math0606435/mwf.tex
1: 
2: % this version has Jones polynomial finiteness moved in
3: % and simplified
4: 
5: \documentstyle[txmac,a4,%leqno,%
6: amssymb,%
7: bibenc,%
8: case,%
9: twoside,%
10: nocaphead2,%
11: epsf,%
12: leqno,%
13: %varthm,%
14: %rotate,%
15: myrot,%
16: %multicol,%
17: mypic,%
18: times,mathptm%
19: ]{article}
20: 
21: \advance\oddsidemargin by -1.9cm
22: \advance\evensidemargin by -1.9cm
23: \advance\textwidth by 3.8cm
24: % \advance\oddsidemargin by -0.3cm
25: % \advance\evensidemargin by -0.3cm
26: % \advance\textwidth by 0.6cm
27: 
28: \def\mynewtheo#1#2{%
29: \newtheorem{@#1}{#2}[section]%
30: \newenvironment{#1}{\begin{@#1}\rm}{\end{@#1}}}
31: 
32: \mynewtheo{lemma}{Lemma}
33: \mynewtheo{exer}{Exercise}
34: \mynewtheo{theo}{Theorem}
35: \mynewtheo{rem}{Remark}
36: \mynewtheo{defi}{Definition}
37: \mynewtheo{conj}{Conjecture}
38: \mynewtheo{corr}{Corollary}
39: \mynewtheo{prop}{Proposition}
40: \mynewtheo{question}{Question}
41: \mynewtheo{exam}{Example}
42: 
43: \makeatletter
44: 
45: \newenvironment{theorem}{\begin{theo}}{\end{theo}}
46: \newenvironment{conjecture}{\begin{conj}}{\end{conj}}
47: \newenvironment{myitems}{%
48: \def\@listi{\leftmargin \leftmargini \parsep 4\p@ \@plus
49: 2\p@ \@minus \p@\topsep1\p@ \@plus 2\p@ \@minus 1\p@
50: \itemsep 4\p@ \@plus 2\p@ \@minus \p@} \begin{itemize}
51: \itemsep\z@\relax \partopsep\z@\relax}{\end{itemize}}
52:  
53: 
54: 
55: \parskip 5pt plus 3pt minus 2pt\relax
56: 
57: \parindent0pt\relax
58: 
59: \pagestyle{headings}
60: 
61: \begin{document}
62: 
63: \input{myeqn.tex}
64: % \input{young.tex}
65: \renewenvironment{eqn}{\begin{equation}}{\end{equation}\ignorespaces}
66: 
67: \def\rottab#1#2#3{
68: \expandafter\advance\csname c@#3\endcsname by -1\relax
69: \centerline{%
70: %\hfill
71: %\fboxsep0pt\relax
72: %\fbox{%
73: \rbox{\centerline{\vbox{\setbox1=\hbox{#1}%
74: \hbox to \wd1{\hfill\vbox{{%
75: %\tracingmacros1
76: \caption{#2}}}\hfill}%
77: \vskip6mm
78: \box1}}%
79: }
80: %}
81: }%
82: }
83: 
84: 
85: % \let\old@diag\diag \def\diag{\special{ps: pstack (====) ==}\old@diag}
86: 
87: \newcommand{\mybin}[2]{\text{$\Bigl(\begin{array}{@{}c@{}}#1\\#2%
88: \end{array}\Bigr)$}}
89: \newcommand{\mybinn}[2]{\text{$\biggl(\begin{array}{@{}c@{}}%
90: #1\\#2\end{array}\biggr)$}}
91: 
92: \def\overtwo#1{\mbox{\small$\mybin{#1}{2}$}}
93: \newcommand{\mybr}[2]{\text{$\Bigl\lfloor\mbox{%
94: \small$\displaystyle\frac{#1}{#2}$}\Bigr\rfloor$}}
95: \def\mybrtwo#1{\mbox{\mybr{#1}{2}}}
96: 
97: \def\myfrac#1#2{\raisebox{0.2em}{\small$#1$}\!/%
98: \!\raisebox{-0.2em}{\small$#2$}}
99: \def\ffrac#1#2{\mbox{\small$\ds\frac{#1}{#2}$}}
100: 
101: \def\noloop{{\diag{0.5cm}{0.5}{1}{\picline{0.25 0}{0.25 1}}}}
102: 
103: \def\vrt#1{{\picfillgraycol{0}\picfilledcircle{#1}{0.09}{}}}
104: 
105: \def\ReidI#1#2{
106:   \diag{0.5cm}{0.9}{1}{
107:     \pictranslate{0.4 0.5}{
108:       \picscale{#11 #21}{
109:         \picmultigraphics[S]{2}{1 -1}{
110: 	  \picmulticurve{-6 1 -1 0}{0.5 -0.5}{0.5 0}{0.1 0.3}{-0.2 0.3}
111: 	} %picmultigraphics
112: 	\piccirclearc{-0.2 0}{0.3}{90 270}
113:       }%picscale
114:     }%pictranslate
115:   }%diag
116: }
117: \def\Pos#1#2{{\diag{#1}{1}{1}{#2
118: \picmultiline{-5 1 -1 0}{0 1}{1 0}
119: \picmultiline{-5 1 -1 0}{0 0}{1 1}
120: }}}
121: \def\Neg#1#2{{\diag{#1}{1}{1}{#2
122: \picmultiline{-5 1 -1 0}{0 0}{1 1}
123: \picmultiline{-5 1 -1 0}{0 1}{1 0}
124: }}}
125: \def\Nul#1#2{{\diag{#1}{1}{1}{#2
126: \piccirclearc{0.5 1.4}{0.7}{-135 -45}
127: \piccirclearc{0.5 -0.4}{0.7}{45 135}
128: }}}
129: \def\Inf#1#2{{\diag{#1}{1}{1}{#2
130: \piccirclearc{0.5 1.4 x}{0.7}{135 -135}
131: \piccirclearc{0.5 -0.4 x}{0.7}{-45 45}
132: }}}
133: \def\pos{\Pos{0.5em}{\piclinewidth{10}}}
134: \def\neg{\Neg{0.5em}{\piclinewidth{10}}}
135: \def\nul{\Nul{0.5em}{\piclinewidth{10}}}
136: \def\inf{\Inf{0.5em}{\piclinewidth{10}}}
137: 
138: \def\noloop{{\diag{0.5cm}{0.5}{1}{\picline{0.25 0}{0.25 1}}}}
139: 
140: \def\ReidI#1#2{
141:   \diag{0.5cm}{0.9}{1}{
142:     \pictranslate{0.4 0.5}{
143:       \picscale{#11 #21}{
144:         \picmultigraphics[S]{2}{1 -1}{
145: 	  \picmulticurve{-6 1 -1 0}{0.5 -0.5}{0.5 0}{0.1 0.3}{-0.2 0.3}
146: 	} %picmultigraphics
147: 	\piccirclearc{-0.2 0}{0.3}{90 270}
148:       }%picscale
149:     }%pictranslate
150:   }%diag
151: }
152: 
153: \def\GD{\szCD{6mm}}
154: \def\szCD#1#2{{\let\@nomath\@gobble\small\diag{#1}{2.4}{2.4}{
155:   \picveclength{0.27}\picvecwidth{0.1}
156:   \pictranslate{1.2 1.2}{
157:     \piccircle{0 0}{1}{}
158:     #2
159: }}}}
160: \def\CD{\szCD{4mm}}
161: 
162: \def\point#1{{\picfillgraycol{0}\picfilledcircle{#1}{0.08}{}}}
163: \def\labpt#1#2#3{\pictranslate{#1}{\point{0 0}\picputtext{#2}{$#3$}}}
164: \def\vrt#1{{\picfillgraycol{0}\picfilledcircle{#1}{0.09}{}}}
165: 
166: \def\chrd#1#2{\picline{1 #1 polar}{1 #2 polar}}
167: \def\arrow#1#2{\picvecline{1 #1 polar}{1 #2 polar}}
168: 
169: \def\labch#1#2#3{\chrd{#1}{#2}\picputtext{1.3 #2 polar}{$#3$}}
170: \def\labar#1#2#3{\arrow{#1}{#2}\picputtext{1.3 #2 polar}{$#3$}}
171: \def\labbr#1#2#3{\arrow{#1}{#2}\picputtext{1.3 #1 polar}{$#3$}}
172: 
173: \def\@dcont{}
174: \def\svCD#1{\ea\glet\csname #1\endcsname\@dcont}
175: \def\rsCD#1{\ea\glet\ea\@dcont\csname #1\endcsname\ea\glet
176: \csname #1\endcsname\relax}
177: 
178: \def\addCD#1{\ea\gdef\ea\@dcont\ea{\@dcont #1}}
179: \def\drawCD#1{\szCD{#1}{\@dcont}}
180: 
181: 
182: \def\epsfs#1#2{{\catcode`\_=11\relax\ifautoepsf\unitxsize#1\relax\else
183: \epsfxsize#1\relax\fi\epsffile{#2.eps}}}
184: \def\epsfsv#1#2{{\vcbox{\epsfs{#1}{#2}}}}
185: \def\vcbox#1{\setbox\@tempboxa=\hbox{#1}\parbox{\wd\@tempboxa}{\box
186:   \@tempboxa}}
187: \newbox\@tempboxb
188: \def\vtbox#1{\setbox\@tempboxa=\hbox{\shortstack[t]{#1}}%
189:   \setbox\@tempboxb=\hbox{2rut}\@tempdima\ht\@tempboxa\relax
190:   \advance\@tempdima-\ht\@tempboxb\relax\raise-\@tempdima\box \@tempboxa
191:   %\dp\@tempboxa=\@tempdima
192:   %\box \@tempboxa
193:   }
194: \def\p{\epsfsv{2cm}}
195: 
196: \def\@test#1#2#3#4{%
197:   \let\@tempa\go@
198:   \@tempdima#1\relax\@tempdimb#3\@tempdima\relax\@tempdima#4\unitxsize\relax
199:   \ifdim \@tempdimb>\z@\relax
200:     \ifdim \@tempdimb<#2%
201:       \def\@tempa{\@test{#1}{#2}}%
202:     \fi
203:   \fi
204:   \@tempa
205: }
206: 
207: \def\go@#1\@end{}
208: \newdimen\unitxsize
209: \newif\ifautoepsf\autoepsftrue
210: 
211: \unitxsize4cm\relax
212: \def\epsfsize#1#2{\epsfxsize\relax\ifautoepsf
213:   {\@test{#1}{#2}{0.1 }{4   }
214: 		{0.2 }{3   }
215: 		{0.3 }{2   }
216: 		{0.4 }{1.7 }
217: 		{0.5 }{1.5 }
218: 		{0.6 }{1.4 }
219: 		{0.7 }{1.3 }
220: 		{0.8 }{1.2 }
221: 		{0.9 }{1.1 }
222: 		{1.1 }{1.  }
223: 		{1.2 }{0.9 }
224: 		{1.4 }{0.8 }
225: 		{1.6 }{0.75}
226: 		{2.  }{0.7 }
227: 		{2.25}{0.6 }
228: 		{3   }{0.55}
229: 		{5   }{0.5 }
230: 		{10  }{0.33}
231: 		{-1  }{0.25}\@end
232: 		\ea}\ea\epsfxsize\the\@tempdima\relax
233: 		\fi
234: 		}
235: 
236: \def\vrt#1{{\picfillgraycol{0}\picfilledcircle{#1}{0.09}{}}}
237: \def\fdot#1{{\picfillgraycol{0}\picfilledcircle{#1}{0.02}{}}}
238: \def\cycl#1#2#3#4{\vrt{#1}\vrt{#2}\vrt{#3}\vrt{#4}
239: \picline{#1}{#2}\picline{#2}{#3}\picline{#3}{#4}\picline{#4}{#1}}
240: \def\xcycl#1#2#3#4#5#6{\vrt{#1}\vrt{#2}\vrt{#3}\vrt{#4}\vrt{#5}
241: \vrt{#6}\picline{#1}{#2}\picline{#2}{#3}\picline{#3}{#4}
242: \picline{#4}{#5}\picline{#5}{#6}\picline{#6}{#1}}
243: 
244: %usage ODD # of  coords, then {}
245: \def\@curvepath#1#2#3{%
246:   \@ifempty{#2}{\piccurveto{#1 }{@stc}{@std}#3}%
247:     {\piccurveto{#1 }{#2 }{#2  #3  0.5 conv}
248:     \@curvepath{#3}}%
249: }
250: \def\curvepath#1#2#3{%
251:   \piccurve{#1 }{#2 }{#2 }{#2  #3  0.5 conv}%
252:   \picPSgraphics{/@stc [ #1  #2  -1 conv ] $ D /@std [ #1  ] $ D }%
253:   \@curvepath{#3}%
254: }
255: 
256: %usage EVEN # of  coords, then {}
257: \def\@opencurvepath#1#2#3{%
258:   \@ifempty{#3}{\piccurveto{#1 }{#1 }{#2 }}%
259:     {\piccurveto{#1 }{#2 }{#2  #3  0.5 conv}\@opencurvepath{#3}}%
260: }
261: \def\opencurvepath#1#2#3{%
262:   \piccurve{#1 }{#2 }{#2 }{#2  #3  0.5 conv}%
263:   \@opencurvepath{#3}%
264: }
265: 
266: 
267: \author{A. Stoimenow\footnotemark[1]\\[2mm]
268: \small Research Institute for Mathematical Sciences, \\
269: \small Kyoto University, Kyoto 606-8502, Japan\\
270: \small e-mail: {\tt stoimeno@kurims.kyoto-u.ac.jp}\\
271: \small WWW: {\hbox{\web|http://www.kurims.kyoto-u.ac.jp/~stoimeno/|}}
272: }
273: 
274: {\def\thefootnote{\fnsymbol{footnote}}
275: \footnotetext[1]{Supported by 21st Century COE Program.}
276: }
277: 
278: \title{\large\bf \uppercase{Properties of closed 3-braids}\\
279: [1mm]\uppercase{and other link braid representations}\\
280: [4mm]
281: \phantom{\small\it This is a preprint. I would be grateful
282: for any comments and corrections %but I prefer you not circulate it
283: !}}
284: 
285: \date{\large Current version: \curv\ \ \ First version:
286: \makedate{22}{9}{2004}}
287: 
288: \maketitle
289: 
290: \makeatletter
291: 
292: \let\vn\varnothing
293: \let\point\pt
294: \let\ay\asymp
295: \let\pa\partial
296: \let\ap\alpha
297: \let\bt\beta
298: \let\be\beta
299: \let\Dl\Delta
300: \let\Gm\Gamma
301: \let\gm\gamma
302: \let\de\delta
303: \let\dl\delta
304: \let\eps\epsilon
305: \let\lm\lambda
306: \let\Lm\Lambda
307: \let\sg\sigma
308: \let\vp\varphi
309: \let\zt\zeta
310: \let\om\omega
311: \let\diagram\diag
312: \let\nb\nabla
313: \let\prt\partial
314: \let\wh\widehat
315: \let\wt\widetilde
316: \def\tW{\wt W}
317: \let\sm\setminus
318: \let\tl\tilde
319: % \def\ncap{\not\mathrel{\cap}}
320: % \def\sgn{\text{\rm sgn}\,}
321: \def\inx{\mathop {\operator@font ind}}
322: \def\spn{\mathop {\operator@font span}\nolimits}
323: \def\dig{\mathop {\operator@font diag}}
324: \def\Mc{\max\cf}
325: \def\Md{\max\deg}
326: \def\md{\min\deg}
327: \def\mc{\min\cf}
328: \def\vol{\text{\rm vol}\,}
329: \def\Ra{\Rightarrow}
330: \def\Lra{\Longrightarrow}
331: \def\lra{\longrightarrow}
332: \def\so{\Rightarrow}
333: \def\So{\Longrightarrow}
334: \def\nin{\not\in}
335: \let\ds\displaystyle
336: \let\llra\longleftrightarrow
337: \let\reference\ref
338: % \def\ul#1{\fbox{$#1$}}
339: \let\ol\overline
340: \let\ul\underline
341: \let\u\underline
342: \let\h\hat
343: % \def\ul#1{\diag{1em}{1}{1}{\picbox{0.5 0.5}{1.0 1.0}{$#1$}}}
344: \let\es\enspace
345: \def\lfra{\leftrightarrow}
346: \def\cf{\text{\rm cf}\,}
347: 
348: \def\TM{$^\text{\raisebox{-0.2em}{${}^\text{TM}$}}$}
349: % \def\lb{\linebreak[0]}
350: % \def\lz{\lb\verb}
351: \def\ssim{\stackrel{\ds \sim}{\vbox{\vskip-0.2em\hbox{$\scriptstyle
352: *$}}}}
353: 
354: 
355: \long\def\@makecaption#1#2{%
356:    % \tm
357:    \vskip \abovecaptionskip 
358:    {\let\label\@gobble
359:    \let\ignorespaces\@empty
360:    \xdef\@tempt{#2}%
361:    }%
362:    \ea\@ifempty\ea{\@tempt}{%
363:    \sbox\@tempboxa{%
364:       \fignr#1#2}%
365:       }{%
366:    \sbox\@tempboxa{%
367:       {\fignr#1:}\capt\ #2}%
368:       }%
369:    \ifdim \wd\@tempboxa >\captionwidth {%
370:       %\centerline{\parbox{\captionwidth}{\unhbox \@tempboxa}}%
371:       \rightskip=\@captionmargin\leftskip=\@captionmargin
372:       \unhbox\@tempboxa\par
373:      }%
374:    \else
375:       \centerline{\box \@tempboxa}%
376:       % \hbox to\captionwidth{\hfil\box\@tempboxa\hfil}%
377:    \fi
378:    \vskip \belowcaptionskip
379:    }%
380: %
381: \def\fignr{\small\sffamily\bfseries}%
382: \def\capt{\small\sffamily}%
383: 
384: % \long\def\@makecaption#1#2{%
385: %    % \tm
386: %    \vskip 10pt
387: %    {\let\label\@gobble
388: %    \let\ignorespaces\@empty
389: %    \xdef\@tempt{#2}%
390: %    %\typeout{`#2'}%
391: %    }%
392: %    \ea\@ifempty\ea{\@tempt}{%
393: %    \setbox\@tempboxa\hbox{%
394: %       \fignr#1#2}%
395: %       }{%
396: %    \setbox\@tempboxa\hbox{%
397: %       {\fignr#1:}\capt\ #2}%
398: %       }%
399: %    \ifdim \wd\@tempboxa >\captionwidth {%
400: %       \rightskip=\@captionmargin\leftskip=\@captionmargin
401: %       \unhbox\@tempboxa\par}%
402: %    \else
403: %       \hbox to\captionwidth{\hfil\box\@tempboxa\hfil}%
404: %    \fi}%
405: % %
406: % \def\fignr{\small\sffamily\bfseries}%
407: % \def\capt{\small\sffamily}%
408: 
409: 
410: \newdimen\@captionmargin\@captionmargin2cm\relax
411: \newdimen\captionwidth\captionwidth0.8\hsize\relax
412: 
413: 
414: \def\eqref#1{(\protect\ref{#1})}
415: 
416: \def\proof{\@ifnextchar[{\@proof}{\@proof[\unskip]}}
417: \def\@proof[#1]{\noindent{\bf Proof #1.}\enspace}
418: 
419: \def\hint{\noindent Hint: }
420: \def\problem{\noindent{\bf Problem.} }
421: % \def\question{\noindent{\bf Question.} }
422: 
423: \def\@mt#1{\ifmmode#1\else$#1$\fi}
424: \def\qed{\hfill\@mt{\Box}}
425: \def\qqed{\hfill\@mt{\Box\enspace\Box}}
426: 
427: % \let\Bbb\bf
428: 
429: \def\cU{{\cal U}}
430: \def\cC{{\cal C}}
431: \def\cP{{\cal P}}
432: \def\fg{{\frak g}}
433: \def\tr{\text{tr\,}}
434: \def\cZ{{\cal Z}}
435: \def\cD{{\cal D}}
436: \def\bR{{\Bbb R}}
437: \def\cE{{\cal E}}
438: \def\bZ{{\Bbb Z}}
439: \def\bN{{\Bbb N}}
440: 
441: 
442: \def\bysame{\same[\kern2cm]\,}
443: 
444: \def\br#1{\left\lfloor#1\right\rfloor}
445: \def\BR#1{\left\lceil#1\right\rceil}
446: 
447: \def\abstractname{}
448: 
449: 
450: \@addtoreset {footnote}{page}
451: 
452: \renewcommand{\section}{%
453:    \@startsection
454:          {section}{1}{\z@}{-2.5ex \@plus -1ex \@minus -.2ex}%
455:                {3ex \@plus.2ex}{\Large\bf}%
456: }
457: 
458: \renewcommand{\subsubsection}{%
459:    \@startsection
460:          {subsubsection}{1}{\z@}{-1.5ex \@plus -1ex \@minus -.2ex}%
461:                {1ex \@plus.2ex}{\large\bf}%
462: }
463: \renewcommand{\@seccntformat}[1]{\csname the#1\endcsname .
464: \quad}
465: 
466: \def\bC{{\Bbb C}}
467: \def\bP{{\Bbb P}}
468: 
469: 
470: \makeatletter
471: 
472: 
473: \let\old@tl\~\def\~{\raisebox{-0.8ex}{\tt\old@tl{}}}
474: \let\lra\longrightarrow
475: \let\sm\setminus
476: \let\eps\varepsilon
477: \let\ex\exists
478: \let\fa\forall
479: \let\ps\supset
480: 
481: \def\rs#1{\raisebox{-0.4em}{$\big|_{#1}$}}
482: 
483: {\let\@noitemerr\relax
484: \vskip-2.7em\kern0pt\begin{abstract}
485: \noindent{\bf Abstract.}\enspace
486: We show that 3-braid links with given (non-zero) Alexander
487: or Jones polynomial are finitely many, and can be effectively
488: determined. We classify among closed 3-braids strongly
489: quasi-positive and fibered ones, and show that 3-braid links have
490: a unique incompressible Seifert surface. We also classify the
491: positive braid words with Morton-Williams-Franks
492: bound 3 and show that closed positive braids of braid index 3 are
493: closed positive 3-braids. For closed braids on more
494: strings, we study the alternating links occurring. In particular
495: we classify those of braid index 4, and show that their Morton-%
496: Williams-Franks inequality is exact. Finally, we use the Burau
497: representation to obtain new braid index criteria, including an
498: efficient 4-braid test.\\[2mm]
499: %
500: {\it Keywords:} link polynomial, positive braid, strongly
501: quasi-positive link, 3-braid link, alternating link, Seifert surface,
502: fiber, incompressible surface, braid index, Burau representation
503: \\[2mm]
504: {\it AMS subject classification:} 57M25 (primary), 20F36, 32S55
505: (secondary)
506: \end{abstract}
507: }\vspace{7mm}
508: 
509: {\parskip0.2mm\tableofcontents}
510: \vspace{7mm}
511: 
512: \section{Introduction}
513: 
514: Originating from the pioneering work of Alexander
515: \cite{Alexander2} and Artin \cite{Artin,Artin2}, braid
516: theory has become intrinsically interwoven with knot theory,
517: and over the years, braid representations of different types
518: have been studied, many of them with motivation coming
519: from fields outside of knot theory, for example dynamical
520: systems \cite{Williams} or 4-dimensional QFTs \cite{Kreimer}.
521: 
522: The systematic study of closed 3-braids was begun by Murasugi
523: \cite{Murasugi2}. Later 3-braid links have been classified
524: in \cite{BirMen}, in the sense that 3-braid representations
525: of the same link are exactly described. This is, up to
526: a few exceptions, mainly just conjugacy. The conjugacy
527: problem of 3-braids has a series of solutions, starting with
528: Schreier's algorithm \cite{Sc}, going over Garside
529: \cite{Garside} (for arbitrary braid groups), Xu \cite{Xu}
530: (and later more generally Birman-Ko-Lee \cite{BKL}), until,
531: for example, a recent algorithm of linear complexity due to
532: Fiedler and Kurlin \cite{FK}. So Birman-Menasco's work allows
533: to decide if two 3-braids have the same closure link.
534: 
535: However, many properties of links (except achirality and
536: invertibility) are not evident from (3-strand) braid
537: representations, and thus to classify 3-braid links with
538: special properties remains a non-trivial task. A first result
539: was given in \cite{Murasugi} for rational links, and then
540: improved in \cite{3br}, where we completed this project for
541: alternating links.
542: 
543: {}From the opposite point of view, a natural question concerning
544: braid representations of links is: 
545: 
546: \begin{question}\label{q1}
547: If a braid representation of a particular type exists, does
548: also one exist with the minimal number of strands (among all
549: braid representations of the link)?
550: \end{question}
551: 
552: The minimal number of
553: strands for a braid representation of a link $L$ is called the
554: \em{braid index} of the link, and will be denoted by $b(L)$.
555: 
556: In \cite{Bennequin}, Bennequin studied in relation to contact
557: structures braid representations by bands, independently considered
558: by Rudolph \cite{Rudolph} in the context of algebraic curves, and
559: more recently in \cite{BKL} from a group theoretic point of view.
560: A band representation naturally spans a Seifert surface of the
561: link. See figure \reference{figbd}.
562:  
563: \begin{figure}[h]
564: \[
565: % \begin{array}{c}
566: \epsfs{5cm}{fig1}
567: % \end{array}
568: \]\\[-14mm]
569: \caption{A braided Seifert surface obtained from a band representation
570: \label{figbd}}
571: \end{figure}
572:  
573: % In figure \reference{fig2}, for the move (a) $\to$ (b), we slide
574: % $B$ and $C$, respectively, along $D$ and $A$, then delete $D$
575: % (by deplumbing a Hopf band), and then slide $C$ back along
576: % $A$. For the move (b) $\to$ (c), we slide $C$ along $B$,
577: % then slide $B$ along $C$. The $k$ bands $A$ can be subsequently
578: % removed by Murasugi desumming a $(2,k)$-torus link fiber surface.
579: % Thus the surface $F$, spanned by $\hat\bt$, turns after
580: % (de)summing Hopf bands into a surface $F'$ consisting of $k$
581: % copies of $F'_0$ and one negative band $N$. See figure \ref{fig3} (a).
582: %  
583: % \begin{figure}[h]
584: % \[
585: % \begin{array}{cc}
586: %  (a) & (b) \\
587: % \end{array}
588: % \]
589: % \caption{\label{fig3}}
590: % \end{figure}
591: %  
592: % In figure \reference{fig3}, we perform the move (a) $\to$ (b), sliding
593: % $B$ along $N$. The last surface $F'$, in figure \reference{fig3} (b),
594: % is Murasugi sum of a fibre surface spanning the $(2,2,-2)$-pretzel
595: % link and $\tl F$, where $\tl F$ consists 
596: % of $k - 1$ copies of $F'_0$ and the band $N$. By induction on
597: % $k$, we see that $F'$ is a fibre surface, and hence $\hat\bt$
598: % is a fibred link. \qed
599: %  
600: % \end{appendix}
601: Bennequin proved in \cite{Bennequin} that a minimal genus band
602: representation always exists on 3-braids. Such surface was called
603: by Birman-Menasco \cite{BirMen2} a Bennequin surface. Later, by
604: examples found by Ko and Lee, and Hirasawa and myself, the
605: existence of Bennequin surfaces on the minimal number of strands
606: was found not to extend to 4-braids. (We will see in this paper,
607: though, among many other things, that it does extend for
608: alternating links.) Since by Rudolph's work \cite{Rudolph},
609: a Bennequin surface exists for any link, on a possibly large
610: number of strands, the answer to the question \reference{q1}
611: is negative for minimal genus band representations. 
612: 
613: A special case of band representations are the positive ones,
614: called strongly quasi-positive in \cite{Rudolph}. For such
615: representations the above examples do not apply. The existence
616: of minimal such representations is a question of Rudolph, whose
617: answer is not known. Our first result is a positive answer to
618: Rudolph's question for 3-braids. 
619: 
620: \begin{theorem}\label{thqp}
621: If $L$ is a strongly quasi-positive link, and has braid index 3,
622: then $L$ is has a strongly quasi-positive 3-braid representation.
623: \end{theorem}
624: 
625: This result is an application of the work in \cite{3br} on
626: Xu's normal form \cite{Xu}. It legitimizes the term ``strongly
627: quasi-positive 3-braid link'', which otherwise would be
628: linguistically sloppy, for it may mean a 3-braid link that is strongly
629: quasi-positive, or the link (which is the closure) of a strongly
630: quasi-positive 3-braid.
631: 
632: 
633: As an (\em{a-priori}-so-seeming) improvement of Bennequin's
634: theorem, Birman-Menasco proved \cite{BirMen2} that
635: for 3-braid links not only some but in fact \em{any} minimal genus
636: surface is a Bennequin surface. Xu used this result and the
637: classification in \cite{BirMen} to conclude that most 3-braid links
638: have a unique minimal genus surface. Unfortunately, she failed to
639: deal (in the oriented sense) with some exceptions, which we complete
640: in section \reference{SA} (see theorem \reference{wu}).
641: 
642: The following work involves a further study of Xu's normal form.
643: % A proof of this result is given by the authors in an appendix.
644: %
645: A natural relation of this form to the skein polynomial was
646: exhibited in \cite{3br}. Here we extend this study to the Alexander
647: polynomial $\Dl$. As a result, we can describe 3-braid links with
648: given $\Dl$, and in particular show no non-trivial 3-braid
649: knot has trivial polynomial. In \cite{Birman}, Birman constructed
650: some pairs of different 3-braid links with the same polynomials
651: and proposed the problem to understand the values of the various
652: link polynomials on 3-braid links. Our work solves (to a large
653: extent at least) Birman's problem for the Alexander polynomial.
654: We also have a solution for the Jones and Brandt-Lickorish-%
655: Millett-Ho $Q$ polynomial. 
656: 
657: Parallel to the Alexander polynomial description, and using
658: a recent result of Hirasawa-Murasugi, we classify among 3-braid
659: links the fibered ones. (This result was proved independently
660: by Yi Ni in \cite{Ni}, apparently unaware of this work, which
661: I did to make widely public immediately.) We will see then
662: that a 3-braid knot is fibered if (and only if) its Alexander
663: polynomial is monic, but that this is \em{not} true for 2- or
664: 3- component links.
665: 
666: The fact that most 3-braid links were found fibered,
667: and this could be proved independently from Bennequin and
668: Birman-Menasco, opened the hope for a more natural approach to
669: some of their results. Finally we were indeed able to complete this
670: project, with the assistance of M. Hirasawa and M. Ishiwata, obtaining
671: 
672: \begin{theorem}\label{thuq}
673: Any 3-braid link has a unique (non-closed) incompressible surface.
674: \end{theorem}
675: 
676: This result subsumes all previous uniqueness theorems. Besides,
677: its proof is entirely different, in that it fully avoids the contact
678: geometry approach of Bennequin and the considerable complexity
679: in Birman-Menasco's subsequent braid foliation work. Instead,
680: we use the sutured manifold theory of Gabai \cite{Gabai,Gabai2}
681: (to deal with the fibered links) and Kobayashi-Kakimizu \cite{%
682: Kobayashi,Kakimizu} (for the remaining cases), as well as some
683: of our preceding work on the Alexander polynomial (to rule out
684: disconnected Seifert surfaces). Since Kobayashi-Kakimizu use
685: a slightly stronger notion of uniqueness (see the beginning of
686: \S\reference{Uq}), theorem \reference{thuq} is an improvement even
687: for minimal genus surfaces. Moreover, its proof underscores the
688: geometric meaning of Xu's form, that remained unclear in \cite{BirMen}.
689: While we still cannot provide an ``easy'' proof of Birman-Menasco's
690: classification theorem (see remark on p.~34 of \cite{BirMen}), our
691: work will likely lead to simplifications in their very lengthy
692: treatment, which makes heavy use of incompressible surfaces.
693: 
694: The study and applications of Xu's form occupy the first main
695: part of the paper. Then some related results are included,
696: whose treatment requires different methods. These
697: results are dealt with in the later sections of the paper.
698: 
699: As a first such topic, more substantial effort will be made to
700: prove an analogous result to theorem \reference{thqp} for
701: positive (in the usual Artin generators) braid representations.
702: 
703: \begin{theorem}\label{thg}
704: If $L$ is a positive braid link, and has braid index 3, then 
705: $L$ is a positive 3-braid link.
706: \end{theorem}
707: 
708: So our work here can be understood to answer question
709: \reference{q1} for braid index 3 and strongly quasi-positive
710: or positive braid links. (The state of quasi-positive links,
711: and of quasi-positive braid representations with regard to 
712: question \reference{q1}, in contrast remains open. There is,
713: however, an algorithm, found by orevkov \cite{Orevkov}
714: to decide if a 3-braid is quasi-positive.)
715: 
716: The proof of theorem \reference{thg} requires a study of positive
717: braids where the braid index bound in the Morton-Williams-Franks
718: inequality \cite{Morton,WilFr} is 3. This study builds on and extends
719: (but also considerably simplifies) the work of Nakamura
720: \cite{Nakamura} for Morton-Williams-Franks bound 2. Then some
721: detailed case distinction and calculation are necessary.
722: In \cite{posex_bcr} it was shown, again, that theorem
723: \reference{thg} is not true for 4-braids: there are two 16 crossing
724: knots that have positive 16 crossing 5-braid representations,
725: but braid index 4 (and consequently only non-positive 4-braid
726: representations). It is in fact for these examples that I was
727: led to investigate about the theorem.
728: 
729: In section \ref{Spos},
730: %we come back to theorem \reference{thg}
731: % and show how to obtain its alternative proof
732: %. As a partial step, which is equally
733: % noteworthy,
734: using the arguments in the proof of theorem \reference{thg}
735: and a criterion of Yokota \cite{Yokota}, we consider 3-braid
736: representations of links that are (diagrammatically) positive.
737: We obtain strong restrictions on such representations,
738: and in particular determine which of the links are not fibered.
739: (See theorem \ref{thposq}.)
740: 
741: % We will show two completely different approaches to
742: % theorem \reference{thg}. The first uses the
743: % Morton-Williams-Franks inequality and tries to identify
744: % the positive braids with Morton-Williams-Franks bound 3,
745: % extending the work of Nakamura \cite{Nakamura} for
746: % bound 2. The proof is slightly technical, and requires
747: % the verification of certain special cases.
748: % Another proof, following a longer, but much
749: % more insightful road, is then given by continuing
750: % the arguments that establish theorem \reference{thqp}.
751: 
752: Another problem proposed by Birman, in \cite{MortonPb},
753: was to understand the relation of alternation of links
754: (as a pre-eminent diagram defined property) and braid
755: representations. Early substantial results on the braid
756: index of alternating links were due to Murasugi \cite{Murasugi},
757: who determined the braid index for rational and fibered
758: alternating links. In \cite{3br} we used Xu's form to classify
759: the braid index 3 alternating links. Here, in \S\reference{Sfg},
760: we will easily recover this result and push it forward to braid
761: index 4. For this we use an argument based on the Jones polynomial,
762: and connect the celebrated Kauffman-Murasugi-Thistlethwaite work
763: \cite{Kauffman2,Murasugi3,Thistle2} to braid representations.
764: The existence of the Bennequin surface of alternating links on a
765: 4 string braid (corollary \reference{cfT}) is among several easy
766: consequences we obtain.
767: 
768: The last section \S\reference{SBurau} discusses some applications
769: of the Squier unitarization \cite{Squier} of the Burau
770: representation, that concern also braids on more strings. A
771: first series of conditions are estimates on the norm of special
772: values of the Alexander and Jones polynomial. They relate to and
773: partly extend estimates of Jones in \cite{Jones}. In particular, we
774: will see that the Alexander polynomial provides conditions for every
775: given braid index. Some properties of the skein polynomial, and a
776: relation to Mahler measure are also discussed.
777: 
778: In the case of a 4-braid, one can go further and (almost)
779: identify the Eigenvalues of the Burau matrix
780: from the Alexander and Jones polynomial (of its closure).
781: Using this, a criterion for braid index 4 is derived.
782: We show examples exhibiting the efficiency of this test, including
783: such where not only the Morton-Williams-Franks inequality itself,
784: but also its 2-cable version fails (and so our test seems the only
785: practicable option).
786: 
787: % and is included here to complete the picture.
788: 
789: In the appendix we collect some work of Hirasawa, Ishiwata and
790: Murasugi, which completes the proof of several of our results.
791: 
792: \section{Preliminaries, basic definitions and conventions}
793: % \noindent{\bf Convention.} 
794: 
795: Basic concepts that appear throughout the paper are summarized.
796: `W.l.o.g.' will abbreviate `without loss of generality';
797: `r.h.s.' will stand for `right hand-side'.
798: 
799: \subsection{Links and link diagrams\label{Sdg}}
800: 
801: Links are represented by diagrams; we assume diagrams are
802: oriented (though sometimes orientation is not relevant).
803: 
804: A crossing $p$ in a link diagram $D$ is called \em{reducible} (or
805: nugatory) if it looks like on the left of figure \reference{figred}.
806: $D$ is called reducible if it has a reducible crossing, else it is
807: called \em{reduced}. The reducing of the reducible crossing $p$
808: is the move depicted on figure \reference{figred}. Each diagram $D$
809: can be (made) reduced by a finite number of these moves.
810: 
811: \begin{figure}[htb]
812: \begin{eqn}\label{eqred}
813: \diag{6mm}{4}{2}{
814:   \picrotate{-90}{\rbraid{-1 2}{1 1.4}}
815:   % \piccirclearc{2 0.5}{1.3}{45 135}
816:   % \piccirclearc{2 1.5}{1.3}{-135 -45}
817:   \picputtext[u]{2 0.7}{$p$}
818:   \picscale{1 -1}{
819:     \picfilledcircle{0.7 -1}{0.8}{$P$}
820:   }
821:   \picfilledcircle{3.3 1}{0.8}{$Q$}
822: } \qquad\lra\qquad
823: \diag{6mm}{4}{2}{
824:   \piccirclearc{2 0.5}{1.3}{45 135}
825:   \piccirclearc{2 1.5}{1.3}{-135 -45}
826:   \picfilledcircle{1 1}{0.8}{$P$}
827:   \picfilledcircle{3 1}{0.8}{$Q$}
828: } 
829: \end{eqn}
830: \caption{\label{figred}}
831: \end{figure}
832: 
833: We assume in the following all diagrams reduced, unless otherwise
834: stated.
835: 
836: The diagram on the right of figure \reference{figtan}
837: is called \em{connected sum} $A\# B$ of the diagrams $A$ and $B$.
838: If a diagram $D$ can be represented as the connected sum of 
839: diagrams $A$ and $B$, such that both $A$ and $B$ have at least one
840: crossing, then $D$ is called \em{disconnected} (or composite), else
841: it is called \em{connected} (or prime). $K$ is \em{prime} if whenever
842: $D=A\# B$ is a composite diagram of $K$, one of $A$ and $B$
843: represent an unknotted arc (but not both; the unknot is
844: not considered to be prime per convention).
845: 
846: By $c(D)$ we denote the
847: \em{number of crossings} of $D$, $n(D)$ the number of components
848: of $D$ (or $K$, $1$ if $K$ is a knot), and $s(D)$ the \em{number
849: of Seifert circles} of $D$. The \em{crossing number} $c(K)$ of a
850: knot or link $K$ is the minimal crossing number of all diagrams
851: $D$ of $K$.
852: $!D$ is the \em{mirror image} of $D$, and $!K$ is the mirror image
853: of $K$. Clearly $g(!D)=g(D)$ and $g(!K)=g(K)$.
854: 
855: \begin{figure}[htb]
856: \[
857: \diag{6mm}{3}{2}{
858:   \piccirclearc{1.8 1}{0.5}{-120 120}
859:   \picfilledcircle{1 1}{0.8}{$A$}
860: }\, \#\,
861: \diag{6mm}{3}{2}{
862:   \piccirclearc{1.2 1}{0.5}{60 300}
863:   \picfilledcircle{2 1}{0.8}{$B$}
864: }\quad =\quad
865: \diag{6mm}{4}{2}{
866:   \piccirclearc{2 0.5}{1.3}{45 135}
867:   \piccirclearc{2 1.5}{1.3}{-135 -45}
868:   \picfilledcircle{1 1}{0.8}{$A$}
869:   \picfilledcircle{3 1}{0.8}{$B$}
870: } 
871: \]
872: \caption{\label{figtan}}
873: \end{figure}
874: 
875: The \em{Seifert graph} $\Gm(D)$ of a diagram $D$ is defined to
876: be the graph whose vertices are the Seifert circles of $D$ and
877: whose edges are the crossings. The \em{reduced Seifert graph}
878: $\Gm'(D)$ is defined by removing multiple copies of an edge
879: between two vertices in $\Gm(D)$, so that a simple edge
880: remains. (See \cite{MurPrz,MurPrz2} for example.)
