\beta)$ is
1889: $G'=g_{1}(G_{1})\cup \cdots \cup g_{t}(G_{t}) \cup h(G_0)$. Hence, we just
1890: need to show that if $j\in\{1,\ldots, t\}\backslash \{i_1,\ldots,i_r\}$,
1891: then every element in $g_j(G_j)$ can be written as a product of elements
1892: in $G$.
1893: 
1894: Take then $j$ as above. There must be some $i_k$ permutable with $j$, so $\beta_{[j]}=\beta_{[i_k]}$
1895: and there is some $\eta\in Z_0(\widehat{\beta})$ such that $\eta({\cal C}_{i_k})={\cal C}_{j}$.
1896: Notice that $G_j$ is a set of generators for $Z(\beta_{[j]})=Z(\beta_{[i_k]})$, so every $\gamma\in G_j$
1897: can be written as a product of elements in $G_{i_k}$. Hence the braid
1898: $\alpha=h(\eta)^{-1} g_{i_k}(\gamma) h(\eta)$ can be written as a product of elements in $G$.
1899: Moreover, one has $\widehat{\alpha}=\widehat{h(\eta)}^{-1} 1 \widehat{h(\eta)} = 1$, and the
1900: only nontrivial interior braids in $\alpha$ are those corresponding to ${\cal C}_j$. Since the
1901: interior braids $h(\eta)_{i_k,l}$ for every $l$  are just powers of $\beta_{[i_k]}=\beta_{[j]}$,
1902: and $\gamma$ commutes with $\beta_{[j]}$, it follows that for every $l$, $\alpha_{j,l}=\gamma$.
1903: Therefore $\alpha=g_j(\gamma)$, so every element in $g_j(G_j)$ can be written as a product of elements
1904: in $G$, thus $G$ is a generating set for $Z(\beta)$.
1905: \end{proof}
1906: 
1907: The generating set we have just defined is, unfortunately, not always the
1908: smallest possible one:
1909: 
1910: \begin{example}\rm
1911: Consider the five string braid
1912: $\beta = \sigma_3 \sigma_4 \sigma_2 \sigma_3 \sigma_1
1913: \sigma_2 \sigma_2 \sigma_3 \sigma_4 \sigma_1 \sigma_2 \sigma_3$ --
1914: the canonical reduction system of this braid has two round circles,
1915: one containing punctures number 1, 2 and 3, the other punctures number
1916: 4 and 5; the tubular braid is just a full twist of the two fat
1917: strings: $\widehat{\beta} = \sigma_1^2$. Moreover, the interior braids of
1918: each tube is trivial. According to theorem \ref{T:main1}, the centralizer
1919: of this braid is
1920: $$
1921: Z(\beta)\cong (B_3 \times B_2) \rtimes PB_2 \cong (B_3 \times \Z)\rtimes \Z
1922: $$
1923: and the generating set constructed in this section has four elements:
1924: two for $B_3$, and one for each factor $\Z$. We now claim that this
1925: generating set is not as small as possible.
1926: 
1927: Indeed, $B_3 \times \Z$ can be generated by only two elements (and thus
1928: $Z(\beta)$ can be generated by three elements). To see this, recall that
1929: the 3-string braid group is isomorphic to the group of the $(2,3)$-torus knot.
1930: Thus $B_3$ has a presentation $\langle y,z\ |\ y^3z^{-2} = 1\rangle$
1931: (with $y=\sigma_1\sigma_2$ and $z=\sigma_1 \sigma_2 \sigma_1$).
1932: Moreover, the factor $\Z$ is generated by $\sigma_4$. Now the two generators
1933: $(y,\sigma_4)$ and $(z,\sigma_4)$ generate $B_3\times \Z$, because
1934: $(1,\sigma_4)$ can be written as $(y,\sigma_4)^3 (z,\sigma_4)^{-2}$.
1935: \end{example}
1936: 
1937: % ------------------------------------------------------------------
1938: 
1939: \section{Some algorithmic aspects}\label{S:alg}
1940: 
1941: The aim of this section is to present the essential ingredients for an
1942: algorithm which, for any given braid, finds a generating set of its
1943: centralizer subgroup that matches the description of the previous sections.