881: 
882: The \em{(Seifert) genus} $g(K)$ resp.\ \em{Euler characteristic}
883: $\chi(K)$ of a knot or link $K$ is said to be the minimal genus
884: resp.\ maximal Euler characteristic of Seifert surface of $K$.
885: For a diagram $D$ of $K$, $g(D)$ is defined to be the genus
886: of the Seifert surface obtained by Seifert's algorithm on $D$,
887: and $\chi(D)$ its Euler characteristic. We have $\chi(D)=s(D)-c(D)$
888: and $2g(D)=2-n(D)-\chi(D)$.
889: 
890: The numbering of knots we use is as in the tables of \cite
891: [appendix]{Rolfsen} for prime knots of crossing number $\le 10$,
892: and as in \cite{KnotScape} for those of crossing number $11$
893: to $16$. KnotScape's numbering is reorganized so that for
894: given crossing number non-alternating knots are appended after
895: alternating ones, instead of using `a' and `n' superscripts.
896: 
897: \subsection{Polynomial link invariants\label{brD}}
898: 
899: Let $X\in\bZ[t,t^{-1}]$. The \em{minimal} or \em{maximal degree} 
900: $\md X$ or $\Md X$ is the minimal resp.\ maximal exponent of $t$ 
901: with non-zero coefficient in $X$. Let $\spn_tX=\Md_tX-\md_tX$.
902: The coefficient in degree $d$ of $t$ in $X$ is denoted $[X]_{t^d}$
903: or $[X]_{d}$. The \em{leading coefficient} $\Mc\,X$ of $X$ is its
904: coefficient in degree $\Md X$. If $X\in\bZ[x_1^{\pm 1},x_2^{\pm 1}]$,
905: then $\Md_{x_1}X$ denotes the maximal degree in $x_1$. Minimal
906: degree and coefficients are defined similarly, and of course
907: $[X]_{x_1^k}$ is regarded as a polynomial in $x_2^{\pm 1}$.
908: 
909: Let $P(v,z)$ be the \em{skein polynomial} \cite{HOMFLY,LickMil}.
910: It is a Laurent polynomial in two variables of oriented knots and
911: links. We use here the convention of \cite{Morton}, i.e. with the
912: polynomial taking the value $1$ on the unknot, having the variables
913: $v$ and $z$ and satisfying the skein relation
914: \begin{eqn}\label{1}
915: v^{-1}\,P\bigl(
916: \diag{5mm}{1}{1}{
917: \picmultivecline{-5 1 -1.0 0}{1 0}{0 1}
918: \picmultivecline{-5 1 -1.0 0}{0 0}{1 1}
919: }
920: \bigr)\,-\,
921: v \,P\bigl(
922: \diag{5mm}{1}{1}{
923: \picmultivecline{-5 1 -1 0}{0 0}{1 1}
924: \picmultivecline{-5 1 -1 0}{1 0}{0 1}
925: }
926: \bigr)\,=\,
927: z\,P\bigl(
928: \diag{5mm}{1}{1}{
929: \piccirclevecarc{1.35 0.5}{0.7}{-230 -130}
930: \piccirclevecarc{-0.35 0.5}{0.7}{310 50}
931: }
932: \bigr)\,.
933: \end{eqn}
934: 
935: We will denote in each triple as in \eqref{1} 
936: the diagrams (from left to right) by $D_+$, $D_-$ and $D_0$.
937: For a diagram $D$ of a link $L$, we will use all of the notations
938: $P(D)=P_D=P_D(l,m)=P(L)$ etc.\ for its skein polynomial, with
939: the self-suggestive meaning of indices and arguments. So we
940: can rewrite \eqref{1} as
941: \begin{eqn}\label{srel}
942: v^{-1}P_+(v,z)-vP_-(v,z) = zP_0(v,z)\,.
943: \end{eqn}
944: 
945: The \em{writhe} is a number ($\pm1$), assigned to any crossing in a
946: link diagram. A crossing as on the left in \eqref{1} has writhe 1 and
947: is called \em{positive}. A crossing as in the middle of \eqref{1} has
948: writhe $-1$ and is called \em{negative}. The writhe of a link diagram
949: is the sum of writhes of all its crossings.
950: 
951: Let $c_{\pm}(D)$ be the number of positive, respectively negative
952: crossings of a diagram $D$, so that $c(D)=c_+(D)+c_-(D)$ and
953: $w(D)=c_+(D)-c_-(D)$. 
954: 
955: The \em{Jones polynomial} \cite{Jones2} $V$, and (one variable)
956: \em{Alexander polynomial} \cite{Alexander} $\Dl$ are obtained
957: from $P$ by variable substitutions
958: \begin{eqn}\label{DP}
959: \Dl(t)=P(1,t^{1/2}-t^{-1/2})\,,
960: \end{eqn}
961: and
962: \begin{eqn}\label{VPsub}
963: V(t)=P(t,t^{1/2}-t^{-1/2})\,.
964: \end{eqn}
965: 
966: Hence these polynomials also satisfy corresponding skein relations.
967: (In algebraic topology, the Alexander polynomial is usually
968: defined only up to units in $\bZ[t,t^{-1}]$; the present
969: normalization is so that $\Dl(t)=\Dl(1/t)$ and $\Dl(1)=1$.)
970: 
971: In very contrast to its relatives, the \em{range} of the 
972: Alexander polynomial (i.e., set of values it takes) is known.
973: Let us call a polynomial $\Dl\in\bZ[t^{1/2},t^{-1/2}]$
974: \em{admissible} if it satisfies for some natural number
975: $n\ge 1$ the three properties\\
976: \mbox{}\ \ (i)\quad $t^{(n-1)/2}\Dl\in \bZ[t^{\pm 1}]$,\\
977: \mbox{}\ (ii)\quad  $\Dl(t)=(-1)^{n-1}\Dl(1/t)$ and \\
978: \mbox{}(iii)\quad   $(t^{1/2}-t^{-1/2})^{n-1}\mid
979:                     \Dl$ for $n>1$, or $\Dl(1)=1$ for $n=1$. \\
980: It is well-known that these are exactly the polynomials that occur
981: as (1-variable) Alexander polynomials of some $n$-component link. 
982: 
983: The \em{Kauffman polynomial} \cite{Kauffman} $F$ is usually defined
984: via a regular isotopy invariant $\Lm(a,z)$ of unoriented links.
985: We use here a slightly different convention for the variables
986: in $F$, differing from \cite{Kauffman,Thistle2} by the interchange
987: of $a$ and $a^{-1}$. Thus in particular we have for a link diagram $D$
988: the relation $F(D)(a,z)=a^{w(D)}\Lm(D)(a,z)$, where $\Lm(D)$ is the
989: writhe-unnormalized
990: version of the polynomial, given in our convention by the properties
991: \begin{eqnarray}
992: & \Lambda\bigl(\Pos{0.5cm}{}\bigr)\ +\ \Lambda\bigl(\Neg{0.5cm}{}\bigr)\ =\ z\ \bigl(\ 
993: \Lambda\bigl(\Nul{0.5cm}{}\bigr)\ +\ \Lambda\bigl(\Inf{0.5cm}{}\bigr)\ \bigr)\,,\label{wseven}\\[2mm]
994: & \label{wseven.5} \Lambda\bigl(\ \ReidI{-}{-}\bigr) = a^{-1}\ \Lambda\bigl(\noloop\bigr);\quad
995: \Lambda\bigl(\ \ReidI{-}{ }\bigr) = a\ \Lambda\bigl(\noloop\bigr)\,,\\[2mm]
996: & \label{wseven.7} \Lambda\bigl(\,\mbox{\Large $\bigcirc$}\,\bigr) = 1\,.
997: \end{eqnarray}
998: 
999: The \em{Brandt-Lickorish-Millett-Ho polynomial} \cite{BLM} $Q$
1000: is given by $Q(z)=F(1,z)$.
1001: 
1002: \subsection{Semiadequacy\label{Ssaq}}
1003: 
1004: An alternative description of $V$ is given by the Kauffman bracket in
1005: \cite{Kauffman2}. We do not need this description directly, but
1006: for self-containedness is it useful to recall the related concept of
1007: semiadequacy that was popularized in \cite{LickThis}.
1008: 
1009: Let $D$ be an unoriented link diagram.
1010: A \em{state} is a choice of splittings of type $A$ or 
1011: $B$ for any single crossing (see figure \ref{figsplit}), 
1012: Let the \em{$A$-state} of $D$ be the state where all
1013: crossings are $A$-spliced; similarly define the $B$-state.
1014: 
1015: We call a diagram \em{$A$-(semi)adequate} if in the $A$-state
1016: no crossing trace (one of the dotted lines in figure
1017: \reference{figsplit}) connects a loop with itself.
1018: Similarly we define $B$-(semi)adequate. A diagram is
1019: \em{semiadequate} if it is $A$- or $B$-semiadequate, and
1020: \em{adequate} if it is simultaneously $A$- and $B$-semiadequate.
1021: A link is adequate/semiadequate if it has an adequate/semiadequate
1022: diagram. (Here $A$-adequate and $B$-adequate is what is called
1023: $+$adequate resp.\ $-$adequate in \cite{Thistle}.)
1024: 
1025: \begin{figure}[htb]
1026: \[
1027: \diag{8mm}{4}{2}{
1028:    \picline{0 0}{4 2}
1029:    \picmultiline{-5.0 1 -1.0 0}{0 2}{4 0}
1030:    \picputtext{2.7 1}{A}
1031:    \picputtext{1.3 1}{A}
1032:    \picputtext{2 1.6}{B}
1033:    \picputtext{2 0.4}{B}
1034: } \quad
1035: \diag{8mm}{4}{2}{
1036:    \pictranslate{2 1}{
1037:        \picmultigraphics[S]{2}{1 -1}{
1038:            \piccurve{-2 1}{-0.3 0.2}{0.3 0.2}{2 1}
1039:        }
1040:        {\piclinedash{0.05 0.05}{0.01}
1041:         \picline{0 -0.4}{0 0.4}
1042:        }
1043:    }
1044:    \picputtext{2.7 1}{A}
1045:    \picputtext{1.3 1}{A}
1046: } \quad
1047: \diag{8mm}{4}{2}{
1048:    \pictranslate{2 1}{
1049:        \picmultigraphics[S]{2}{-1 1}{
1050:            \piccurve{2 -1}{0.1 -0.5}{0.1 0.5}{2 1}
1051:        }
1052:        {\piclinedash{0.05 0.05}{0.01}
1053:         \picline{0 -0.6 x}{0 0.6 x}
1054:        }
1055:    }
1056:    \picputtext{2 1.6}{B}
1057:    \picputtext{2 0.4}{B}
1058: }
1059: \]
1060: \caption{\label{figsplit}The A- and B-corners of a
1061: crossing, and its both splittings. The corner A (resp. B)
1062: is the one passed by the overcrossing strand when rotated 
1063: counterclockwise (resp. clockwise) towards the undercrossing 
1064: strand. A type A (resp.\ B) splitting is obtained by connecting 
1065: the A (resp.\ B) corners of the crossing. It is useful to
1066: put a ``trace'' of each splitted crossing as an arc connecting
1067: the loops at the splitted spot.}
1068: \end{figure}
1069: 
1070: We also need a part of the semiadequacy formulas for the Jones
1071: polynomial. As in \cite{ntriv}, for an non-negative integer $i$, we
1072: write $V_i$ for $[V]_{\md V+i}$ and $\bar V_i=[V]_{\Md V-i}$. (These 
1073: are the $i+1$-st and $i+1$-st last coefficient of $V$ in its
1074: coefficient list.) For an A- (resp. B-) semiadequate diagram, 
1075: in \cite{ntriv}, and independently in \cite{DasLin}, formulas were 
1076: obtained for $V_{1,2}$ (resp. $\bar V_{1,2}$).
1077: 
1078: \parbox[b]{12.5cm}{
1079: \parskip 5pt plus 3pt minus 2pt\relax
1080: {}From the $A$-state $A(D)$ of a diagram $D$ we define two
1081: graphs.  The first graph, we call it the \em{$A$-graph}
1082: $G(A)=G(A(D))$ has vertices for each loop in $A(D)$, and
1083: an edge between each pair of loops that are connected
1084: by a trace of at least one crossing in $D$. If there are
1085: at least two such traces, we call the edge \em{multiple}.  
1086: 
1087: Let $\bigtriangleup(D)=\bigtriangleup(A(D))$ be the number of
1088: cycles of length 3 (\em{triangles}) in $G(A(D))$.  
1089: 
1090: The second graph, we call it the \em{intertwining graph}
1091: $IG(A)=IG(A(D))$, has vertices for each multiple edge
1092: in $A(D)$, and an edge in $IG(A)$ is drawn between each
1093: pair of vertices in $IG(A)$, whose corresponding multiple
1094: traces in $A(D)$ contain traces of the form shown on the right.
1095: }
1096: \enspace\parbox[t]{4.0cm}{
1097: \vbox to 4.0cm{\vss
1098: \[
1099: \diag{10mm}{4}{4}{
1100:   \picline{2 0}{2 4}
1101:   {\piclinedash{0.1 0.1}{0.01}
1102:    \picline{1 0.8}{2 0.8}
1103:    \picline{3 1.6}{2 1.6}
1104:    \picline{1 2.4}{2 2.4}
1105:    \picline{3 3.2}{2 3.2}
1106:   }
1107:   \picfilledellipse{1.6 1 x}{0.3 1.2}{}
1108:   \picfilledellipse{2.4 3 x}{0.3 1.2}{}
1109: }
1110: \]\vss}
1111: }
1112: 
1113: \begin{theorem}(\cite{ntriv,DasLin})\label{T02}
1114: Assume that $D$ is a connected $A$-adequate diagram. Then
1115: $V_0V_1=\chi(G(A))-1$ and
1116: \[
1117: V_0V_2=\mybin{2-\chi(G(A))}{2}+\chi(IG(A))-\bigtriangleup(A(D))\,.
1118: \]
1119: \end{theorem}
1120: 
1121: Here $\chi$ is the Euler characteristic (number of vertices minus
1122: number of edges); $V_0=\pm 1$ by \cite{LickThis}.
1123: Analogous formulas hold for $B$-adequate diagrams.
1124: 
1125: \subsection{Braids and braid words\label{brs}}
1126: 
1127: The $n$-string \em{braid group} $B_n$ is considered generated by
1128: the Artin \em{standard generators} $\sg_i$ for $i=1,\dots,n-1$. These
1129: are subject to relations of the type $[\sg_i,\sg_j]=1$ for $|i-j|>1$,
1130: which we call \em{commutativity relations} (the bracket denotes the
1131: commutator) and $\sg_{i+1}\sg_i\sg_{i+1}= \sg_i\sg_{i+1}\sg_i$,
1132: which we call \em{Yang-Baxter} (or shortly YB) \rm{relations}.
1133: 
1134: We will make one noteworthy modification of this notation.
1135: In the following $\sg_3$ will stand for the usual Artin generator
1136: for braids of $4$ or more strands (such braids are considered
1137: explicitly only in sections \reference{S2.2} and \reference{SBurau}),
1138: while on $3$-braids (considered in all other sections) it will denote
1139: the ``band'' generator $\sg_2\sg_1\sg_2^{-1}=\sg_1^{-1}\sg_2\sg_1$.
1140: 
1141: It will be often convenient in braid words to write $\pm i$ for
1142: $\sg_i^{\pm 1}$. For example, $[21(33-4)^232]$ is an alternative
1143: writing for $\sg_2\sg_1(\sg_3^2\sg_4^{-1})^2\sg_3\sg_2$.
1144: 
1145: The following definition summarizes basic terminology of
1146: braid words used throughout the paper.
1147: 
1148: \begin{defi}\label{dfext}
1149: Choose a word 
1150: \begin{eqn}\label{wd}
1151: \bt=\prod_{i=1}^n\sg_{k_i}^{l_i}
1152: \end{eqn}
1153: with $k_i\ne k_{i+1}$ and $l_i\ne 0$. We understand such words
1154: \em{in cyclic order}.
1155: 
1156: Call $\sg_{k_i}^{l_i}$ the \em{syllables}
1157: of $\bt$. For such a syllable, let $k_i$ be called the \em{index}
1158: of the syllable, and $l_i$ its \em{exponent} or \em{length}.
1159: We call a syllable $\sg_{k_i}^{l_i}$ \em{non-trivial} if $|l_i|>1$
1160: and \em{trivial} if $|l_i|=1$.
1161: 
1162: We say $n$ is the \em{syllable length} of $\bt$ in \eqref{wd}.
1163: The \em{(word) length} $c(\bt)$ of $\bt$ is $\sum_{i=1}^n |l_i|$, and
1164: $c_{\pm}(\bt)=\sum_{\pm l_i>0}\pm l_i$ are the \em{positive/negative
1165: length} of $\bt$. A word is \em{positive} if $c_-(\bt)=0$, or
1166: equivalently, if all $l_i>0$. The \em{exponent sum} of $\bt$ is
1167: defined to be $[\bt]=\sum_{i=1}^n l_i$. The \em{index sum} of
1168: $\bt$ is $\sum_{i=1}^n k_i\cdot |l_i|$ (i.e.
1169: each letter, not just syllable in $\bt$, contributes to that sum).
1170: 
1171: For $3$-braid words, the $k_i$ interchange between $1$ and $2$.
1172: Thus the vector of the $l_i$, considered up to cyclic permutations,
1173: determines the conjugacy class unambiguously. We call it
1174: the \em{Schreier vector}.
1175: 
1176: Let $\bt$ be a positive word.
1177: We call $\tl\bt$ an \em{extension} of $\bt$ if $\tl\bt$ is obtained by
1178: replacing some (possibly no) trivial syllables in $\bt$ by non-trivial
1179: ones of the same index. Contrarily, we call $\bt$ a \em{syllable
1180: reduction} of $\tl\bt$. We call $\bt$ \em{non-singular} if $[\bt]_k
1181: >1$ for all $k=1,\dots,n-1$, where $[\bt]_k=\sum_{k_i=k}l_i$ is the
1182: exponent sum of $\sg_k$ in $\bt$. If $[\bt]_k=1$, we say that the
1183: syllable of index $k$ in $\bt$ is \em{isolated} or \em{reducible}.
1184: \end{defi}
1185: 
1186: To avoid confusion, it seems useful to clarify \em{a priori}
1187: the following use of symbols (even though we recall it at
1188: appropriate places later).
1189: 
1190: A comma separated list of integers
1191: will stand for a \em{sequence of syllable indices} of a braid word. 
1192: The `at' sign `@', written after such an index means that the
1193: corresponding syllable is trivial, while by an exclamation mark
1194: `$!$' we indicate that the syllable is non-trivial. (As the
1195: exponent for non-trivial syllables will be immaterial, it is
1196: enough to distinguish only whether the syllable is trivial or not.)
1197: If none of $!$ and @ is specified, we do not exclude explicitly
1198: any of either types. 
1199: 
1200: A bracketed but non-comma separated list of integers will stand
1201: for a braid word. An asterisk `$*$' put after a letter (number)
1202: in such a word means that this letter may be repeated
1203: (it need not be repeated, but it must not be omitted).
1204: So a (possibly trivial) index-2 syllable can be written
1205: as $22*$. The expression `$[23]$' should mean a letter which
1206: is either `2' or `3', and `$[23]+$' means a possibly empty
1207: sequence of letters `$2$' and `$3$'.
1208: 
1209: \subsection{Braid representations of links\label{brt}}
1210: 
1211: By a theorem of Alexander \cite{Alexander}, any link is the \em{closure}
1212: $\hat\bt$ of a braid $\bt$. The \em{braid index} $b(L)$ of a link $L$
1213: is the smallest number of strands of a braid $\be$ whose closure
1214: $\hat\be$ is $L$. See \cite{Morton,WilFr,Murasugi}. Such $\bt$ are
1215: also called \em{braid representations} of $L$. The closure operation
1216: gives for a particular braid word $\bt$ a link diagram $D=\hat\bt$.
1217: Then we have for example $w(\hat\bt)=[\bt]$, $c(\bt)=c(\hat\bt)$,
1218: $c_\pm(\bt)=c_\pm(\hat\bt)$, and $s(\hat\bt)$
1219: is the number of strings of $\bt$ (i.e. $n$ for $\bt\in B_n$).
1220: 
1221: Many properties of braid words we will deal with
1222: relate to the corresponding properties of their link diagrams.
1223: For example, a braid word is called \em{positive} if it contains no
1224: $\sg_i^{-1}$, or in other words, its closure diagram is positive.
1225: A braid is positive if it has a positive braid word.
1226: In a similar fashion, we say that braid is ($A$/$B$-)adequate if
1227: it has an ($A$/$B$-)adequate word representation, and a word is
1228: adequate if the link diagram obtained by its closure is adequate. 
1229: 
1230: If a braid word $\bt$ is written as $\sg_1^{\pm 1}\ap\sg_1^p\ap'$,
1231: where $p\in\bZ$ and none of $\ap$ and $\ap'$ contains a syllable of
1232: index $1$, then the diagram admits a \em{flype}, which exchanges
1233: the syllables of index $1$ in $\bt$, so that we obtain $\sg_1^p
1234: \ap\sg_1^{\pm 1}\ap'$. This operation preserves the isotopy
1235: type of the closure link $\hat\bt$, but in general changes
1236: the braid conjugacy class. The phenomenon is explained in
1237: \cite{BirMen}. In the context of general link diagrams,
1238: the flype has been studied also extensively, most prominently
1239: in \cite{MenThis}.
1240: 
1241: Alternatively to the standard Artin generators, one
1242: considers also a representation of the braid groups by
1243: means of an extended set of generators (and their inverses)
1244: \[
1245: \sg_{i,j}^{\pm 1}=\sg_i\dots\sg_{j-2}\sg_{j-1}^{\pm 1}
1246: \sg_{j-2}^{-1}\dots\sg_i^{-1}
1247: \]
1248: for $1\le i<j\le n$. Note that 
1249: \begin{eqn}\label{op}
1250: \sg_i=\sg_{i,i+1}\,.
1251: \end{eqn}
1252: A representation of a braid $\bt$, and its closure link $L=\hat\bt$,
1253: as word in $\sg_{i,j}^{\pm 1}$ is called a \em{band representation}
1254: \cite{BKL}. A band representation of $\bt$ spans naturally a
1255: Seifert surface of the link $L$ as in figure \ref{figbd}: one
1256: glues disks into the strands, and connects them by half-twisted
1257: bands along the $\sg_{i,j}$. The resulting surface is called
1258: \em{braided Seifert surface} of $L$. A minimal genus Seifert
1259: surface of $L$, which is a braided Seifert surface, is also
1260: called a \em{Bennequin surface} \cite{BirMen2}.
1261: 
1262: In this paper we will deal exclusively with band representations
1263: in $B_3$. Then we have three band generators $\sg_{i,i\bmod 3+1}$
1264: (where $i=1,2,3$, and `$\bmod$' is taken with values between $0$ and
1265: $2$). With \eqref{op}, we have $\sg_1=\sg_{1,2}$ and $\sg_2=\sg_{2,3}$,
1266: and with the special meaning of $\sg_3\in B_3$ introduced above,
1267: $\sg_3=\bar\sg_{1,3}$, where bar denotes the mirror image.
1268: (This mirroring is used here for technical reasons related to
1269: Xu's normal form, as explained below.)
1270: 
1271: If a band representation contains only positively half-twisted bands
1272: (i.e. no $\sg_{i,j}^{-1}$ occur), it is called \em{band-positive}
1273: or \em{strongly quasi-positive}. A link with a strongly quasi-positive
1274: band (braid) representation is called strongly quasi-positive.
1275: Such links have an importance in connection to algebraic curves;
1276: see \cite{Rudolph2}.
1277: 
1278: Using the skein polynomial, define a quantity by
1279: \begin{eqn}\label{mwfb}
1280: MWF(L)\,=\,\frac 12\,\big(\Md_vP-\md_vP\big)+1\,.
1281: \end{eqn}
1282: The \em{Morton-Williams-Franks braid index inequality}
1283: \cite{Morton,WilFr} (abbreviated as MWF) states that
1284: \begin{eqn}
1285: b(L)\ge MWF(L)
1286: \end{eqn}
1287: for every link $L$. This inequality is often exact (i.e. an
1288: equality). The study of links where it is exact or not
1289: has occupied a significant part of previous literature.
1290: Most noteworthy is the work of Murasugi \cite{Murasugi}
1291: and Murasugi-Przytycki \cite{MurPrz2}.
1292: 
1293: The Morton-Williams-Franks 
1294: results from two other inequalities, due to Morton,
1295: namely that for a diagram $D$, we have
1296: \begin{eqn}\label{mq}
1297: 1-s(D)+w(D)\,\le\,\md_lP(D)\,\le\,\Md_lP(D)\,\le\,s(D)-1+w(D)\,.
1298: \end{eqn}
1299: Williams-Franks showed these inequalities for the case of braid
1300: representations (i.e. when $D=\hat\bt$ for some braid $\bt$).
1301: Later it was observed from the
1302: algorithm of Yamada \cite{Yamada} and Vogel \cite{Vogel}
1303: that the braid version is actually equivalent to (and not just
1304: a special case of) the diagram version.
1305: 
1306: These inequalities were later improved in \cite{MurPrz2} in a way 
1307: that allows to settle the braid index problem for many links
1308: (see \S\ref{Sfg} or also \cite{Ohyama}). 
1309: 
1310: In the special case of $3$-braids (and with the special meaning
1311: of $\sg_3$ as described above), Xu \cite{Xu} gives a normal form
1312: of a conjugacy class in $\sg_{1,2,3}$. By Xu's algorithm, each
1313: $\bt\in B_3$ can be written in one of the two forms
1314: {\def\theenumi{\Alph{enumi}}
1315: \def\labelenumi{(\theenumi)}
1316: \begin{enumerate}
1317: \item\label{itA} $[21]^kR$ or $L^{-1}[21]^{-k}$ ($k\ge 0$), or
1318: \item\label{itB} $L^{-1}R$,
1319: \end{enumerate}
1320: }where $L$ and $R$ are positive words in $\sg_{1,2,3}$ with (cyclically)
1321: non-decreasing indices (i.e. each $\sg_i$ is followed by $\sg_i$
1322: or $\sg_{i\bmod 3+1}$, with `$\bmod$' taken with values between $0$
1323: and $2$). Since the form \reference{itB} must be cyclically
1324: reduced, we may assume that $L$ and $R$ do not start or end with
1325: the same letter. This form is the shortest word in $\sg_{1,2,3}$
1326: of a conjugacy class. By Bennequin's aforementioned result, the
1327: braided surface is then a minimal genus (or Bennequin) surface.
1328: 
1329: \subsection{Gau\ss{} sum invariants}
1330: 
1331: We recall briefly the definition of Gau\ss{} sum invariants.
1332: They were introduced first in \cite{Fied} for braids, and later
1333: \cite{Fied2,VirPol} for knots. It is known that all
1334: they give formulas for Vassiliev invariants.
1335: 
1336: \begin{defi}(\protect\cite{Fied2})\ 
1337: A Gau\ss{} diagram of a knot diagram is an oriented circle with
1338: arrows connecting points on it mapped to a crossing and
1339: oriented from the preimage of the undercrossing (underpass)
1340: to the preimage of the overcrossing (overpass).
1341: \end{defi}
1342: 
1343: We will call a pair of crossings whose arrows intersect in
1344: the Gau\ss{} diagram a \em{linked} pair.
1345: 
1346: \begin{exam}
1347: As an example, figure \ref{fig6_2} shows the knot $6_2$ in its
1348: commonly known projection and the corresponding Gau\ss{} diagram.
1349: \end{exam}
1350: 
1351: \begin{figure}[htb]
1352: {%\small
1353: \[
1354: \begin{array}{c@{\qquad}c}
1355:      \diag{2cm}{2.0}{2.0}{
1356:   \picputtext[dl]{0 0}{\autoepsffalse\epsfs{4cm}{6_2}}
1357:   \picputtext[u]{1.25 1.65}{$1$} 
1358:   \picputtext[u]{0.8 1.65}{$2$} 
1359:   \picputtext[u]{0.6 0.6}{$3$} 
1360:   \picputtext[u]{1.0 0.35}{$4$} 
1361:   \picputtext[u]{1.3 0.9}{$5$} 
1362:   \picputtext[u]{0.9 1.3}{$6$} 
1363: 	}
1364:  & 
1365: \diag{1.6cm}{2.4}{2.4}{\pictranslate{1.2 1.2}{
1366: \piclinewidth{60}
1367: \piccircle{0 0}{1}{}
1368: \labar{240}{ 30}{1}
1369: \labar{ 60}{210}{2}
1370: \labar{300}{ 90}{3}
1371: \labar{120}{330}{4}
1372: \labar{  0}{150}{5}
1373: \labar{180}{270}{6}
1374: } % \pictranslate
1375: } % \diag
1376: \end{array}
1377: \]}
1378: \caption{The standard diagram of the knot $6_2$ and its Gau\ss{}
1379: diagram.\label{fig6_2}}
1380: \end{figure}
1381: 
1382: The simplest (non-trivial) Vassiliev knot invariant is the
1383: \em{Casson invariant} $v_2$, with $v_2=\Dl''(1)/2=-V''(1)/6$,
1384: for which Polyak-Viro
1385: \cite{VirPol,VirPol2} gave the simple Gau\ss{} sum formula
1386: \begin{eqn}\label{pv2_}
1387: v_2\,=\,\GD{\arrow{225}{45}
1388: \arrow{315}{135}
1389: \picfillgraycol{0}\picfilledcircle{1 90 polar}{0.09}{}
1390: }\,.
1391: \end{eqn}
1392: Here the point on the circle corresponds to a point on the
1393: knot diagram, to be placed arbitrarily except on a crossing.
1394: (The expression does not alter with the position of the
1395: basepoint; we will hence have, and need, the freedom to place
1396: it conveniently.)
1397: 
1398: We will use the symmetrized version of \eqref{pv2_}
1399: w.r.t. taking the mirror image of the knot diagram:
1400: \begin{eqn}\label{pv2}
1401: v_2\,=\,\frac{1}{2}\es\left(
1402: \GD{\arrow{225}{45}
1403: \arrow{315}{135}
1404: \picfillgraycol{0}\picfilledcircle{1 90 polar}{0.09}{}
1405: }+
1406: \GD{\arrow{45}{225}
1407: \arrow{135}{315}
1408: \picfillgraycol{0}\picfilledcircle{1 90 polar}{0.09}{}
1409: }\right)\,.
1410: \end{eqn}
1411: 
1412: \section{Xu's form and Seifert surfaces\label{SA}}
1413: 
1414: We studied the relation of Xu's algorithm and the skein
1415: polynomial in \cite{3br}, and here we will go further to
1416: connect fiberedness, the and Alexander and Jones polynomial
1417: to Xu's form.
1418: 
1419: \subsection{Strongly quasi-positive links among links of braid index 3}
1420: 
1421: Theorem \reference{thqp} follows relatively easily from
1422: the work in \cite{3br}, but it is a good starting point for
1423: the later more substantial arguments basing on Xu's form.
1424: 
1425: \proof[of theorem \reference{thqp}]
1426: The inequalities of \eqref{mq} for the $v$-degree
1427: of $P$ applied on a positive band representation show that,
1428: with $P=P(L)$ and $\chi=\chi(L)$,
1429: \begin{eqn}\label{x}
1430: \md_vP\ge 1-\chi\,.
1431: \end{eqn}
1432: Because of \cite[proposition 21]{LickMil}, which now says
1433: \begin{eqn}\label{p21}
1434: P(v,v^{-1}-v)=1\,,
1435: \end{eqn}
1436: we have $\md_lP\le
1437: \Md_zP$, and by \cite{3br} $\Md_zP=1-\chi$. So from \eqref{x}
1438: we obtain 
1439: \begin{eqn}\label{y}
1440: \md_lP=1-\chi\,. 
1441: \end{eqn}
1442: 
1443: It is known that the minimal degree term in $z$ of the skein polynomial
1444: of a $n$-component link is divisible by $(v-v^{-1})^{n-1}$. (In
1445: \cite{Kneissler} in fact all occurring terms are classified.) So
1446: if $MWF (K)=1$, then $K$ is a knot. Now, the identity \eqref{p21}
1447: implies that if $MWF=1$ for some knot $K$, then $P(K)=1$. For
1448: any non-trivial knot $1-\chi=2g>0$, so for any non-trivial strongly
1449: quasi-positive knot from \eqref{x} we have $\md_vP>0$, and so $P\ne 1$.
1450: 
1451: So a strongly quasi-positive link cannot have $P=1$, and always
1452: $MWF\ge 2$.  % (The split links are easy to deal with.)
1453: Therefore, when the braid index is 3, we have
1454: $MWF\in \{2, 3\}$. Then \eqref{y} and the inequalities \eqref{mq}
1455: for the $v$-degree of $P$ show that a 3-braid representation $\bt$
1456: has exponent sum $[\bt]=3-\chi$, unless $MWF=2$ and $[\bt]=1-\chi$.
1457: 
1458: In former case, we can find a minimal genus band representation
1459: from $\bt$ by Xu's algorithm \cite{Xu}, and this representation
1460: must be positive. In latter case, we will have one negative band,
1461: and have Xu's form $L^{-1}R$. Here $L$ and $R$ are positive words in
1462: the letters $\sg_1,\sg_2,\sg_3=\sg_2\sg_1\sg_2^{-1}$ with $\sg_i$
1463: followed by $\sg_i$ or $\sg_{i\bmod 3+1}$ (as described
1464: in \S\reference{brt}), and $L$ is a single letter.
1465: 
1466: W.l.o.g. assume $L=\sg_3$. If the first or last letter in $R$
1467: is a $\sg_3$ then we can cancel two bands, and have a positive
1468: representation. If some $\sg_3$ occurs in $R$, then we can write
1469: $L^{-1}R=\sg_3^{-1}\ap\sg_3\ap'$ where $\ap,\ap'$ are positive
1470: band words. Such a representation is quasi-positive with a Seifert
1471: ribbon \cite{Rudolph} of smaller genus. However, it is known that
1472: for strongly quasi-positive links the genus and 4-genus coincide
1473: (see for example \cite{posqp}). If $R$ has no $\sg_3$, then
1474: we have up to cyclic letter permutation a braid of the form
1475: $\sg_1^k\sg_2^l\sg_1^{-1}\sg_2^{-1}$ with $k\ge 0$, $l>0$. This
1476: braid is easily seen to reduce to (a positive one on) two strands.
1477: \qed
1478: 
1479: \begin{rem}\label{rqw}
1480: It is not true that all 3-braids that have the skein polynomial of
1481: $(2,n)$-torus links have positive 3-band representations. Birman's
1482: \cite{Birman} construction (see definition \reference{DEFB} below)
1483: yields examples like $12_{2037}$ in \cite{KnotScape} with one
1484: negative band. This construction gives in fact all non-obvious
1485: 3-braid $MWF=2$ examples, a circumstance shown in \cite{ntriv} by
1486: applying theorem \ref{thq} (and reportedly in previous unpublished
1487: work of El-Rifai). The result is explained and used below in the
1488: proof of theorem \reference{thJC}.
1489: \end{rem}
1490: 
1491: \subsection{Uniqueness of minimal genus Seifert surfaces\label{Uq}}
1492: 
1493: The discussion here came about from the desire to complete Xu's
1494: uniqueness theorem for Seifert surfaces of 3-braid links. While
1495: our result will be further improved later using work to follow,
1496: we need to introduce some notation and basic tools. We will use
1497: the work of Kobayashi \cite{Kobayashi}, implying that the property
1498: a surface to be a unique minimal genus surface is invariant under
1499: Hopf (de)plumbing.
1500: 
1501: Subsequently Mikami Hirasawa advised me about a subtlety
1502: concerning the notion of `uniqueness' which must be explained.
1503: Xu's work considers Seifert surfaces unique in the sense isotopic
1504: to each other, if we \em{may move} the link by the isotopy.
1505: In particular, such isotopy may interchange link components.
1506: However, we demand the isotopy to preserve component orientation.
1507: (For unoriented isotopy Xu's result is complete.) Contrarily,
1508: Kobayashi's setting assumes uniqueness in the sense that
1509: Seifert surfaces are isotopic to each other \em{fixing} the link.
1510: Hirasawa explained that the definitions are not equivalent, and
1511: that a unique (if we may move the link) minimal genus surface
1512: may get not unique when one (de)plumbs a single Hopf band. To
1513: account for this discrepancy, we should establish proper language.
1514: 
1515: \begin{defi}
1516: Let us use the term \em{unique} for
1517: Kobayashi's notion of uniqueness (up to isotopy fixing the
1518: link), and let us call Xu's notion of uniqueness (up to
1519: isotopy which may move the link or permute components,
1520: but preserves orientation) \em{weakly unique}.
1521: \end{defi}
1522: 
1523: Here we state the following extension of Xu's uniqueness theorem.
1524: 
1525: \begin{theorem}\label{wu}
1526: Every 3-braid link has a weakly unique minimal genus Seifert surface.
1527: \end{theorem}
1528: 
1529: 
1530: \proof Birman-Menasco showed that any minimal genus Seifert surface
1531: of a 3-braid link is isotopic to a Bennequin surface. Then, Xu
1532: showed that every conjugacy class of 3-braids carries a canonical
1533: (up to oriented isotopy) Bennequin surface. The conjugacy classes
1534: of 3-braids with given closure link were classified in \cite{BirMen}.
1535: Most links admit a single conjugacy class, and we are done, as in
1536: \cite{Xu}. The exceptional cases are easy to deal with, except the
1537: ``flype admitting'' braids 
1538: \begin{eqn}\label{fl3}
1539: \sg_1^{\pm p}\sg_2^{\pm q}\sg_1^{\pm r}\sg_2^{\pm 1}\,.
1540: \end{eqn}
1541: The flype interchanges $\pm q$ and $\pm 1$, and (in general) gives
1542: a different conjugacy class, which differs from the original one
1543: by orientation. So we have (as in \cite{Xu}), two Bennequin surfaces
1544: isotopic only up to orientation. We will settle this case now also
1545: for oriented isotopy, that is, show that these surfaces are
1546: isotopic to themselves with the opposite orientation.
1547: 
1548: Recall that for a fibered link, a minimal genus surface is the
1549: same as a fiber surface, and such a surface is unique (Neuwirth-%
1550: Stallings theorem). Moreover, by work of Gabai \cite{Gabai,Gabai2,%
1551: Gabai3} and Kobayashi \cite{Kobayashi} the properties of a surface
1552: to ba a minimal genus surface, a unique minimal genus surface or a
1553: fiber surface are invariant under Murasugi (de)sum with a fiber
1554: surface. In particular, this invariance holds for (de)plumbing
1555: a Hopf band (which is understood to be an unknotted annulus with 
1556: one full, positive or negative,
1557: twist). Now we note that in the cases in \eqref{fl3} where $p=r$,
1558: the flype is trivial (i.e. realized by a conjugacy), so that the
1559: surface is weakly unique. However, since Kobayashi's theorem may
1560: fail for weakly unique surfaces, we cannot reduce our surfaces to
1561: this case. We will use Hopf (de)plumbings to recur all cases to fiber
1562: surfaces or a 2-full twisted annulus. Then we understand that our
1563: surfaces are unique, and in particular weakly unique. The type of
1564: Hopf (de)plumbings we will apply is to interconvert all powers
1565: of a given band generator of given sign (for example $\sg_3^k$ is
1566: equivalent to $\sg_3^{-1}$ for each $k<0$).
1567: 
1568: W.l.o.g. assume in \eqref{fl3} that we have $-$ in $\pm 1$ and that none
1569: of $p,q,r$ is $0$ (the other cases are easy). We assume $p,q,r>0$
1570: and vary the signs before $p,q,r$ properly. Also, since the flype
1571: interchanges $\pm p$ and $\pm r$, we may assume $\pm p\ge \pm r$.
1572: So we exclude the sign choice $(-p,+r)$.
1573: 
1574: \begin{caselist}
1575: \case $p,q,r>0$. We can write \eqref{fl3} up to cyclic permutation
1576: as $(\sg_2^{-1}\sg_1\sg_2)^p\sg_2^{q-1}\sg_1^r$, and conjugating with
1577: $\sg_2\sg_1\sg_2$ we have $\sg_3^p\sg_1^{q-1}\sg_2^r$. Since $p,r>0$,
1578: this is Hopf plumbing equivalent to $\sg_3\sg_1\sg_2$ or $\sg_3\sg_2$.
1579: These cases are a disk and a 2-full twisted annulus for
1580: the reverse $(2,4)$-torus link, and we are done.\label{cs1}
1581: 
1582: \case $p,r>0$, $-q<0$. We have an alternating braid, which is
1583: fibered.
1584: 
1585: \case $p,q,r<0$. We have a negative braid, which is fibered.
1586: 
1587: \case $-p,-r<0$, $q>0$. We have a word of the form 
1588: $\sg_2^{q}\sg_1^{-r}\sg_2^{-1}\sg_1^{-p}$. If one of
1589: $p$ and $r$ is $1$, then we have a $(2,\,.\,)$-torus link.
1590: If $q=1$, then we can go over to the mirror image, and land in
1591: case \reference{cs1}. Now, when $p,q,r\ge 2$, the
1592: minimal word in Xu's form is $2^{q-2}(-3)(-2)^{r-1}(-1)^{p-1}$.
1593: Since $p-1$ and $r-1$ are non-zero, its surface is plumbing
1594: equivalent to the one of $[2-3-2-1]$ (if $q>2$) or $[-3-2-1]$
1595: (if $q=2$). These are the annulus for the reverse $(2,-4)$-torus link,
1596: and the fiber of the $(-2,-2,2)$-pretzel link, and we are done. 
1597: 
1598: \case $p>0$, $-r<0$, $q>0$.
1599: We have 
1600: \[
1601: \sg_1^p\sg_2^q\sg_1^{-r}\sg_2^{-1}=\sg_1^p\sg_2^{q-1}
1602: (\sg_2\sg_1^{-1}\sg_2^{-1})^r=\sg_1^p\sg_2^{q-1}\sg_3^{-r}\,.
1603: \]
1604: Such a surface is plumbing equivalent to the one for $\sg_1\sg_3^{-1}$
1605: (if $q=1$) resp. $\sg_1\sg_2\sg_3^{-1}$ (if $q>1$), which are the
1606: fibers of the unknot and Hopf link resp.
1607: 
1608: \case $p>0$, $-r<0$, $-q<0$. We have 
1609: \[
1610: \sg_1^{-r}\sg_2^{-1}\sg_1^p\sg_2^{-q}=\sg_1^{1-r}(\sg_1^{-1}\sg_2^
1611: {-1}\sg_1)\sg_1^{p-1}\sg_2^{-q}=\sg_1^{1-r}\sg_3^{-1}\sg_1^{p-1}
1612: \sg_2^{-q}\,.
1613: \]
1614: If $-r=-1$ then we have a $(2,p-q-1)$-torus link. If $-r<-1$, then
1615: the band surface from the right word above is plumbing equivalent to the
1616: one for $\sg_1^{-1}\sg_3^{-1}\sg_2^{-1}$ (if $p=1$) or 
1617: $\sg_1^{-1}\sg_3^{-1}\sg_1\sg_2^{-1}$ (if $p>1$), which are the
1618: annulus for the reverse $(2,4)$-torus link, and the fiber of the
1619: $(2,2,-2)$-pretzel link. \qed
1620: \end{caselist}
1621: 
1622: \subsection{Fiberedness\label{ppp}}
1623: 
1624: For the rest of the paper we normalize
1625: $\Dl$ so that $\Dl(1)=1$ and $\Dl(t)=\Dl(1/t)$.
1626: 
1627: \begin{theorem}\label{Thq}
1628: Let $L$ be a strongly quasi-positive 3-braid link. Then the following
1629: are equivalent:
1630: \def\labelenumi{\theenumi)}
1631: \begin{enumerate}
1632: \item $L$'s minimal genus surface is a Hopf plumbing,
1633: \item $L$'s minimal genus surface is a fiber surface,
1634: \item $\Md\Dl(L)=1-\chi(L)$ and $\Mc\Dl(L)=\pm 1$,
1635: \item $L$'s Xu normal form is \em{not} $R$, with syllable
1636: length of $R$ divisible by $3$. In other words, the Xu normal
1637: form is not an extension (in the sense of definition
1638: \ref{dfext}) of $[(123)^k]$ for $k>0$.
1639: \item Some minimal band form of $L$ contains
1640: $\sg_1^k\sg_2^l\sg_1^m$ or $\sg_2^k\sg_1^l\sg_2^m$
1641: as subword for $k,l,m>0$, or is $\sg_1^k\sg_2^l$.
1642: \end{enumerate}
1643: \end{theorem}
1644: 
1645: \proof  $(1)\So (2)\So (3)$ are clear.
1646: 
1647: $(5)\So (1)$. Assume after deplumbing, all letters occur in single
1648: power, and up to conjugacy the word starts with $\sg_1\sg_2\sg_1$
1649: or $\sg_2\sg_1\sg_2$. By adjusting one of the two, we can have a
1650: $\sg_1^2$ or $\sg_2^2$ if the next letter is $\sg_1$ or $\sg_2$. In
1651: that case we deplumb a Hopf band. If the next letter is $\sg_3$, then
1652: we have $\sg_1\sg_2\sg_1\sg_3=\sg_1\sg_2^2\sg_1$, can can also deplumb
1653: a Hopf band. Then we reduce the surface for that of $\sg_1\sg_2\sg_1$
1654: which is the Hopf band.
1655: 
1656: $(3)\So (4)$. We prove the contrary. Assume $(4)$ does not hold. 
1657: Band-positive surfaces are always of minimal genus, so that the
1658: properties we investigate are invariant under Hopf (de)plumbings.
1659: Under applying skein relations at non-trivial syllables we are
1660: left with powers of $\sg_1\sg_2\sg_3$. Apply the skein relation
1661: for $\Dl$ at the last band. Then $\bt_{-}$ and $\bt_0$ are both
1662: of minimal length. We already proved, in $(5)\so (1)$, that
1663: $\bt_0$ is fibered, so $\Md\Dl=1-\chi(\bt_0)$ and $\Mc\Dl=\pm 1$.
1664: The same holds for $\bt_{-}$ by \cite[proposition 2]{3br}, since 
1665: $\bt_{-}$ is of Xu's minimal form $L^{-1}R$ and is not positive.
1666: So the terms in degree $1-\chi(\bt)$ of $\Dl(\bt)$ either cancel, or
1667: give $\pm 2$. 
1668: 
1669: $(4)\So (5)$. We prove the contrary. Assume $(5)$ does not hold.
1670: If we do not have a word $\bt=R$ in Xu's form of length divisible by
1671: $3$, another option would be to have a word with cyclically decreasing
1672: indices. But note that $[321]=[121]$ is the Hopf band, while
1673: $[321321]=[211211]$ contains a $121$, too, and so we
1674: are done. Otherwise the index array of
1675: the syllables of $\bt$ must contain the same entry with distance $2$,
1676: and such a word is conjugate to the ones excluded.
1677: \qed
1678: 
1679: \begin{theorem}\label{xLR}
1680: Any closed braid of Xu's form $L^{-1}R$ is fibered.
1681: \end{theorem}
1682: 
1683: The proof uses some work of Hirasawa-Murasugi.
1684: A consequence of their result is the following lemma,
1685: which we require. It is proved in appendix \reference{HM}.
1686: 
1687: \begin{lemma}\label{lki}
1688: The links $L_k$, given by the closed 3-braids $[(123)^k\,-2]$ for $k>0$,
1689: are fibered.
1690: \end{lemma}
1691: 
1692: % \proof Consider the knots $L_k'$ given by the closed 3-braid
1693: % words $[(12)^{3k+1}(-2)^{3k+l}]$, where $l=2$ for
1694: % $k$ even, and $l=3$ for $k$ odd. The Xu normal form is found using
1695: % the identity $2^3(123)=(12)^3=(123)2^3$ to be
1696: % \[
1697: % (12)^{3k+1}(-2)^{3k+l}\to 12(123)^k(-2)^l\to 21223(123)^{k-1}(-2)^l\to
1698: % 1223(123)^{k-1}(-2)^{l-1}\,,
1699: % \]
1700: % and this surface is Hopf plumbing equivalent to that of $L_k$.
1701: % 
1702: % We know from \cite{3br} (as explained further in \cite{ntriv})
1703: % that if $\bt\in B_3$ is not strongly quasi-positive, then
1704: % $2\Md\Dl=1-\chi$ and $\Mc\Dl=\pm 1$.
1705: % 
1706: % Now, since $L_k'$ has Xu normal form of the type $L^{-1}R$, we
1707: % find that $\Mc\Dl(L_k')=\pm 1$. Then, by the result of
1708: % Hirasawa-Murasugi $L_k'$ is fibered, and hence so is $L_k$. \qed
1709: 
1710: 
1711: \proof[of theorem \reference{xLR}]
1712: We use induction on the length of $L^{-1}R$ and for fixed
1713: length on the exponent sum. Under Hopf deplumbings assume all
1714: syllables in $L$ and $R$ are trivial.
1715: Assume up to mirroring that $L$ is not shorter than $R$.
1716: Permute by conjugacy $R$ to
1717: the left, and permute the indices so that $L$ starts with $-3$. 
1718: 
1719: The following transformations also offer a Hopf deplumbing
1720: \[
1721: 1-3-2-1\to -2 1 -2 -1\to -2 -2 -1 2\to -2 -1 2
1722: \]
1723: and
1724: \[
1725: 1 2 -3\to -2 1 1
1726: \]
1727: These reductions fail both if either $L$ and $R$ have length at most
1728: $2$, or $R$ has length $1$. (Remember $L$ is not shorter than $R$.)
1729: In former case one checks directly that one has a disk, Hopf band or
1730: connected sum of two Hopf bands. We consider latter case.
1731: 
1732: By conjugacy permute the indices so that $R=1$; also assume $L$ has
1733: length at least $3$. If $L$ starts and ends with $-2$ then we do
1734: \[
1735: 1 -2 -1 \dots -2 \to -2 3 -1 \dots -2\to 3 -1 \dots -2 -2
1736: \]
1737: (where the right transformation is a conjugacy) and deplumb a
1738: Hopf band. If $L$ starts with $-3$ then we transform as before
1739: \[
1740: 1 -3 -2 -1\to -2 1 -2 -1\to -2 -2 -1 -2
1741: \]
1742: and deplumb a Hopf band. So $L$ starts with $-2$ and ends on $-3$.
1743: Then the mirror image of $\bt$ is up to conjugacy of the form
1744: $[(123)^k\,-2]=[(1 2 2 1)^k (-2)^{k+1}]$, which we dealt with in
1745: the lemma before.  \qed
1746: 
1747: Combinedly, we obtain
1748: 
1749: \begin{corr}
1750: Let $W$ be a 3-braid link. Then the following are equivalent:
1751: \def\labelenumi{\theenumi)}
1752: \begin{enumerate}
1753: \item $W$ is fibered,
1754: \item $\Md\Dl(W)=1-\chi(W)$ and $\Mc\Dl(W)=\pm 1$,
1755: \item $W$'s Xu form is not of the type $L^{-1}$ or $R$,
1756: with syllable length of $L$ or $R$ divisible by $3$. \qed
1757: \end{enumerate}
1758: \end{corr}
1759: 
1760: \begin{exam}
1761: The routine verification, with Mikami Hirasawa, of the tables in
1762: \cite{KnotScape} for fibered knots, has shown non-fibered knots 
1763: with monic Alexander polynomial of degree matching the genus
1764: start at 12 crossings. One of these knots, $12_{1752}$, has
1765: braid index $4$, so that (as expected) the corollary does not
1766: hold for 4-braids.
1767: \end{exam}
1768: 
1769: \begin{corr}
1770: Any non-split closed 3-braid $\hat\bt$ with $|[\bt]|\le 2$ is fibered,
1771: in particular so is any slice 3-braid knot. \qed
1772: \end{corr}
1773: 
1774: Again $6_1$ and $9_{46}$ show the second part is not true for 4-braids.
1775: The first statement in the corollary extends simultaneously
1776: the property of amphicheiral knots, which follows also
1777: from \cite{BirMen}, since amphicheiral 3-braid
1778: knots are closed alternating 3-braids.
1779: 
1780: Originally the insight about a unique minimal genus surface motivated
1781: the fibered 3-braid link classification. Still this insight lacks
1782: asset as to the somewhat improper way it emerges. Note for example,
1783: that the stronger version of \cite{BirMen2} of Bennequin's theorem
1784: enters decisively into \cite{Xu} and the proof of theorem \ref{wu},
1785: but then paradoxically latter imply that Birman-Menasco's formulation
1786: is actually equivalent to, and not really an improvement of,
1787: Bennequin's theorem. Also, Birman-Menasco's classification
1788: \cite{BirMen} priorizes Schreier's conjugacy algorithm and lacks
1789: any geometric interpretation of Xu's normal form, while such an
1790: interpretation becomes evident in our setting. This provided strong
1791: motivation for theorem \reference{thuq}. Its proof is completed
1792: by dealing with the non-fibered cases in appendix \reference{HI}.
1793: It requires also a part of the further detailed consideration of the
1794: Alexander polynomial that follows next.
1795: 
1796: \section{Polynomial invariants}
1797: 
1798: \subsection{Alexander polynomial}
1799: 
1800: In \cite{Birman}, Birman proposed (but considered as very difficult)
1801: the problem to classify 3-braid
1802: links with given polynomials. In \cite{3br} we dealt with the skein
1803: polynomial. Now we can extend our results to the Alexander polynomial
1804: (with the convention in the beginning of \S\reference{ppp}).
1805: The following discussion gives a fairly exact description how to
1806: find the 3-braid links, if such exist, for any possible admissible
1807: (as specified in \S\reference{brD}) polynomial.
1808: 
1809: A solution for the Jones polynomial is presented afterwards.
1810: % (though not explained in detail)
1811: 
1812: \begin{lemma}
1813: If $\bt$ is strongly quasi-positive and fibered, then $\Mc\Dl=+1$.
1814: \end{lemma}
1815: 
1816: (Here it is essential to work with the leading, not trailing
1817: coefficient of $\Dl$ and with strongly quasi-positive links and
1818: not their mirror images.)
1819: 
1820: \proof We know $\md_vP=\Md_zP=1-\chi$. So $[P]_{z^{1-\chi}}$ has
1821: a term in degree $v^{1-\chi}$. The coefficient must be $+1$
1822: because of \eqref{p21} and because it is the only coefficient that
1823: contributes to the absolute term in $P(v,v^{-1}-v)$. Now from
1824: the classification of leading $z$-terms of $P$ in \cite{3br}
1825: it follows that $[P]_{z^{1-\chi}}$ can have at most one further term,
1826: with coefficient $\pm 1$. If such term exists, the substitution
1827: \eqref{DP} would either cancel the terms
1828: in degree $t^{(1-\chi)/2}$ in $\Dl(t)$, or give coefficient $\pm 2$,
1829: so our link cannot be fibered. Thus a second term does not exist,
1830: and the claim follows from \eqref{DP}. \qed
1831: 
1832: \begin{lemma}\label{ty}
1833: Let $\bt'=[(123)^k]$ be an even power $k$ of $[123]$.
1834: Assume $\bt$ has Xu normal form $R$, and after syllable
1835: reduction becomes $\bt'$, but $\bt\ne\bt'$ itself (i.e. some
1836: syllable in $\bt$ is non-trivial). Then $\hat\bt$ satisfies
1837: $\Md\Dl=-1-\chi$; moreover $\Mc\Dl>0$ and is equal to the 
1838: number of non-trivial syllables in $\bt$.
1839: \end{lemma}
1840: 
1841: \proof We use for fixed $k$ induction on the exponent sum.
1842: If exactly one syllable is non-trivial with exponent $2$,
1843: then applying the skein relation at the exponent-2
1844: syllable shows that $\bt=\bt_+$ inherits the Alexander
1845: polynomial of $\bt_-$ with positive sign (since $\bt_0=\bt'$,
1846: whose closure has zero polynomial). 
1847: Now $\bt_-$ is positive and has an index array with a subsequence
1848: of the form $xyx$, and so is fibered by theorem \reference{Thq}.
1849: Then by lemma \reference{ty}, $\Mc\Dl=+1$ and
1850: $\Md\Dl=1-\chi(\hat\bt_-)=-1-\chi(\hat\bt)$.
1851: 
1852: If $[\bt]-[(123)^k]>1$, then applying the skein relation
1853: at any non-trivial exponent $2$ syllable gives (with positive
1854: sign) the Alexander polynomials of two closed braids
1855: $\bt_0$ and $\bt_-$, former of which is fibered and
1856: latter of which has the requested property by induction.
1857: Then the maximal terms in degree $1-\chi(\hat\bt_-)=
1858: -\chi(\hat\bt_0)$ are positive and do not cancel.
1859: The one of $\bt_-$ is $+1$, while the one of $\bt_0$
1860: by induction one less than the number of non-trivial 
1861: syllables in $\bt$ (since in $\bt_0$ one more syllable
1862: becomes trivial).
1863: 
1864: If some syllable in $\bt$ has exponent $>2$ then $\bt_-$ is
1865: not minimal, and the degree and leading coefficient of $\Dl$
1866: are inherited (with positive sign) from $\bt_0$. \qed
1867: 
1868: \begin{lemma}\label{tz}
1869: Assume $\bt$ has Xu normal form $R$, which after syllable
1870: reduction becomes an odd power $k$ of $[123]$.
1871: Then $\Md\Dl=1-\chi$, and $\Mc\Dl=+2$.
1872: \end{lemma}
1873: 
1874: \proof First we prove the claim if $\bt=[(123)^k]$.
1875: We know that $\Md\Dl=1-\chi$ and $\Mc\Dl=\pm 2$,
1876: so we must exclude $\Mc\Dl=-2$. Applying the skein relation gives
1877: the polynomials of $\bt_0$ and $\bt_-$ with positive sign.
1878: $\bt_0$ is positive and fibered as before, so $\Mc\Dl=+1$.
1879: Then clearly $\bt_-$ (which is of the form $L^{-1}R$ and also
1880: fibered) cannot have $\Mc\Dl=-3$. 
1881: 
1882: If $\bt\ne [(123)^k]$, then it has a non-trivial syllable.
1883: Applying the skein relation at a letter in that syllable
1884: we find that $\bt_-$ reduces. So the leading term comes
1885: from $\bt_0$, and with positive sign. \qed
1886: 
1887: 
1888: 
1889: \begin{theorem}\label{THA}
1890: Fix some admissible Alexander polynomial $\Dl$. Then
1891: \def\labelenumi{\theenumi)}
1892: \begin{enumerate}
1893: \item If $\Dl=0$, then the 3-braid links with such polynomial
1894: are the split links and the closures of (incl.\ negative)
1895: even powers of $[123]$.
1896: \item If $\Dl\ne 0$, there are only finitely many 3-braid links
1897: with this $\Dl$. They all have $1-\chi=2\Md\Dl$ or $-1-\chi=2\Md\Dl$.
1898: In latter case they are up to mirroring strongly quasi-positive.
1899: \item If $\Mc\Dl\le -2$, then no 3-braid knot or 3-component
1900: link has such Alexander polynomial, and any 2-component link
1901: is strongly quasi-negative.
1902: \item If $\Mc\Dl\ge +2$, any 3-braid knot or 3-component link
1903: with such Alexander polynomial is strongly quasi-positive or a
1904: mirror image of it. Any 2-component link is strongly quasi-positive.
1905: \item If $|\Mc\Dl|>2$, then any 3-braid link 
1906: with such Alexander polynomial has $2\Md\Dl=-1-\chi$.
1907: \end{enumerate}
1908: \end{theorem}
1909: 
1910: \begin{defi}\label{DEFB}
1911: Let for $\bt\in B_3$ with $6\mid [\bt]$, the \em{Birman dual}
1912: $\bt^*$ be defined by $\bt^{-1}\dl^{2[\bt]/3}$, where
1913: $\dl=[121]$ and $\dl^2$ generates the center of $B_3$.
1914: \end{defi}
1915: 
1916: Birman \cite{Birman} shows that $\hat\bt$ and $\hat\bt^*$
1917: have the same skein polynomial. This observation relates to
1918: our explanation at a couple of places, for example, in remark
1919: \reference{rqw}, and also in the below arguments.
1920: 
1921: \proof[of theorem \reference{THA}]
1922: Let us exclude \em{a priori} trivial and split links.  The claims
1923: follow from the discussion of $2\Md\Dl$ and $\Mc\Dl$ in cases.
1924: 
1925: We know by theorem \reference{xLR} (or by \cite{3br,ntriv}, as noted
1926: before) that if $\bt\in B_3$ is not (up to mirroring) strongly
1927: quasi-positive, then $2\Md\Dl=1-\chi$ and $\Mc\Dl=\pm 1$.
1928: 
1929: It remains to deal with the Xu form $R$. The form $L$ is just the
1930: mirror image, and mirroring preserves the Alexander polynomial for
1931: knots and 3-component links and alters the sign for 2-component
1932: links.
1933: 
1934: If making trivial all syllables in $\bt$, the new word
1935: $\bt'$ is not a power of $[123]$, then we proved that $\hat\bt$
1936: is fibered, so again $2\Md\Dl=1-\chi$ and $\Mc\Dl=\pm 1$.
1937: 
1938: If $\bt'$ is an even power of $[123]$, then by Birman duality we
1939: conclude that $\Dl(\hat\bt')=0$. If $\bt\ne \bt'$, then one uses
1940: lemma \reference{ty}.
1941: % induction on the syllables exponents. If $\bt$ and $\bt'$
1942: % differ only by a letter, then applying the skein relation at
1943: % the exponent-2 syllable shows that $\bt=\bt_+$ inherits the Alexander
1944: % polynomial of $\bt_-$ (since $\bt_0=\bt'$). This has form $L^{-1}R$
1945: % and so $2\Md\Dl=1-\chi(\hat\bt_-)=-1-\chi(\hat\bt)$ and $\Mc\Dl=\pm 1$.
1946: % Then the degree and leading term are successively propagated.
1947: Thus $2\Md\Dl=-1-\chi$.
1948: % and also here $\Mc\Dl=\pm 1$.
1949: 
1950: If $\bt'$ is an odd power of $[123]$, then use the observation in
1951: lemma \reference{tz}  (or make one syllable to exponent $4$ and
1952: use Birman duality) to conclude that $2\Md\Dl=1-\chi$ and
1953: $\Mc\Dl=+2$. \qed
1954: 
1955: \begin{exam}
1956: In certain situations this theorem gives the most rapid test to
1957: exclude closed 3-braids. For example, the knot $13_{6149}$ has
1958: $MWF=3$, but seeing that $\Dl$ has $\Mc\Dl=-2$ we immediately
1959: conclude that it cannot be a closed 3-braid.
1960: \end{exam}
1961: 
1962: \begin{corr}
1963: There are only finitely many 3-braid links with given $\Md\Dl$
1964: (provided $\Dl\ne 0$). \qed
1965: \end{corr}
1966: 
1967: It is actually true (as we will prove below) that for links with
1968: \em{any} bounded braid index there are only finitely many different
1969: Alexander polynomials of given degree admitted. However, such
1970: Alexander polynomials may be admitted by infinitely many different
1971: links (of that braid index). See \S\reference{s4.2} for some remarks.
1972: % Using the non-faithfulness of
1973: % the Burau representation on $\ge 5$ strands, it should not
1974: % be too hard to find such families. The 4-strand case however
1975: % seems open.
1976: % 
1977: % \begin{question}
1978: % Are there only finitely many 4-braid links with given $\Dl\ne 0$
1979: % (or equivalently $\Md\Dl$)?
1980: % \end{question}
1981: 
1982: \begin{corr}\label{cbd}
1983: Non-split 3-braid links bound no disconnected Seifert surfaces (with
1984: no closed components). In particular there are no non-trivial 3-braid
1985: boundary links.
1986: \end{corr}
1987: 
1988: \proof For such links the Alexander polynomial is zero. The only
1989: non-split 3-braid links $L$ of zero polynomial are closures of
1990: $[(123)^k]$ for even $k$. These cases are easily ruled out by linking
1991: numbers. Any pair of components of $L$ has non-zero linking number.
1992: So a connected component of any Seifert surface $S$ of $L$ must have
1993: at least two boundary components of $L$, and if $S$ is disconnected,
1994: $L$ has at least 4 components, which is clearly not the case. \qed
1995: 
1996: The examples that falsify this claim for 4-braids are again
1997: easy: consider the (closure links of) words in $[1\pm 2-1]$ and
1998: $[2\pm 3-2]$.
1999: 
2000: \begin{corr}
2001: If $\Dl$ is the Alexander polynomial of a 3-braid link, then
2002: $|\Mc\Dl|\le \Md\Dl+2$. For a knot $|\Mc\Dl|\le\Md\Dl+1$.
2003: \end{corr}
2004: 
2005: \proof If $|\Mc\Dl|\le 2$ then we are easily done. 
2006: ($\Dl=\pm 2$ cannot occur for a knot.) Otherwise we have
2007: up to mirroring a strongly quasi-positive word $\bt$ reducing to
2008: an even power of $[123]$. Now $|\Mc\Dl|$ counts non-trivial syllables,
2009: so $[\bt]=3-\chi\ge 2|\Mc\Dl|$, and $-1-\chi=2\Md\Dl$. For a knot
2010: the number of syllables with exponent $\ne 2$ is at least $2$,
2011: so $[\bt]-2\ge 2|\Mc\Dl|$. \qed
2012: 
2013: Note in particular that the proof shows that $|\Mc\Dl|$
2014: can be any given natural number, and how to find the link
2015: that realizes this number. We emphasize this here, because
2016: later we will prove a contrary statement in the case of
2017: \em{alternating} links for \em{every} arbitrary braid index
2018: (see corollary \reference{cSG}).
2019: 
2020: \begin{corr}
2021: A 3-braid \em{knot} is fibered if and only if $\Mc\Dl=\pm 1$.
2022: \end{corr}
2023: 
2024: \proof It remains to explain why no 3-braid knot has $\Mc\Dl=\pm 1$
2025: but $\Md\Dl<1-\chi$. Latter condition would imply that we have
2026: up to mirroring a strongly quasi-positive word $\bt$ reducing to
2027: an even power of $[123]$, and former condition that $\bt$ has only one
2028: non-trivial syllable. But $[(123)^{2k}]$ has 3-component closure,
2029: and making one syllable non-trivial cannot give a knot. \qed
2030: 
2031: In particular, it is worth noting
2032: 
2033: \begin{corr}
2034: No non-trivial 3-braid knot has trivial Alexander polynomial. \qed
2035: \end{corr}
2036: 
2037: Again, the two 11 crossing knots immediately show that this is
2038: not true for 4-braids.
2039: 
2040: \begin{exam}\label{x29}
2041: We can also easily determine the 3-braid links for some small degree
2042: Alexander polynomials. For example, we see that no other 3-braid
2043: knot has the polynomial of $3_1$, $4_1$ or $5_2$. Similarly we can
2044: check that no 3-braid knot has the polynomial of $9_{42}$ and
2045: $9_{49}$ (which shows that these knots have braid index 4), a
2046: fact we will derive in the last section using entirely different
2047: representation theory arguments. 
2048: \end{exam}
2049: 
2050: \subsection{\label{Js}Jones polynomial}
2051: 
2052: The control of the Jones polynomial on 3-braid links was the
2053: object of main attention in \cite{Birman}. We can accomplish
2054: this with a similar argument to $\Dl$. The result we obtain
2055: can be conveniently described in our setting and is as follows:
2056: 
2057: \begin{theorem}\label{Hh}
2058: Let $L$ be a non-split 3-braid link, and $L=\hat\bt$ with $\bt\in
2059: B_3$. Then
2060: \begin{eqn}\label{g}
2061: \spn V(L)\,\le\,4-\chi(L)\,.
2062: \end{eqn}
2063: Equality holds if and only if $L$ is strongly quasi-signed (i.e.
2064: -positive or -negative, or equivalently $|[\bt]|=3-\chi(L)$), and
2065: not fibered. More specifically, the following holds:
2066: \def\labelenumi{\theenumi)}
2067: \begin{enumerate}
2068: \item If $L$ is strongly quasi-positive, then 
2069: $\ds\quad
2070: \md V\,=\,\frac{1-\chi}{2}\quad\mbox{and}\quad\mc V\,=\,\pm 1\,.
2071: $\\
2072: Analogously, if $L$ is strongly quasi-negative, then
2073: $\ds\quad
2074: \Md V\,=\,\frac{\chi-1}{2}\quad\mbox{and}\quad\Mc V\,=\,\pm 1$\,.
2075: \item If $L$ is strongly quasi-positive and fibered, then
2076: $\spn V(L)\,\le\,3-\chi(L)$\,.
2077: If $L$ is strongly quasi-positive and not fibered, then \eqref{g}
2078: is an equality and $\Mc V\,=\,\pm 1$. (The properties for strongly
2079: quasi-negative are analogous.)
2080: \item If $L$ is not strongly quasi-signed and $|[\bt]|<1-\chi(L)$,
2081: then 
2082: \[
2083: \mc V(L)\,=\,\pm 1\,, \quad\Mc V(L)\,=\,\pm 1\quad\mbox{and}\quad
2084: \spn V(L)\,=\,3-\chi(L)\,.
2085: \]
2086: \item If $L$ is not strongly quasi-signed and $[\bt]=1-\chi(L)$, then
2087: \[
2088: \Mc V(L)\,=\,\pm 1\quad\mbox{and}\quad
2089: \Md V(L)\,=\,\frac{5}{2}-\frac{3}{2}\chi(L)\,.
2090: \]
2091: Moreover, $\md V\ge \ds\frac{-1-\chi}{2}$, and if equality holds,
2092: then $\mc V=\pm 2$. (The case $[\bt]=\chi(L)-1$ is analogous.)
2093: \end{enumerate}
2094: \end{theorem}
2095: 
2096: Apart from solving Birman's problem how to determine 3-braids
2097: with given Jones polynomial, theorem \ref{Hh} easily implies that
2098: no non-trivial 3-braid link has trivial (i.e. unlink) polynomial. We
2099: defer the discussion of the non-triviality of the Jones polynomial
2100: to \cite{ntriv}, where we work in the much more general context of
2101: semiadequate links. In that paper we will show that semiadequate
2102: links have non-trivial Jones polynomial. This result in fact motivated
2103: theorem \reference{thq}, which then provides a different conclusion
2104: about the non-triviality of the polynomial. A further application
2105: will be the classification of the 3-braid links with unsharp
2106: Morton-Williams-Franks inequality (mentioned in remark \ref{rqw}).
2107: 
2108: \begin{corr}\label{fintg}
2109: For a given Jones polynomial $V$ (actually the pair
2110: $(\md V,\Md V)$ is enough), there are at most three
2111: values of $\chi(L)$ of a 3-braid link $L$ with $V(L)=V$.
2112: If $\md V\cdot \Md V\le 0$, then $\chi(L)$ is unique.
2113: \end{corr}
2114: 
2115: \proof The theorem shows that the value of $\chi$ is determined
2116: by one of $\Md V$, $\md V$ or $\spn V=\Md V-\md V$.
2117: In particular for a pair $(\md V,\Md V)$ there exist at most three
2118: values of $\chi$ of 3-braid links with a Jones polynomial realizing
2119: this pair. If $\md V\cdot \Md V\le 0$, then the options that a 3-braid
2120: $\bt$ with $V(\hat\bt)=V$ is strongly quasi-signed or almost
2121: quasi-signed are excluded (up to a few simple cases that can be
2122: checked directly), so $\chi(L)$ is determined (unambiguously) by
2123: $\spn V$. \qed
2124: 
2125: In particular, since 3-braid links of given $\chi$
2126: are only finitely many, we have
2127: 
2128: \begin{corr}\label{finth}
2129: There are only finitely many closed 3-braids with the same
2130: Jones polynomial, actually with the same pair $(\md V,\Md V)$. \qed
2131: \end{corr}
2132: 
2133: For example, one
2134: easily sees that no 3-braid knot has the polynomial of $9_{42}$.
2135: Similarly, no other 3-braid knot has the polynomial of the figure-8-knot
2136: (there is, however, a 4-braid knot with such polynomial, $11_{386}$).
2137: 
2138: That there are only finitely many closed 3-braids with the same
2139: skein polynomial was known from \cite{3br}. For the Jones polynomial
2140: one should note that infinite families were constructed
2141: by Traczyk \cite{Traczyk} if one allows polynomials up to units.
2142: Traczyk's examples show that for fibered
2143: (strongly quasi-)positive links $\spn V$ may remain the same
2144: while $\chi\to-\infty$, so one cannot expect a full (lower) 
2145: control on $\spn V$ from $\chi$. From these links, one obtains
2146: by connected sum infinite families with the same
2147: polynomial for 5-braids. (The status of 4-braids remains unclear.)
2148: Also Kanenobu \cite{Kanenobu2} constructed finite families of 3-braids
2149: of any arbitrary size, so that our result is the maximal possible.
2150: 
2151: A question that surfaces naturally with these remarks in mind is 
2152: 
2153: \begin{question}
2154: Does for 3-braid links the Jones polynomial (or Alexander polynomial)
2155: determine the skein polynomial? In other words, do any two
2156: 3-braid links with the same $V$ (or $\Dl$) have also equal $P$?
2157: \end{question}
2158: 
2159: We have at least the following partial result, whose proof we
2160: postpone after the proof of theorem \reference{Hh}.
2161: 
2162: \begin{corr}\label{CVP}
2163: A given Jones polynomial $V=V(L)$ is realized for 3-braid links
2164: $L$ by at most three different skein polynomials $P(L)$. If
2165: $\md V\cdot \Md V\le 0$, then $V(L)$ determines $P(L)$.
2166: \end{corr}
2167: 
2168: In the general case one cannot expect a positive answer to the
2169: above question. At least for links there is now the method
2170: of \cite{EKT} available, which should yield large families of
2171: links with the same Jones but different skein (or Alexander)
2172: polynomial. For constructing families with the same Alexander
2173: (but different Jones or skein) polynomial, further techniques
2174: are available, applicable also for knots, like the non-faithfulness
2175: of the Burau representation (see \cite{Bigelow}) or tangle surgeries
2176: (see \cite{Bleiler}).
2177: 
2178: % \begin{corr}\label{fintg}
2179: % For a given Jones polynomial $V$ (actually the pair
2180: % $(\md V,\Md V)$ is enough), there are at most three
2181: % values of $\chi(L)$ of a 3-braid link $L$ with $V(L)=V$.
2182: % If $\md V\cdot \Md V\le 0$, then $\chi(L)$ is unique.
2183: % \end{corr}
2184: % 
2185: % \proof The theorem shows that the value of $\chi$ is determined
2186: % by one of $\Md V$, $\md V$ or $\spn V=\Md V-\md V$.
2187: % 
2188: % In particular for a pair $(\md V,\Md V)$ there exist at most three
2189: % % values of $\chi$ of 3-braid links with a Jones polynomial realizing
2190: % this pair; such links are only finitely many. \qed
2191: 
2192: For the proof of theorem \ref{Hh} we use the previous work in
2193: \cite{3br} (that in
2194: particular answered Birman's question) on the skein polynomial. We
2195: apply again the result in \cite{3br} that $\Md_zP=1-\chi$ for closed
2196: 3-braids. As in that paper, we distinguish the cases of band-%
2197: positive, band-negative 3-braids, and such of Xu's form $L^{-1}R$. 
2198: 
2199: By the Morton-Williams-Franks inequalities \eqref{mq},
2200: a 3-braid $\bt$ of exponent sum (writhe) $[\bt]$ has $v$-degrees of
2201: $P$ in $[\bt]-2$ we call \em{left degree}, $[\bt]$ we call \em{middle
2202: degree} and $[\bt]+2$ we call \em{right degree}. The terms of
2203: $[P]_{z^k}$ for some $k$ in these degrees will be called left,
2204: middle and right terms\footnote{Note that here the brackets for
2205: polynomials and for braids have a completely different meaning.}. 
2206: 
2207: \begin{lemma}\label{lemg}\mbox{}\\
2208: (a) If $\bt$ is band-positive, then $\Mc_zP$ has $v$-terms in the left
2209: degree and possibly in the middle degree. If $\bt$ is band-negative,
2210: $\Mc_zP$ has $v$-terms in the right degree and possibly in the
2211: middle degree. In either situation a term in the middle degree occurs
2212: if and only if $\hat\bt$ is not fibered.\\
2213: (b) If $\bt$ is $L^{-1}R$
2214: then $\Mc_zP$ has $v$-terms in the middle degree only.\\
2215: In both cases all non-zero occurring coefficients are $\pm 1$.
2216: \end{lemma}
2217: 
2218: \proof In (a) we prove only the first claim (the second
2219: claim is analogous). Let $\chi=\chi(\hat\bt)$. Since
2220: $\md_lP\le \Md_zP=1-\chi$ and $1-\chi=[\bt]-2\ge 
2221: \md_lP$ by MWF, we have $\md_lP=\Md_zP$, and then \eqref{p21}
2222: implies that $\Mc_zP$ has a $v$-term in the left
2223: degree, with coefficient $\pm 1$. From \cite[Theorem 3]{3br} we
2224: have then that it has no right-degree term, and that if it has
2225: a middle-degree term, the coefficient is $\pm 1$. Now the
2226: previous work and the substitution $v=1$, $z=t^{1/2}-t^{-1/2}$
2227: for $\Dl$ in \eqref{DP} easily show that the middle term
2228: occurs if and only if $\hat\bt$ is not fibered.
2229: 
2230: Now consider (b). If $\Mc_zP$ has a left-degree term, then using
2231: the substitution \eqref{p21}, we saw that $\bt$ is band-positive.
2232: Otherwise, $[\bt]<3-\chi$, so the contribution of the coefficient
2233: of $z^{1-\chi}v^{[\bt]-2}$ in \eqref{p21}, which is not cancelled,
2234: is not in degree $0$. Analogously one argues if $\Mc_zP$ has a
2235: right-degree term. \qed
2236: 
2237: Note that keeping track of $[\bt]$ and distinguishing between
2238: left and middle term is important here. By remark \ref{rqw},
2239: we have links with equal polynomials such that the left
2240: term of the one is the middle term of the other.
2241: 
2242: \begin{lemma}\label{lemldg}
2243: (a) If $\bt$ is $L^{-1}R$ and $L$ has exponent sum $[L]>1$, then
2244: $[P]_{z^{-1-\chi}}$ has a left-degree term, and the coefficient
2245: is $\pm 1$. \\
2246: (b) If $[L]=1$ and
2247: $[P]_{z^{-1-\chi}}$ has a left-degree term, then the coefficient
2248: is $\pm 2$. \\
2249: The two analogous statements hold replacing $L$ by $R$
2250: and left-degree term by right-degree term.
2251: \end{lemma}
2252: 
2253: \proof We proved already that $\Mc_zP$ has only a middle-degree term.
2254: So it must remain under the substitution of \eqref{p21} if and only
2255: if $[\bt]=1-\chi$. The cases (a) and (b) occur when this term must
2256: be cancelled, and complemented to $1$, respectively. \qed
2257: 
2258: \proof[of theorem \reference{Hh}]
2259: The statements follow mainly by putting together the last two
2260: lemmas and looking in which degrees the
2261: non-cancelling contributions of the coefficients of $P$
2262: occur under the substitution \eqref{VPsub}.
2263: 
2264: If part (a) of lemma \reference{lemg} applies, we established already
2265: (in theorem \reference{thqp})
2266: that the positive band form of $\bt$ is equivalent to the strong
2267: quasi-positivity of $\hat\bt$. In this case $\md V$ comes from
2268: the left-degree term in $\Md_zP$. If $\hat\bt$ is not
2269: fibered, then $[P]_{z^{-1-\chi}}$ must have a right-degree
2270: term (with coefficient $\pm 1$) to cancel the middle-degree
2271: term of $\Md_zP$ under \eqref{p21}.