1944: Since, for any braid $\beta$ and any $k\in \Z$, the centralizer subgroups of
1945: $\beta$ and $\beta \Delta^{2k}$ coincide, we can always assume that $\beta$
1946: is positive.
1947: 
1948: We start by mentioning that algorithms that perform the Nielsen-Thurston
1949: classification, and give the invariant folitations in the {\pA} case (in the
1950: form of train tracks), are available -- notably, there are Bestvina-Haendel's
1951: \cite{BestvinaHaendel} and of Los' \cite{Los} algorithms; and computer
1952: implementations are available on the web \cite{Brinkmann, Hall}.
1953: 
1954: We recall briefly the idea of the two automatic structures on braid groups
1955: that are relevant for us: for the first one, given by Garside \cite{Garside}
1956: and Thurston \cite{ThBnaut} (and refined by El-Rifai and Morton \cite{EM}),
1957: we think of $D_n$ has having the $n$ punctures lined up on the real line
1958: in the disk $D$. For the second one, given by Birman, Ko, and Lee \cite{BKL},
1959: we think of $B_n$ as having the $n$ punctures regularly spaced on the circle
1960: of radius $1$. Apart from that, the structures are exactly analogue.
1961: In the Garside-Thurston structure, there is a canonical way to write $\beta$
1962: as a product of divisors of $\Delta$, namely by pushing each crossing between
1963: two strings into a factor as far to the  left as possible. This normal form
1964: is called the {\sl left greedy normal form}. For instance, in this normal
1965: form all factors which are {\it equal} to $\Delta$ (not just divisors of it)
1966: are grouped together at the very left of the product decomposition.
1967: Analogously, Birman-Ko-Lee write
1968: each braid as a product of divisors of $\delta$ in a left-greedy way.
1969: If $\beta$ is a positive braid, then its {\sl super summit set} is the
1970: subset of all elements $\alpha$ of its conjugacy class which satisfy the
1971: following conditions:
1972: \begin{enumerate}
1973: \item[(i)] $\alpha$ is positive,
1974: \item[(ii)] the writing of $\alpha$ in left greedy normal form has as
1975: few factors as possible among all elements satisfying (i),
1976: \item[(iii)]  the writing of of $\alpha$ in left greedy normal form has
1977: as many factors on the left as possible equal to $\Delta$ (or $\delta$),
1978: among all elements satisfying (i) and (ii).
1979: \end{enumerate}
1980: Two positive elements of $B_n$ are conjugate if and only if their super
1981: summit sets coincide. Given $\beta\in B_n$ there is an algorithm, given
1982: in~\cite{FrGM_conj} (which is an improvement of the algorithm in~\cite{EM}),
1983: to compute its super summit set. It is as follows: first we repeatedly
1984: {\sl cycle} $\beta$ (i.e.~move the first factor different from $\Delta$,
1985: respectively $\delta$, to the end and calculate the left greedy form of the
1986: resulting braid), until this process runs into a loop. At this point we are
1987: guaranteed to have achieved condition (ii) above. Then we repeatedly
1988: {\sl decycle} (i.e.~move the last factor to the front
1989: and calculate the left greedy form of the resulting braid) until we run
1990: into a loop. Then all elements of this loop belong to the super summit set.
1991: Afterwards, all other elements of the super summit set can be found
1992: recursively by conjugating already known elements by (suitable) divisors of
1993: $\Delta$ (respectively $\delta$), and retaining the result if it belongs to
1994: the super summit set.
1995: 
1996: This algorithm for computing the super summit set is necessary for our
1997: purposes. Now suppose we are given a braid $\beta\in B_n$ and we want to
1998: compute its centralizer. First we need to determine if $\beta$ is periodic,
1999: reducible or {\pA}, and then we can use the results in this paper.
2000: 
2001: \begin{remark}
2002: Very recently, V. Gebhardt~\cite{Gebhardt} presented a better algorithm for
2003: the conjugacy problem in braid groups. He defined the {\em ultra summit set},
2004: which is in general much smaller than the super summit set described here.
2005: \end{remark}
2006: 
2007: % ..................................................................