2272: 
2273: If part (b) of lemma \reference{lemg} applies, we use the
2274: further information of lemma \ref{lemldg}.
2275: In case (a) of lemma \ref{lemldg}, the left and right
2276: terms in $[P]_{z^{-1-\chi}}$ determine the degrees
2277: and edge coefficients of $V$.
2278: 
2279: In case (b) of lemma \ref{lemldg},
2280: the maximal term in $V$ comes from the right-degree term in
2281: $[P]_{z^{-1-\chi}}$. A coefficient in $t^{(-1-\chi)/2}$
2282: may come only from a left-degree term in $[P]_{z^{-1-\chi}}$,
2283: which, if occurring, is with coefficient $\pm 2$. \qed
2284: 
2285: This proof underscores the significance of \eqref{VPsub}
2286: as a tool for studying the Jones polynomial. So far it
2287: seems to have been useful just for calculating specific
2288: Jones polynomials from $P$. In \S\reference{Sfg} we will
2289: see further results that come out of considering this
2290: substitution.
2291: 
2292: We note the following equalities that follow from the proof of
2293: theorem \ref{Hh}. These will be needed in the study of the Q
2294: polynomial, and are also helpful for corollary \reference{CVP}.
2295: 
2296: \begin{lemma}\label{lemma_}
2297: If $\bt$ is a not strongly quasi-signed 3-braid of exponent sum $e$
2298: and $V$, $\chi$ the Jones polynomial resp. Euler characteristic
2299: of its closure, then
2300: \begin{eqnarray}
2301: \label{one_}\Md V-e & = & \frac{1-\chi}{2}+1 \\
2302: \label{two_}\md V-e & = & -\frac{1-\chi}{2}-1 
2303: \end{eqnarray}
2304: \em{Exceptions} are \eqref{two_} if $\bt$ is strongly
2305: almost quasi-positive, and \eqref{one_} if $\bt$ is strongly
2306: almost quasi-negative. \qed
2307: \end{lemma}
2308: 
2309: \proof[of corollary \reference{CVP}]
2310: Depending on whether a 3-braid representation $\bt$ is strongly
2311: quasi-signed, almost quasi-signed, or none of both, $V$ determines
2312: $e=[\bt]$ via $\chi(\hat\bt)$ and/or \eqref{one_} or \eqref{two_}.
2313: With $V$ and $e$, one can recover the trace of the Burau matrix
2314: $\psi_2(\bt)$, and from that also $P(\hat\bt)$ (see \cite{Birman} or
2315: \eqref{V_3} below). Again, if $\md V\cdot \Md V\le 0$, then the
2316: options that $\bt$ is strongly quasi-signed or almost quasi-signed
2317: are ruled out easily. \qed
2318: 
2319: \subsection{Q polynomial}
2320: 
2321: We extend the scope of the previous results to the
2322: Brandt-Lickorish-Millett-Ho polynomial. The main aim here is to prove
2323: 
2324: \begin{theo}\label{thQ}
2325: Only finitely many non-trivial 3-braid links have given $Q$
2326: polynomial, and none has trivial (i.e., unlink) polynomial.
2327: \end{theo}
2328: 
2329: Note that by Kanenobu's work \cite{Kanenobu3}, again
2330: there are finite families of arbitrary large size,
2331: so we claim again a sort of contrary result.
2332: 
2333: While this result may be considered less relevant
2334: that its analoga for $\Dl$ and $V$, its proof displays
2335: the largest variety of tools necessary to apply, and 
2336: shows how the various approaches to link invariants
2337: (skein relations, state models and representation theory), 
2338: which are often considered in isolation, can usefully 
2339: complement each other. Indeed, $Q$ seems in general 
2340: more difficult to treat than $V$. Apart from Kidwell's 
2341: results \cite{Kidwell} for alternating links, and those
2342: in \cite{pos} for positive \em{knots}, neither the 
2343: non-triviality nor the finiteness property seem to have 
2344: been known previously for any other class of links.
2345: 
2346: The proof makes use of the full extent of the study of 
2347: Xu's form. Various additional ingredients will be necessary.
2348: One such is J. Murakami's $Q$-$V$ formula, found in
2349: \cite{Murakami} (by representation theory, and proved also
2350: later by Kanenobu \cite{Kanenobu} using skein relations).
2351: It allows to recur in many cases the problem from $Q$ to $V$.
2352: When dealing with $V$, we use beside the previous discussion
2353: the (Kauffman bracket based) formulas of \S\ref{Ssaq}.
2354: 
2355: In cases Murakami's formula is not helpful, we apply the
2356: Polyak-Viro formula \eqref{pv2} for Casson's knot invariant
2357: $v_2=\myfrac{1}{2}\Dl''(1)$,
2358: and the following formula, (likewise skein theoretic and) due
2359: to Kanenobu \cite{Kanenobu4}, that relates $v_2$ to $Q$.
2360: 
2361: \begin{theo}(\cite{Kanenobu4})
2362: For a link $L$ with $n$ components $K_i$, $i=1,\dots,n$, we have
2363: \begin{eqn}\label{Kanf}
2364: Q_L'(-2)=3(-2)^n\sum_iv_2(K_i)+3(-2)^{n-2}\sum_{i<j}lk(K_i,K_j)^2
2365: +(n-1)(-2)^{n-3}\,,
2366: \end{eqn}
2367: where $lk$ is the linking number.
2368: \end{theo}
2369: 
2370: Using this formula, we prove first a special case of theorem
2371: \ref{thQ}.
2372: 
2373: \begin{prop}\label{pgw}
2374: Only finitely many (non-trivial) strongly quasi-positive 3-braid
2375: links have given $Q$ polynomial, and none has trivial polynomial.
2376: \end{prop}
2377: 
2378: For the case of knots, we have the following more specific statement,
2379: that has also some independent meaning.
2380: 
2381: \begin{prop}\label{pgq}
2382: If $K$ is a strongly quasi-positive 3-braid knot, then $v_2(K)\ge
2383: g(K)$.
2384: \end{prop}
2385: 
2386: Note that in \cite{pos} we proved the same inequality for a
2387: general \em{positive} knot. Since the pretzel knots with
2388: trivial Alexander polynomial are strongly quasi-positive, nothing
2389: like this holds for a general strongly quasi-positive knot, though.
2390: 
2391: We need some preparations. Let us first fix a convention. We number
2392: braid strands by 1,2,3 from left to right in the bottom of the
2393: braid by and compose words from bottom to top. We propagate strand
2394: number through the crossings. In particular strand( number)s may 
2395: appear permuted in the middle or on top of the braid. To refer to the 
2396: ordering of the strands with regard ot the local position in the braid
2397: diagram, we speak of left, middle and right strand. So, for example, 
2398: if strands 1 and 2 enter below a $\sg_1$ (strand 1 as left strand and
2399: strand 2 as middle strand), then they exit above in the order 2\ 1 
2400: (strand 2 as left strand and strand 1 as middle strand). 
2401: 
2402: Let for a (not necessarily pure, and not closed) braid word the
2403: linking numbers of strand $i$, $j$ numbered as explained above be
2404: the sum of the writhes of all crossings these two strands pass.
2405: (Note that for links this differs by the additional factor 1/2,
2406: which is not relevant, though.)
2407: 
2408: \begin{lemma}\label{lmlk}
2409: Let $\bt$ be a strongly quasi-positive 3-braid word
2410: (not necessarily pure, and not closed). Then the
2411: linking numbers $l_{ij}$ between pairs $(i,j)$ of 
2412: strands satisfy $l_{13}+l_{23}>0$, unless $\bt$ is split.
2413: \end{lemma}
2414: 
2415: \proof The sum of writhes of crossings a strand passes
2416: through each of $\sg_{1,2,3}$ is non-negative. So
2417: $l_{13}+l_{23}\ge 0$. If $l_{13}+l_{23}=0$, then
2418: strand 3 passes only from below as left or middle
2419: strand into a $\sg_3=\sg_1^{-1}\sg_2\sg_1$. But since it
2420: starts (at the bottom of the braid) as a rightmost 
2421: strand, this means that is passes no crossing, so $\bt$ is split. 
2422: \qed
2423: 
2424: We prove proposition \ref{pgq}
2425: by induction on the length of the positive word
2426: in $\sg_{1,2,3}$. Assume $\bt$ is written as such a word.
2427: We can w.l.o.g. cyclically permute the indices. 
2428: 
2429: \begin{lemma}
2430: If $\hat\bt$ is a strongly quasi-positive 3-braid, then a positive 
2431: word $\bt$ can be chosen so that it contains a non-trivial syllable,
2432: \em{unless} $\bt$ is a power of $[123]$. In particular, it is 
2433: always possible if $\hat\bt$ is a knot, or a fibered link.
2434: \end{lemma}
2435: 
2436: \proof Choose Xu's form. It is $[12]^k\bt'$, where $k\ge 0$ and $\bt'$ 
2437: has cyclically non-decreasing indices. This word $\bt$ has a non-trivial
2438: syllable up to cyclic permutations of the letters, unless $\bt'$
2439: is a power of $[123]$. In this case, if $k>0$, we apply a YB relation 
2440: at the initial $121$ in $\bt$, and are done. \qed
2441: 
2442: \proof[of proposition \ref{pgq}] Let $\hat\bt$ be a knot. 
2443: We can w.l.o.g. now assume $\bt$ is written as a word with a 
2444: non-trivial syllable $s$. Now we apply the skein relation
2445: of $\Dl$ at a letter of $s$. Let $K=K,K_-,K_0$ be the skein
2446: triple, and $\bt_{\pm,0}$ the corresponding braids. Also we can 
2447: cyclically permute the indices in $\bt_{\pm,0}$. We choose them so 
2448: that the strand fixed by the permutation of $\bt_0$ is number 3. We 
2449: know from the skein relation of $\Dl$ 
2450: that $v_2(K)-v_2(K_-)=lk(L,M)$ where $L,M$ are the components 
2451: of $K_0$. Now that linking number is positive by lemma \ref{lmlk} 
2452: (unless $\bt_0$ is a split braid, but this is impossible because $s$
2453: is non-trivial). By induction on $g(K)$, the claim of proposition
2454: \ref{pgq} follows. \qed
2455: 
2456: 
2457: 
2458: \proof[of proposition \reference{pgw}] We distinguish three
2459: cases depending on the number of components of $\hat\bt$.
2460: 
2461: \begin{caselist}
2462: \case 1 component. In this case $Q'(-2)$, by Kanenobu's formula
2463: \eqref{Kanf},
2464: is a multiple of Casson's invariant $v_2$. The claim we wish to 
2465: show then is implied by the estimate in proposition \reference{pgq}.
2466: 
2467: \case 2 components. In this case Kanenobu's formula 
2468: involves only the square of the linking number of both components 
2469: $K,L$ of the closure link of $\bt$ and $v_2$ of the 2-string subbraid 
2470: component $K$. Looking at Xu's normal form, we see that if $\bt$ is 
2471: a strongly quasi-positive 3-braid, then writing $\bt$ as a word
2472: in $\sg_{1,2,3}$, and then expanding $\sg_3=\sg_1^{-1}\sg_2\sg_1$,
2473: we obtain a word in $\sg_{1,2}$ of some length $c$ with at most 
2474: $c/5$ negative letters. Since thus $\bt$ has at most $c/5$ negative
2475: crossings, it is easy to see that not both $v_2(K)$ and $lk(K,L)$
2476: can be zero, and one grows unboundedly when $c$ grows. So we are done.
2477: 
2478: \case 3 components. In this case Kanenobu's formula \eqref{Kanf} 
2479: reduces to the square sum of the linking number of pairs components.
2480: Let $l_{12},l_{13},l_{23}$ be these linking numbers. Again it is 
2481: clear that not all of $l_{12},l_{13},l_{23}=0$, and that when 
2482: $l_{12}+l_{13}+l_{23}\ge 2c/5$ grows unboundedly, then so will
2483: $l_{12}^2+l_{13}^2+l_{23}^2$, so again we are done.
2484: \end{caselist}
2485: \qed
2486: 
2487: Now we move on to settle the remaining cases in theorem \ref{thQ}.
2488: Here the tools differ considerably. Beside the study of Xu's form,
2489: we need the help of the following theorem, proved in \cite{ntriv}.
2490: 
2491: \begin{theorem}\label{thq}
2492: Any $3$-braid is up to conjugacy $A$-adequate or $B$-adequate.
2493: \end{theorem}
2494: 
2495: We proved also the following
2496: 
2497: \begin{theorem}
2498: A $3$-braid word is $B$-adequate if and only if it is (a) a
2499: negative word, or (b) it contains no $[121]$ as subword, and negative
2500: entries in the Schreier vector appear isolated. (That is, in
2501: cyclic order they are preceded and followed by a positive entry.)
2502: \end{theorem}
2503: 
2504: Using this criterion we prove now
2505: 
2506: \begin{lemma}\label{oia}
2507: Assume $\bt$ is strongly $\le 2$-almost positive and of exponent 
2508: sum $e\ge 0$. Then $\bt$ is $B$-adequate.
2509: \end{lemma}
2510: 
2511: \proof
2512: If $\bt$ is strongly quasi-positive or almost strongly
2513: quasi-positive, then it is of Xu's form $L^{-1}R$
2514: where $L$ is a single letter. By direct observation
2515: we can verify that when permuting indices cyclicly
2516: properly and writing out $\sg_3$ in $\sg_{1,2}$, then
2517: the resulting word has no $[121]$ subword, and negative
2518: entries in the Schreier vector appear isolated. (In case
2519: of $[(312)^k31-2]$, we must cancel 2 crossings first.)
2520: 
2521: If $\bt$ is 2-almost strongly quasi-positive, then then
2522: it is of Xu's form $L^{-1}R$ where $L$ has 2 letters.
2523: We can assume that $L=[12]$ or $L=[1^2]$, then we write down
2524: the subword of $L^{-1}R$ consisting of the first and last
2525: letters of $R$ and $L$ (herein $k\ge 1$):
2526: \[
2527: \begin{array}{ll}
2528: [[23] \dots 33-2-1] & [[23] \dots 22-1-1] \\{}
2529: [[23] \dots 23-2-1] & [[23] \dots 31^k2-1-1] \\{}
2530: [[23] \dots 31^k-2-1] & [[23] \dots 33-1-1] \\{}
2531: & [[23] \dots 23-1-1]
2532: \end{array}
2533: \]
2534: (Here `[23]' means as in \S\ref{brs} that the word is to begin with
2535: `2' or `3'.) The reductions and $B$-adequacy test are done
2536: case-by-case. \qed
2537: 
2538: Using the formulas in \S\reference{Ssaq}, we prove
2539: 
2540: \begin{lemma}\label{yu}
2541: If $L=\hat\bt$ is a link, which is the closure of an $A$-semiadequate 
2542: $3$-braid $\bt$, and $V_0V_1=-1$, then $V_0V_2=1$ or $2$.
2543: \end{lemma}
2544: 
2545: \proof By \cite{ntriv} a A-semiadequate $3$ braid is either
2546: positive, in which case $V_1=0$, or has no $[-1-2-1]$ and
2547: positive entries are isolated in the Schreier vector.
2548: In \cite{ntriv} we described the words  for latter
2549: type which have $V_1=0$. They are (up to braid relations
2550: and cyclic letter permutations) of the form%
2551: \footnote{Note that these are exactly the 3-braids which are
2552: reducible in the terminology of dynamic properties;
2553: this reducibility has nothing to do, though, with the
2554: reducibility in Markov's theorem.}
2555: \[
2556: [1^{-2}2^{-2}1^{-2}\dots 1^{-2}2^{-1}1^p2^{-1}]\,.
2557: \]
2558: 
2559: Analogously to that study, we can see that if $V_0V_1=-1$,
2560: then $\bt$ has a word which is obtained from the above type
2561: by replacing some $1^{-2}$ by $1^{-k}$ 
2562: for $k\ge 3$. The intertwining graph of
2563: the $A$-state is a path (all vertices have valence 2, except two
2564: of valence 1), so $\chi(IG)=1$. We have $\bigtriangleup
2565: =1$ if $k=3$, and otherwise $\bigtriangleup=0$. So by
2566: theorem \reference{T02}, we have $V_0V_2=1$ for
2567: $k=3$, and $V_0V_2=2$ otherwise. \qed
2568: 
2569: To explain what such a property has to do with $Q$, 
2570: now we must introduce J. Murakami's formula \cite{Murakami}.
2571: (See also Kanenobu \cite[Theorem 2]{Kanenobu}.)
2572: 
2573: Let $i=\sqrt{-1}$, $u=\sqrt{-t}$ and $x=u+u^{-1}$. Let further%
2574: \footnote{Note that $V$ is here what is written as $J$, and not $V$, 
2575: in \cite{Kanenobu}.} for a braid $\bt$ of exponent sum $e$,
2576: \begin{eqn}\label{chi}
2577: \chi(\bt,t)=i^eu^{-2e}V_{\hat\bt}(t)+u^{-e}(x^2-2)\,.
2578: \end{eqn}
2579: 
2580: Then Murakami's formula is
2581: 
2582: \begin{theo}(J. Murakami)\label{TMF}
2583: If $L$ is a closure of a 3-braid $\bt$ of exponent sum $e$,
2584: then
2585: \begin{eqn}\label{sonjo}
2586: Q(L,x)\,=\,\chi(\bt,\sqrt{t})^2-1\,+\,\frac{2(x^2+x-1)}{x^2(x^2-3)}
2587: (u^e+e^{-e})+\frac{-x^4-2x^3+3x^2+4x-4}{x^2(x^2-3)}\chi(\bt,t)\,.
2588: \end{eqn}
2589: \end{theo}
2590: 
2591: \proof[of theorem \ref{thQ}] We can w.l.o.g. (taking the mirror image)
2592: assume that $e\ge 0$, and (excluding trivial to check special cases)
2593: that $\chi\le 0$. Clearing denominators and absolute terms in
2594: \eqref{sonjo}, we find
2595: \begin{eqnarray}
2596: \label{Q1} Q_1(L,x):=x^2(x^2-3)(Q(L,x)+1) & = &
2597: \bigl[ i^e(-t)^{-e}V(t)+(-t)^{-e/2}(-t-\frac{1}{t}) \bigr]
2598: \cdot (-x^4-2x^3+3x^2+4x-4) \\
2599: \nonumber & &  + \bigl[ i^e(-t)^{-e/2}V(-\sqrt{-t})+(-t)^{-e/4}x \bigr]^2
2600: \cdot x^2(x^2-3) \\
2601: \nonumber & & +2(x^2+x-1)(\sqrt{-t}^e+\sqrt{-t}^{-e})\,.
2602: \end{eqnarray}
2603: 
2604: Rearranging, we need to show that the sum of the following 5
2605: expressions,
2606: regarded as a polynomial in $t^{\pm 1/2}$, has either arbitrarily
2607: small minimal degree or arbitrarily large maximal degree.
2608: (Note that by \eqref{sonjo} this sum must be self-conjugate in 
2609: $t$ up to coefficient signs, which is not at all evident directly.)
2610: 
2611: \begin{eqnarray*}
2612: T'_1 & = & i^e(-t)^{-e}V(t)\cdot (-x^4-2x^3+3x^2+4x-4) \\
2613: T'_2 & = & t^{-e}\bigl[V(-\sqrt{-t})\bigr]^2\cdot x^2(x^2-3) \\
2614: T'_3 & = & 2i^e(-t)^{-3e/4}V(-\sqrt{-t})x^3(x^2-3) \\
2615: T'_4 & = & (-t)^{-e/2}\bigl[(x^2-2)(-x^4-2x^3+3x^2+4x-4)+x^4(x^2-3)+2(x^2+x-1)
2616:  \bigr] \\
2617: T'_5 & = & (-t)^{e/2}\bigl[2(x^2+x-1) \bigr] 
2618: \end{eqnarray*}
2619: 
2620: Switching $-t\to t$ to simplify the expressions, and regrouping terms, 
2621: we have
2622: \begin{eqnarray*}
2623: T_1 & = & i^et^{-e}V(-t)\cdot (-1\ -2\ -1\ -2\ [-4]\ -2\ -1\ -2\ -1) \\
2624: T_2 & = & (-t)^{-e}\bigl[V(-\sqrt{t})\bigr]^2\cdot (1\ 0\ 1\ 0\ [0]\ 0\ 1\ 0\ 1) \\
2625: T_3 & = & 2i^et^{-3e/4}V(-\sqrt{t})\cdot (1\ 0\ 2\ 0\ 1\ [0]\ 1\ 0\ 2\ 0\ 1) \\
2626: T_4 & = & t^{-e/2}\bigl[-2(x-1)(x^2-3)(x^2-1) \bigr] \\
2627: T_5 & = & t^{e/2}\bigl[2(x^2+x-1) \bigr] 
2628: \end{eqnarray*}
2629: Now $u=t^{1/2}$ and $x=t^{1/2}+t^{-1/2}$. The last factors for $T_
2630: {1,2,3}$ are given as a list of coefficients, with the absolute term
2631: put in brackets. Note that these are polynomials in $\sqrt{t}$, so for
2632: example, the first two polynomials have degree $2$ in $t$.
2633: 
2634: Since we are concerned with cancellations of the leading and trailing
2635: coefficients of $T_k$, let us compile their minimal $m_k=\md_tT_k$ and
2636: maximal degrees $M_k=\Md_tT_k$ (note that by taking degrees w.r.t.
2637: $t$, and not $\sqrt{t}$, the $m_k$, $M_k$ are only half-integers).
2638: \[
2639: \begin{array}{c|c|c}
2640: \raisebox{-1em}{\ry{2.3em}} k & m_k & M_k \\
2641: \hline
2642: \raisebox{-1em}{\ry{2.5em}}1 & \md V-e-2 & \Md V-e+2 \\[2mm]
2643: 2 & \md V-e-2 & \Md V-e+2 \\[2mm]
2644: 3 & -\ffrac{3e}{4}+\ffrac{\md V}{2}-\ffrac{5}{2} & -\ffrac{3e}{4}+
2645:     \ffrac{\Md V}{2}+\ffrac{5}{2}\\[2mm]
2646: 4 & -\frac{e}{2}-\frac{5}{2} & -\frac{e}{2}+\frac{5}{2} \\[2mm]
2647: 5 & \frac{e}{2}-1 & \frac{e}{2}+1\raisebox{-1em}{\ry{1.7em}}
2648: \end{array}
2649: \]
2650: (Here again $\md V=\md_t V(t)$ is a half-integer, and similarly
2651: $\Md V$.)
2652: 
2653: Our aim will be to determine what of the $M_k$ is the largest, and 
2654: to show that the coefficients of the contributing $T_k$ do not cancel.
2655: An important case where we exclude (problematic) cancellation is
2656: 
2657: \begin{lemma}\label{lmcan}
2658: Assume that 
2659: \begin{eqnarray}
2660: \label{GG} M_1=M_2 & \ge & M_k+3/2\quad\mbox{ and } \\
2661: \label{gg} m_1=m_2 & \le & m_k-3/2
2662: \end{eqnarray}
2663: for $k=3,4,5$.
2664: Then $\Md_t Q_1(L,x)\ge M_1-1$. (Here $Q_1(L,x)$ refers to \eqref{Q1}.)
2665: If $L$ the closure of a 3-braid which is $A$-adequate resp. $B$-adequate
2666: then assuming \eqref{gg} resp.\ \eqref{GG} alone is sufficient.
2667: \end{lemma}
2668: 
2669: \proof By theorem \ref{thq}, $L$ is the closure of a 3-braid $\bt$ which 
2670: is either $A$-adequate or $B$-adequate. We consider the $B$-adequate case 
2671: and show that $M_1=M_2\ge M_k+3/2$ implies $\Md_t Q_1(L,x)\ge M_1-1$.
2672: If $\bt$ is $A$-adequate, we obtain similarly $\md_t Q_1(L,x)\le 
2673: m_1+1$, and the result follows by the (anti)symmetry of $Q_1$.
2674: 
2675: Now if $M_1=M_2\ge M_k+3/2$, the last three coefficients of $M_{1,2}$
2676: are not cancelled from other $M_k$ (remember these are polynomials
2677: in $\sqrt{t}$, and coefficients are meant for such polynomials), 
2678: so it is enough to show that they do not (completely) cancel 
2679: among each other. Now a look at the formulas for $T_{1,2}$ shows
2680: that if these three coefficients are all to cancel, then (apart from
2681: proper sign coincidences) we must have $\bar V_1\bar V_0=-1$ and 
2682: $\bar V_2=0$. (Take also into account the switch $-t\to t$.)
2683: However, this situation was ruled out by lemma \ref{yu}. \qed
2684: 
2685: We assumed $e\ge 0$; we also excluded the case $e=3-\chi$ of
2686: strongly quasi-positive braids. To apply the lemma, we need to
2687: establish \eqref{gg} and/or \eqref{GG}. For this we use now
2688: lemma \reference{lemma_}.
2689: 
2690: Clearly, a strongly almost quasi-negative braid should not 
2691: be considered for $e\ge 0$; the almost quasi-positive braid
2692: require a small extra argument, which is given below. Now
2693: \begin{eqn}\label{hj}
2694: \Md V-e\,=\,\frac{1-\chi}{2}+1\,\ge\,\frac{e}{2}+1\,,
2695: \end{eqn}
2696: so $M_3\ge M_4$, and we can discard $M_4$. Also from \eqref{hj}
2697: we obtain
2698: \[
2699: M_1-M_5-1=\Md V-\frac{3e}{2}\ge 1\,,
2700: \]
2701: so $M_1>M_5+1$, and can neglect $M_5$. Similarly, if
2702: $\bt$ is not almost strongly quasi-positive, we have with 
2703: \eqref{two_} and $\md V<e/2$ that
2704: \begin{eqn}\label{mmm}
2705: m_3< m_4<m_5\,,
2706: \end{eqn}
2707: so $m_{4,5}$ are also irrelevant.
2708: 
2709: So we need to deal only with $m_k,M_k$ for $k\le 3$. Moreover, we 
2710: see from \eqref{one_} that $M_1\ge (1-\chi)/2+3$. So whenever we can 
2711: apply lemma \ref{lmcan}, we have for $\chi\le 0$ that $\Md Q_1\ge 
2712: M_1-1>2$, so $\Md_z Q>0$, and $Q$ is not an unlink 
2713: polynomial. Moreover, we have for a sequence of links
2714: from $\chi\to -\infty$ also $\Md M_1\to \infty$, so
2715: $\Md Q\to \infty$, and a given (even just degree of a)
2716: $Q$ polynomial occurs only finitely many times, as desired.
2717: A similar argument applies if we use \eqref{two_}.
2718: 
2719: To apply lemma \ref{lmcan}, we need to check $M_3<M_1-1$ and $m_3>m_1+
2720: 1$. Note that if $\bt$ is almost strongly quasi-positive, then \eqref
2721: {two_} may not hold. So may not be able to apply lemma \ref{lmcan}
2722: directly. However, we remedied this by showing in lemma \ref{oia}
2723: that $\bt$ is $B$-adequate. Then \eqref{one_} holds, and it is enough
2724: to use $M_3<M_1-1$ (the condition \eqref{mmm} becomes also irrelevant).
2725: 
2726: Now using \eqref{one_} and \eqref{two_}
2727: \begin{eqnarray}
2728: \nonumber
2729: m_1-m_3+1 & = & \md V -e -2 +\frac{3e}{4}-\frac{\md V}{2}+\frac{5}{2}+1
2730: \,=\,\frac{\md V}{2}-\frac{e}{4}+\frac{3}{2} \\
2731: \nonumber
2732: & = & \frac{e}{2}-\frac{1-\chi}{4}-\frac{1}{2}-\frac{e}{4}+\frac{3}{2}\,=\,
2733: \frac{e}{4}-\frac{1-\chi}{4}+1\,, \\
2734: \label{iua}
2735: M_1-M_3-1 & = & \frac{\Md V}{2}-\frac{e}{4}-\frac{3}{2} \\
2736: \nonumber
2737: & = & \frac{e}{2}+\frac{1-\chi}{4}+\frac{1}{2}-\frac{e}{4}-\frac{3}{2}\,=\,
2738:       \frac{1-\chi}{4}+\frac{e}{4}-1\,.
2739: \end{eqnarray}
2740: 
2741: Then \eqref{GG} becomes equivalent to
2742: \[
2743: \frac{1-\chi}{4}+\frac{e}{4}>1\,.
2744: \]
2745: So, assuming $\chi\le 0$ and $e\le 1-\chi$, for \eqref{GG} 
2746: we are left to consider only $(e,\chi)=(0,-1)$, $(0,-3)$,
2747: $(1,0)$, $(1,-2)$, or $(2,-1)$, which are trivial cases to check. 
2748: 
2749: For \eqref{gg} we must deal with $A$-adequate but not $B$-adequate
2750: braids with $e\ge -3-\chi$, i.e. $\le 3$-almost strongly
2751: quasi-positive braids. For $\le 2$-almost strongly
2752: quasi-positive braids we can use lemma \reference{oia}.
2753: While one may extend the argument there to $3$-almost strongly
2754: quasi-positive braids, it would require a longer
2755: case-by-case analysis. We provide instead a different argument.
2756: 
2757: We try to modify the argument for proposition \reference{pgw}.
2758: In case $\hat\bt$ is a link of 2 or 3 components, we can argue again
2759: as there, and are left with the cases $e=0,1$ of $6$ and $7$ bands,
2760: which are easy to deal with.
2761: 
2762: So assume $\hat\bt$ is a knot.
2763: Now from \eqref{iua}, we have $m_3=m_1+1$, so if still some of the
2764: first two coefficients of $T_1$, $T_2$ do not cancel, we have
2765: $\md_t Q_1\le m_1+\myfrac{1}{2}$, and are done.
2766: Otherwise again $V_1V_0=-1$. So by the proof of lemma \reference{yu},
2767: \[
2768: \bt=[1^{-k}2^{-2}1^{-2}\dots 1^{-2}2^{-1}1^l2^{-1}]\,,
2769: \]
2770: with $k\ge 3$ and $l\ge 1$. Since $\hat\bt$ is a knot, $k$ and $l$
2771: are easily observed to be odd. Now a $\le 3$-almost positive
2772: word in Xu's form can be chosen, when written out as a word in
2773: $\sg_{1,2}$ of $c$ letters, to have $c_-\le c/5+3$ negative
2774: letters/crossings. If we reduce the word by cancelling
2775: crossings, this inequality remains true.
2776: 
2777: We estimate the Casson invariant $v_2$ of $\hat\bt$ using the Gau\ss{}
2778: diagram formula of Polyak-Viro \eqref{pv2}. Let us put the
2779: basepoint right after the group of $l$ positive crossings (which
2780: of either strands is immaterial). 
2781: 
2782: Now $l$ is odd and the $l$ positive crossings are pairwise linked,
2783: so their contribution to the Gau\ss{} diagram sum \eqref{pv2} is,
2784: independently on the location of the basepoint, equal to $v_2(T_{2,l})
2785: =\ffrac{l^2-1}{8}$, where $T_{2,l}$ is the $(2,l)$-torus knot. The
2786: analogous claim is true for the group of $k$ negative crossings.
2787: 
2788: If some of the positive crossings is linked with a negative crossing,
2789: it must be a negative crossing of a $\sg_2^{-1}$. Now by our choice
2790: of basepoint, in
2791: each syllable $\sg_2^{-2}$ only one of the two crossings gives a pair
2792: that contributes to \eqref{pv2}, and the contribution is always
2793: $-1/2$. Putting this together, we have
2794: \begin{eqn}\label{oiz}
2795: v_2\,\ge\,\frac{l^2-1}{8}+\frac{k^2-1}{8}-\frac{c_--k+2}{4}\cdot l
2796: \cdot \frac{1}{2}\,.
2797: \end{eqn}
2798: Assuming $c\ge 10$ (the other cases are checked directly), we have
2799: \[
2800: l\,\ge\frac{4c}{5}-3\,\ge\,\frac{c}{5}+3\,\ge\,c_-=c-l\,.
2801: \]
2802: So the r.h.s. in \eqref{oiz} is minimized for $l\,\ge\ffrac{4c}{5}-3$
2803: when putting $l=\ffrac{4c}{5}-3$. Then we have, with $k\ge 3$ and
2804: $c_-=c-l$, the estimate
2805: \[
2806: v_2\,\ge\,\frac{3c^2}{50}-\frac{29c}{40}+\frac{11}{4}\,,
2807: \]
2808: which is positive for $c\ge 10$, and grows when $c\to \infty$. 
2809: 
2810: With this theorem \ref{thQ} is proved. \qed
2811: 
2812: \begin{rem}
2813: It is interesting whether $\Md Q\to\infty$ also for the strongly 
2814: quasi-positive braids, but in Murakami's formula massive cancellations 
2815: become possible and $Q_1$ cannot be easily controlled.
2816: At least one can prove using theorem \ref{thq} and
2817: some results of Thistlethwaite in \cite{Thistle} that
2818: there are only finitely many 3-braid links with given
2819: $\Md_zF$, where $F$ is the Kauffman polynomial.
2820: \end{rem}
2821: 
2822: \section{Positivity of 3-braid links}
2823: 
2824: \subsection{Positive braid links\label{S2.2}}
2825: 
2826: \subsubsection{The Morton-Williams-Franks bound}
2827: 
2828: For the proof of theorem \reference{thg} we will need to study
2829: the behaviour of the bound \eqref{mwfb} in the Morton-Williams-Franks
2830: inequality (we abbreviate as MWF) on positive
2831: braids. This was begun by Nakamura \cite{Nakamura},
2832: who settled the case $MWF=2$ in the suggestive way: such braids
2833: represent only the $(2,n)$-torus links. (The case $MWF=1$ is
2834: trivial.) We will introduce a method that considerably simplifies
2835: his proof (but still makes use of some of his ideas), and
2836: then go on to deal with $MWF=3$. The example of non-sharp
2837: MWF inequality, $13_{9365}$ in \cite{KnotScape} (the connected
2838: 2-cable of the trefoil), given in \cite{MorSho}, is in fact
2839: only among a small family of exceptional cases.
2840: 
2841: \begin{theorem}\label{thmwf}
2842: If $b(\hat\bt)>MWF(\hat\bt)=3$, then $\bt$ reduces to a 4-braid,
2843: and is given by one of the following forms (assuming that $3*$
2844: denotes a sequence of at least one letter $3$, and $11*$
2845: resp.\ $22*$ sequences of at least two letters $1$ or $2$):\\
2846: \[
2847: [22*3*122*11*23211*], \quad\mbox{and}\quad [22*3122*11*23*211*].
2848: \]
2849: \end{theorem}
2850: 
2851: 
2852: We will reduce the proof to a finite number of words to
2853: check, which is done by calculation using the program of
2854: \cite{MorSho}. Since a direct computation is more reliable
2855: than an increasingly difficult mathematical argument, we have
2856: not tried to minimize the calculation by all means. However,
2857: we point out that for the sake of theorem \reference{thg}
2858: alone (rather than its refinement, theorem \reference{thmwf}),
2859: the following weaker statement is sufficient, for which
2860: a considerable part of the case-by-case calculations can
2861: be dropped. This corollary requires the notion of
2862: semiadequacy \cite{LickThis}, and can be deduced from theorem
2863: \reference{thmwf} by direct check of the exceptional words.
2864: (We will sometimes write $MWF(\bt)$ for $MWF(\hat\bt)$.)
2865: 
2866: \begin{corr}\label{cort}
2867: If $\bt$ is a positive braid word, and $MWF(\bt)=3$, then $\bt$
2868: reduces (up to Markov equivalence) to a positive 4-braid word $\bt'$,
2869: and the diagram $\hat\bt'$ is not $B$-adequate.
2870: \end{corr}
2871: 
2872: Here the notions of $A$-adequate and $B$-adequate for diagrams
2873: and braids are as explained in \S\ref{brD} and \S\ref{brt}. We
2874: note, as a consequence of \cite{Thistle}, that a braid is ($A$/%
2875: $B$-)adequate if and only if some, or equivalently any,
2876: minimal length word of a braid in its conjugacy class it is so.
2877: 
2878: We will thus prove theorem \reference{thg} only using corollary
2879: \reference{cort}, and indicate in the proof of theorem \reference{thmwf}
2880: the point where the corollary follows (and the rest of the argument
2881: is not needed). The argument that elegantly replaces the remaining
2882: case-by-case checks requires theorem \reference{thq}. Since (by
2883: taking again the full extent of our proof) theorem \reference{thq}
2884: is nonetheless not indispensable, we permit ourselves to defer its
2885: proof to a separate paper \cite{ntriv}.
2886: 
2887: 
2888: % \proof It is easy to see that each positive irreducible $\bt\in B_4$
2889: % with a $\sg_2$ can be written as $\sg_2^{k_1}\sg_1^{l_1}\sg_3^{m_1}\dots
2890: % \sg_2^{k_n}\sg_1^{l_n}\sg_3^{m_n}$, where $m_i>0$ and $l_i+m_i>0$.
2891: % 
2892: % Now if $\bt$ contains no YB relation, then all if $l_i+m_i>1$,
2893: % and $n>1$ or $l_1,m_1>1$.
2894: % 
2895: % Now if all syllables of $\sg_2$ are non-trivial, we can remove
2896: % in the positive resuloution tree and have the split union of
2897: % two non-trivial (because $\bt$ irreducible) $(2,k)$-torus links,
2898: % so $MWF=4$$. If ....
2899: % \proof Check directly. ($B$-adequacy is invariant under regular isotopy,
2900: % so any other 4-braid word giving the same knot is not $B$-adequate
2901: % either.) \qed
2902: 
2903: \proof[of theorem \reference{thg}]
2904: If $MWF(\bt)\le 2$, then we are done. So assume $MWF(\bt)=3$.
2905: Assume first $L$ has an $A$-adequate 3-braid. Since $A$-adequate
2906: diagrams minimize the number of negative crossings, and $L$ has
2907: a positive (braid) diagram, the $A$-adequate 3-braid diagram is
2908: positive, and we are done. So let $L$ have a $B$-adequate 3-braid
2909: diagram $\hat\bt'$. Now by corollary \reference{cort}, we find
2910: that $L$ reduces to a positive 4-braid $\bt$
2911: with $c_+(\bt)$ positive crossings, and $\bt$ is not $B$-adequate.
2912: Since $\bt'$ is $B$-adequate and $B$-adequate diagrams
2913: minimize the number of positive crossings, $c_+(\bt')<c_+(\bt)$.
2914: On the other hand, $MWF(L)=3$ and the inequalities of MWF for the
2915: $v$-degree of $P$ show that $\bt'$ must have exponent sum
2916: $[\bt']=c_+(\bt)-1$. Since $[\bt']\le c_+(\bt')$, we must have
2917: equality, so $\bt'$ is positive. \qed
2918: 
2919: Note the following easy and useful consequence of theorem 
2920: \reference{thg}:
2921: 
2922: \begin{corr}
2923: A link which is a closure of a positive braid of at most 4 strings,
2924: has a minimal crossing diagram as a closed positive braid, and a
2925: minimal string positive braid representation.
2926: \end{corr}
2927: 
2928: \proof The case of the braid representation is straightforward,
2929: and it implies the minimal crossing diagram statement by
2930: looking at $\Md_zP$ and using Morton's inequalities. \qed
2931: 
2932: The examples in \cite{posex_bcr}, mentioned after theorem \ref{thg},
2933: show that the corollary is not true in case of positive 5-braids at
2934: least for the positive minimal braid representation. So far no examples
2935: are known where no minimal crossing positive braid diagram exists
2936: (it was known to exist from \cite{WilFr,Murasugi} for closed
2937: positive braids with a full twist, which include the torus links,
2938: and from \cite{posex_bcr} for positive braid knots of at most
2939: 16 crossings), but the pathologies for minimal strings hint to caution.
2940: It was shown in \cite{posex_bcr} that one, and in \cite{ntriv} that
2941: infinitely many fibered positive knots
2942: have no minimal crossing positive diagram.
2943: 
2944: \subsubsection{Maximal subwords}
2945: 
2946: Here we start the technical considerations needed to prove 
2947: theorem \reference{thmwf}. We consider the form \eqref{wd},
2948: now with all $l_i>0$.
2949: 
2950: \begin{defi}
2951: We define \em{summit syllables} in \eqref{wd}:
2952: \def\theenumi{\alph{enumi}}
2953: \def\labelenumi{\theenumi)}
2954: \begin{enumerate}
2955: \item All $\sg_{n-1}^{l_i}$ are summit syllables, and
2956: \item if $\ap=\sg_k^{l_i}$ and $\ap'=\sg_k^{l_j}$ are summit
2957: syllables, with no $\sg_{k'}^{l'}$ for $k'\ge k$ occurring between
2958: $\ap$ and $\ap'$, then all $\sg_{k-1}^{l'}$ occurring between $\ap$
2959: and $\ap'$ are summit syllables.
2960: \end{enumerate}
2961: \end{defi}
2962: 
2963: Note that, according to definition \reference{dfext}, we
2964: consider syllables in cyclic order. The relation ``between'' in
2965: the above definition should also be understood in that sense:
2966: a syllable occurring after the last index $i$ syllable
2967: $\ap$ and/or before the first index $i$ syllable $\ap'$ is
2968: considered to be between $\ap$ and $\ap'$.
2969: 
2970: For the following considerations it is (not necessary but) helpful
2971: to visualize $\bt$ by the braid scheme explained in \cite{bind}.
2972: 
2973: Separate $\bt$ in
2974: \eqref{wd} into subwords $\ap_1\dots \ap_n$, such that $\ap_i$
2975: contains only syllables of odd or even index, and this parity
2976: changes between $\ap_i$ and $\ap_{i+1}$. Then for a syllable
2977: $\sg_k^l$ occurring in $\ap_i$, put the integer $l$ at the point
2978: $(k,i)\in \bN\times \bN\subset \bR^2$ in the plane. Here $(k,i)$
2979: is the point in the $i$-th row and $k$-th column, with rows
2980: numbered (as in Cartesian coordinates) from bottom to top and
2981: columns from left to right.
2982: 
2983: One obtains a certain checkerboard pattern of integers we call
2984: \em{braid scheme} of $\bt$. (If we do not put any integer on
2985: a point $(k,i)$, we assume its ``content'' is zero, or it is
2986: ``empty''. So for all non-empty points $(k,i)$ in the scheme,
2987: $i+k$ is always even or always odd.)
2988: 
2989: One can \em{reduce} the
2990: scheme by moving an integer $l$ at $(k,i)$ to $(k,i-2)$ if $i>2$
2991: and the points $(k\pm 1,i-1)$ are empty. We call the scheme
2992: reduced if it does not admit any such move. Then in a reduced
2993: scheme, summit syllables of $\bt$ are those, whose entries in
2994: the scheme are ``on top'' when viewing the scheme from the left.
2995: {}From this viewangle the following ``geographic'' choice of
2996: terminology becomes more plausible.
2997: 
2998: 
2999: \begin{defi}
3000: Summit syllables still have a cyclic order from \eqref{wd}. We call
3001: the subword $\bt'$ of $\bt$ in \eqref{wd} made of summit syllables
3002: the \em{maximal subword}. The subword made of non-summit syllables
3003: (i.e. the subword obtained by deleting in $\bt$ all syllables in
3004: $\bt'$) is called \em{non-maximal subword}. 
3005: 
3006: Note that neighbored summit syllables have indices $k_i$ differing
3007: by $\pm 1$. We say that a summit syllable is \em{minimal} resp.
3008: \em{maximal} if its both neighbors have higher resp.\ lower index.
3009: 
3010: We call $\bt$ \em{summit reduced} if all its minimal summit syllables
3011: are non-trivial.  We call $\bt$ \em{index reduced} if it is
3012: non-singular and its index sum $\sum_{i=1}^n k_i\cdot l_i$ cannot be
3013: reduced by a Yang-Baxter relation, i.e. $\bt$ contains no
3014: $\sg_{i+1}\sg_i\sg_{i+1}$ as subword.
3015: \end{defi}
3016: 
3017: \begin{lemma}
3018: Index reduced $\So$ summit reduced. In particular a summit reduced
3019: form always exists.\qed
3020: \end{lemma}
3021: 
3022: Recall that a positive resolution tree is a rooted tree with directed 
3023: edges,
3024: whose vertices (nodes) contain positive braid words, the root labelled
3025: by $\bt$. Every vertex has exactly one incoming edge, except the
3026: root that has none, and zero or two outoutgoing edges. In former
3027: case it is labelled by an unlink (terminal node).
3028: In latter case it is labelled by
3029: a word of the form $\ap\sg_i^2\ap'$, with $\ap,\ap'$ positive
3030: words, and the two vertices connected by the outgoing edges 
3031: are labelled by $\ap\sg_i\ap'$ and $\ap\ap'$, or positive words
3032: obtained therefrom by Markov equivalence (isotopy of the closure link).
3033: 
3034: In \cite{Nakamura} the following fact was observed, and used
3035: decisively, and we shall do the same here.
3036: 
3037: \begin{theorem}(Nakamura \cite{Nakamura})
3038: $MWF(\bt)$ is the maximal number of components of a (link in a) node
3039: in a positive resolution tree for $\bt$.
3040: \end{theorem}
3041: 
3042: In particular, $MWF$ is monotonous (does not decrease) under word
3043: extension, and does not depend on the exponent of non-trivial
3044: syllables.
3045: 
3046: \begin{lemma}\label{lemh}
3047: If $\bt$ is \em{summit reduced}, and $\bt'$ is obtained from $\bt$
3048: by removing all summit syllables, then there is a positive resolution
3049: tree for $\bt$ that contains $\bt'$ as a node.
3050: \end{lemma}
3051: 
3052: \proof Since minimal syllables are non-trivial, one can delete them
3053: in the resolution tree. The two neighbors in the maximal subword
3054: join to a non-trivial new minimal syllable, and so one iterates the
3055: procedure. \qed
3056: 
3057: Since all $\sg_{n-1}^{l'}$, $\sg_{n-2}^{l'}$ in \eqref{wd} occur as
3058: summit syllables, $\bt'$ has split last two strands, and so we have
3059: a quick proof of Nakamura's main result.
3060: 
3061: \begin{corr}(Nakamura \cite{Nakamura})
3062: Any summit reduced positive word on $n\ge 3$ strands has $MWF\ge 3$.
3063: \qed
3064: \end{corr}
3065: 
3066: In particular any positive braid representation of a $(2,n)$-torus
3067: link can be reduced to the standard one by index-decreasing YB
3068: relations and removals of nugatory crossings.  
3069: 
3070: \subsubsection{The proof of Theorem \reference{thmwf}:
3071: Initial simplifications\label{4.0}}
3072: 
3073: The following fact is well-known:
3074: 
3075: \begin{theorem}\label{th4.1}
3076: $MWF(\bt)=1$ if an only if $[\bt]_i=1$ for $i=1,\dots,n-1$.
3077: \end{theorem}
3078: 
3079: Now for $MWF=3$ it suffices to ensure that (either we can reduce the
3080: braids) or can find words, whose non-maximal subwords do not
3081: give an unknot. \em{For the rest of the section we assume that
3082: $\bt'$ gives the unknot.}
3083: 
3084: We will work by induction on the number of strands, and for
3085: fixed number of strands on the index sum. So we consider a
3086: positive braid word $\bt$, and assume w.l.o.g. it has the
3087: smallest index sum among positive braid representatives of
3088: its closure link for the same number of strands. For such
3089: $\bt$, we will either reduce it (by at least one strand or
3090: crossing), or show $MWF\ge 4$.
3091: 
3092: Most braids $\bt$ will be easily dealt with, but there remain certain
3093: families of words that require a case-by-case study. We decided not
3094: to omit too many of the (tedious) details of this part, in order
3095: to keep the proof followable, even if it may not contribute to
3096: its (esthetic) appearance.
3097: 
3098: Note that in order to prove $MWF(\bt)\ge 4$, it suffices to
3099: go over to a (link in a) suitably chosen node in a positive
3100: resolution tree for $\bt$ and show $MWF\ge 4$ for this node. In
3101: particular, we can remove from $\bt$ all syllables of index $\le
3102: k$ and the $k$ resulting left isolated strands.
3103: 
3104: The case of reductions is more delicate. In some situations
3105: we can describe them directly, but this is not always the case.
3106: Then we proceeded as follows. First we took generic examples, in
3107: making all syllables non-trivial whose triviality we have not argued
3108: about. We adjust parities so that the closure is a knot, and checked
3109: using KnotScape \cite{KnotScape} that the braid reduces (by at least one
3110: strand/crossing).
3111: 
3112: Later we wrote a computer program that seeks reductions by keeping
3113: given crossings rigid. Such reductions would commute with replacing
3114: rigid crossings by any tangle, in particular by any non-trivial
3115: braid word syllable. (Non-trivial syllables behave
3116: similarly to  rigid vertices, and suggest that the reduction
3117: is likely to work in general.) This way we can find reductions for
3118: infinite families of braids on a given number of strands. By turning
3119: all summit index-1 syllables into rigid crossings, one can also
3120: handle the braids that occur for an increasing number of strands.
3121: 
3122: The technical details of the application of this program are,
3123: however, tedious and little insightful. Instead we content
3124: ourselves with giving the examples we processed with KnotScape.
3125: % The alternative proof of theorem \reference{thg} given later
3126: % is another reason why we felt not motivated to make this
3127: % way the approach here even more technical.
3128: 
3129: 
3130: \begin{defi}
3131: A \em{valley} resp. \em{mountain} is a subword of the maximal word
3132: that starts with same index syllables and contains only one minimal
3133: resp. maximal syllable. This syllable is called the \em{bottom} of
3134: the valley resp.\ \em{summit} or \em{top} of the mountain.
3135: The index of the bottom/top is the \em{depth} resp.\ \em{height}.
3136: \end{defi}
3137: 
3138: We assume there are at least two mountains of maximal height (i.e.
3139: $n-1$). Otherwise we have a split component or a $(2,n)$-torus
3140: connected component, or a reducible braid and can work by induction
3141: on the number of strands. Similarly at least one valley has
3142: depth $1$, otherwise $\sg_1$ in $\bt'$ remains reducible in $\bt$.
3143: 
3144: The following operation will be somewhat important, and we will
3145: call it ``filling the valley''.
3146: 
3147: \begin{lemma}(``filling the valley'')
3148: Any valley can be removed from the maximal subword in the positive
3149: resolution tree.
3150: \end{lemma}
3151: 
3152: \proof Same as for lemma \reference{lemh}. \qed
3153: 
3154: \begin{lemma}
3155: If a mountain $M$ is not of maximal height
3156: (i.e. $n-1$), then $MWF(\bt)\ge 4$.
3157: \end{lemma}
3158: 
3159: \proof Let $k<n-1$ be the height of $M$.
3160: We assumed there are at least two mountains of maximal height. So
3161: now w.l.o.g. assume some, say
3162: the left, of the neighbored mountains of $M$ has height $k'>k$.
3163: Fill the two valleys around $M$ starting with $\sg_{k-1}$.
3164: Then the maximal subword has a syllable index sequence
3165: $k+1,k,k-1,k,k-1,k$. Make the second and third syllable trivial
3166: (if not already), and apply YB relations, moving the fourth
3167: syllable to the left: $k+1,\ul{k-1},k,k-1(,k-1),k$. The result is a
3168: summit reduced word, in which a new syllable of index $k-1$
3169: (the underlined one) was removed and it became non-maximal. Hence the
3170: non-maximal subword has exponent sum $[\bt']_{k-1}>1$, and so
3171: $MWF\ge 4$. \qed
3172: 
3173: \begin{lemma}\label{lem2val}
3174: If $\bt$ has $>2$ valleys of depth at most $n-3$, then $MWF(\bt)\ge 4$.
3175: \end{lemma}
3176: 
3177: \proof % It suffices to check it for 4-braids.
3178: It suffices to check for 3 mountains (as one can fill
3179: separate valleys) and 4-braids (as one can fill valleys by
3180: levels as in the proof of lemma \reference{lemh} and the remark
3181: after theorem \reference{th4.1}). This is just
3182: the word [1232112321121321], which is easily checked (to
3183: have $MWF=4$). \qed
3184: 
3185: \subsubsection{Two mountains\label{S4.1}}
3186: 
3187: We assume in \S\reference{S4.1} and \S\reference{S4.2} that $n\ge 5$.
3188: The case of $4$-braids is considered later in \S\reference{S4.3}.
3189: We refer to \S\reference{brs} for the use of notation we will employ.
3190: 
3191: We assume first $\bt$ has two mountains. By the previous remarks they
3192: are both of height $n-1$.
3193: 
3194: So now consider words with syllable index sequence
3195: \begin{eqnarray}\nonumber
3196: & 1,2,\dots,n-2,n-1,\quad
3197: p_1@,\dots,p_k@,\quad
3198: n-2,n-3,\dots,g+1,g!,\quad \\
3199: \label{xx} &
3200: g+1,\dots,n-2,n-1,\quad
3201: q_1@,\dots,q_l@,\quad
3202: n-2,n-3,\dots,2,1
3203: \end{eqnarray}
3204: such that $l+k=n-3$ and
3205: $\{p_1,\dots,p_k,q_1,\dots,q_l\}=\{1,\dots,n-3\}$.
3206: We will distinguish only between non-trivial and trivial syllables
3207: (in former case exponent is immaterial). For non-trivial syllables
3208: we write an exclamation mark
3209: after the index, for trivial ones an `at' (@) sign.
3210: If none of $!$ and @ is specified, we do not exclude explicitly any of
3211: either types. We write $\bt_1,\dots,\bt_6$ for the subwords separated by
3212: space in \eqref{xx}.
3213: 
3214: Assume w.l.o.g. (up to reversing the braid's orientation) that
3215: some $p_i$ is $1$, and let up to commutativity the $p_1,\dots,p_k$
3216: subword be written as $h,h-1,\dots,1,p_1',\dots,p_{k-h}'$ (the 
3217: $p_i'$ contain the indices above $h$ occurring as $p_i$).
3218: 
3219: Assume the maximal syllable of $\sg_{h+1}$ in $\bt_1$ is non-trivial.
3220: Then one can write $\bt$ as word with a subword of index sequence 
3221: \begin{eqn}\label{siks}
3222: 1,\dots,h+1!,h@,\dots,1@\,. 
3223: \end{eqn}
3224: If now $h<n-3$, we can make in \eqref{siks} all syllables
3225: trivial except $h+1$, which we make of exponent $2$, then
3226: split a loop (component of a link in a node of a positive
3227: resolution tree) by removing all the syllables in \eqref{siks}
3228: together with the terminating `$1$' in \eqref{xx}. We obtain
3229: \begin{eqnarray*}
3230: & h+2,\dots,n-2,n-1,\quad
3231: p_1'@,\dots,p_{k-h}'@,\quad
3232: n-2,n-3,\dots,g+1,g!,\quad \\
3233: & g+1,\dots,n-2,n-1,\quad
3234: q_1@,\dots,q_l@,\quad
3235: n-2,n-3,\dots,2.
3236: \end{eqnarray*}
3237: (Here no syllables of index 1 occur, and the split loop is the
3238: isolated leftmost strand.) Then we fill the valley starting with the
3239: index-$n-1$-summits, splitting another loop (the rightmost strand),
3240: \[
3241: h+2,\dots,n-3,n-2,\quad p_1'@,\dots,p_{k-h}'@,\quad q_1@,\dots,q_l@,
3242: \quad n-2,n-3,\dots,2,
3243: \]
3244: and are left with a
3245: word that has at least two $\sg_{n-2}$. So $MWF\ge 4$. 
3246: 
3247: \begin{caselist}
3248: \case Assume the smaller valley has depth $g>1$.
3249: The $\bt$ has two index-1 syllables, a trivial (non-summit)
3250: and a non-trivial (summit) one. By a flype one can exchange
3251: them, and so have a non-summit reduced word. Then one
3252: can change $\bt$ to a word of smaller index sum, and so
3253: we are done by induction.
3254: 
3255: % So we have $h=n-3$ (in particular $\bt_5$ is trivial).
3256: % Now if $g>2$, then the $p_l=1$ occurring at the
3257: % end in $\bt_2$, can be moved to the end of $\bt$ right
3258: % before the $\dots,2,1$ of $\bt_6$,
3259: % and by reversal of the word we have a situation with $h=1$.
3260: % We can exclude $g>2$ by assuming the present $h$ is the minimal
3261: % possible.
3262: % 
3263: % Thus $g=2$. Then we have a braid of the form
3264: % \begin{eqn}\label{seven}
3265: % 1*,2,\dots,n-2*,\quad
3266: % n-3@,\dots,1@,\quad
3267: % n-1,\dots,3,2*,\quad
3268: % 3,\dots,n-1,\quad
3269: % n-2,\dots,2\,.
3270: % \end{eqn}
3271: % According to the spacing group the syllables into subwords
3272: % $\bt'_{1}$ to $\bt'_5$.
3273: % 
3274: % By making all syllables trivial except the index-$n-2$ syllable in
3275: % $\bt'_1$ and the index-2 syllable in $\bt'_3$, we have a 3-component
3276: % link. Two components have no self-crossings (the first and last strand).
3277: % The third component has self-crossings, namely the syllables in
3278: % $\bt'_5$. If in $\bt$ some of the syllables in $\bt'_5$ were
3279: % non-trivial, we could now delete it, and have a 4-component link,
3280: % so that $MWF(\bt)\ge 4$.
3281: % So in \eqref{seven} the syllables in $\bt'_5$ are trivial.
3282: % 
3283: % % Now, by drawing a closed braid picture, we see that when the
3284: % % $n-2$-index syllable in $\bt'_5$ is trivial, and so is the
3285: % % $n-1$-index syllable in $\bt_3$, then the braid is Markov reducible.
3286: % % Thus the $n-1$-index syllable in $\bt_3$ is non-trivial.
3287: % % 
3288: % % Now remove the $n-1$-index syllable and also the other in $\bt_5$.
3289: % % Call the bew braid $\tl\bt$. It has 4 non-trivial syllables:
3290: % % in cycl order of inex $1$, $n-2$, $2$, $n-2$. 
3291: % % 
3292: % % When we make trivial the last two non-trivial syllables, the closure 
3293: % % becomes again a 3-component link, whose one component's self crossings
3294: % % are in $\bt_3$, except the $2*$. By the previous argument all syllables
3295: % % in $\bt_3$ must be trivial, except the $2*$.
3296: % 
3297: % Now, we argue that all syllables in $\bt'_3$ must be trivial, 
3298: % except the $n-1$ and $2*$. Namely, assume otherwise the $b$-index
3299: % syllable is non-trivial for $2<b<n-1$. Then remove all index $\le
3300: % b-2$ syllables in $\bt$. First let $b<n-2$. Let $\bt_i''$
3301: % be the result of $\bt_i'$ (for $1\le i\le 5$) under this
3302: % deletion. The $\sg_{b-1}$ in $\bt_1''$ now forms with $\bt_2''$
3303: % the letter-order-reversed subword \eqref{siks} (with $h=1$).
3304: % By the same loop split argument following \eqref{siks} we have
3305: % $MWF\ge 4$, since a $\sg_{n-2}$ remains in $\bt_2''$ and
3306: % $\bt_4''$. (Here $1=h<n-3$, or $n\ge 5$, is needed.)
3307: % If $b=n-2$, then we obtain after deleting index $\le b-2$
3308: % syllables a 5-braid. Making all syllables we can trivial,
3309: % we have the word 112332143323432. By direct check it has $MWF=
3310: % 4$, and so $MWF(\bt)\ge 4$. 
3311: % 
3312: % Similarly we have, as assumed, that all
3313: % syllables in $\bt'_2$ are trivial. So we have
3314: % \[
3315: % 1*,2,\dots,n-3,n-2*,\quad n-3@,\dots,1@,\quad
3316: % n-1,n-2@,\dots,3@,2*,\quad 3,\dots,n-1,\quad n-2@,\dots,2@
3317: % \]
3318: % 
3319: % By testing the 5-braids, we see that $MWF\ge 4$ if both
3320: % index $n-1$-syllables are non-trivial. If one is trivial,
3321: % one has to verify now that such braids reduce to positive
3322: % braids on a smaller number of strands (and then work by
3323: % induction). We checked this up to $n=9$, and it suggests
3324: % a reduction is possible for higher $n$.
3325: 
3326: \case Now $g=1$. We write $\bt_1,\dots, \bt_6$ for the
3327: 6 subwords separated by spacing in \eqref{xx}.
3328: By a similar argument as after \eqref{siks} we can argue that
3329: if one can reorder the syllables in $\bt_{2,5}$ so that
3330: $\bt$ has a subword with an index sequence 
3331: \begin{eqn}\label{(eight)}
3332: k,k+1,\dots,h-1,h!,h-1,\dots,k\,,
3333: \end{eqn}
3334: with the first or last
3335: $h-k$ syllables belonging to $\bt_{2,5}$ and the others
3336: to $\bt_{1,3,4,6}$, and $h<n-2$, then $MWF\ge 4$.
3337: 
3338: W.l.o.g. assume $1\in \bt_5$ (which is meant to abbreviate that
3339: $\bt_5$ contains an index-$1$ syllable).
3340: 
3341: \begin{caselist}
3342: \case
3343: Now if $2\not\in\bt_5$ (so $2\in\bt_2$) then the
3344: $2$-index syllables in $\bt_{4,6}$ are trivial
3345: (because we have otherwise \eqref{(eight)} with $k=1$, $h=2$).
3346: 
3347: We distinguish several cases by the subwords of
3348: (non-summit) syllables of index $1,2$ and $3$ in $\bt_{2,5}$.
3349: We separate the subwords between $\bt_{2}$ and $\bt_5$
3350: by a vertical line `$|$'. (We assume here that $n\ge 6$.
3351: The case $n=5$ must be handled by a separate, but
3352: simplified, argument.) Note that by symmetries we can
3353: exchange the words of $1,2$ and $3$ left and right from `$|$'
3354: and also (simultaneously) reverse both, and can also use the
3355: commutativity of $1$ and $3$. Then we are left with
3356: the following cases.
3357: 
3358: \begin{caselist}
3359: \case
3360: $2|31$. If $3\in \bt_5$, then both $3$ in $\bt_{1,3}$ are trivial.
3361: (Otherwise, we would have \eqref{(eight)} for $k=2$ and $h=3$.)
3362: % \begin{verbatim}
3363: Below we give pairs of words, the first obtained by extending all
3364: admissible syllables to be non-trivial, and the second one by
3365: extending the first word to one with knot closure, which was then
3366: checked to reduce (by at least one strand, not necessarily to a
3367: 3 braid).
3368: \\
3369: $[1122344  2 55665544322111   2334455 3441 66554433211]$, \\
3370: $[11222344 2 5556655444322111 2334455 3441 666554433211]$ reduces. \\
3371: % \end{verbatim}
3372: \case
3373: If $3\in \bt_2$, then one of both $3$-index syllables in
3374: $\bt_{1,3}$ must be trivial. (It is the syllable in $\bt_3$ if `$3$'
3375: occurs before `$2$' in $\bt_2$, or the syllable in $\bt_1$
3376: otherwise.)
3377: 
3378: \begin{caselist}
3379: \case
3380: $ 23|1$.
3381: % \begin{verbatim}
3382: \\
3383: $[1122344 23 556655443322111    2334455 441 66554433211]$, \\
3384: $[1122344 23 556665554443322111 2334455 441 66554433211]$ reduces. \\
3385: % \end{verbatim}
3386: \case $ 32|1$
3387: % \begin{verbatim}
3388: \\
3389: $[11223344  32 55665544322111   2334455 441 66554433211]$, \\
3390: $[112233444 32 5556665544322111 2334455 441 66554433211]$ reduces.  \\
3391: % \end{verbatim}
3392: \end{caselist}
3393: \end{caselist}
3394: 
3395: \case
3396: Now assume $2\in\bt_5$. Since $1\in\bt_5$, now the exclusion
3397: of \eqref{(eight)} shows that only one of the $2$-index syllables
3398: in $\bt_{4,6}$ is trivial, but we will show that also one in
3399: $\bt_{1,3}$ is. Namely, by making the proper index-$1$ (summit)
3400: syllable to exponent 2, and one of the $2$-index syllables of $\bt_
3401: {1,3}$ trivial, one can slide by braid relations the $1$-index
3402: syllable $\sg_1=X$ from $\bt_5$ to $\bt_2$.
3403: By applying the previous argument, both
3404: $2$-index syllables in (the now modified) $\bt_{1,3}$ are trivial.
3405: One of them was previously made trivial to slide $X$ in, but the
3406: condition on the other one persists for the original braid.
3407: 
3408: 
3409: In all situations,
3410: make all the other syllables in $\bt_{1,3,4,6}$ non-trivial and
3411: check using KnotScape that the braid reduces. 
3412: 
3413: % MAKE THIS EXACT!!!
3414: 
3415: \begin{caselist}
3416: \case $ 3| 12$. In this case the above argument shows that the
3417: syllable $2\in\bt_1$ is trivial.
3418: % \begin{verbatim}
3419: \\
3420: $[11233445566  344 55443322111 234455   12 66554432211]$, \\
3421: $[112334455666 344 55443322111 23444555 12 665544322111]$ reduces. \\
3422: % \end{verbatim}
3423: \case $ 3| 21$. Here $2\in\bt_3$ is trivial.
3424: % \begin{verbatim}
3425: \\
3426: $[112233445566 344 5544332111 2234455 21 6655443211]$, \\
3427: $[112233445566 344 5544332111 2234455 21 66655544432111]$ reduces. \\
3428: % \end{verbatim}
3429: \case $| 123$. As before $2\in\bt_1$ is trivial.
3430: % \begin{verbatim}
3431: \\
3432: $[11233445566 55443322111 234455 123 665544332211]$, \\
3433: $[11233445566 55443322111 234455 123 666555444332211]$ reduces. \\
3434: % \end{verbatim}
3435: \case $| 132$.
3436: % \begin{verbatim}
3437: \\
3438: $[11233445566 55443322111 2334455 132 66554432211]$, \\
3439: $[11233445566 55443322111 2334455 132 66655544432211]$ reduces. \\
3440: % \end{verbatim}
3441: \end{caselist}
3442: \end{caselist}
3443: 
3444: \end{caselist}
3445: 
3446: \subsubsection{More than two mountains\label{S4.2}}
3447: 
3448: % VERY VAGUE AND MUCH NEEDS TO BE DONE USING COMPUTATIONS !!!
3449: To deal with the general case, 
3450: now we make the following modifications. We call a
3451: summit syllable sequence with indices $n-2$ and $n-1$
3452: terminated on both sides by $n-1$'s a modified
3453: mountain or \em{plateau}. We have again by lemma
3454: \reference{lem2val} only two valleys of depth $<n-2$,
3455: or alternatively only two plateaus (now instead of mountains).
3456: The case of more than two mountains is thus mainly a
3457: adaptation of the case of two mountains, replacing
3458: mountains by plateaus.
3459: 
3460: % Also we allow $\bt$ and the maximal summit subword
3461: % to throw in occurrance of non-trivial $\sg_{n-1}$ syllables.
3462: 
3463: % The words that admit reductions apply as in the previous
3464: % two-mountains-case,
3465: % 
3466: % 2 deep, 1 shallow valley
3467: % 123211232213211 reduces
3468: % 123121123223211 reduces 
3469: % 
3470: % CHECK CAREFULLY !!!
3471: % 
3472: % , but for proving $MWF\ge 4$ the
3473: % following modifications of the argument are needed.
3474: 
3475: Again we may assume non-maximal subwords have exactly one (and trivial)
3476: syllable per index. The elimination of the maximal subwords can be done
3477: similarly. % only that when arriving at $\dots,n,n-1@,n,n-1@,n$
3478: 
3479: We distinguish two cases as in the above study of the 2-mountain
3480: words, depending on the depth $g$ of the second valley (the
3481: other valley has depth $1$ by the same argument as above).
3482: 
3483: \begin{caselist}
3484: \case $g>1$. We use the previous flyping argument.
3485: % In the first case above the same arguments show $g=2$ and (after
3486: % deleting the index-$n-2$ syllables in the plateaus) that
3487: % the syllables in $\bt_{3,5}$ in the form \eqref{seven} are trivial,
3488: % except the index-$n-2$ syllables in both and the index-$2$ and
3489: % $n-1$ syllable in $\bt_{3}$.
3490: % (Otherwise, fill one valley, remove a loop from \eqref{hh}, and
3491: % something non-trivial remains.) Also $n-2\in\bt_1$ may now be trivial.
3492: % 
3493: % {$112233\hat 4554455\ \ 321\ 5544\ul 32\ 2233445544\ 5544
3494: % \ul 3\ul 2$}
3495: % 
3496: % % CHECK ANYTH REDUCES !!!
3497: % One has to check that such braids reduce. For example, \\
3498: % % \begin{verbatim}
3499: % 11223345544455 321 55544322 22333445544 5544432 reduces, and \\
3500: % 1122334445544455 321 55544322 22333445544 5544432 reduces. 
3501: % \end{verbatim}
3502: 
3503: \case $g=1$.
3504: In the second case we had restrictions on exponents of syllables
3505: with index $2$ and $3$ occurring in the maximal subwords from
3506: the position of syllables with index $1$ and $2$ occurring in
3507: the non-maximal subwords. The restrictions on $2$-(index) syllables
3508: from $1$-(index) syllables remain. So do the restrictions on
3509: $3$-syllables from $2$-syllables unless we have $\le 5$
3510: strands. The argument is the same: one can still pull
3511: out two loops and has at least two letters of $\sg_{n-2}$.
3512: % (additionally discarding all additional non-trivial $\sg_{n-1}$
3513: % syllables)
3514: 
3515: Now we check restrictions on $3$-index syllables for $5$ strands
3516: and reducibility. We have up to extensions a finite number of
3517: special braids to verify. Clearly, extensions are never
3518: admissible for non-summit syllables, and always admissible for
3519: summit index-1-syllables (since they are all non-trivial). 
3520: Also extensions of subwords $(n-2\ n-2\ n-1)^k$ occurring
3521: repeatedly are redundant, since doubling the letter $n-1$ is the
3522: same as deleting the $n-2\ n-2$ for the next $k$. We
3523: will verify that the property $MWF\ge 4$ resp. reducibility
3524: does not depend on the value of $k$ as soon as $k>0$.
3525: We will thus deal only with the other extensions.
3526: 
3527: \begin{caselist}
3528: \case The non-summit $2$-index syllable is between two valleys
3529: of depth $n-2=3$ - this is handled as before: \\
3530: $[1234334123343211234321]$ has $MWF=4$, \\
3531: $[1234334213343211234321]$ has $MWF=4$, \\
3532: $[1234334233432112314321]$ has $MWF=4$.
3533: 
3534: \case The non-summit $2$-index syllable is between
3535: one depth $n-2=3$ valley and one depth $1$ valley. These are words
3536: of the form $[123412(334)^k3211234(334)^l321]$, \\
3537: $[123421(334)^k3211234(334)^l321]$,\\
3538: $[12342(334)^k32112341(334)^l321]$ \\
3539: for $k>0$, $l\ge 0$, and their extensions.
3540: 
3541: \begin{caselist}
3542: \case $[123412(334)^k3211234(334)^l321]$ and extensions.
3543: 
3544: \begin{caselist}
3545: \case $l=0$. To display the extendability of syllables,
3546: in the following notation the necessarily trivial syllables are
3547: hatted, while a possibly trivial syllable is underlined.
3548: 
3549: $[1\hat 2\hat 3\ul 412(334)^k\ul 3\ul 211\hat 2\hat 3\hat 4\ul 3\ul 21]$. 
3550: 
3551: When $l=0$, then making non-trivial any single of the hatted syllables
3552: makes $MWF=4$ already for $k=1$, while for any $k>0$ making any
3553: combination (possibly all) of the underlined syllables non-trivial
3554: gives $MWF=3$.
3555: 
3556: Now again check that braids reduce, for example: \\ 
3557: $[1234123343343211234321]$ has $MWF=3$, \\
3558: $[12341233433433432111234321]$ reduces to 3 strands, \\
3559: $[1234441233433433221123433221]$ reduces to 3 strands.
3560: 
3561: 
3562: \case 
3563: When $l=1$, the already for $k=1$ we have without extensions
3564: $[1234123343211234334321]$ and $MWF=4$.
3565: \end{caselist}
3566: 
3567: \case $[123421(334)^k3211234(334)^l321]$ and extensions.
3568: 
3569: \begin{caselist}
3570: \case When $l=0$, we have for all $k$ the extendability
3571: 
3572: $[1\ul 2\hat 3\ul 421(334)^k\ul 3\hat 211\ul 2\hat 3\hat 4\ul 3\hat 21]$,
3573: 
3574: with the same explanation as before.
3575: 
3576: For example, $[122344213343343343321122343321]$ reduces to 3 strands.
3577: 
3578: \case When $l\ge 1$, then already for $k=1$ and no extensions
3579: $MWF([1234213343211234334321])=4$.
3580: 
3581: \end{caselist}
3582: 
3583: \case $[12342(334)^k32112341(334)^l321]$ and extensions.
3584: 
3585: \begin{caselist}
3586: \case When $l=0$, we have
3587: 
3588: $[1\ul 2\hat 3\ul 42(334)^k\ul 3\ul 211\hat 2\hat 3\hat 41\ul 3\hat 21]$.
3589: 
3590: For example, $[122344233433433433221123413321]$ reduces to 3 strands.
3591: 
3592: \case When $l>0$, already for $k=1$, $l=1$ without extension
3593: we have $[1234233432112341334321]$ and $MWF=4$.
3594: \end{caselist}
3595: \end{caselist}
3596: 
3597: \case The non-summit $2$-index syllable is between
3598: two depth $1$ valleys. These are (up to symmetry) words of the form\\
3599: $[1234(334)^k321123421321]$ and \\
3600: $[12341(334)^k32112342321]$, \\
3601: for $k>0$, and their extensions (here necessarily $l=0$).
3602: 
3603: \begin{caselist}
3604: \case $[1234(334)^k32112342132]$ and extensions:
3605: 
3606: $[1\ul 2\ul 3\ul 4(334)^k\ul 3\hat 211\ul 2\hat 3\hat 421\hat 3\hat 21]$.
3607: 
3608: Again for $k=1$ making non-trivial all underlined syllables 
3609: gives $MWF=3$, while making non-trivial any of the hatted
3610: syllables gives $MWF=4$. 
3611: 
3612: For example,
3613: $[12 3  4 3343       2112 3 214321]$ has $MWF=3$, and its extensions\\
3614: $[122333443343       211223 214321]$ and \\
3615: $[12233 4433433433433211223421 321]$ \\
3616: were checked to reduce to 3 strands.
3617: 
3618: The same is the outcome for $k>1$.
3619: 
3620: \case $[12341(334)^k32112342321]$ and extensions:
3621: 
3622: $[1\h 2\u{34}1(334)^k\u 3\h 211\u 2\h 3\h 42\h 3\u 21]$
3623: 
3624: The reducibility cases follow analogously. % CHECK !!
3625: For example, $[123344133433433433211223423221]$ reduces to 3 strands.
3626: 
3627: % For $l>0$ we have always $MWF=4$.
3628: 
3629: \end{caselist}
3630: 
3631: \end{caselist}
3632: \end{caselist}
3633: 
3634: \subsubsection{4-braids\label{S4.3}}
3635: 
3636: If some mountain is not of height 3, or $>2$ valleys of
3637: depth $1$ exist, then we are done as before (see \S\reference{4.0}).
3638: 
3639: So the maximal subword is of the form
3640: $1,2,3,(2!,3)^p,2,1,1,2,3,(2!,3)^n,2,1$, and the non-maximal
3641: subword is a single $\sg_1^1$. We separate the summit syllables
3642: and their letters by the summit syllables of index $1$ into
3643: a \em{left} and \em{right plateau}. Assume w.l.o.g. the (non-summit)
3644: $\sg_1^1$ is in the left plateau. The word `in' is to mean that
3645: in cyclic order of the syllables of $\bt$ the syllable $\sg_1^1$
3646: can be written to occur just before or after a syllable with
3647: index $n-1$ that belongs to the left plateau. This means that
3648: we can write $\bt$ as
3649: \begin{eqn}\label{blmn}
3650: \bt_{l,m,n}=[1\vtbox{2\\.}3(223)^n1(223)^m\vtbox{2\\..} 1123(223)^l21]
3651: \end{eqn}
3652: with $n,m,l\ge 0$, or some of its extensions. Note that MWF will
3653: be monotonous in $m,n,l$, i.e. $MWF(\bt_{l,m,n+1})\ge
3654: MWF(\bt_{l,m,n})$ etc. Using symmetry assume $n\ge m$.
3655: 
3656: One can check already at this stage that such words are not
3657: $B$-adequate. So we obtain corollary \reference{cort}, and
3658: for the proof of theorem \reference{thg} the rest of the
3659: argument here can be replaced by the application of theorem
3660: \reference{thq}. Note that $B$-adequacy is invariant
3661: under isotopy preserving writhe and crossing number, so
3662: any other positive 4-braid word giving the same link is
3663: not $B$-adequate either.
3664: 
3665: If $n+l>0$, then the $2$-index syllable $\vtbox{2\\..}$
3666: in \eqref{blmn} must be trivial.
3667: % there are $2*$ of the left plateau on the left (right) of
3668: % $\sg_1^1$ or some in the right plateau, then the right
3669: % (left) $2$-index syllable of that mountain must be trivial.
3670: Otherwise remove all $2!$ in $(2!,3)^m$ (if any), and split two loops
3671: as explained after \eqref{siks}. The `$2!$' in $(2!,3)^l$ or
3672: $(2!,3)^n$ remain, and so $MWF=4$.
3673: With a similar argument we see that if $m+l>0$, then the $2$-index
3674: syllable $\vtbox{2\\.}$ is trivial.
3675: 
3676: We distinguish 3 cases depending on whether these arguments
3677: apply or not.
3678: 
3679: \begin{caselist}
3680: \case Both non-triviality arguments apply. So we have a family of
3681: words $\bt_{l,m,n}=[123(223)^n1(223)^m21123(223)^l21]$ with $n,m>0$,
3682: or $l>0$ and their extensions, and both $\vtbox{2\\..}$ and
3683: $\vtbox{2\\.}$ are trivial.
3684: 
3685: \begin{caselist}
3686: \case $l=0$. We assumed $n,m>0$, and already for $n=m=1$,
3687: the word $[12322312232112321]$, we have $MWF=4$.
3688: 
3689: \case $l>0$.
3690: 
3691: \begin{caselist}
3692: \case $m=n=0$. These are extensions of $[1231211123(223)^l21]$,
3693: and the admissibility is found to be:
3694: 
3695: $[1\h 2\h 31\h 211\ul{23}(223)^l\ul 21]$.
3696: 
3697: The two letters `$2$' in the left plateau cannot be doubled
3698: ($[123122112322321]$, $[122312112322321]$), neither the `$3$' \\
3699: ($[123132112322321]$), since $MWF=4$ already for $l=1$.
3700: 
3701: Without extension, $l=1$ ($[123121112322321]$) and $l=3$
3702: ($[123121112322322322321]$) reduce.
3703: % 
3704: % $[1231211223223223221]$ reduces.
3705: % 
3706: % % So so do (sloppy) $123121123[23]+21$ in regexp notation.
3707: % 
3708: % $[123121112223223223223221]$ has $MWF=3$. \\
3709: % $[1231211122232232232232221]$ reduces. \\
3710: So we find that $[12312112*3[23]+2*1]$ reduce.
3711: (Recall that, while `$2*$' in
3712: a braid word should mean at least one letter `$2$', the term `$[23]+$'
3713: should mean a possibly empty sequence of letters `$2$' and `$3$'.
3714: We distinguish braid words from index sequences by not putting
3715: commas between the numbers.)
3716: 
3717: \case $m+n>0$; this reduces to the case of $m=n=0$ with some of the
3718: twos or the three in the left plateau doubled, where we found $MWF=4$.
3719: 
3720: \end{caselist}
3721: 
3722: \end{caselist}
3723: 
3724: 
3725: \case In the case one of the non-triviality conditions on
3726: $\vtbox{2\\..}$ and $\vtbox{2\\.}$ does not apply,
3727: we have $123(223)^n12112321$ with $n>0$ and its extensions.
3728: (Now the right plateau is a mountain.)