2008: 
2009: \sh{Periodic elements}
2010: 
2011: Deciding whether a given element $\beta$ of $B_n$ is periodic is very
2012: easy: one calculates the $n-1$st and the $n$th power of $\beta$. Then $\beta$
2013: is periodic if and only if one of the two results is a power of $\Delta^2$.
2014: 
2015: If $\beta^{n-1}=\Delta^{2k}$ for some $k\in \N$, then $\beta$ is conjugate
2016: to $\gamma_{(n)}^k$ (as can be easily seen from lemma~\ref{L:per_gamma_delta}),
2017: and a conjugating element can be found explicitly using either of the
2018: two standard algorithms. Similarly, if $\beta^n=\Delta^{2k}$, then $\beta$
2019: is conjugate to $\delta_{(n)}^k$, and either algorithm yields an
2020: explicit conjugating element. In either case, one can find explicitly
2021: a generating set of the centralizer subgroup with only two elements, using
2022: propositions \ref{P:centrdelta'} or \ref{P:centrgamma}.
2023: 
2024: % ..................................................................
2025: 
2026: \sh{Finding reducing curves of reducible elements}
2027: 
2028: After establishing that an element $\beta$ of $B_n$ is not periodic, we need
2029: to check whether it is reducible, and if it is, we want to find explicitly
2030: an invariant multicurve. This is, in fact, a standard part of
2031: Bestvina-Haendel's \cite{BestvinaHaendel} and of Los' \cite{Los}
2032: algorithms.
2033: 
2034: We want to point out one particulary elegant alternative, which is due to
2035: Benardete, Gutierrez and Nitecki \cite{BGN2} (see also~\cite{BGN}). We think of
2036: $D_n$ as having the $n$ punctures lined up horizontally, and we look at
2037: Garside-Thurston's left greedy normal form. The key observation from
2038: \cite{BGN2} is the following: suppose that $C$ is an invariant multicurve of
2039: a braid $\beta$, and that the normal form of $\beta$ is
2040: $\beta=\beta_1\cdot\ldots\cdot \beta_k$, where $\beta_1,\ldots,\beta_k\in B_n$
2041: are divisors
2042: of $\Delta$. Moreover, suppose that all components of $C$ are {\sl round}
2043: (i.e.~actual geometric circles in $D_n$). Then we have not only that
2044: $\beta_1\cdot\ldots\cdot\beta_n(C)=C$, but also that all components of all the
2045: multicurves $\beta_1\cdot\ldots\cdot\beta_i(C)$ are round for $i=1,\ldots,k$.
2046: 
2047: As remarked in \cite{BGN2} this implies as a corollary that invariant
2048: multicurves are visible as round curves in the super summit set of $\beta$,
2049: and in particular the reducibilty of a braid is easily detectable from the
2050: super summit set. To prove the corollary we note that $\beta$ has a conjugate
2051: in which all components of the curve system $C$ are round; moreover,
2052: $\beta$ and its conjugate have the same super summit set. Now cycling and
2053: decycling this conjugate does not change the fact that there  is a {\it round}
2054: invariant curve system, by the key observation above. At the end of the
2055: cycling/decycling procedure we have found elements of the super summit set
2056: which contain the desired round invariant curves.
2057: 
2058: Now it is shown in~\cite{BGN2} how to determine if a
2059: given braid preserves a system of disjoint round curves. And there is a finite
2060: number of these systems. Moreover, since for each element of the super summit
2061: set we know how it can be conjugated to obtain $\beta$, we can find explicitly
2062: all curves that belong to a reduction system for $\beta$. We can then easily
2063: determine, by its definition, which of these curves belong to the canonical
2064: reduction system of $\beta$. That is, we can compute the canonical reduction
2065: system of $\beta$.
2066: 
2067: By the results in this paper, $Z(\beta)$ is then a semi-direct product of two
2068: groups that can be computed by induction on the number of strings. Hence, it
2069: only remains to study the case when $\beta$ is {\pA}.
2070: 
2071: 
2072: % ..................................................................