3729: 
3730: % \begin{caselist}
3731: % \case
3732: The case $n=3$ ($[123223223223121112321]$) simplifies. \\
3733: % So so do (sloppy) $123[23]+312112321$ in regexp notation.
3734: % Also 12223223223223121112321, so so do (sloppy) $12*[23]+312112321$.
3735: $[122332232232232231211222321]$ simplifies.
3736: 
3737: 
3738: However, $MWF([123223122112321])=4$, so the right $2$-index syllable
3739: in the left plateau ($\vtbox{2\\..}$ in \eqref{blmn}) must be
3740: trivial.
3741: 
3742: The right 2 and the 3 in the right plateau must be trivial: \\
3743: $[1232231211123221]$, \\ $[1232231211123321]$ have $MWF=4$.
3744: 
3745: But the left `$2$' of the left and right plateau may not be trivial:
3746: $[12223223223223121112222321]$ has $MWF=3$. \\
3747: $[122232232232231211122222321]$ simplifies. \\
3748: So $[12*[23]+312112*321]$ simplifies. % ?
3749: 
3750: We arrive at the form $[1\u{23}(223)^n1\h 211\u 2\h 3\h 21]$.
3751: 
3752: % \dots
3753: % \end{caselist}
3754: 
3755: \case
3756: In case both non-triviality conditions do not apply, we have
3757: $[12312112321]$ and its extensions. (So both plateaus are
3758: mountains.)
3759: 
3760: Since non-triviality is nonetheless possible, we may have non-trivial
3761: $2$-index syllables in the left plateau.
3762: We distinguish three cases again according to whether
3763: $\vtbox{2\\..}$ and $\vtbox{2\\.}$ are trivial or not.
3764: 
3765: 
3766: \begin{caselist}
3767: \case Both $2$-index syllables are trivial: $[12312112321]$.
3768: 
3769: % 12312112321
3770: 
3771: By direct check: \\
3772: % 122312112321 $MWF=3$
3773: $[123312112321]$ has $MWF=3$, \\
3774: % 123122112321 $MWF=3$
3775: $[123121122321]$ has $MWF=3$, \\
3776: $[123121123321]$ has $MWF=3$, \\
3777: $[123121123221]$ has $MWF=3$. \\
3778: 
3779: % 12233122112321 $MWF=3$
3780: % 122331221122321 $MWF=4$
3781: % 122331221123321 $MWF=4$
3782: % 122331221123221 $MWF=4$
3783: 
3784: $[12312112233221]$ has $MWF=3$, \\
3785: % 122312112233221 $MWF=4$
3786: $[123312112233221]$ has $MWF=4$. \\
3787: % 123122112233221 $MWF=4$
3788: 
3789: So one can extend the right mountain's `$2$'s and one of the
3790: left or right mountain's `$3$'s, but not both `$3$'s.
3791: 
3792: $[123121122233221]$ reduces, \\
3793: $[123312112223221]$ reduces, \\
3794: $[1233121122233221]$ has $MWF=4$.
3795: 
3796: So a reducing check is to be made on $[11*231211*22*3*22*]$.
3797: % and 11*22*3*122*11*232
3798: 
3799: $[123121112233221]$ reduces, \\
3800: $[123121112222333322221]$ reduces, \\
3801: $[1233312111222322221]$ reduces.
3802: 
3803: % \begin{caselist}
3804: % \case
3805: % 11*231211*22*3*22*
3806: % 
3807: % \case 11*22*3*122*11*232
3808: % 
3809: % 122331222112321 DOES NOT REDUCE !!!
3810: % 122331222112321 connected-2-cabled MWF =7 !!!!!
3811: % 
3812: % So we have exceptions 11*22*3*122*11*232 !
3813: % (Or 22*3*122*11*23211*, which is the
3814: % first family in theorem \reference{thmwf}.)
3815: % 
3816: % 
3817: % 12331222112321 2-cabled MWF =6 !!!!!
3818: % 112331222112321 reduces
3819: % 11233331222112321 reduces
3820: % 
3821: % 12231222112321 2-cabled MWF =7 !!!!!
3822: % 1223122112321 2-cabled MWF =7 !!!!!
3823: % 
3824: % So we have exceptions 11*22*3122*11*232 !
3825: % (Or 22*3122*11*23211*, which is a special case of the
3826: % first and second family in theorem \reference{thmwf}.)
3827: % 
3828: % 1223312112321 2-cabled MWF =6 !!!!!
3829: % 12222333312112321 reduces
3830: 
3831: \case
3832: One $2$-index syllable is non-trivial. This is the word
3833: $[123122112321]$, with $MWF=3$ (the case
3834: $[122312112321]$ is symmetric).
3835: 
3836: We have the following extensions:
3837: 
3838: $[1233122112321]$ has $MWF=3$ (and reduces), \\
3839: $[1231221122321]$ has $MWF=4$, \\
3840: $[1231221123221]$ has $MWF=3$ (and reduces), \\
3841: $[1231221123321]$ has $MWF=3$ (and reduces).
3842: 
3843: The following combined extensions are to check:
3844: $[12331221123321]$ has $MWF=4$, \\
3845: $[12331221122321]$ has $MWF=4$, \\
3846: $[12312211223321]$ has $MWF=4$. \\
3847: 
3848: Thus we are left to deal with
3849: $[1233*122*112321]$,\\
3850: $[123122*112322*1]$,\\
3851: $[123122*11233*21]$,\\
3852: and check that they all reduce:
3853: 
3854: $[12333312222112321]$ reduces,\\
3855: $[12312222112322221]$ reduces,\\
3856: $[12312222112333321]$ reduces.
3857: 
3858: \case Both $2$-index syllables are non-trivial:
3859: $[1223122112321]$. This is a braid word $\bt_0$ for $13_{9465}$.
3860: We have the following extensions:
3861: 
3862: $[12233122112321]$ has $MWF=3$, \\
3863: $[12231221122321]$ has $MWF=4$, \\
3864: $[12231221123221]$ has $MWF=4$, \\
3865: $[12231221123321]$ has $MWF=3$.
3866: 
3867: The only common extension of the two $MWF=3$ extensions is \\
3868: $[122331221123321]$, which has $MWF=4$.
3869: 
3870: So it remains to verify that (the closures of)
3871: $[22*3*122*11*23211*]$ and $[22*3122*11*23*211*]$ have braid index 4.
3872: 
3873: % 1222233312211232111 no reduce
3874: % 12223122211123321 no reduce
3875: 
3876: For this we use the two-cabled MWF inequality. Let for a braid
3877: $\bt\in B_n$, the ``two-cabled'' braid $(\bt)_2\in B_{2n}$ be
3878: obtained from $\bt$ by replacing in \eqref{wd} each $\sg_i$
3879: by $\sg_{2i}\sg_{2i-1}\sg_{2i+1}\sg_{2i}$. Then $(\bt)_2$
3880: is a braid representation of the (blackboard framed) two-cable
3881: link $L_2$ of the closure $L=\hat\bt$ of $\bt$. We consider now
3882: for $\bt$ the above braid $\bt_0$.
3883: 
3884: We know, from the computations described in \cite{MorSho,WilFr},
3885: that $MWF((13_{9465})_2)=7$. This is in fact true also for the
3886: connected cable (the one with braid representation $(\bt_0)_2
3887: \cdot\sg_1$). Now we claim that reducing or resolving a clasp
3888: (changing a $\sg_i^2$ into a $\sg_i$ or deleting it) does not
3889: reduce the two-cabled MWF bound. The 2-cable of a clasp can be
3890: resolved by resolving 4 clasps. The 2-cable of a crossing in a
3891: 2-cabled clasp can be resolved by resolving one clasp and changing
3892: twice $\sg_i^2\to \sg_i$. Finally, the 2-cable of an isolated 
3893: $\sg_i$ can be reduced into two internal twists of the
3894: doubled original strand. Such twists can be collected for
3895: every doubled component, resolved for each doubled component to
3896: one, and joined if doubled components are joined by reducing
3897: a doubled crossing in a doubled clasp. So the two-cabled MWF
3898: reduces to the one of the connected cable of $13_{9465}$ and
3899: we are done.
3900: \end{caselist}
3901: \end{caselist}
3902: 
3903: The proof of theorem \reference{thmwf} is now completed.
3904: 
3905: \subsection{Positive links\label{Spos}}
3906: 
3907: In this section, we will refine the arguments proving
3908: theorem \reference{thqp} to restrict the possible 
3909: 3-braid representations of positive links. 
3910: % and then, % in an alternative way, classify
3911: % positive braid links of braid index 3.
3912: % 
3913: Our positivity considerations will make use of the criterion
3914: of Yokota \cite{Yokota}, and the Kauffman polynomial $F$.
3915: We recall the properties \eqref{wseven}~--~\eqref{wseven.7}\,
3916: that determine $F$ and its writhe-unnormalized version $\Lm$.
3917: 
3918: Note that for $P$ one can similarly define a regular isotopy
3919: invariant
3920: \begin{eqn}\label{dPtl}
3921: \tl P(D)(a,z)=(ia)^{-w(D)}P(D)(ia,iz)\,,
3922: \end{eqn}
3923: with $i=\sqrt{-1}$. Then $\tl P$ satisfies similar relations to
3924: \eqref{wseven}~-- \eqref{wseven.7}. The difference to $\Lm$ is that
3925: $\tl P$ is defined on oriented link diagrams, and that the term making
3926: orientation incompatible on the right of \eqref{wseven} is missing.
3927: 
3928: \begin{theorem}\label{thF}(Yokota \cite{Yokota})
3929: If $L$ is a positive link, then 
3930: \[
3931: \md_a F(L)\,=\,\md_v P(L)\,=\,1-\chi(L)\,,
3932: \]
3933: and
3934: \[
3935: [F(L)]_{a^{1-\chi(L)}}=[P(L)(ia,iz)]_{a^{1-\chi(L)}}\,.
3936: \]
3937: \end{theorem}
3938: 
3939: We also require an extension of braids to the context of $F$.
3940: This was described in \cite{BW} and \cite{Murakami}, but we use
3941: only the generators of the algebra defined there. Strings will be
3942: assumed numbered from left to right and words will be composed
3943: from bottom to top. We write $\sg_i$ for a braid generator, where
3944: strand $i$ from the lower left corner, passing over strand $i+1$,
3945: goes to the upper right corner.
3946: 
3947: We add elements $\dl_i$ of the following form:
3948: \[
3949: \diag{4mm}{6}{2.5}{
3950:   \pictranslate{0 0.5}{
3951:   \picline{0 0}{0 2}
3952:   \picmultigraphics{3}{0.2 0}{
3953:     \picfilledcircle{0.55 1}{0.03}{}
3954:   }
3955:   \picline{6 0}{6 2}
3956:   \picmultigraphics{3}{0.2 0}{
3957:     \picfilledcircle{5.05 1}{0.03}{}
3958:   }
3959:   \picline{1.5 0}{1.5 2}
3960:   \picline{4.5 0}{4.5 2}
3961:   \piccirclearc{3 0}{0.5}{0 180}
3962:   \piccirclearc{3 2}{0.5}{180 0}
3963:   \picputtext{2.5 -.3}{\footnotesize $i$}
3964:   \picputtext{3.5 -.3}{\footnotesize $i+1$}
3965:   }
3966: }\es.
3967: \]
3968: By hat we denote the usual closure operation.
3969: 
3970: A word in the described generators gives rise to an unoriented tangle
3971: diagram that turns into an unoriented link diagram under closure. 
3972: If the word has no $\dl_i$ then this diagram can be oriented to
3973: give a(n oriented) closed braid diagram. We will assume this
3974: orientation is chosen. Otherwise, a coherent orientation is not
3975: generally possible. In particular, the sign of exponents of $\sg_i$
3976: in this context may not coincide with the sign of the corresponding
3977: crossings after some (or even any) orientation choice of the diagram.
3978: For kinks (the diagram fragments occurring in \eqref{wseven.5} on the
3979: left hand-sides),
3980: however, a sign is definable since any possible component
3981: orientation chosen gives rise to the same (skein) sign. So we will be
3982: able (and we will need) to distinguish between positive (in the
3983: left equation of \eqref{wseven.5}) and negative
3984: kinks (in the right one).
3985: 
3986: \begin{lemma}\label{lemma1}
3987: Let 
3988: \begin{eqn}\label{Df}
3989: D=\wh{\es}\,\left(\,
3990: \sg_1^{k_{1,1}}\sg_2^{k_{1,2}}\dots \sg_1^{k_{1,n_1}}\sg_2^{-1}\quad
3991: \sg_1^{k_{2,1}}\dots \sg_1^{k_{2,n_2}}\sg_2^{-1}\quad\dots
3992: \quad \sg_l^{k_{l,1}}\dots \sg_1^{k_{l,n_l}}\dl_2\,\right)
3993: \end{eqn}
3994: where $n_i\ge 1$ odd and $k_{i,j}\ge 2$ when $1<j<n_i$ and
3995: $k_{i,j}\ge 1$ when $j=1$ or $n_i$.
3996: Then $\md_a\Lm(D)=-l$.
3997: \end{lemma}
3998: 
3999: \proof For $l=1$ see lemma \reference{lemma4} below. Then use induction
4000: on $l$. Change a crossing of a $\sg_2^{-1}$ in $D=D_-$. Then:
4001: 
4002: $D_+$ can be reduced by the same argument as in the proof of
4003: lemma 5.2 of \cite{ntriv} until we have a form \eqref{Df} with
4004: smaller $l$, and a non-zero number of negative kinks added.
4005: Then since negative kinks shift the $a$-degree of $\Lm$ up,
4006: we have $\md_a\Lm(D_+)>-l$.
4007: 
4008: $D_0$ has $\md_a\Lm(D_0)=1-l>-l$ by induction.
4009: 
4010: $D_\infty=D_{l_1}\# D_{l_2}$ with $l_1+l_2=l$, so $\md_a\Lm(D_
4011: \infty)=-l_1-l_2=-l$, and $\md_a\Lm(D)$ is inherited from
4012: $\Lm(D_\infty)$.
4013: \qed
4014: 
4015: \begin{lemma}\label{lemma2}
4016: Let 
4017: \begin{eqn}\label{Df2}
4018: D=\wh{\es}\,\left(\,\sg_1^{k_{1,1}}\sg_2^{k_{1,2}}\dots
4019: \sg_1^{k_{1,n_1}}\sg_2^{-1}\quad
4020: \sg_1^{k_{2,1}}\dots \sg_1^{k_{2,n_2}}\sg_2^{-1}\quad\dots
4021: \quad \sg_l^{k_{l,1}}\dots \sg_1^{k_{l,n_l}}\sg_2^{-1}\,\right)
4022: \end{eqn}
4023: where $n_i\ge 1$ odd and $k_{i,j}\ge 2$ when $1<j<n_i$ and
4024: $k_{i,j}\ge 1$ when $j=1$ or $n_i$.
4025: Then $\md_a\Lm(D)\ge -2$ when $l\le 2$ and $\md_a\Lm(D)=-l$
4026: when $\l\ge 3$.
4027: \end{lemma}
4028: 
4029: \proof Assume first we proved the result for $l\le 2$, and that
4030: $l>2$. We argue by induction on $l$.
4031: 
4032: Apply the $\Lm$-relation at a $\sg_2^{-1}$ crossing in $D=D_-$.
4033: Then $D_0$ has $\md_a\Lm=1-l$ by induction. $D_+$ simplifies
4034: as in the proof of lemma 5.2 of \cite{ntriv}. This simplification
4035: only removes negative letters. It can be iterated until one
4036: of two situations occurs. It can (a) happen that all negative
4037: letters disappear. Then we have a positive braid and $\md_a\Lm=-2>-l$
4038: by Yokota's result. Or it can (b) occur that no $\sg_1\sg_2\sg_1$
4039: or $\sg_2\sg_1\sg_2$ occur as subwords. Then $k_{i,j}\ge 2$
4040: for $1<j<n_i$, and the number $l$ of negative crossings has
4041: decreased strictly. So we have by induction $\md_a\Lm(D_+)>-l$.
4042: Finally we must deal with the $D_\infty$ term. This follows
4043: from lemma \reference{lemma1}.
4044: 
4045: It remains to justify the claim $\md_a\Lm(D)\ge -2$ when $l\le 2$.
4046: This is done exactly with the same argument, only that now we
4047: observe that all of $\Lm(D_{0,+,\infty})$ have $a$-degree $\ge -2$. \qed
4048: 
4049: \begin{lemma}\label{lemma3}
4050: The closed braids in \eqref{Df2} are not positive for $l\ge 2$.
4051: \end{lemma}
4052: 
4053: \proof The representation \eqref{Df2} clearly gives rise to
4054: a positive band representation by replacing $\sg_2\sg_1^{k_{i,n_i}}
4055: \sg_2^{-1}$ by $\sg_3^{k_{i,n_i}}$. So $[\bt]=3-\chi(\hat\bt)$.
4056: If $l>2$, then we have
4057: \[
4058: \md_a F(D)\,=\,w(D)\,+\,\md_a \Lm(D)\,=\,3-\chi(\hat\bt)+\md_a \Lm(D)\,
4059: <\,1-\chi(\hat\bt)\,
4060: \]
4061: but $\md_v P(D)=1-\chi(\hat\bt)$ as before, so $\md_a F\ne \md_v P$
4062: and we are done by theorem \reference{thF}.
4063: 
4064: If $l=1$, then if $n_1=1$, we have a $(2,k_{1,1})$-torus
4065: link, and for $n_1=3$ we have the $(1,k_{1,1},k_{1,2},k_{1,3})
4066: $-pretzel link. Otherwise we apply the relations for $\tl P$
4067: and $\Lm$ at the negative crossing. Then $\md_a\Lm(D_{+,0})=
4068: \md_a\tl P(D_{+,0})=-2$, and
4069: $[\Lm(D_{+,0})]_{a^{-2}}=[\tl P(D_{+,0})]_{a^{-2}}$
4070: by theorem \reference{thF} since $D_{+,0}$ are positive braids.
4071: That the extra term $\Lm(D_\infty)$ has no contribution to
4072: $a^{-2}$ follows from lemma \reference{lemma4} below. So
4073: $D$ satisfies Yokota's conditions in theorem \reference{thF}.
4074: (We do not always know if $D$ depicts a positive link; see
4075: remark \reference{rty}.)
4076: 
4077: If $l=2$ then resolve a negative crossing in $D=D_-$ via the
4078: relation $\tl P_-=z \tl P_0-\tl P_+$ and via the $\Lm$-relation
4079: $\Lm_-=z\Lm_0+z\Lm_\infty-\Lm_+$. Now $[\tl P_{+,0}]_{a^{-2}}=
4080: [\Lm_{+,0}]_{a^{-2}}$. This follows from the above argument
4081: for $l=1$. The additional term $\Lm_\infty$ has $a$-degree
4082: $-2$ by lemma \reference{lemma1}, so $[\Lm_-]_{a^{-2}}\ne
4083: [\tl P_-]_{a^{-2}}$. Again by theorem \reference{thF},
4084: $D=D_-$ can therefore not belong to a positive link.  \qed
4085: 
4086: \begin{rem}
4087: Ishikawa asks in \cite{Ishikawa} whether strongly
4088: quasi-positive knots satisfy the equality $TB=2g_s-1$,
4089: where $TB$ is the maximal Thurston-Bennequin invariant.
4090: The proof of the lemma shows that there are infinitely many
4091: strongly quasi-positive 3-braid knots with $TB<2g_s-1$:
4092: take any of the knots with $l>2$. (There are also many other
4093: such knots, like the counterexamples to Morton's conjecture
4094: in \cite{posex_bcr}.)
4095: \end{rem}
4096: 
4097: \begin{lemma}\label{lemma4}
4098: If 
4099: \[
4100: D=D_{[n]}=\wh{\es} (\sg_1^{k_1}\sg_2^{k_2}\dots\sg_1^{k_n}\dl_2)
4101: \]
4102: with $n\ge 1$ odd, $k_1,k_n\ge 1$ and $k_i\ge 2$ when $i=2,\dots,n-1$,
4103: then $\md_a \Lm(D)=-1$. Writing $\ol{k}=(k_1,\dots,k_n)$ and
4104: \[
4105: \tl w(\ol{k})\,=\,(k_1-1)+\sum_{l=2}^{n}(k_l-2)\,, 
4106: \]
4107: we have $\Md_z [\Lm(D)]_{a^{-1}}\,=\,\tl w(\ol{k})$.
4108: \end{lemma}
4109: 
4110: \proof If $n=1$ we check directly (we have a reduced diagram
4111: of the $(2,k_1)$-torus link), so let $n\ge 3$.
4112: 
4113: Consider first three special forms of $\ol{k}$.
4114: 
4115: If $\ol{k}=(12^*1)$ (with $2^*$ being a sequence of `$2$'),
4116: then $D$ is regularly isotopic to a trivial
4117: 2-component link diagram. If $\ol{k}=(12^*)$ or $(2^*1)$, then $D$
4118: is regularly isotopic to an unknot diagram with one positive kink.
4119: In both situations the claims follow directly.
4120: 
4121: Now let $\ol{k}=(2^*)$. We orient $D$ so as to become negative.
4122: Then $D$ depicts the $(2,-n-1)$-torus
4123: link. We can evaluate $\Lm$ on $L$ from $F(L)$ by
4124: normalization. For its mirror image $!L$,
4125: we can use theorem \reference{thF} to conclude that
4126: $\md_a F(!L)=n$. Now it is also
4127: known that $\spn_a F(!L)=c(!L)$, and $c(!L)=n+1$, so
4128: \[
4129: \Md_a F(!L)=\md_a F(!L)+\spn_a F(!L)=n+(n+1)=2n+1\,.
4130: \]
4131: Thus $\md_a F(L)=-1-2n$, and since $w(D)=-2n$, we have
4132: \[
4133: \md_a \Lm(D)=-w(D)+\md_a F(D)=-(-2n)-1-2n=-1\,.
4134: \]
4135: That $\Md_z[\Lm(D)]_{a^{-1}}=1$ can also be obtained by
4136: direct calculation.
4137: 
4138: If $\ol{k}$ is not of these special types, then $k_l\ge 3$ for some
4139: $1\le l\le n$. We resolve a positive crossing in $\sg_j^{k_l}$.
4140: We have (with $D=D_+$)
4141: \begin{eqn}\label{*}
4142: \Lm (D_+)\,=\,z\Lm (D_0)\,+z\Lm(D_\infty)-\Lm(D_-)\,.
4143: \end{eqn}
4144: Here $D_0$ has $\tl w$ by one less and comes with a $z$-factor,
4145: so it is enough to show that $\Lm(D_-)$ and $\Lm(D_\infty)$
4146: do not contribute.
4147: 
4148: If $k_l>3$, then $D_-$ is of the required form and has $\tl w$
4149: by two less, so by induction $\Lm(D_-)$ has too small $z$-degree
4150: in $[\Lm]_{a^{-1}}$.
4151: 
4152: Now consider $k_l=3$. As in the proof of lemma 5.2 of \cite{ntriv} we
4153: can move
4154: by braid relations (regular isotopy) a $\sg_1$ in $\sg_1\sg_2\sg_1$
4155: until we get a $\sg_i\sg_i^{-1}$ (and then cancel). Here we
4156: replaced $\sg_i^{-1}$ by $\dl_i$, so that such $\sg_i$ right before
4157: a $\dl_i$ becomes a kink, which is negative. By repeating this
4158: transformation, we obtain a form that has no $\sg_i\sg_{i\pm 1}\sg_i$
4159: (i.e. $k_i>1$ for $1<i<n$) and a certain non-zero number of
4160: negative kinks collected at both ends. The negative kinks
4161: shift the degree in $a$ of $\Lm$ up, so by induction $D_-$
4162: has no contribution to $[\Lm]_{a^{-1}}$.
4163: 
4164: It remains to deal with the term of $D_\infty$ in \eqref{*}.
4165: It is the connected sum $D_{[n_1]}\# D_{[n_2]}$ with $n_1+n_2=n-1$,
4166: and with $k_l-1>1$ negative kinks. So by induction $\md_a\Lm(D_\infty)
4167: =k_l-1+(-1)+(-1)>-1$, and $\Lm(D_\infty)$ gives no
4168: contribution to $[\Lm(D_+)]_{a^{-1}}$. \qed
4169: 
4170: \begin{theorem}\label{thposq}
4171: If a 3-braid link is positive, then it is the closure
4172: of a positive or almost positive 3-braid. Along these links
4173: the non-fibered ones are exactly the $(1,p,q,r)$-pretzel
4174: links.
4175: \end{theorem}
4176: 
4177: \proof Consider the first claim.
4178: % If a 3-braid is almost positive, and does not
4179: % reduce to a positive one, then it is of the form
4180: % \eqref{Df2} with $l=1$. % We observed in the proof
4181: % of lemma \reference{lemma3}, that such links are positive.
4182: % 
4183: Since $L$ is positive, it is strongly
4184: quasi-positive, and so has a positive 3-braid band representation
4185: $\bt$ by theorem \reference{thqp}. So $\bt$ is of Xu's form $R$
4186: or $(21)^kR$ (with $k>0$) up to extensions. If $\bt$ contains
4187: $\sg_1\sg_2\sg_1$ or $\sg_2\sg_1\sg_2$ we can reduce it
4188: as before, until (a) it becomes positive, or (b)
4189: it still has a positive band representation, but it
4190: does not contain $\sg_1\sg_2\sg_1$ or $\sg_2\sg_1\sg_2$.
4191: In case (a) we are done. In case (b) we observe that
4192: $\bt$ is of the form \eqref{Df2}, apply lemma
4193: \reference{lemma3}, and conclude that $l\le 1$.
4194: 
4195: It remains to argue which links are not fibered.
4196: Positive braids are always fibered, and by direct
4197: observation for an almost positive braid we have
4198: Xu's form $R$ or $(21)^kR$ depending on whether
4199: in \eqref{Df2} (with $l=1$) we have $n_1=3$ or
4200: $n_1>3$. We proved in theorem \reference{Thq}
4201: that the forms $(21)^kR$ (for $k>0$) give fibered
4202: closure links. For $n_1=3$ we have as before the
4203: $(1,p,q,r)$-pretzel links. \qed
4204: 
4205: \begin{rem}\label{rty}
4206: Unfortunately, we cannot completely determine which of the almost
4207: positive braid representations with $l=1$ and $n_1>3$ in \eqref{Df2}
4208: give positive links. For example the Perko knot $10_{161}$ has
4209: such a representation, and it is positive. The Perko move (see
4210: \cite{HTW}), turning the closed braid diagram of $10_{161}$
4211: into a positive diagram, applies for more general examples.
4212: However, some knots, like $14_{46862}$, do not seem subjectable
4213: to this or similar moves, and their positivity status remains
4214: unclear at this point.
4215: \end{rem}
4216: 
4217: Similarly, one would hope to prove that
4218: among these links none is a positive braid link (and this way
4219: to obtain a different, but much more insightful proof of theorem
4220: \reference{thg}). Using the polynomials, one can exclude certain
4221: families, for example all links of $n_1=5$, but a complete argument
4222: again does not seem possible.
4223: 
4224: % we have a link, which belongs to a family
4225: % of the Perko knot $10_{161}$; latter arises when $n_1=5$, $k_{1,*}=
4226: % (1,2,3,2,1)$. The Perko move (see \cite{HTW}) turns this diagram
4227: % into a positive diagram, and the same move applies for the more
4228: % general examples.
4229: 
4230: % At least we can derive theorem \reference{thg} (again),
4231: % now possibly in a more favorable way.
4232: % 
4233: % \proof[of theorem \reference{thg}]
4234: % The one direction is trivial. Assume thus $L$ is braid positive and
4235: % of braid index $3$. In particular, $L$ is positive, so by theorem
4236: % \reference{thposq}, it is a closed positive or almost positive 
4237: % 3-braid $\bt$. Assume (trivially) we have latter case. $L$ is
4238: % also fibered, so $n_1>3$ in \eqref{Df2} (with $l=1$).
4239: % 
4240: % Now we use the skein polynomial $P$ and claim
4241: %  
4242: % \begin{lemma}\label{lemma6}
4243: % We have for such closed braids $L=\hat\bt$
4244: % that $[P]_{z^{-1-\chi}v^{3-\chi}}=0$. %and $V_1=V_2=0$, where
4245: % % $n(L)$ is the number of components of $L$.
4246: % \end{lemma}
4247: %  
4248: % It is easy to see that no positive braid has such
4249: % polynomial. The term in $[P]_{(zv)^{1-\chi}}$
4250: % comes from a terminal node in the positive resolution
4251: % tree in which all crossings are smoothed out until
4252: % a braid as in theorem \reference{th4.1} remains.
4253: % By considering a node in the tree in which all crossings
4254: % are smoothed out except exactly one, we have a Hopf
4255: % link, and a Hopf link has $[P]_{z^{-1}v^{3}}=1$.
4256: % Since its contribution is not cancelled, a non-trivial
4257: % positive braid must have $[P]_{z^{-1-\chi}v^{3-\chi}}\ne 0$.
4258: % \qed
4259: % 
4260: % % By \cite{posqp}, we have for a (non-trivial, non-split)
4261: % % positive braid link that $V_2\ne 0$, so $L$ is not the
4262: % % closure of apos braid, and we are done. \qed
4263: % % 
4264: % \proof[of lemma \reference{lemma6}]
4265: % The minimal possible word {\tt 1 2 2 1 1 2 2 1 -2} is found to have
4266: % $P$ of the stated type by direct calculation. Now we use induction
4267: % on the word length and apply the skein relation for $P$ at
4268: % some positive letter in a non-trivial syllable. We work
4269: % again with the form \eqref{Df2} (with $l=1$). If $n_1=5$ we
4270: % choose $i$ so that the syllable has exponent $k_{1,i}\ge 3$
4271: % if $1<i<n_1$ or $k_{1,i}\ge 2$ if $i=1$ or $n_1$. We have 
4272: % \begin{eqn}\label{prel}
4273: % P(D_+)=v^2P(D_-)+vzP(D_0)\,.
4274: % \end{eqn}
4275: % Now all three occurring links are strongly quasi-positive.
4276: % So 
4277: % \[
4278: % \md_v P(D_-)+2=\md_v P(D_0)+1=\md_v P(D_+)\, .
4279: % \]
4280: % % (This can be concluded from the explanations in
4281: % % \cite{posqp}.) Also the terms $t^2V_(D_-)$ and $t^{3/2}
4282: % % V_(D_0)$ have $V_0=\pm 1$ (of oppsoite sign!) and $V_1=0$
4283: % % Because either braids either reduce to a positive braid,
4284: % 
4285: % The choice of $k_i$ for $n_1=5$ prevents
4286: % us from descending into the undesriable situation that
4287: % $D_-$ or $D_0$ has $n_1=3$. 
4288: % $D_0$ is always of the required form.
4289: % So in particular 
4290: % \begin{eqn}\label{pd0}
4291: % [P(D_0)]_{v^{-1-\chi(D_0)}z^{3-\chi(D_0)}}=0
4292: % \end{eqn}
4293: % by induction.
4294: % 
4295: % $D_-$ may reduce to a positive braid if $k_i=3$ and $1<i<n_1$ or $k_i
4296: % =2$ for $i=1$ or $n_1$. Otherwise it
4297: % is also of the previous form with $n_1>3$.
4298: % In either case $D_-$ is fibered. Now it follows from the
4299: % description of $[P]_{1-\chi}$ in \cite{3br} (more explicitly
4300: % given in \S\reference{SA}) that fibered strongly quasi-positive
4301: % 3-braids have
4302: % \begin{eqn}\label{pd-}
4303: % [P(D_-)]_{z^{1-\chi(D_-)}v^{3-\chi(D_-)}}=0\,.
4304: % \end{eqn}
4305: % Putting now \eqref{pd0} and \eqref{pd-} into
4306: % \eqref{prel} gives the result. \qed
4307: 
4308: \section{Studying alternating links by braid index\label{Sfg}}
4309: 
4310: The combination of the identity \eqref{p21} and the skein-Jones
4311: substitution \eqref{VPsub} was already used in \S\reference{Js}
4312: to translate the determination of the 3-braid link genus from
4313: $P$ to $V$. A similar line of thought will now enable us to
4314: extend the other main result in \cite{3br}, the description
4315: of alternating links of braid index 3. This result was
4316: motivated by the work of Murasugi \cite{Murasugi},
4317: and Birman's problem in \cite{MortonPb} how to
4318: relate braid representations and diagrammatic properties
4319: of links. We will see how via \eqref{VPsub} and 
4320: the famous Kauffman-Murasugi-Thistlethwaite theorem
4321: \cite{Kauffman2,Murasugi3,Thistle2} the Jones polynomial
4322: enters in a new way into the braid representation picture.
4323: The argument will lead to the braid index 3
4324: result surprisingly easily, and then also to the
4325: classification for braid index 4 (which seems out of scope
4326: with the methods in \cite{3br} alone). We also obtain a
4327: good description of the general (braid index) case.
4328: 
4329: Our starting point is the following general result concerning the
4330: MWF-bound \eqref{mwfb}. A diagram is called special if it
4331: has no separating Seifert circles; see \cite{Cromwell}.
4332: The number of Seifert circles of $D$ is denoted by $s(D)$.
4333: 
4334: \begin{theorem}\label{_ty}
4335: Assume $L$ is a non-trivial non-split alternating link, and $MWF(L)=
4336: k$. Then an alternating (reduced) diagram $D$ of $L$ has $s(D)\le 2k-
4337: 2$ Seifert circles, and equality holds only if $D$ is special.
4338: \end{theorem}
4339: 
4340: \proof As $L$ is non-trivial and non-split, we have $1-\chi(L)> 0$.
4341: It is also well known, that $\Md_zP(L)=1-\chi(L)$ (see
4342: \cite{Cromwell}). Now $\spn_v P(L)\le 2k-2$ by assumption. Under
4343: the substitution in \eqref{mwfb} this translates to 
4344: \begin{eqn}\label{sV}
4345: \spn V(L)\le 1-\chi(L)+2k-2\,.
4346: \end{eqn}
4347: On the other hand, by \cite{Kauffman2,Murasugi3,Thistle2},
4348: $\spn V(L)=c(D)$, and also $1-\chi(L)=c(D)-s(D)+1$.
4349: So
4350: \begin{eqn}\label{ff}
4351: c(D)\le c(D)-s(D)+1+2k-2\,,
4352: \end{eqn}
4353: and then $s(D)\le 2k-1$. Now if this is an equality, then
4354: so is \eqref{sV}. Then one easily sees that $P(L)$ must have
4355: non-zero coefficients in both monomials $z^{\Md_zP}v^{\Md_vP}$
4356: and $z^{\Md_zP}v^{\md_vP}$. Now under the substitution in \eqref{p21},
4357: both these monomials give a non-cancelling contribution,
4358: and one of them is not in ($v$-)degree 0, so the identity
4359: \eqref{p21} cannot hold. Moreover, if \eqref{sV} fails
4360: just by one, then still one of the two coefficients must
4361: be non-zero. In order its (non-cancelling) contribution on
4362: the left of \eqref{p21} to be in degree $0$, we see that
4363: either $\Md_zP=\md_vP$, or $\Md_zP=-\Md_vP$. Using \cite
4364: {Cromwell}, one concludes then that these are precisely the cases of a
4365: (positively or negatively) special alternating link. \qed
4366: 
4367: Using \cite{Yamada} and \cite{Vogel} we have some simple
4368: estimate on the unsharpness of MWF for alternating links.
4369: 
4370: \begin{corr}
4371: For an alternating non-trivial non-split link $L$, we have
4372: $b(L)\le 2MWF(L)-2$. \qed
4373: \end{corr}
4374: 
4375: Of course, for many (in particular alternating) links
4376: $b(L)=MWF(L)$ (that is, MWF is exact), or at least $b-MWF$
4377: is small. So the above estimate should be considered as
4378: a worst-case-analysis. Even if not strikingly sharp,
4379: it is still far from trivial, in view of what we
4380: already know can occur for non-alternating knots. Namely,
4381: using the construction in \cite{Kanenobu2} and the work
4382: in \cite{BirMen3} (see remark \reference{rfn} below),
4383: one can find sequences of knots $(K_i)$ for which $MWF$
4384: (in fact the full $P$ polynomial) is constant, but
4385: $b(K_i)\to\infty$.
4386: 
4387: Another observation is that one can now extend the outcome
4388: of the work in \cite{SV} by replacing crossing number of
4389: an alternating knot by its braid index.
4390: 
4391: \begin{corr}
4392: Let $n_{g,b}$ denote the number of alternating knots of
4393: braid index $b$ and genus $g$. Then $n_{g,b}$ is finite.
4394: Moreover, for $g$ fixed, we have that
4395: $\lim_{b\to \infty} n_{g,b}/b^{6g-4}=C_g$ is a constant.
4396: \end{corr}
4397: 
4398: \proof The finiteness of $n_{g,b}$ follows by \eqref{ff}.
4399: When $\chi(L)$ is fixed, and $c(L)\to\infty$, one easily
4400: sees from \eqref{ff} that $MWF(L)$ behaves asymptotically
4401: (up to an $O(1)$, i.e. bounded, term)
4402: at least like $c(D)/2$. Now Ohyama's result \cite{Ohyama}
4403: implies that $b(L)$ (and $MWF(L)$ as well) behave asymptotically
4404: exactly as $c(L)/2$. So from \cite{SV} we have the result.
4405: \qed
4406: 
4407: \begin{rem}\label{rfn}
4408: Of course we could also gain, as in \cite{SV}, an estimate on the
4409: $C_g$ and its asymptotics for $g\to \infty$, it would just multiply
4410: by $2^{6g-4}$. One should also note that the finiteness of
4411: $n_{g,b}$, which one sees from \eqref{ff}, is not necessarily
4412: clear \em{a priori}. In fact, however, Birman-Menasco proved
4413: \cite{BirMen3} that $n_{g,b}$ is a finite number even for \em{general}
4414: (i.e. without restriction to alternating) knots. Their methods seem,
4415: though, quite unhelpful to estimate these numbers properly.
4416: \end{rem}
4417: 
4418: Theorem \ref{_ty} immediately leads to the first slight sharpening
4419: of the description of alternating links of braid index 3 in
4420: \cite{3br}. (The case of $MWF=2$ is even more obvious, and omitted.)
4421: 
4422: \begin{corr}\label{opi}
4423: An alternating link has $MWF=3$ if and only if it has braid index 3.
4424: \end{corr}
4425: 
4426: \proof If $MWF(L)\le 3$, then the alternating diagram has at
4427: most 4 Seifert circles, and exactly 4 only if it is special.
4428: Apart from connected sums (which are easily handled), we obtain
4429: the diagrams of closed alternating 3-braids and the $(p,q,r,s)$-
4430: pretzel diagrams. By direct calculation of $P$ we saw in
4431: \cite{3br} that if $\min(p,q,r,s)\ge 2$, then $MWF\ge 4$,
4432: and that otherwise the pretzel link has braid index 3. \qed
4433: 
4434: The case of 4-braids is now not too much more difficult.
4435: % (even though
4436: % it was completely out of scope with the method in \cite{3br}).
4437: 
4438: \begin{theorem}\label{tty}
4439: Let $L$ be a prime non-split alternating link. The following 3
4440: conditions are equivalent:
4441: \begin{enumerate}
4442: \item\label{Xa} $MWF(L)=4$
4443: \item\label{Xb} $b(L)=4$
4444: \item\label{Xc} $L$ is one of the links, whose reduced 
4445: alternating diagrams are described (up to mirror images) as follows
4446: \begin{enumerate}
4447: \item\label{Ya} The Murasugi (or connected) sum of three
4448:   $(2,n_i)$-torus links (with $|n_i|>1$),
4449: \item\label{Yb} The Murasugi (or connected) sum of a $(2,n)$-torus
4450:   link with a $(p,q,r,s)$-pretzel link, with one of $p,q,r,s$ equal
4451:   to 1, or its mirror image, or
4452: \item\label{Yc} a special diagram whose Seifert graph
4453: (see \S\ref{Sdg}) is as shown in figure \reference{figSG}.
4454: \end{enumerate}
4455: \end{enumerate}
4456: \end{theorem}
4457: 
4458: \proof $\reference{Xb}\So \reference{Xa}$. This follows from
4459: corollary \ref{opi}.