2073: 
2074: \sh{Pseudo-Anosov elements: commutation with $\delta_{(n)}^k$}
2075: 
2076: Suppose that our braid $\beta$ fails the tests of periodicity and reducibility,
2077: hence it is known to be {\pA}. We need to check if it commutes with a
2078: periodic braids other than powers of $\Delta^2$.
2079: 
2080: We shall think of $D_n$ as having its $n$ punctures uniformly distributed
2081: over the circle of radius $1$, and we consider Birman-Ko-Lee's
2082: left-greedy normal form. We want to decide
2083: algorithmically whether $\beta$ is conjugate to a braid $\alpha$ with
2084: the property that $\alpha$ commutes with $\delta_{(n)}^k$ for some
2085: positive integer $k<n$. If it is, we want to know the conjugating braid
2086: explicitly. The following result yields such an algorithm.
2087: %
2088: \begin{prop}\label{P:commdelta}
2089: Suppose that a {\pA} braid $\beta$ has a conjugate which commutes with
2090: $\delta_{(n)}^k$ for some integer $k$. Then there exists an element $\alpha$
2091: of the super summit set of $\beta$ which has the property that $\alpha$,
2092: and in fact every factor of the left greedy normal form of $\alpha$,
2093: commutes with $\delta_{(n)}^k$.
2094: \end{prop}
2095: %
2096: \begin{proof} Let $\beta'$ be a conjugate of $\beta$ which commutes with
2097: $\delta_{(n)}^k$. If $\beta'=\beta'_1\cdot\ldots\cdot\beta'_r$ is the
2098: left-greedy normal form of $\beta'$, then each factor $\beta'_i$ is a divisor
2099: of $\delta_{(n)}$ which is $\frac{2\pi k}{n}$-symmetric. This follows from
2100: the fact that the very definition of the left-greedy normal form is
2101: completely rotation symmetric. More precisely, the fact that two
2102: consecutive factors $\beta'_i \beta'_{i+1}$ determine a left-greedy normal
2103: form is not modified by rotating them. Hence, the product
2104: $(\delta_{(n)}^{-k} \beta'_1 \delta_{(n)}^k ) \cdots
2105: (\delta_{(n)}^{-k} \beta'_r\delta_{(n)}^k)$ is in left-greedy normal form.
2106: Since this product equals $\delta_{(n)}^{-k} \beta'\delta_{(n)}^k=\beta'$,
2107: whose left-greedy normal form is $\beta'_1\cdots \beta'_k$, we obtain that
2108: $\delta_{(n)}^{-1} \beta'_i \delta_{(n)}=\beta'_i$, for $i=1,\ldots,r$.
2109: 
2110: Using the same argument inductively, we see that the cycling and decycling
2111: procedure only ever creates braids in left greedy normal form in which all
2112: factors are $\frac{2\pi k}{n}$-symmetric.
2113: \end{proof}
2114: 
2115: Now we notice that it is very easy to decide if a given divisor of $\delta$ (in
2116: the Birman-Ko-Lee context) is invariant under a given rotation. Hence one
2117: can determine if a braid commutes with an (explicitly computable) conjugate of
2118: $\delta_{(n)}^k$ by looking at the elements of its super summit set.
2119: 
2120: 
2121: 
2122: % ..................................................................
2123: 
2124: \sh{Pseudo-Anosov elements: commutation with $\gamma_{(n)}^k$}
2125: 
2126:  Now we want to determine if a given {\pA} braid commutes with a conjugate
2127: of $\gamma_{(n)}^k$, for a given positive integer $k<n-1$. This is only
2128: possible if there is some index $i\in \{1,\ldots,n\}$ such that $\beta$
2129: preserves $P_i$, as can be easily seen by looking at the corresponding
2130: permutations.
2131: 
2132: Call ${\cal P}_i=\{\{P_i\},\{P_1,\ldots,P_{i-1},P_{i+1},\ldots,P_n\}\}$, a
2133: partition of $\{P_1,\ldots,P_n\}$. Then $\beta$ should belong to
2134: $B_{{\cal P}_i}$. There is a natural map
2135: $f_i\co B_{{\cal P}_i} \rightarrow B_{n-1}$ which consists of forgetting
2136: the $i$th string. Notice that, if a braid $\alpha$ commutes with
2137: $\gamma_{(n)}^k$ (where $P_1$ is considered to be the central point of
2138: $D_{(n)}$) then $f_1(\alpha)$ commutes with
2139: $f_1(\gamma_{(n)}^k)=\delta_{(n-1)}^k$.