4460: 
4461: $\reference{Xc}\So \reference{Xb}$. That the links in
4462: \reference{Xc} have braid index at least 4 follows from the
4463: description of the links with $b(L)\le 3$ in \cite{3br}
4464: (which also comes out of the proof of corollary \eqref{opi}).
4465: It is also not too hard to check that for all these links
4466: $b=4$ by exhibiting a diagram $D$ with $s(D)=4$. The most
4467: systematic way seems to apply the graph index inequality of
4468: Murasugi-Przytycki \cite{MurPrz2} (see also \cite{Ohyama}).
4469: 
4470: $\reference{Xa}\So \reference{Xc}$. By applying theorem
4471: \reference{_ty}, we need to deal with non-special
4472: diagrams of at most 5 and special diagrams of at most 6
4473: Seifert circles.
4474: 
4475: First consider the non-special diagrams.
4476: 
4477: For the fibered links (the reduced Seifert graph is a tree),
4478: Murasugi's result \cite{Murasugi} leads directly to case
4479: \reference{Ya}. For non-fibered links, the Seifert graph
4480: must have a cycle, which must be of length at least 4 (the
4481: Seifert graph is bipartite). Then the only option
4482: that remains is case \reference{Yb}. We must still
4483: argue why one of $p,q,r,s$ must be $\pm 1$. This can
4484: be done using Murasugi-Przytycki's work, but one easily
4485: sees it also by a direct skein theoretic argument, which
4486: we explain.
4487: 
4488: Let $L(p,q,r,s)$ be the corresponding link and $P(p,q,r,s)$
4489: its skein polynomial. Look first at $p=0$, $q,r,s\ge 2$. Then
4490: $L(p,q,r,s)$ a connected sum of two $(2,n)$-torus
4491: links, and a closed alternating 3-braid. So
4492: $MWF(L(p,q,r,s))=5$. Now since the case $p=1$, $q,r,s\ge 2$ has
4493: $MWF(L(1,q,r,s))\le b(L(1,q,r,s))=4$, the skein relation \eqref{srel}
4494: easily shows that the maximal degree coefficient $\Mc_vP(p,q,r,s)$
4495: of $v$ in $P$ (which is a polynomial in $z$) for $p=2$ is inherited
4496: from $p=0$. Then further applications of \eqref{srel} show that
4497: the $z$-degree of $\Mc_vP(p,q,r,s)$ increases with $p$, so 
4498: in particular this term never vanishes, and so $MWF(L(p,q,r,s))=5$.
4499: 
4500: \begin{figure}[htb]
4501: {\small
4502: \[
4503: \begin{array}{c}
4504:   \diag{6mm}{3}{2}{
4505:   \pictranslate{0 0.5 x}{
4506:     \diag{6mm}{3}{2}{
4507:       \cycl{0.5 2}{0.5 0.5}{2 0.5}{2 2}
4508:     }
4509:   }
4510:   \picputtext{1.75 0}{\scriptsize (all multiplicities $\ge 2$)}
4511:   }
4512:   \qquad
4513:   \diag{6mm}{3}{2}{
4514:     \cycl{1.3 2}{0 1}{1.1 0}{1.5 1}
4515:     \cycl{1.3 2}{2.8 1}{1.1 0}{1.5 1}
4516:     \picputtext{2.5 0.4}{$1$}
4517:   }
4518:   \qquad
4519:   \diag{6mm}{3}{2}{
4520:     \pictranslate{0 0.5}{
4521:     \xcycl{0 0}{1.5 0}{3 0}{3 1.5}{1.5 1.5}{0 1.5}
4522:     \picputtext{0.75 -0.3 x}{$1$}
4523:     \picputtext{0.75 -0.3}{$1$}
4524:     }
4525:   }
4526:   \qquad
4527:   \diag{6mm}{3}{2}{
4528:     \pictranslate{0 0.5}{
4529:     \xcycl{0 0}{1.5 0}{3 0}{3 1.5}{1.5 1.5}{0 1.5}
4530:     \picline{1.5 1.5}{1.5 0}
4531:     \picputtext{2.25 -0.3}{$1$}
4532:     \picputtext{0.75 -0.3}{$1$}
4533:     }
4534:   }
4535:   \qquad
4536:   \diag{6mm}{4}{2.5}{
4537:     \pictranslate{0 0.5}{
4538:     \xcycl{0 0}{1.5 0}{3 0}{3 1.5}{1.5 1.5}{0 1.5}
4539:     \picline{1.5 1.5}{1.5 0}
4540:     \picputtext{2.25 1.8}{$1$}
4541:     \picputtext{0.75 -0.3}{$1$}
4542:     \opencurvepath{0 1.5}{0 2}{0.5 2.5}{3 2.5}{4 2}{4 0.5}{3.5 0}{3 0}{}
4543:     }
4544:   }
4545:   \\[5mm]
4546: \end{array}
4547: \]
4548: }
4549: \caption{The reduced Seifert graphs of the alternating diagrams
4550: of special alternating links of braid index 4. Simple edges
4551: have their multiplicity (1) attached, and the other edges (of
4552: multiplicity one or more) are unlabelled. In the first graph,
4553: exceptionally, all edges have multiplicity at least 2.\label{figSG}}
4554: \end{figure}
4555: 
4556: 
4557: Now consider the special diagrams. For them one considers the
4558: Seifert graph, and needs to write
4559: down all bipartite planar graphs on at most 6 vertices, which
4560: have no cut vertex. Since the diagram $D$ is special, the placement
4561: of multiple copies of an edge give diagrams equivalent up to
4562: flypes, so it is enough to consider simple graphs (the reduced
4563: Seifert graph), and have the multiplicities of an edge written
4564: as its label. The graphs can be easily compiled using the observation
4565: that they must contain a cycle of length 4 or 6; see figure
4566: \reference{figSG}. By direct inspection we see that the edge
4567: multiplicities must be as specified in the figure. (In fact if an
4568: edge is multiple it turns out irrelevant what its multiplicity
4569: is, so in this case we just omit the label.) We rule out the
4570: remaining multiplicities by a skein theoretic calculation,
4571: similar to the one explained for case \reference{Yb}.
4572: 
4573: Let w.l.o.g. (up to mirroring) $D$ be positive. For each edge
4574: $e$ in $\Gm(D)$ of variable multiplicity $\ge i$ (where $i=1,2$)
4575: we calculate the skein polynomial of the diagram that
4576: corresponds to $\Gm$ for multiplicities $i$ and $i+1$ of $e$.
4577: That is, if $\Gm(D)$ has $l$ edges of variable multiplicity,
4578: we have $2^l$ polynomials to calculate. Then we check for
4579: each such set of $2^l$ polynomials that $Q=P_{v^{9-\chi(D)}}$
4580: is non-zero, and $\Md_zP-\Md_zQ$ as well as $\Mc_zQ$ is
4581: constant within this set of $2^l$ polynomials. Then
4582: by \eqref{srel} this property is inherited to diagrams
4583: $D$ whose $\Gm(D)$ have edges of higher multiplicity,
4584: and in particular $MWF\ge 5$.
4585: 
4586: There is one more graph,
4587: \[
4588: \diag{6mm}{3}{2}{
4589:   \cycl{2.0 2}{0 1}{1.8 0}{1.2 1}
4590:   \cycl{2.0 2}{2.4 1}{1.8 0}{3.5 1}
4591: }
4592: \]
4593: not included in figure \reference{figSG}. In that
4594: case (by the method we just explained we verify that)
4595: $MWF\ge 5$ for all non-zero edge multiplicities. \qed
4596: 
4597: In \cite{MurPrz2}, Murasugi-Przytycki define a certain
4598: quantity $\inx(D)$, assigned to a link diagram $D$,
4599: called \em{index}. (We omit here the detailed discussion;
4600: one can consult also \cite{Ohyama} or \cite{gener}.)
4601: Their motivation was to give an upper estimate 
4602: \begin{eqn}\label{mp}
4603: b(L)\le s(D)-\inx(D)
4604: \end{eqn}
4605: for the braid index of the underlying link $L$. Their origin
4606: of \eqref{mp} consists in an appropriate move (see
4607: figure 8.2 in \cite{MurPrz}) which reduces
4608: the number of Seifert circles of the diagram.
4609: Murasugi-Przytycki conjectured that for an alternating
4610: diagram $D$, the inequality \eqref{mp} is exact. This
4611: conjecture is also confirmed for alternating links up
4612: to braid index 4. 
4613: 
4614: Based on the Murasugi-Przytycki procedure, we can re-enter
4615: the Bennequin surface topic.
4616: 
4617: \begin{corr}\label{cfT}
4618: Alternating links of braid index at most 4 carry a Bennequin
4619: surface on a minimal string braid.
4620: \end{corr}
4621: 
4622: \proof In \cite{gener} it was explained how to apply
4623: restrictedly the Seifert circle reduction move of
4624: Murasugi-Przytycki (and then those of Yamada \cite{Yamada})
4625: so as to obtain a braided surface. This modified reduction
4626: is easily checked to lead to the minimal number of strings for the
4627: links in question. \qed
4628: 
4629: % (It was verified in \cite{gener}
4630: % for alternating knots of genus up to 4.)
4631: 
4632: \begin{corr}
4633: Let $L$ be an alternating link of braid index $4$. Then
4634: $|\Mc\Dl(L)|\le 5$, and if $|\Mc\Dl(L)|>2$, then $L$ is
4635: special alternating.
4636: \end{corr}
4637: 
4638: \proof It is well known, that $|\Mc\Dl(L)|$ is multiplicative
4639: under Murasugi sum and for a special diagram depends only
4640: on the reduced Seifert graph. The result follows by calculation
4641: for the specific types. \qed
4642: 
4643: The proof of this corollary, and the extension of the multiplicativity
4644: of $\Mc\Dl$ to $\Mc_zP$ for diagrammatic Murasugi sum \cite{MurPrz}
4645: demonstrates also the following more general principle:
4646: 
4647: \begin{corr}\label{cSG}
4648: For any given braid index, there are only finitely many values of
4649: $\Mc\Dl$ and $\Mc_zP$ among Alexander and skein polynomials of
4650: alternating links of that braid index. \qed
4651: \end{corr}
4652: 
4653: We saw that for $\Dl$ this statement is wrong for non-alternating
4654: links even among 3-braids. On the other hand, by \cite{3br},
4655: it is true for $P$, and we do not know if it remains true for
4656: (closed) braids on more strings. (One could also ask if
4657: infinitely many leading coefficients of $\Dl$ occur if $\Md\Dl
4658: =1-\chi$, but we se no deeper meaning in this question, so will
4659: not dwell further upon it here.)
4660: 
4661: The following is also worth observing. Call a subclass
4662: $\cC'$ of a class $\cC$ of links \em{generic} in $\cC$ if
4663: \[
4664: \lim_{n\to\infty}\,\frac{\#\,\{L\in \cC'\,:\,c(L)=n\,\}}
4665: {\#\,\{L\in \cC\,:\,c(L)=n\,\}}\,=\,1\,.
4666: \]
4667: 
4668: \begin{corr}
4669: The number of special alternating links of given braid index
4670: grows polynomially in the crossing number. In particular,
4671: a generic alternating link of given braid index is not
4672: special alternating.
4673: \end{corr}
4674: 
4675: \proof A special alternating link is determined by the Seifert
4676: ($=$ checkerboard) graph of its alternating diagram. The number
4677: of such graphs with a fixed number of vertices grows polynomially
4678: in the number of edges. The second claim in the corollary
4679: follows because it is easy to see that in contrast the number
4680: of non-special alternating links grows exponentially (due to
4681: exponentially many, in the crossing number, ways to perform the
4682: Murasugi sum at a separating Seifert circle of the alternating
4683: diagram). \qed
4684: 
4685: \begin{rem}
4686: Note that in contrast we showed in \cite{SV} that a generic
4687: alternating knot (and the case of links is analogous) of given
4688: \em{genus is} special alternating. This shows from yet another
4689: point of view the opposition between genus and braid index.
4690: \end{rem}
4691: 
4692: Another immediate and useful consequence of theorem \ref{tty}
4693: and the preceding remarks is
4694: 
4695: \begin{corr}\label{cft}
4696: If an alternating link $L$ has $MWF(L)\le 4$ (in particular
4697: if $b(L)\le 4$), then the MWF inequality is exact
4698: (i.e. an equality) for $L$. \qed
4699: \end{corr}
4700: 
4701: This gives a nice complement to the MWF exactness results in
4702: \cite{Murasugi,MurPrz2}. (As another such amplification, we 
4703: proved the case of \em{knots} and genus $\le 4$ in \cite{gener}.)
4704: 
4705: Note that at MWF bound 5 we hit already at the Murasugi-Przytycki
4706: examples \cite{MurPrz2} of non-exact MWF (with $b=6$). So corollary
4707: \reference{cft} is not true for $MWF\ge 5$ or $b\ge 6$. We do
4708: not know about the case $b=5$. However, ruling out braid index
4709: $5$ for the Murasugi-Przytycki family is a serious computational
4710: problem (only two specific members were dealt with, a 4-component
4711: 15 crossing link and an 18 crossing knot; see \cite[\S 19]
4712: {MurPrz2}). Already with this circumstance in mind,
4713: one cannot expect to easily extend the corollary (or theorem
4714: \reference{tty}) for $b=5$ either, even if it may be true.
4715: 
4716: On the other hand, leaving these troublesome exceptions aside,
4717: the above discussion should fairly clearly explain how
4718: the general picture continues for alternating links with MWF
4719: bound 5 and more.
4720: 
4721: \section{Applications of the representation theory\label{SBurau}}
4722: 
4723: So far the representation theory behind the skein, Jones, and
4724: Alexander polynomial was not used. We will give some applications
4725: of it now. The theory is well explained in \cite{Jones}. We will
4726: use Jones's conventions, unless otherwise specified.
4727: 
4728: \subsection{The Jones conjecture\label{SJS}}
4729: 
4730: There is a conjecture, often attributed to Jones (who speculated
4731: on it at least for knots and in it its weaker form, as given below;
4732: see 357 l.-6 of his paper), stating that a minimal
4733: string braid representation has a unique exponent sum.
4734: 
4735: \begin{conj}(Jones's conjecture)\label{jcc}
4736: \def\labelenumi{\theenumi)}
4737: \begin{enumerate}
4738: \item (weaker version)\label{weV}
4739: If $\bt,\bt'\in B_n$ satisfy $\hat\bt=\hat\bt'=L$ and $n=b(L)$,
4740: then $[\bt]=[\bt']=:w_{\min}(L)$.
4741: \item (stronger version) Part \reference{weV} holds, and if
4742: $\bt''\in B_n'$ for $n'>n$ has $\hat\bt''=L$, then 
4743: \begin{eqn}\label{stz}
4744: \big|[\bt'']-w_{\min}(L)\big|\le n'-n\,.
4745: \end{eqn}
4746: \end{enumerate}
4747: \end{conj}
4748: 
4749: It was observed in \cite{posex_bcr} that counterexamples
4750: to the conjecture would make MWF and \em{all} of its cabled
4751: versions unsharp. (In \cite{posex_bcr} the weaker version was
4752: focussed on, but the same arguments address the stronger version
4753: too.) Thus, for example, corollary \reference{cft} can also
4754: be regarded as a partial solution to Jones's conjecture.
4755: Similarly, the work in \S\reference{S2.2} shows:
4756: 
4757: \begin{corr}
4758: If a link $L$ is the closure of a positive braid, and $b(L)\le 4$,
4759: then the strong Jones conjecture is true for $L$. 
4760: \end{corr}
4761: 
4762: \proof We use theorem \reference{thmwf}, with the remark that
4763: the exceptional braid words therein were shown to have braid
4764: index 4 using the 2-cabled MWF. \qed
4765: 
4766: For 3-braids the weaker version was known to be true again
4767: from Birman-Menasco's classification result. We will show now
4768: the stronger version, as a consequence of the description in
4769: \cite{3br} of 3-braid links with unsharp MWF inequality (and thus
4770: with a much simpler proof than appealing to Birman-Menasco). See
4771: also remark \reference{rJC}.
4772: 
4773: \begin{theorem}\label{thJC}
4774: The stronger version of Jones's conjecture holds for 3-braid links.
4775: \end{theorem}
4776: 
4777: \noindent{\bf Convention.}
4778: The letter $\Dl$, with an integer subscript $n$, is used in
4779: \S\ref{SJS},
4780: in deviation from other sections, exclusively for the half-twist
4781: braid on $n$ strings, and \em{not} for the Alexander polynomial.
4782: (The Alexander polynomial will not appear in \S\reference{SJS}.)
4783: 
4784: \proof[of theorem \reference{thJC}]
4785: As explained, it suffices to deal only with the 3-braid links
4786: of unsharp MWF inequality. In \cite{3br} these links were
4787: described fully. If $\mu=\sg_1^{6a\pm 1}\sg_2^{\mp 1}$,
4788: then for the Birman dual $\bt=\mu^*$ of $\mu$ (see definition
4789: \ref{DEFB}) we have $\hat\mu^*\ne \hat\mu$ when $6a\pm 1>6$, but
4790: $K=\hat\bt$ has the skein polynomial of the $(2,6a\pm 1)$-torus
4791: knot $\hat \mu$. In \cite{3br} is was proved that these knots $K$
4792: form the full list of 3-braid links with unsharp MWF inequality.
4793:  
4794: Consider first %(with a slight change of notation)
4795: the knots $K=\hat\bt$ with $\bt=\mu^*$ for
4796: $\mu=\sg_1^{6a+1}\sg_2^{-1}$ and $a>0$. It is easy
4797: to bring $\bt$ into Xu's normal form, and to observe
4798: that it has exactly one negative band. This exhibits
4799: a Seifert ribbon (see \cite{Rudolph}) of smaller genus
4800: than $g(K)$, i.e. $g_s(K)<g(K)$ (see also the proof of
4801: theorem \reference{thqp}). Now from Xu's form one sees $w_{\min}
4802: (K)=[\bt]=2g(K)$. Take this for a moment as a definition of
4803: $w_{\min}(K)$; this writing will be justified when we
4804: show that the writhe is unique. (One could also quote
4805: Birman-Menasco here, but the complexity of their argument
4806: is unnecessary.) So 
4807: \begin{eqn}\label{iopa}
4808: 1-\chi_s(L)<w_{\min}(L)\,.
4809: \end{eqn}
4810: Using \eqref{mq},
4811: we see that if \eqref{stz} is violated for some $\bt''\in B_{n'}$,
4812: then $[\bt'']=w_{\min}(L)+2+(n'-3)$. However, by the Rudolph-%
4813: Bennequin inequality \cite{Rudolph2} (see also \cite{pos}),
4814: we would have then 
4815: \[
4816: 1-\chi_s(L)\ge [\bt'']-n'+1=w_{\min}(L)\,,
4817: \]
4818: in contradiction to \eqref{iopa}.
4819: 
4820: The effort focusses on the other braids $\bt=\mu^*$
4821: for $\mu=\sg_1^{6a-1}\sg_2$.
4822: In this case we consider $!K$, whose Xu form has only negative
4823: bands. The inequalities \eqref{mq} show that we must rule out a
4824: braid representation $\bt''\in B_{n'}$ of $!K$ with $[\bt'']=w_{\min}
4825: (!K)+2+(n'-3)$, where (we set in the same \em{a posteriori} to
4826: be justified manner) $w_{\min}(!K)=-2g(K)-2$.
4827: 
4828: We prohibit $\bt''$ by evaluating the 2-cable skein polynomial
4829: $P(\hat\gm_a)$ of $!K$ for
4830: \begin{eqn}\label{gma}
4831: \gm_a\,=\,[4354]\cdot [2132]^{6a-1}\,(\Dl_6^2)^{-2a}\,\in\,B_6,
4832: \end{eqn}
4833: with $a>1$ and $\Dl_6^2=[12345]^6$ generating the center of $B_6$,
4834: and showing that the polynomial has non-zero terms in $v$-degree
4835: $[\gm_a]-5$ or $[\gm_a]-3$. This may appear a banality, but in
4836: fact requires a substantial use of Jones' work in \cite{Jones}.
4837: 
4838: Consider the more general 6-braids
4839: \[
4840: \bt_{k,l}\,:=\,[4354]\cdot [2132]^k\,\Dl_6^{2l}\,.
4841: \]
4842: (We are of course only interested in the special case $k=6a-1$,
4843: $l=-2a$ and $a>1$, but it is useful to treat the 2-parameter family
4844: first.)
4845: 
4846: The skein polynomial of $\hat\bt_{k,l}$ can be evaluated using
4847: the representation theory in \cite{Jones}. We adopt the convention
4848: that \em{all references} to lemmas, page and paragraph numbers,
4849: equations of the form ($x.y$), etc. for the rest of this proof
4850: are understood to be \em{to Jones's paper}, unless noted otherwise.
4851: 
4852: Jones's % paper gives a clear and detailed account on the
4853: % calculation of the skein polynomial. His
4854: version of the skein polynomial, 
4855: \[
4856: X(q,\lm)=P(v,z)\mbox{\quad with $v=\sqrt{\lm q}$\ \ and
4857: \ $z=\sqrt{q}-1/\sqrt{q}$\,},
4858: \]
4859: can be evaluated on a closed
4860: $n$-braid $\bt$ by a weighted sum, with weights $W_Y$ in $\lm$ and
4861: $q$, of traces (in $q$) of irreducible representations (irreps) $\pi_Y$
4862: of $B_n$, indexed by Young diagrams (or tableaux\footnote{Here
4863: consistently `tableau' is used as a synonym for (Young) diagram,
4864: i.e. with no additional information attached to it.}) $Y$, or
4865: equivalently by partitions of $n$. (A partition of $n$ is a
4866: tuple $(n_1,\dots,n_k)$ with $n_k>0$, $n_i\ge n_{i+1}$ and
4867: $\sum n_i=n$.) We may identify the Young diagram with its
4868: partition, counting partitions in horizontal rows. For
4869: example the partition $(4)=\Young{4}$ means one row (trivial
4870: representation), while $(1111)=\Young{1,1,1,1}$ means one column
4871: (parity representation); thus $\pi_{3,1}=\pi_{\Young[1.5mm]
4872: {3,1}}$. Since the calculation of $W_Y$ was given in
4873: \cite{Jones}, we will not repeat it in detail. We will
4874: also deal (mostly) with $Y$ where the calculation of
4875: $\pi_Y$ is explained in \cite{Jones}.
4876: 
4877: Using Definition 6.1 (of \cite{Jones}), (5.5) and the formula
4878: for $W_Y(q,\lm)$ from p. 347 top, one has for a 6-braid $\bt$
4879: of exponent sum $e=[\bt]$,
4880: \begin{eqn}\label{STR}
4881: X_{\hat\bt}(q,\lm)\,=\,-\sqrt{\lm}^{e-5}\,\sum_{Y\vdash 6}
4882: \wt W_Y(q,\lm)\,\tr \pi_Y(\bt)\,.
4883: \end{eqn}
4884: Here $\wt W_Y$ denote the slightly rescaled weights
4885: \[
4886: \wt W_Y(q,\lm)\,=\,\frac{R_{Y}(q,\lm)}{Q_Y(q)}\,\cdot\,
4887: \frac{1-q}{1-\lm q}\,,
4888: \]
4889: with $R_Y$ being specified on p.\ 347 after Figure 5.6, and
4890: $Q_Y(q)$ being the hook length product term from p. 346 middle.
4891: The symbol `$\mathord{\vdash}$' is taken from partition theorists
4892: and is used to mean here that $Y$ is a Young tableau of $6$ boxes
4893: (or equivalently a partition of $6$).
4894: In order to avoid denominators it is useful to multiply
4895: \eqref{STR} by $(1-q^2)\cdot\dots\cdot(1-q^6)$, so
4896: \begin{eqn}\label{STR2}
4897: -(1-q^2)\cdot\dots\cdot(1-q^6)X_{\hat\bt}\,=\,
4898: \sqrt{\lm}^{e-5}\,\sum_{Y\vdash 6}
4899: \wh W_Y(q,\lm)\,\tr \pi_Y(\bt)\,,
4900: \end{eqn}
4901: with 
4902: \[
4903: \wh W_Y(q,\lm)\,=\,\frac{R_Y(q,\lm)}{1-\lm q}\,\cdot\,
4904: \frac{(1-q)(1-q^2)\cdot\dots\cdot(1-q^6)}{Q_Y(q)}\,
4905: \]
4906: becoming a Laurent polynomial in $\lm$ and $q$\,.
4907: 
4908: Note that $\pi_Y$ are representations that involve only the
4909: variable $q$, not $\lm$. The obvious question is how to
4910: evaluate their traces.
4911: 
4912: Consider first 
4913: \[
4914: \bt_k=\bt_{k,0}\,=\,[4354]\,[2312]^k\,.
4915: \]
4916: 
4917: There are 11 Young diagrams $Y$ of 6 boxes. Now $\bt_k\in B_4\subset
4918: B_6$, and according to p. 340 top, the restriction of $\pi_Y$ to $B_4$
4919: splits into a direct sum of representations $\pi_{Y'}$ indexed by
4920: 4-box Young diagrams $Y'$, which we name for shorthand
4921: \begin{eqn}\label{shh}
4922: A=\Young{1,1,1,1}\quad
4923: B=\Young{2,1,1}\quad
4924: C=\Young{2,2}\quad
4925: D=\Young{3,1}\quad
4926: E=\Young{4}\,.
4927: \end{eqn}
4928: 
4929: The representations $\pi_{Y'}$ are clarified completely in \S 8
4930: (see also \S\reference{SUn} in \em{this} paper below). Denote by
4931: $\psi_{n-1}$ the (reduced) $(n-1)$-dimensional representation
4932: of $B_n$; see \S 2. Let us also write $-1$ for the parity
4933: representation $(-1)^{[\,.\,]}$, and $-\rho=-1\otimes \rho$ for the
4934: direct (tensor, or Kronecker) product of a representation $\rho$ with
4935: the parity. Then $\pi_A=-1$, $\pi_E$ is given by $q^{[\,.\,]}$,
4936: $\pi_B=-\psi_3$ (Note 5.7; the sign disappears here, though,
4937: because for us always $e=[\bt]$ is even), $\pi_D=-\bigwedge^2\psi_3$
4938: and $\pi_E=\pi_{\Young[1.5mm]{2,1}}\circ\ol{\ry{0.6em}\es}$,
4939: with $\ol{\es\ry{0.6em}}$ being the homomorphism $B_4\to B_3$
4940: given by $\ol{\sg_{1,2,3}\ry{0.6em}}=\sg_{1,2,1}$ (p. 355).
4941: Also $\pi_{\Young[1.5mm]{2,1}}=-\psi_2$.
4942: 
4943: Let us write below $\ol{Y}$ for the \em{transposed} (or \em{dual})
4944: Young diagram to $Y$, given by exchanging rows and columns.
4945: For example, for $Y=\Young{3,1}$, we have $\ol{Y}=\Young{2,1,1}$.
4946: The relation between $\pi_Y$ and $\pi_{\ol{Y}}$ is given in Note
4947: 4.6.
4948: 
4949: Now $\psi_3(\ap)$ and $\psi_2(\bar\ap)$ for 
4950: \[
4951: \ap=[2132]
4952: \]
4953: are easy to
4954: calculate (see \S 2). We find former's eigenvalues to be $\pm t$ and
4955: $-t^2$, and latter's $t$ and $t^3$. In particular, both matrices are
4956: diagonalizable (because the eigenvalues are distinct for generic $t$).
4957: Setting $t=q$ as in Note 5.7, then we
4958: have the following table of eigenvalues of the $\pi_{Y'}(\ap)$
4959: \[
4960: \begin{array}{|c|c|c|c|c|}
4961: \hline
4962: \ry{1.3em}A & B & C & D & E \\[1mm]
4963: \hline
4964: \ry{1.5em}1& -q   & q^3 & -q^3 & q^4 \\
4965: 	 &  q   & q   &  q^3 &     \\
4966: 	 & -q^2 &     & -q^2 &     \\[2mm]
4967: \hline
4968: \end{array}
4969: \kern2cm
4970: \begin{array}{c|*{7}{|c}}
4971: \ry{1.2em}i & 1 & 2 & 3 & 4 & 5 & 6 & 7 \\[1.5mm]
4972: \hline  
4973: \ry{1.2em}\dl_i & 1 & q^4 & q & -q & -q^2 & q^3 & -q^3\\[1mm]
4974: \end{array}
4975: \]
4976: Let us number the 7 possible eigenvalues of $\pi_Y'(\ap)$ by
4977: $\dl_i$ as shown on the right. 
4978: 
4979: Thus $\pi_Y(\ap)$ are all diagonalizable, with eigenvalues $\dl_i$.
4980: The multiplicities of $\pi_{Y'}$ in $\pi_Y$ are also easy to
4981: calculate. They are the number of descending paths from $Y'$ to
4982: $Y$ in the Figure 3.3 (continued one more row to the bottom).
4983: Table \reference{tab1} shows the multiplicities of $\pi_{Y'}$
4984: (in the shorthand of \eqref{shh}) in $\pi_Y$ and the resulting ones,
4985: which we write $M(i,Y)$, of the eigenvalues $\dl_i$ in $\pi_Y(\ap)$.
4986: 
4987: \begin{table}[ptb]
4988: \captionwidth\vsize\relax
4989: \newpage
4990: \vbox to \textheight{\vfil
4991: \rottab{ 
4992: \hbox to \vsize{%\hss%\footnotesize
4993: \hss
4994: % \setbox\@tempboxa=\hbox{%
4995: % \textwidth\vsize\relax
4996: $\displaystyle
4997: \begin{array}{|c||*{11}{c|}}
4998: \hline\ry{1.0cm}%
4999: Y & \Young{1,1,1,1,1,1} & % 1
5000: \Young{2,1,1,1,1} &       % 2
5001: \Young{2,2,1,1} &         % 3
5002: \Young{2,2,2} &           % 4
5003: \Young{3,1,1,1} &         % 5
5004: \Young{3,2,1} &           % 6
5005: \Young{3,3} &             % 7
5006: \Young{4,1,1} &           % 8
5007: \Young{4,2} &             % 9
5008: \Young{5,1} &             % 10
5009: \Young{6} \\[6mm]         % 11
5010: \hline\ry{1.6em}%
5011: Y' & A & 2A+B & A+2B+C & B+C & A+2B+D & 2B+2C+2D & C+D & B+2D+E & C+2D+E
5012: & D+2E & E \\[2mm]
5013: \hline\ry{1.6em}%
5014: %  PARTITION #7 & #8 ARE SWOPPED FROM MY NOTES
5015: %      #1 #2  #3  #4  #5  #6  #8  #7  #9  #10 #11
5016: 1    & 1 & 2 & 1 &   & 1 &   &   &   &   &   &   \\
5017: q^4  &   &   &   &   &   &   &   & 1 & 1 & 2 & 1 \\
5018: q    &   & 1 & 3 & 2 & 2 & 4 & 1 & 1 & 1 &   &   \\
5019: -q   &   & 1 & 2 & 1 & 2 & 2 &   & 1 &   &   &   \\
5020: -q^2 &   & 1 & 2 & 1 & 3 & 4 & 1 & 3 & 2 & 1 &   \\
5021: q^3  &   &   & 1 & 1 & 1 & 4 & 2 & 2 & 3 & 1 &   \\
5022: -q^3 &   &   &   &   & 1 & 2 & 1 & 2 & 2 & 1 &   \\[2mm]
5023: \hline\ry{1.6em}%
5024: d    & 1 & 5 & 9 & 5 & 10& 16& 5 &10 & 9 & 5 & 1 \\
5025: r    & 0 & 1 & 3 & 2 & 4 & 8 & 3 & 6 & 6 & 4 & 1 \\
5026: 30r/d& 0 & 6 &10 &12 & 12& 15& 18& 18& 20& 24& 30\\[2mm]
5027: \hline
5028: \end{array}%
5029: $%
5030: % }%
5031: % \@tempdima0.3\vsize\relax\advance\@tempdima by -0.3\wd\@tempboxa\relax
5032: % \kern\@tempdima
5033: % \hss
5034: % \copy\@tempboxa
5035: \hss}%
5036: }{This table displays the Young tableaux $Y$ of 6 boxes, the
5037: decomposition into $\pi_{Y'}$ for 4-box Young tableaux $Y'$ of the
5038: sub-representation of $B_4$ in $\pi_Y$ (writing $Y'$ for $\pi_{Y'}$),
5039: the multiplicities $M(i,Y)$
5040: of the 7 possible eigenvalues $\dl_i$ of $\pi_Y(\ap)$, and the
5041: quantities $d$ ($=\dim \pi_Y$) and $r$ (rank of the idempotent
5042: $e_i$) occurring in Jones's lemma 9.3.\label{tab1}}{table}
5043: \vss}
5044: \newpage
5045: \end{table}
5046: 
5047: Since with $\pi_{Y'}(\ap)$ also all $\pi_Y(\ap)$ are
5048: diagonalizable, we have
5049: \begin{eqn}\label{XX}
5050: \tr \pi_Y(\bt_k)\,=\,\tr \pi_Y(\ap^k\cdot [4354])\,=\,
5051: \sum_{i=1}^7\,c(i,Y)\,\dl_i^k\,,
5052: \end{eqn}
5053: with $c(i,Y)$ given as follows. Consider $\pi_Y([4354])$ in
5054: the basis of $\pi_Y$ that diagonalizes $\pi_Y(\ap)$. Then to
5055: obtain $c(i,Y)$, sum the $(j,j)$-entries of the matrix
5056: of $\pi_Y([4354])$ over rows/columns $j$, for which the
5057: $(j,j)$-entry of $\pi_Y(\ap)$ is $\dl_i$.
5058: 
5059: So the problem to evaluate $\tr \pi_Y(\bt_k)$ transforms into
5060: the one to determine $c(i,Y)$. There are \em{a priori} 77
5061: of those, given by combining $7$ eigenvalues $\dl_i$ with
5062: 11 Young diagrams $Y$. However, one immediately notes that
5063: clearly $c(i,Y)=0$ when $M(i,Y)=0$. This leaves 45 of the 77 values.
5064: 
5065: If we can calculate $\tr \pi_Y$ for general 6-braids, then we
5066: can use \eqref{XX} as a linear equation for $c(i,Y)$.
5067: If $\pi_Y(\ap)$ has $l_Y$ different eigenvalues, we can
5068: determine $c(i,Y)$ for that $Y$ by calculating the l.h.s. of
5069: \eqref{XX} for $l_Y$ different values of $k$ and solving for $c(i,Y)$.
5070: (Always $l_Y\le 6$, as evident from table \reference{tab1}.)
5071: 
5072: In case $Y$ is one of the one-hook diagrams like $\Young{4,1,1}$,
5073: then $\pi_Y$ is by p. 354 bottom (a tensor product of a parity,
5074: which disappears at even exponent sum, with) an exterior power
5075: of $\psi_5$. Thus, as in \S 7 (of Jones's paper), one can evaluate
5076: $\tr \pi_Y(\bt_k)$ from the characteristic polynomial of the Burau
5077: matrix $\psi_5(\bt_k)$. We determined this way the corresponding
5078: $c(i,Y)$, and 
5079: verified them with a few extra values of $k$ in \eqref{XX}. While
5080: it is clear that all $c(i,Y)$ should be rational expressions in $q$,
5081: we expected them to be in fact Laurent polynomials. So we were a bit
5082: startled by the denominators $1+q^2$. However, according to p. 343
5083: top, the Hecke algebra may degenerate at roots of unity $q$, which
5084: justifies at least cyclotomic polynomials as denominators.
5085: 
5086: \begin{table}[ptb]
5087: \captionwidth\vsize\relax
5088: \newpage
5089: \vbox to \textheight{\vfil
5090: \rottab{ 
5091: \hbox to \vsize{%\hss%\footnotesize
5092: \hss
5093: % \setbox\@tempboxa=\hbox{%
5094: % \textwidth\vsize\relax
5095: $\displaystyle
5096: \begin{array}{|c||*{11}{c|}}
5097: \hline\ry{1.0cm}%
5098: Y & \Young{1,1,1,1,1,1} & % 1
5099: \Young{2,1,1,1,1} &       % 2
5100: \Young{2,2,1,1} &         % 3
5101: \Young{2,2,2} &           % 4
5102: \Young{3,1,1,1} &         % 5
5103: \Young{3,2,1} &           % 6
5104: \Young{3,3} &             % 7
5105: \Young{4,1,1} &           % 8
5106: \Young{4,2} &             % 9
5107: \Young{5,1} &             % 10
5108: \Young{6} \\[6mm]         % 11
5109: \hline\ry{1.6em}%
5110: %  PARTITION #7 & #8 ARE SWOPPED FROM MY NOTES
5111: %      #1 #2  #3  #4  #5  #6  #8  #7  #9  #10 #11
5112: 1 & 1 & -\left( \frac{q^4}{1 + q^2} \right)  & \frac{q^7}
5113:    {\left( 1 + q^2 \right) \,\left( 1 + q + q^2 \right) } & 0 & 0 & 0 & 0 & \
5114: 0 & 0 & 0 & 0 \\
5115: %
5116: q^4  & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \frac{q}
5117:    {\left( 1 + q^2 \right) \,\left( 1 + q + q^2 \right) } & -\left( \frac{q^2}{1 + q^2} \right)  & q^4 \\
5118: %
5119: q    & 0 & \frac{1}{2} & \frac{-\left( 
5120:        \left( -1 + q \right) \,q \right) }{2} & q + \frac{q^3}{2} & \frac{q^2 + 2\,q^4}{-2 - 2\,q^2} & \frac{q\,\left( 1 - 4\,q + q^2 \right) }
5121:     {2} - c(6,\Young[1.5mm]{ 3,2,1} ) & 0 & \frac{q^6}{2 + 2\,q^2} & \frac{q^6}
5122:    {1 + q + q^2} & 0 & 0 \\[2mm]
5123: %
5124: -q &   0 & \frac{1}{2} & \frac{-\left( 
5125:        \left( -1 + q \right) \,q \right) }{2} & \frac{q^3}{2} & \frac{q^2 + 
5126:      2\,q^4}{-2 - 2\,q^2} & \frac{{\left( 1 - q \right) }^2\,q}{2} - 
5127:    \frac{q^2}{1 + q} - c(6,\Young[1.5mm]{ 3,2,1} ) & 0 & \frac{q^6}
5128:    {2 + 2\,q^2} & 0 & 0 & 0 \\
5129: %
5130: -q^2 &   0 & \frac{1}{1 + q^2} & \frac{q\,
5131:      \left( 1 + \left( -1 + q \right) \,q \right) }{1 + q^2} & -q^2 & 
5132: -q^2 & q\,\left( 1 + \left( -1 + q \right) \,q \right)  & -q^2 & -q^2 & \frac{q^3\,\left( 1 - q + q^2 \right) }{1 + q^2} & \frac{q^6}
5133:    {1 + q^2} & 0 \\[2mm]
5134:    %
5135: q^3 &   0 & 0 & \frac{1}{1 + q + q^2} & 0 & \frac{1}
5136:    {2 + 2\,q^2} & c(6,\Young[1.5mm]{ 3,2,1} ) & \frac{q}{2} + q^3 & \frac{-\left( q^2\,
5137:        \left( 2 + q^2 \right)  \right) }{2\,\left( 1 + q^2 \right) } & \frac{\left( -1 + q \right) \,q^2}{2} & \frac{q^4}
5138:    {2} & 0 \\[3mm]
5139: -q^3 &   0 & 0 & 0 & 0 & \frac{1}{2 + 2\,q^2} & \frac{q^2}{1 + q} + 
5140:    c(6,\Young[1.5mm]{ 3,2,1} ) & \frac{q}{2} & \frac{-\left( q^2\,
5141:        \left( 2 + q^2 \right)  \right) }{2\,\left( 1 + q^2 \right) } & \frac{\left( -1 + q \right) \,q^2}{2} & \frac{q^4}{2} & 0 \\[2mm]
5142: \hline
5143: \end{array}%
5144: $%
5145: % }%
5146: % \@tempdima0.3\vsize\relax\advance\@tempdima by -0.3\wd\@tempboxa\relax
5147: % \kern\@tempdima
5148: % \hss
5149: % \copy\@tempboxa
5150: \hss}%
5151: }{This table shows the values $c(i,Y)$ of \eqref{XX} for Young tableaux
5152: $Y$ of 6 boxes and the 7 possible eigenvalues $\dl_i$ of $\pi_Y(\ap)$.