2140: 
2141: Hence we have a necessary condition that must be satisfied. If $\beta$
2142: preserves a puncture $P_i$, then we conjugate it to some $\alpha$ that
2143: preserves $P_1$, and we test whether a conjugate of  $f_1(\alpha)$ commutes
2144: with $\delta_{(n-1)}^k$ for some $k<n-1$. If this does not happen, for
2145: $i=1,\ldots, n$, then no conjugate of $\beta$ commutes with $\gamma_{(n)}^k$.
2146: 
2147: This necessary condition is of course not sufficient. A sufficient
2148: and testable condition is now given by the following result. Recall that, by
2149: corollary~\ref{C:addstring}, there is an isomorphism
2150: $\chi=(\bar{\theta}^*)^{-1}\theta^*$ from
2151: $Z(\delta_{(n-1)}^k)$ to $Z(\gamma_{(n)}^k)$, given by adding a trivial
2152: string at the centre of $D_{n-1}$. Notice that, if
2153: $\zeta\in Z(\gamma_{(n)}^k)$, then $\chi(f_1(\zeta))=\zeta$. Then one has:
2154: 
2155: \begin{prop}
2156: Suppose that $\alpha\in B_n$ preserves $P_1$, and $\tld\alpha=f_1(\alpha)$
2157: commutes with $\delta_{(n-1)}^k$. Then the following two statements are
2158: equivalent.
2159: \begin{enumerate}
2160: \item[(i)] $\alpha$ is conjugate to an element $\zeta$ of $B_n$
2161: which commutes with $\gamma_{(n)}^k$, and the conjugating homeomorphism
2162: preserves $P_1$.
2163: \item[(ii)]  $\alpha$ is conjugate to $\chi(\tld\alpha)$.
2164: \end{enumerate}
2165: \end{prop}
2166: 
2167: 
2168: \begin{proof}
2169: The implication (ii)$\Rightarrow$(i) is immediate, by choosing
2170: $\zeta := \chi(\tld\alpha)$.
2171: 
2172: For the implication (i)$\Rightarrow$(ii), we suppose that (i) holds, that is,
2173: there is an element $\eta\in B_{{\cal P}_1}$ such that
2174: $\eta^{-1} \alpha \eta= \zeta$, where $\zeta\in Z(\gamma_{(n)}^k)$. We can
2175: apply $f_1$ to all these elements, denoting $\tld\eta=f_1(\eta)$ and
2176: $\tld\zeta=f_1(\zeta)$.
2177: This yields  $(\tld\eta)^{-1} \tld\alpha \:\tld\eta = \tld\zeta$,
2178: where $\tld\alpha, \tld\zeta \in Z(\delta_{(n-1)}^k)$.
2179: 
2180: If we show that $\tld\eta\in Z(\delta_{(n-1)}^k)$, then we can apply $\chi$ to
2181: all factors, obtaining
2182: $\chi(\tld\eta)^{-1} \chi(\tld\alpha) \:\chi(\tld\eta) =
2183: \chi(\tld\zeta)=\zeta$, hence $\chi(\tld\alpha)$ is conjugate to $\zeta$
2184: which is conjugate to $\alpha$, and the result follows.
2185: 
2186: Let us then show that $\tld\eta$ commutes with $\delta_{(n-1)}^k$.
2187: Notice that $\zeta$ is a {\pA} braid that commutes with $\gamma_{(n)}^k$.
2188: Hence it preserves a projective foliation ${\cal F}_{\zeta}$, which is
2189: invariant under a rotation by an angle of $\frac{2\pi k}{n-1}$. But in this
2190: case $\tld\zeta$ also preserves ${\cal F}_{\zeta}$, with the same
2191: stretch factor, hence it is also {\pA}. Since $\tld\alpha$ is conjugated
2192: to $\tld\zeta$, then it is {\pA} as well, and we call ${\cal F}_{\tld\alpha}$
2193: its corresponding projective foliation (which is also invariant under the
2194: same rotation, since $\tld\alpha$ commutes with $\delta_{(n-1)}^k$).
2195: Since $(\tld\eta)^{-1} \tld\alpha \:\tld\eta = \tld\zeta$, we have that
2196: $\tld\eta$ sends ${\cal F}_{\tld\alpha}$ to ${\cal F}_{\zeta}$.
2197: 
2198: Now consider the braid $d=\tld\eta^{-1} \delta_{(n-1)}^k \tld\eta$.
2199: It is conjugate to $\delta_{(n-1)}^k$, and hence periodic.
2200: Moreover, it preserves ${\cal F}_{\zeta}$, so
2201: it commutes with $\tld\zeta$. But the periodic elements in the centralizer
2202: of $\tld\zeta$ form a cyclic group containing $\delta_{(n-1)}^k$, and
2203: $\delta_{(n-1)}^k$ is the only element having exponent sum $(n-2)k$. Since
2204: $d$ has exactly the same exponent sum, it follows that $d=\delta_{(n-1)}^k$.
2205: Hence $\tld\eta$ commutes with $\delta_{(n-1)}$, and the result follows.
2206: \end{proof}
2207: 
2208: An algorithm for testing whether a braid $\beta$ is conjugate to a braid
2209: which commutes with $\gamma_{(n)}^k$ is now easy to construct: for each of
2210: the $n$ punctures test whether the puncture is fixed by $\beta$, and whether
2211: forgetting this puncture yields a braid which is conjugate to a braid
2212: $\tld{\alpha}$ that commutes with $\delta_{(n)}^k$. (We know how to do
2213: this, by the results of the previous subsection). For each puncture
2214: that does satisfy this property, test whether $\chi(\tld\alpha)$  (which is
2215: obtained from $\tld{\alpha}$ by adding a ``trivial'' string in the
2216: centre), is conjugate to $\beta$. If, for one of the punctures, this is the
2217: case, then the answer is ``yes'', otherwise ``no''.
2218: 
2219: % ..................................................................
2220: 
2221: \sh{Pseudo-Anosov elements: finding roots}
2222: 
2223: It remains to describe a last step for computing a generating set for
2224: $Z(\beta)$, when $\beta$ is {\pA}. We assume that we have already computed
2225: the subgroup $\langle \rho \rangle$ of periodic braids commuting with $\beta$.
2226: Then we can multiply $\beta$ by a suitable power of $\rho$, to obtain a braid
2227: $b$ that preserves the singular leaves of the projective foliations
2228: corresponding to $\beta$. Then we know that
2229: $Z(\beta)=\langle \alpha \rangle \times \langle \rho \rangle$, where $\alpha$
2230: is the smallest possible root of $b$.
2231: 
2232: The last problem, therefore, is to determine whether a given {\pA} braid $b$
2233: has a $k$th root, for given $k$, and to compute that root. This problem has
2234: been solved in~\cite{Sty} (generalised to all Garside groups in \cite{Sibert}).
2235: Moreover, since the number of possible values of $k$ is finite (we are
2236: assuming that $b$ is positive), we have an algorithm for computing $\alpha$,
2237: thus a generating set for $Z(\beta)$.
2238: 
2239: {\bf Acknowledgements } We are grateful to a number of people for discussions
2240: and valuable ideas that greatly contributed to this research. The examples of
2241: Nikolai V.\ Ivanov \cite{IvanovTalk,IvanovRecent}, which we learned about
2242: through discussions with Mustafa Korkmaz, were an important inspiration and
2243: greatly helped us clarify our ideas. It was thus from Ivanov (via Korkmaz)
2244: that we learned that the number of generators may have to grow quadratically
2245: with the number of strings, contradicting a conjecture in \cite{GMFr}. We are
2246: very grateful to Sang Jin Lee who later, but independently of Ivanov, came up
2247: with his examples, conjectured that they represent the worst case, and kindly
2248: communicated these ideas to us by email. (Hessam Hamidi-Teherani found the
2249: same examples as Lee immediately after listening to Ivanov's talk, but we
2250: didn't learn this until very recently.) We also thank David Bessis for useful
2251: discussions, and Joan Birman for telling us about the references~\cite{BGN}
2252: and \cite{BGN2}.