5153: \label{tab2}}{table}
5154: \vss}
5155: \newpage
5156: \end{table}
5157: 
5158: For $Y=\Young{2,2,2}$, $\pi_Y$ was written down directly on p. 362,
5159: and for its transposed (dual) Young diagram $\Young{3,3}$, one
5160: uses Note 4.6. 
5161: 
5162: There remain 3 representations, for $Y=\Young{3,2,1}$,
5163: $\Young{4,2}$, and its dual $\Young{2,2,1,1}$, with a total of
5164: 15 unknown $c(i,Y)$. These were more complicated to find, since
5165: we knew of no way to evaluate $\tr \pi_Y$ directly. To help
5166: ourselves, first observe that we have, for fixed $Y$, the trace
5167: identities
5168: \begin{eqn}\label{tri}
5169: \sum_{i=1}^7\,c(i,Y)\,=\,\tr \pi_Y([4354])\,=\,
5170: \tr \pi_Y([2132])\,=\,\sum_{i=1}^7 M(i,Y)\dl_i\,,
5171: \end{eqn}
5172: which again give a linear condition on the $c(i,Y)$. To find
5173: further identities, we used \eqref{STR2} in a ``backward'' manner.
5174: We calculated for small\footnote{In the parametrization $P(v,z)$,
5175: used by Morton-Short, unlike for $X$, the coefficients of
5176: $P(\hat\bt_{k,l})$ become quickly large and produce machine
5177: size integer overflows. In particular, we could not calculate
5178: correctly polynomials for $|l|>2$.} $k,l$ ($k$ odd, $|k|\le 5$,
5179: $|l|\le 1$) the polynomial on the left using Morton-Short's
5180: program \cite{MorSho2}. We substituted the known $c(i,Y)$ on
5181: the right of \eqref{STR2}, obtaining thus linear conditions
5182: for the yet unknown $c(i,Y)$. To determine the coefficients,
5183: it remains to understand the effect on $\tr\pi_Y$ of multiplying
5184: with the full twist $\Dl_6^2$. This was, however, done also by
5185: Jones in \S 9, lemma 9.3:
5186: \begin{eqn}\label{l93}
5187: \pi_Y(\Dl_n^2)\,=\,q^{rn(n-1)/d}\,Id_{\pi_Y}\,.
5188: \end{eqn}
5189: For given $Y$, the number $r=r_Y$ is calculated as in lemma 9.1
5190: from figure 3.3, and $d=d_Y=\dim \pi_Y$ more easily by the hook
5191: length formula on p. 341. Call $e_Y=n(n-1)r/d$ (with $n=6$) the
5192: exponent\footnote{This is not to be confused with the variable
5193: $e$, which we use for exponent sum of a braid, or with the
5194: idempotent $e_i$ from Jones's lemma 9.1.} of $q$ on the right
5195: hand-side of \eqref{l93}. These values are given in table
5196: \ref{tab1}. Two simple checks are $\sum\limits_{Y\vdash n}d_Y^2=
5197: n!$ (because the multiplicity of each irrep in the Hecke algebra
5198: equals its dimension), and that $e_Y$ are integers (Remark on p. 358
5199: bottom) and satisfy $e_{Y}+e_{\ol{Y}}=n(n-1)$ for all $Y$
5200: (because of Note 4.6). So \eqref{STR2} gets
5201: \begin{eqn}\label{STR3}
5202: -
5203: (1-q)\cdot\dots\cdot(1-q^6)\,X(\hat\bt_{k,l})(q,\lm)\,=\,
5204: \sqrt{\lm}^{e-5}\,\sum_{Y\vdash 6}\,\wh W_Y(q,\lm)\,
5205: q^{l\cdot e_Y}\,\sum_{i=1}^7\,c(i,Y)\,\dl_i^k\,,
5206: \end{eqn}
5207: with $e=[\bt_{k,l}]=4+4k+30l$.
5208: 
5209: Actually, each polynomial
5210: $X(\hat\bt_{k,l})$ gives 6 equations for $c(i,Y)$, because
5211: there are 6 relevant $\lm$-coefficients on both hand-sides
5212: of \eqref{STR2} (in degrees $\ffrac{e-5}{2}$, $\ffrac{e-3}{2}$,
5213: \dots, $\ffrac{e+5}{2}$\,; it is helpful to multiply again by
5214: $(1+q^2)$ to get disposed of the denominators of the known
5215: $c(i,Y)$). We have with 18 polynomials 111 equations (6 equations per
5216: polynomial plus the three relevant trace equations \eqref{tri}).
5217: Still the resulting system was too hard to solve by computer, using 
5218: MATHEMATI\-CA\TM{} \cite{Wolfram}, since its coefficients are
5219: (Laurent) polynomials in $q$, with dozens of terms each.
5220: 
5221: However, substituting some (rational) values of $q$, the system
5222: can be solved immediately. We used this to check first the rank of
5223: the matrix (i.e. which equations are linearly redundant). Again we
5224: were surprised that for 15 variables $c(i,Y)$ the rank was only
5225: $14$. This, however, can be explained from our restraint to odd
5226: $k$ (which we chose for some, purely technical,
5227: component number concerns in the calculation
5228: with Morton-Short's program). Whenever two opposite eigenvalues
5229: $\dl_i$ and $\dl_{i'}=-\dl_i$ occur in $\pi_Y(\ap)$, the equations
5230: \eqref{STR2} for odd $k$ can detect only $c(i,Y)-c(i',Y)$. The
5231: trace equations \eqref{tri}, which involve $c(i,Y)$ and $c(i',Y)$
5232: with the same sign, remedy the shortcoming for $\Young{4,2}$ and
5233: its transposed diagram, but for $\Young{3,2,1}$ we have two pairs
5234: $(i,i')$ of opposite eigenvalues, so we still lose one dimension.
5235: This is, however, not really a problem, because in the
5236: braids $\gm_a=\bt_{6a-1,-2a}$ of \eqref{gma} we
5237: need for our proof, $k=6a-1$ is always odd, so we need only
5238: $c(i,Y)-c(i',Y)$ to evaluate the polynomial in \eqref{STR2}.
5239: 
5240: We used the special evaluations to select equations that give
5241: the full matrix rank, and to guess the formula for general $q$
5242: for some of the $c(i,Y)$. (The ones we already found suggest that
5243: these formulas should not be so complicated.) Substituting these
5244: (yet potential) solutions too, gives an even simpler linear system
5245: for the still unknown $c(i,Y)$, which then could be solved in $q$.
5246: 
5247: Since we know that our matrix has rank 14, it is enough to check
5248: our solution (up to the 1-dimentional ambiguity, which will
5249: disappear in \eqref{STR2}) with the 111 equations we have. The
5250: result was confirmed, and is shown in table \ref{tab2}.
5251: 
5252: With all $c(i,Y)$ determined, the main work is done.
5253: So far we can evaluate $X(\hat\bt_{k,l})$ for odd $k$.
5254: Multiplying \eqref{STR3} by $Z=(1+q)(1+q^2)(1+q+q^2)$ to clear
5255: all denominators in the $c(i,Y)$, normalizing and taking
5256: coefficients in $\lm^m$ for $m=0,\dots,5$, we have
5257: (with $e_Y$ being the exponents of $q$ in \eqref{l93} and
5258: $e=[\bt_{k,l}]$)
5259: \begin{eqn}\label{OPI}
5260: \left[-\frac{1}{\sqrt{\lm}^{e-5}}\,\cdot\,Z\,\cdot\,
5261: (1-q)\cdot\dots\cdot(1-q^6)\,
5262: X_{\hat\bt_{k,l}}(q,\lm)\,\right]_{\lm^m}\,=\,
5263: \sum_{Y\vdash 6}\,\sum_{i=1}^7\,
5264: (Z\cdot c(i,Y))\,\dl_i^k\,
5265: q^{l\cdot e_Y}\,\bigl[\wh W_Y(q,\lm)\,\bigr]_{\lm^m}\,.
5266: \end{eqn}
5267: 
5268: Recall that for our proof it is enough to show that this term
5269: becomes non-zero for $m=0$ or $m=1$, when $k=6a-1$ and $l=-2a$
5270: for an integer $a>1$ (and $\bt_{k,l}=\gm_a$ in \eqref{gma}).
5271: Now, for odd $k$ (and fixed $Y$), we can group
5272: the sum over $7$ terms $Z\cdot c(i,Y)$ into 5 terms $\tl c(i,Y)$
5273: accounting for $\dl_3=-\dl_4$ and $\dl_6=-\dl_7$, and thus
5274: excluding $i=4,7$. So the above sum in \eqref{OPI} becomes
5275: \begin{eqn}\label{yy}
5276: \sum_{Y\vdash 6}\,\sum_{\scbox{\shortstack{$i=1$\\$i\ne 4$}}}^6
5277: \tl c(i,Y)\dl_i^k\,q^{l\cdot e_Y}\,\bigl[\wh W_Y(q,\lm)\,\bigr]
5278: _{\lm^m}\,.
5279: \end{eqn}
5280: (Now in the $\tl c(i,Y)$, the 1-degree ambiguity of $c(i,Y)$ for
5281: $Y=\Young{3,2,1}$, as explained above, cancels out.)
5282: 
5283: Among the 55 possible $(i,Y)$ (with $i\ne 4,7$), only 21
5284: of the $\tl c(i,Y)$ are non-zero. It turns out that for $m=0$,
5285: when $k=6a-1$, $l=-2a$ and $a>1$, there is a unique term among
5286: the 21 summands in \eqref{yy} whose minimal degree in $q$ is the
5287: smallest (it is $14-36a$). Thus $[P(\hat\bt_{k,l})]_{v^{e-5}}\ne 0$.
5288: We calculated the polynomial with Morton-Short's program for $a=2$
5289: (where the calculation was still feasible), and it confirmed that
5290: all 6 $v$-terms appear. We also calulated that for $m=0$ and
5291: $a=0,1$ there are two terms of smallest minimal degree. This
5292: better ought to be so, because in that case $\hat\gm_{a}$ are just
5293: a 2-cable of the unknot and $!5_1$, resp., and the coefficients
5294: for $m=0,1$ in \eqref{yy} must be $0$, which we checked once
5295: more separately. (Also we found that for $m=1$ there are two
5296: terms of smallest minimal degree in \eqref{yy} for all $a\ge 0$.)
5297: With this the proof of theorem \reference{thJC} is complete.
5298: 
5299: Let us finally say that the computer part of the calculation owed
5300: a lot to the use of MATHEMATI\-CA. While, if done properly, it could
5301: be carried out in a few minutes, it required a week of work to
5302: find the way of skillfully programming MATHEMATI\-CA
5303: to do all the separate steps in an efficient way. \qed
5304: 
5305: \begin{rem}
5306: Note that one could handle the cases $\mu=\sg_1^{6a+1}\sg_2^{-1}$
5307: from the beginning of the proof also using the representation
5308: theoretic argument, by looking at $m=4,5$ for $a<0$. We waived on
5309: this investigation, though, since the proof for $a>0$ was laborious
5310: enough.
5311: \end{rem}
5312: 
5313: \subsection{Unitarity of the Burau representation\label{SUn}}
5314: 
5315: In the following $q$ and $t$ are unit norm complex numbers.
5316: We define $\arg (e^{is}):=s\bmod 2\pi$ for $s\in\bR$.
5317: We continue using the formalism of Young tableaux,
5318: the representations $\pi_Y$ and the notations $X(q,\lm)$
5319: and $W_Y$ of the proof of theorem \reference{thJC}.
5320: 
5321: For the Alexander polynomial $\Dl$ (we resume the
5322: notational convention from before \S\ref{SJS}), as well as
5323: for $3$- and $4$-braids, the
5324: representations $\pi_Y$ are given by Burau representations. We note
5325: (again, and more explicitly) the following descriptions of $\pi_Y$
5326: in terms of the Burau representation $\psi_n$ given in \cite{Jones}.
5327: Again the indexing is chosen so that $\psi_{n-1}$ is the reduced
5328: $(n-1)$-dimensional representation of $B_n$, and by $-\rho$ we denote
5329: the direct product of $\rho$ with the parity representation.
5330: 
5331: \Youngunitlength1.5mm
5332: As before, $e$ stands for the exponent sum of a braid $\bt$.
5333: For $3$-braids we have the following properties (with
5334: reference to the explanation in \cite{Jones}):
5335: \begin{enumerate}
5336: \item $\pi_{\Young{2,1}}(\bt)(q)=(-1)^e\psi_2(\bt)(q)=
5337: q^e\psi_2(\bt^{-1})(q)$ (because of
5338: row-column symmetry; see Note 4.6).
5339: \item $\pi_{\Young{1,1,1}}(\bt)=(-1)^e$ and
5340: $\pi_{\Young{3}}(\bt)=q^e$ (Note 4.7)
5341: \end{enumerate}
5342: 
5343: For $4$-braids we have:
5344: \begin{enumerate}
5345: \item $\pi_{\Young{2,1,1}}=-\psi_3$. So
5346: $\pi_{\Young{2,1,1}}(\bt)=(-1)^e\psi_3(\bt)$. (Note 5.7)
5347: \item $\pi_{\Young{3,1}}(\bt)(q)=q^e\psi_3(\bt^{-1})(q)$ (because of
5348: row-column symmetry; see Note 4.6). Also 
5349: \[
5350: \pi_{\Young{3,1}}=-\pi_{\Young{2,1,1}}\wedge
5351: \pi_{\Young{2,1,1}}=-\psi_3\wedge \psi_3\,,
5352: \]
5353: where wedge denotes antisymmetric product (see p.354 bottom).
5354: \item $\pi_{\Young{2,2}}(\bt)=(-1)^e\psi_2(\bar\bt)$, where bar denotes
5355: the homomorphism from $B_4$ to $B_3$ given by $\bar\sg_{1,2,3}=\sg_{1,
5356: 2,1}$. (p. 355)
5357: \item $\pi_{\Young{1,1,1,1}}(\bt)=(-1)^e$ and
5358: $\pi_{\Young{4}}(\bt)=q^e$ (Note 4.7)
5359: \end{enumerate}
5360: 
5361: Now Squier observes in \cite{Squier}, that $\psi_i(\bt^{-1})(q)$
5362: and $\psi_i(\bt)(q^{-1})$ are conjugate and so have the same
5363: trace. So by the self-symmetry of $\pi_{\Young{2,1}}$ we have
5364: $\tr \pi_{\Young{2,1}}(t), \tr \pi_{\Young{2,2}}(t)\in (-t)^{e/2}\bR$.
5365: Similarly $(-t)^{-e/2}\tr \pi_{\Young{2,1,1}}(t)$ and $(-t)^{-e/2}\tr
5366: \pi_{\Young{3,1}}(t)$ are conjugate complex numbers. 
5367: These properties will be important below.
5368: 
5369: \begin{rem}
5370: Squier uses a different convention for
5371: $\psi_i$ from Jones. He transposes and changes sign in
5372: matrix entries with odd row-column sum (i.e., conjugates by
5373: $\dig(1,-1,1,-1,\dots)$). This, however, does not affect
5374: our arguments. 
5375: \end{rem}
5376: 
5377: The key point in arguments below is Squier's result.
5378: We write $M^*$ for the conjugate transposed of a matrix $M$.
5379: (That is, $M^*_{i,j}=\ol{M}_{j,i}$.)
5380: 
5381: \begin{theorem}(Squier \cite{Squier})
5382: For any $n\ge 1$ there exists a Hermitian matrix $J=J^{[n+1]}$ and a
5383: regular matrix $M$, such that with $J_0=M^*JM$ we have $\psi_n^*J_0
5384: \psi_n=J_0$.
5385: \end{theorem}
5386: 
5387: In particular, $J$ is degenerate or definite iff $J_0$ is so.
5388: Moreover,
5389: \[
5390: J_{i,j}\,=\,\left\{\,\begin{array}{c@{\quad}l}
5391: -1 & \mbox {if\ $|i-j|=1$} \\
5392: \sqrt{t}+1/\sqrt{t} & \mbox {if\ $i=j$} \\
5393: 0 & \mbox {otherwise}
5394: \end{array}\right.\,.
5395: \]
5396: 
5397: It is easy to see that if $t=1$, then $J$ is positive definite.
5398: Now definiteness is an open condition, so for $t$ close to $1$,
5399: it is still valid. 
5400: % By using the criterion of positivity of the principle minors,
5401: One can determine when $J$ loses this
5402: property. % in simple cases. It is easy to verify that for $3$-braids
5403: % this occurs in $e^{\pm 2\pi i/3}$, while for $4$-braids in $\pm i$.
5404: % In general we have
5405: 
5406: \begin{prop}\label{p4.1}
5407: The Squier form $J^{[n]}$ on $B_n$ degenerates exactly in the $n$-th
5408: roots of unity. In particular, it is positive definite exactly
5409: when $|\arg t|<2\pi /n$.
5410: \end{prop}
5411: 
5412: \proof Denote by $J^{[n]}$ the form corresponding to $n$-braids,
5413: i.e. the one given by restricting $J$ to the first $n-1$
5414: rows and columns. It is not too hard to calculate the determinant
5415: of $J^{[n]}$. By development in the last row,
5416: \[
5417: \det J^{[n]}\,=\,(\sqrt{t}+1/\sqrt{t})\det J^{[n-1]}-\det J^{[n-2]}\,,
5418: \]
5419: whence
5420: \[
5421: \det J^{[n]}\,=\,\frac{t^n-1}{\sqrt{t}-1/\sqrt{t}}\,\cdot\,
5422: \frac{1}{(\sqrt{t})^n}\,.
5423: \]
5424: Then the claim follows easily. To see definiteness use the
5425: (positivity of) the principal minor criterion. Since $e^{\pm 2\pi i/n}$
5426: is a simple zero of $\det J^{[n]}$, the determinant must become
5427: negative for $|\arg t|\in(2\pi/n,4\pi/n)$. Then applying this
5428: argument to all $n'<n$ shows that $J^{[n]}$ is not positive definite
5429: for $|\arg t|\ge 2\pi/n$. \qed
5430: 
5431: 
5432: So on the arcs of $S^1$ that connect the primitive $n$-th root of
5433: unity to $1$, we have that $J$ is positive definite. Now
5434: if $J$ is such, it can be written as $Q^*Q$, and then
5435: conjugating $\psi_i$ by $QM$ we obtain a $U(n-1)$-representation.
5436: This means in particular that all eigenvalues of $\psi_i$
5437: have unit norm. We will below derive implications of this
5438: circumstance for the link polynomials.
5439: 
5440: \subsection{Norm estimates\label{s4.2}}
5441: 
5442: The Jones polynomial $V$ can be specified, for our
5443: purposes, by $V(t)=X(t,t)$. In the following, which
5444: root of complex numbers is taken is irrelevant,
5445: important is though that it be kept fixed in subsequent
5446: calculations. By $\Re$ we denote the real part of
5447: a complex number.
5448: 
5449: \begin{theorem}\label{th1}
5450: If $|t|=1$, $\Re t>0$ and $\bt$ is a $4$-braid, then 
5451: $\big |V_{\hat\bt}(t)\big|\,\le\,(2\Re \sqrt{t})^3$.
5452: If $\bt$ is a $3$-braid and $\Re t>-1/2$, then
5453: $\big |V_{\hat\bt}(t)\big|\,\le\,(2\Re \sqrt{t})^2$.
5454: \end{theorem}
5455: 
5456: \proof We have from \cite{Jones} that if $\bt\in B_4$ with
5457: $[\bt]=e$, then
5458: \begin{eqn}\label{Vf}
5459: V_{\hat\bt}(t)\,=\,\left(-\sqrt{t}\right)^{e-3}
5460: \left[\,\frac{t(1-t^3)}{1-t^2}\,\tr \psi_3+\frac{t^2}{1+t}
5461: \tr \bar\psi_2\,+\,\frac{1-t^5}{1-t^2}\,\right]\,,
5462: \end{eqn}
5463: where $\bar\psi_2$ is the composition of $\psi_2$ with
5464: $\bar{ }\,:\,B_4\to B_3$. Taking norms and using that
5465: $\bar\psi_2$ and $\psi_3$ are unitary, we find
5466: \begin{eqn}\label{X}
5467: \big |V(t)\big|\,\le 3\left|\frac{1-t^3}{1-t^2}\right|+
5468: \frac{2}{|1+t|}+\left|\frac{1-t^5}{1-t^2}\right|\,.
5469: \end{eqn}
5470: It is now a routine (but somewhat tedious) calculation to
5471: verify that the r.h.s. is equal to $(2\Re \sqrt{t})^3$
5472: for $|t|=1$, $\Re t>0$. 
5473: 
5474: For $\bt\in B_3$ we have similarly
5475: \begin{eqn}\label{V_3}
5476: V_{\hat\bt}(t)\,=\,\left(-\sqrt{t}\right)^{e-2}
5477: \left[\,t\cdot \tr\psi_2\,+\,(1+t^2)\,\right]\,,
5478: \end{eqn}
5479: and the result follows using $|\tr\psi_2|\le 2$. \qed
5480: 
5481: This theorem generalizes Jones's result \cite[proposition 15.3]{Jones}
5482: for $n\le 4$, where he considers $t=e^{2\pi i/k}$, $k\ge 5$. In fact
5483: the comparison to (and established coincidence with) Jones's
5484: estimate led to the simplification of the r.h.s. of \eqref{X}.
5485: In \cite{posex_bcr} we noted that Jones's estimate can be
5486: better than MWF when $MWF=3$, but by connected sum one
5487: can give an example for $MWF=4$. Again it appears that for
5488: $\bt\in B_4$ and $t\ne e^{\pm \pi i/3}$ the set $\{\,|V_{\hat
5489: \bt}(t)|\,\}$ is dense in $[0,(2\Re \sqrt{t})^3]$, and similarly it
5490: is in $[0,(2\Re \sqrt{t})^2]$ for $\bt\in B_3$ and $t\ne e^{\pm \pi
5491: i/3},e^{\pm \pi i/5}$. (See also the remarks at the end of \S 12
5492: in \cite{Jones}.)
5493: 
5494: \begin{conjecture}
5495: If $\bt\in B_n$ and $|t|=1$ with $|\arg t|<2\pi /n$ then
5496: $|V(t)|\,\le\,(2\Re \sqrt{t})^{n-1}$.
5497: \end{conjecture}
5498: 
5499: In the case of the Alexander polynomial $\Dl(t)=X(t,1/t)$, we
5500: can say something on general braids.
5501: 
5502: \begin{theorem}
5503: For each $n\ge 2$, %there is an $\eps_n>0$ such that
5504: if $|t|=1$ and $|\arg t|\le 2\pi/n$ and $\bt\in B_n$ then
5505: \[
5506: \big |\Dl_{\hat\bt}(t)\big|\,\le\,\frac{2^{n-1}|1-t|}{|1-t^n|}\,.
5507: \]
5508: \end{theorem}
5509: 
5510: \proof
5511: % Let $\eps_n>0$ be so that $\ds\frac{|1-t^n|}{|1-t|}>n-0.1$ and
5512: $J_0$ is positive definite when $|t|=1$ and $|\arg t|<2\pi/n$.
5513: Then
5514: \[
5515: \big |\Dl_{\hat\bt}(t)\big|\,\frac{|1-t|}{|1-t^n|}\,=\,
5516: \big |\det (1-\psi_{n-1}(\bt))\big|\,,
5517: \]
5518: and all eigenvalues of $1-\psi_{n-1}$ have norm $\le 2$.
5519: The case $|\arg t|=2\pi/n$ follows by continuity. \qed
5520: 
5521: \begin{corr}
5522: For all $n,k$ the set
5523: \[
5524: \big\{\,\Dl(\hat\bt)\,:\,\bt\in B_n\,,\ \ \deg\Dl\le k\,\big\}
5525: \]
5526: is finite. That is, among closed braids of given number of strands
5527: only finitely many Alexander polynomials of given degree occur.
5528: \end{corr}
5529: 
5530: \proof $\Dl$ is determined by $\Dl(t_i)$ for $k$ different
5531: $t_i$ with $|t_i|=1$, $0<\arg t_i<2\pi/n$, by means of a 
5532: linear transformation using the (regular Vandermonde)
5533: matrix $M=(t_i^j)_{i,j=1}^n$.
5534: So $||[\Dl(t)]_{t^j}||\,\le\,||M^{-1}||\cdot ||\Dl(t_i)||$\,.
5535: \qed
5536: 
5537: This result should be put in contrast to the various constructions
5538: of knots with any given Alexander polynomial. For example
5539: a recent construction of Nakamura \cite{Nakamura2} allows to
5540: realize the degree of the polynomial by the (actually braidzel)
5541: genus of the knot. (We were subsequently independently
5542: able to further specialize this result to canonical genus.) 
5543: % do not know the following 
5544: % 
5545: % \begin{question}
5546: % Is any Alexander polynomial admitted by a knot with canonical
5547: % genus equal to the degree of the polynomial?
5548: % \end{question}
5549: % 
5550: A different construction of Fujii 
5551: \cite{Fujii} shows that knots with 3 bridges admit all
5552: Alexander polynomials. So the situation between braid and
5553: plat closures is completely different.
5554: 
5555: Compare also Birman-Menasco's result in \cite{BirMen3}, mentioned
5556: (for knots) in remark \ref{rfn}, that there are only finitely
5557: many closed braids of given number of strands with given genus. 
5558: Note that we do not claim that only finitely many 
5559: closed braids of given number of strands with given Alexander
5560: polynomial (degree) occur. For $3$-braids it is true, but
5561: from $5$-braids on the non-faithfulness of the Burau
5562: representation should (in principle, modulo the evaluability
5563: of another invariant) make it possible to construct
5564: infinite families of links with the same (for example, trivial)
5565: polynomial. It makes some sense to ask about the status of 4-braids.
5566: 
5567: \begin{question}
5568: Are there only finitely many 
5569: closed 4-braids of given Alexander polynomial (degree)?
5570: \end{question}
5571: 
5572: When working with $\Dl$, for $3$- and $4$-braids we can be
5573: more explicit.
5574: 
5575: \begin{corr}\label{pp1}
5576: If $\bt\in B_4$, then
5577: \begin{eqn}\label{Xx}
5578: \big|\,\Dl_{\hat\bt}(t)\,\big|\,\le\,\frac{8\,|1-t|}{|1-t^4|}\,,
5579: \end{eqn}
5580: when $|t|=1$ and $\Re t>0$. If $\bt\in B_3$ and $\Re t>-1/2$, then
5581: \begin{myeqn}{\qed}
5582: \big|\,\Dl_{\hat\bt}(t)\,\big|\,\le\,\frac{4\,|1-t|}{|1-t^3|}\,.
5583: \end{myeqn}
5584: \end{corr}
5585: 
5586: Putting $t=e^{2\pi i/5}$ in \eqref{Xx} we have $\big|\,\Dl(e^{2\pi
5587: i/5})\,\big|\,\le\,8\,.$ This improves the bound $10.47\dots$ in
5588: \cite{posex_bcr} suggested to replace Jones's (incorrect) value 6.5
5589: in \cite{Jones2}. For 3-braid \em{knots} Jones gives in
5590: \cite[proposition 15.2]{Jones} the better bound 3 when $t=i$, using
5591: the property $V(i)=\pm 1$.
5592: 
5593: \begin{exam}\label{4249}
5594: The simplest knots with $MWF=4$ which can be excluded from being a
5595: 4-braid using corollary \reference{pp1} are $13_{8385}$ (where one
5596: can use $t=e^{2\pi i/9}$) and $14_{37492}$ (with $t=e^{2\pi i/8}$).
5597: For $3$-braids we can deal with the known examples $9_{42}$
5598: and $9_{49}$. (This gives now an alternative proof that
5599: there are no 3-braid knots with such Alexander polynomial, as
5600: we explained in example \reference{x29}.)
5601: \end{exam}
5602: 
5603: If we know the exponent sum of $\bt$ we can do better.
5604: 
5605: \begin{prop}\label{pp2}
5606: If $\bt\in B_4$, $e=[\bt]$ and $t$ as before, then
5607: \[
5608: \left|\left(
5609: -\sqrt{t}\right)^{e-3}\Dl_{\hat\bt}(t)\frac{1-t^4}{1-t}-1
5610: +(-t)^e\,\right|\,\le 6.
5611: \]
5612: If $\bt\in B_3$ and $\Re t>-1/2$, then
5613: \begin{myeqn}{\qed}
5614: \left|\left(
5615: -\sqrt{t}\right)^{e-2}\Dl_{\hat\bt}(t)\frac{1-t^3}{1-t}-1
5616: -(-t)^e\,\right|\,\le 2\,.
5617: \end{myeqn}
5618: \end{prop}
5619: 
5620: \begin{exam}\label{lxd}
5621: Since one can determine the possible $e$ via MWF from $P$,
5622: one can apply proposition \reference{pp2} for given $P$. In
5623: the $3$-braid case we can exclude $10_{150}$ this way. The
5624: remaining two 10 crossing knots with unsharp MWF, $10_{132}$
5625: and $10_{156}$, fail~-- understandably, since they share the
5626: skein polynomials of (the closed 3-braids) $5_1$ and $8_{16}$
5627: resp. (See the table\footnote{The second duplication was
5628: noted in the remarks after Jones's table, but not referred
5629: to correctly in its last column.} in \cite{Jones}.) For
5630: $4$-braids several new 14 crossing knots can be ruled
5631: out, for example $14_{21199}$.
5632: \end{exam}
5633: 
5634: 
5635: \begin{rem}
5636: The use of $P$ to restrict the possible values of $e$ is
5637: usually most effective, but not indispensable. There are
5638: other conditions on $e$, originating from Bennequin's
5639: work \cite{Bennequin}, that can be more applicable in
5640: certain cases where the calculation of $P$ is tedious.
5641: Also, when $t$ is a root of unity of order $n$, the tests
5642: depend only on $e\bmod 2n$.
5643: \end{rem}
5644: 
5645: \begin{rem}
5646: Note that the quantity $\chi$ in \eqref{chi} is equal to $u\,
5647: \tr \psi_2$, where $u=\sqrt{-t}+\ds\frac{1}{\sqrt{-t}}$ and
5648: $\psi_2$ is the Burau matrix of $B_3$ (see the explanation in
5649: \cite{Kanenobu}). Thus one can obtain
5650: similar estimates for values of $Q$ on 3-braids.
5651: \end{rem}
5652: 
5653: 
5654: \subsection{Skein polynomial}
5655: 
5656: Now it is natural to look at the full 2-variable skein polynomial
5657: $X$. We have, as in \eqref{STR}, for $\bt\in B_4$ of exponent
5658: sum $e$,
5659: \[
5660: X_{\hat\bt}(q,\lm)\,=\,-\sqrt{\lm}^{e-3}\,
5661: \sum_{Y\vdash 4}\,\tr\pi_Y(q)\,\tW_Y(q,\lm)\,,
5662: \]
5663: where the weights $\tW_Y$ are given in $\lm$-coefficients by
5664: table \reference{table1}. (They are all polynomials in $\lm$ of
5665: degree $3$, with coefficients being rational expressions in
5666: $q$.) Now, with given $e$ and $P$, we have $4$ equations
5667: (the coefficients in $\lm$) in $3$ unknowns (the traces of
5668: $\pi_{211}$, $\pi_{31}$ and $\pi_{22}$; we use here the
5669: partition notation for the subscripts). However, the
5670: restriction of the matrix in table \reference{table1} to
5671: the columns of $\pi_{211}$, $\pi_{31}$ and $\pi_{22}$
5672: has rank $2$. This means that two of the $X_i=[X]_{\lm^
5673: {i+(e-3)/2}}$ for $0\le i\le 3$ determine the other two.
5674: One could believe now to use
5675: this as a 4-braid test. However, these two relations
5676: result from the general substitutions $\lm=1$ and $\lm=1/q^2$
5677: that turn $X$ into the component parity count or $1$.
5678: These substitutions kill all trace weights except of the trivial
5679: or parity representation, and for these representations the weights
5680: become also independent on the braid group. Thus the
5681: relations between the $X_i$ will hold whenever $MWF\le 4$,
5682: and are useless as a 4-braid test. In a similar vein, one has 
5683: 
5684: \newbox\@tempboxb
5685: \setbox\@tempboxb=\hbox{
5686: \capt The weights of the traces contributing to the
5687: $\lm$-coefficients of $X$ for a $4$-braid. Each table entry
5688: must be multiplied with the factor on the right in the
5689: first row to obtain the contribution of the corresponding trace
5690: to $-\sqrt{\lm}^{e-3}$ times the power of $-\lm$ in the
5691: first column. In the first row the symbol $[i]$ denotes $1-q^i$.}
5692: 
5693: \begin{table}
5694: \[
5695: \begin{array}{c*5{|c}}
5696: & 
5697: \ds \tW_{1111}=\frac{1}{[2][3][4]}\times &
5698: \ds \tW_{211}=\frac{q}{[2][1][4]}\times &
5699: \ds \tW_{22}=\frac{q^2}{[2][2][3]}\times &
5700: \ds \tW_{31}=\frac{q^3}{[2][1][4]}\times &
5701: \ds \tW_{4}=\frac{q^6}{[2][3][4]}\times \\[5mm]
5702: \hline
5703: \ry{6mm}
5704: \ds 1 & 1 & 1 & 1 & 1 & 1 \\[1mm]
5705: -\lm  & q^2+q^3+q^4 & 1+q^2+q^3   & 1+q+q^2   & 1+1/q+q^2 & 1+1/q+1/q^2 \\[2mm]
5706: \lm^2 & q^5+q^6+q^7 & q^3+q^2+q^5 & q+q^2+q^3 & 1/q+q+q^2 & 1/q+1/q^2+1/q^3 \\[2mm]
5707: -\lm^3& q^9         & q^5         & q^3       & q         & 1/q^3 \\[2mm]
5708: \end{array}
5709: \]
5710: \caption{\unhbox\@tempboxb
5711: \label{table1}}
5712: \end{table}
5713: 
5714: \begin{prop}\label{pp7.6}
5715: If a braid $\bt$ has $MWF(\bt)\le 4$, then $V(\hat\bt)$ and
5716: $\Dl(\hat\bt)$ together with the exponent sum $[\bt]=e$,
5717: determine $P(\hat\bt)$.
5718: \end{prop}
5719: 
5720: \proof From $e$, $V$ and $\Dl$ we know
5721: \[
5722: \tl X_a\,=\,\sum_{i=0}^3\,X_i(q)\,q^{a(i+(e-3)/2)}\,,
5723: \]
5724: for $a\in\{-2,-1,0,1\}$. (For $a=-2$ and $a=0$ we have the
5725: trivializing substitutions, $a=-1$ corresponds to $\Dl$ and
5726: $a=1$ to $V$.) So one can recover $X_i$ from $\tl X_a$
5727: and $e$. (One can write down an explicit formula easily.) \qed
5728: 
5729: This condition is thus equally unhelpful as a 4-braid
5730: test. For a similar reason, I expect (though I have not
5731: rigorously derived) an explanation of the (experimentally
5732: observed) failure of Jones's conditions \cite[\S 8]{Jones}
5733: (see formula (8.10)) to obstruct to a 4-braid.
5734: 
5735: \begin{exam}
5736: The knot $11_{386}$, known from \cite{LickMil}, has the
5737: Jones polynomial of the figure-8-knot. So $11_{386}$ and its mirror
5738: image show that the dependence of $P$ on $e$ in proposition
5739: \reference{pp7.6} is essential.
5740: \end{exam}
5741: 
5742: \begin{question}
5743: Are there 5-braids $\bt_{1,2}$ (at least one of which has 
5744: $MWF=5$), with the same exponent sum,
5745: $V(\hat\bt)$ and $\Dl(\hat\bt)$, but different $P(\hat\bt)$?
5746: \end{question}
5747: 
5748: The lack of such examples, after some check in the knot tables,
5749: is at least not fully explainable. Only 6-braids could be found.
5750: 
5751: \begin{exam}
5752: The knots $16_{443392}$ and $!15_{223693}$ have the same $V$
5753: and $\Dl$ but different $P$ polynomial. They have $MWF=5$ resp. $4$,
5754: with $5$- resp. $4$-braid representations of exponent sum $2$ resp.
5755: $-1$, so one obtains 6-braids of exponent sum $1$ by stabilization.
5756: \end{exam}
5757: 
5758: One can now, as before, go over to the norms in each
5759: row in table \reference{table1}, or of arbitrary linear
5760: combinations of such rows. Then for $q$ where $J_0$ is
5761: definite we obtain again estimates on $|X_i(q)|$, or of
5762: $|X(q,\lm)|$ for any non-zero complex number $\lm$. (In
5763: particular, for $\lm=q$ we obtain theorem \reference{th1} and
5764: for $\lm=1/q$ corollary \reference{pp1}.) Although they still
5765: contain the Jones and Alexander polynomial conditions (which
5766: were observed both non-trivial in comparison to MWF), such
5767: skein polynomial norm estimates have not proved in practice,
5768: as a 4-braid test, an efficient improvement over their special
5769: cases and MWF. Below we will explain how to do much better.
5770: 
5771: Knots like $10_{132}$ and $10_{156}$ in example \ref{lxd} show
5772: a disadvantage of our test, resulting from not taking into account
5773: information of other invariants. On the opposite hand, the
5774: mere use of $\Dl$ or $P$ reduces calculation complexity, and
5775: excludes any potential further knot with such polynomial.
5776: In the case of $P$ this gives yet a different way to
5777: answer negatively Birman's question if one can realize any
5778: skein polynomial by a link making MWF sharp (see \cite
5779: {posex_bcr,3br}).
5780: 
5781: Using the Brandt-Lickorish-Millett-Ho polynomial, Murakami
5782: \cite{Murakami} and later Kanenobu \cite{Kanenobu} gave with
5783: theorem \reference{TMF} a more
5784: efficient (in excluding examples, though less in calculation
5785: complexity) test for a 3-braid. But the work in \cite{3br} and
5786: in the previous sections of this paper makes the study of 
5787: polynomials of 3-braids anyway less relevant. $4$-braids
5788: become much more of interest, and in this case, after MWF,
5789: the problem to find applicable conditions has been largely
5790: unsettled for quite a while.
5791: 
5792: \subsection{Recovering the Burau trace}
5793: 
5794: \subsubsection{Conditions on the eigenvalues}
5795: 
5796: Using just norms clearly weakens the conditions considerably,
5797: and so one would like to identify the Burau eigenvalues directly.
5798: However, the relations between the $X_i$ do not allow to recover 
5799: by simple algebraic means the individual traces from $P$. 
5800: 
5801: Using Squier's unitarity, there is an analytic way to recover
5802: the Burau trace of 4-braids (at least for generic $t$ and
5803: up to finite indeterminacy). Since by the previous remarks
5804: the use of $P$ is not essential, we will describe the
5805: procedure given $\Dl$, $V$ and $e$. This gives the most
5806: significant practical enhancement to the above 4-braid tests.
5807: 
5808: In the following we fix a 4-braid $\bt$ of exponent sum $e$, whose
5809: closure link has Jones polynomial $V$ and Alexander polynomial $\Dl$.
5810: We let $t$ be a unit norm complex number with non-negative real part.
5811: 
5812: We use that
5813: \[
5814: \tr \psi_3\wedge \psi_3 = (-t)^{e/2}\ol{\tr \psi_3(-t)^{-e/2}}\,.
5815: \]
5816: Then we have
5817: \[
5818: \Dl(t)\,=\,\frac{1-t}{1-t^4}\,\left(-\frac{1}{\sqrt{t}}\right)^{e-3}
5819: \cdot\,\left[ 1-\tr \psi_3 + (-t)^{e/2}\ol{\tr \psi_3(-t)^{-e/2}}-
5820: (-t)^e\,\right]\,.
5821: \]
5822: So, with $i=\sqrt{-1}$,
5823: \[
5824: \dl\,:=-\frac{1}{2i}(-t)^{-e/2}
5825: \left[ \Dl(t)\frac{1-t^4}{1-t}\left(-\frac{1}{\sqrt{t}}\right)^{3-e}
5826: -1 + (-t)^e\,\right]
5827: \]
5828: is a real number. Now when $\lm_{1,2,3}$ are the eigenvalues of
5829: $\psi