2253: 
2254: 
2255: 
2256: % ------------------------------------------------------------------
2257: 
2258: \begin{thebibliography}{999}
2259: 
2260: \bibitem{Arnold} V. I. Arnold, The cohomology ring of the group of dyed
2261: braids, Mat.Zametki, V. 5 (1969), 227-231.
2262: 
2263: \bibitem{BGN} D.~Benardete, M.~Gutierrez, Z.~Nitecki, A combinatorial
2264: approach to reducibility of mapping classes, Contemporary Math. 150
2265: (1993), 1--31.
2266: 
2267: \bibitem{BGN2} D.~Benardete, M.~Gutierrez, Z.~Nitecki, Braids and the
2268: Nielsen-Thurston classification, J. Knot Theory and its Ramifications
2269: 4 (1995), 549-618.
2270: 
2271: \bibitem{BDM} D.~Bessis, F.~Digne, J.~Michel, Springer theory in braid
2272: groups and the Birman-Ko-Lee monoid, Pacific J. Math. 205 (2002), no. 2,
2273: 287--309.
2274: 
2275: \bibitem{BestvinaHaendel} M.~Bestvina, M.~Haendel, Train-tracks for surface
2276: homeomorphisms, Topology 34 (1995), 109--140.
2277: 
2278: \bibitem{Bir} J.~Birman, Braids, links, and mapping class groups, Annals
2279: of Math.~Studies, 82, Princeton University Press 1975.
2280: 
2281: \bibitem{BKL} J.~Birman, K.~H.~Ko, S.~J.~Lee, A new approach to the word
2282: and conjugacy problems in the braid groups, Adv. Math. 139 (1998), 322--353.
2283: 
2284: \bibitem{BLM} J.~Birman, A.~Lubotzky and J.~McCarthy,
2285: Abelian and solvable subgroups of the mapping class groups, Duke Math.J.\ 50
2286: (1983), 1107-1120.
2287: 
2288: \bibitem{Brinkmann} P.~Brinkmann, An implementation of the Bestvina-Handel
2289: algorithm for surface homeomorphisms,  Experiment. Math. 9 (2000), 235--240.
2290: Computer program available at
2291: {\tt www.math.uiuc.edu/\~{ }brinkman/software/train/}
2292: 
2293: \bibitem{coko} A. Constantin, B. Kolev, The theorem of
2294: Ker\'{e}kj\'{a}rt\'{o} on periodic homeomorphisms of the disc and the
2295: sphere, Enseign. Math (2) 40 (1994) No 3-4, 193-204.
2296: %Preprint (2003) {\tt arXiv:math.GN/0303256}.
2297: 
2298: \bibitem{Eilenberg} S.~Eilenberg, Sur les transformations p\'eriodiques de
2299: la surface de sph\`ere, Fund.~Math.~22 (1934), 28--41.
2300: 
2301: \bibitem{EM} E.~A.~El-Rifai, H.~R.~Morton, Algorithms for positive braids,
2302: Quart.~J. Math. Oxford Ser.~(2) 45 (1994), no.~180, 479--497.
2303: 
2304: \bibitem{FLP} A.~Fathi, F.~Laudenbach, V.~Poenaru, Travaux de Thurston sur
2305: les surfaces - seminaire Orsay, Asterisque 66--67, Soci\'et\'e Math.~de
2306: France, 1991.
2307: 
2308: \bibitem{FRZ} R. Fenn, D. Rolfsen, J. Zhu, Centralisers in the braid group and
2309: singular braid monoid, l'Enseignment Math\'ematique 42 (1996), 75--96.
2310: 
2311: \bibitem{FrGM_conj} N.~Franco, J.~Gonz\'alez-Meneses, Conjugacy problem for braid
2312: groups and Garside groups. To appear in Journal of Algebra.
2313: 
2314: \bibitem{GMFr} N.~Franco, J.~Gonz\'alez-Meneses,  Computation of centralizers
2315: in braid groups and Garside groups. To appear in Revista Matem\'atica
2316: Iberoamericana.
2317: 
2318: \bibitem{Garside} F.~A.~Garside, The braid group and other groups,
2319: Quart.~J.~Math.~Oxford 20 (1969), 235--254.
2320: 
2321: \bibitem{Gebhardt} V. Gebhardt, A New Approach to the Conjugacy Problem in
2322: Garside Groups, Preprint (2003). {\tt arXiv:math.GT/0306199}.
2323: 
2324: \bibitem{Hall} T. Hall, Computer implementation of Bestvina-Handel algorithm
2325: available at
2326: {\tt www.liv.ac.uk/maths/PURE/MIN\_SET/CONTENT/members/T\_Hall.html}
2327: 
2328: \bibitem{Ivanov} N.~V.~Ivanov, Subgroups of Teichm\"uller modular groups,
2329: Translations of mathematical monographs vol.~115 (1992), AMS.
2330: 
2331: \bibitem{IvanovTalk} N.~V.~Ivanov, talk at the special session ``Mapping
2332: class groups and the geometric theory of Teichm\"uller spaces'' at the
2333: $974^{\rm th}$ meeting of the AMS, Ann Harbour, MI, March 1--3, 2002.
2334: 
2335: \bibitem{IvanovRecent} N.~V.~Ivanov, Examples of centralizers in the Artin
2336: braid groups, Preprint (2003), {\tt arXiv:math.GT/0306418}
2337: 
2338: \bibitem{deK} B.~de Ker\'ekj\'art\'o, \"Uber die periodischen Transformationen
2339: der Kreisscheibe und der Kugelfl\"ache, Math.~Annalen 80 (1919), 3--7.
2340: 
2341: \bibitem{Los} J.~Los, Pseudo-Anosov maps and invariant train tracks in the
2342: disc: a finite algorithm, Proc. London Math. Soc. (3) 66 (1993), 400--430.
2343: 
2344: \bibitem{Makanin} G.~S.~Makanin, On normalizers in the braid group,
2345: Mat.Sb.~86 (128), 1971, 171--179.
2346: 
2347: \bibitem{Manfredini} S.~Manfredini, Some subgroups of Artin's braid group.
2348: Special issue on braid groups and related topics (Jerusalem, 1995).
2349: Topology Appl. 78 (1997), no. 1-2, 123--142.
2350: 
2351: \bibitem{Orevkov} S.~Yu.~Orevkov, Quasipositivity test via unitary
2352: representations of braid groups and its applications to real algebraic curves,
2353: J. Knot Theory Ramifications~10 (2001), no. 7, 1005--1023.
2354: 
2355: \bibitem{ParisRolfsen} L.~Paris, D.~Rolfsen, Geometric subgroups of
2356: surface braid groups, Ann. Inst. Fourier 49 (1999), 101-156
2357: 
2358: \bibitem{PeHa} R.~C.~Penner, J.~L.~Harer,  Combinatorics of train tracks,
2359: Annals of Mathematics Studies, 125. Princeton University Press,
2360: Princeton, NJ, 1992.
2361: 
2362: \bibitem{Sibert} H.~Sibert, Extraction of roots in Garside groups,
2363: Comm. Algebra 30 (2002), no. 6, 2915--2927.
2364: 
2365: \bibitem{Sty}  V.~B.~Sty\v snev,  Izv. Akad. Nauk SSSR Ser. Mat. 42 (1978),
2366: no. 5, 1120--1131, 1183.
2367: 
2368: \bibitem{Thclass} W. P. Thurston,  On the geometry and dynamics of
2369: diffeomorphisms of surfaces, Bull. Amer. Math. Soc. (N.S.) 19 (1988),
2370: 417--431.
2371: 
2372: \bibitem{ThBnaut} W. P. Thurston, Braid Groups, Chapter 9 of ``Word
2373: processing in groups",
2374: D.~B.~A.~Epstein, J.~W.~Cannon, D.~F.~Holt, S.~V.~F.~Levy, M.~S.~Paterson and
2375: W.~P.~Thurston, Jones and Bartlett Publishers, Boston, MA, 1992.
2376: 
2377: \end{thebibliography}
2378: 
2379: \end{document}
2380: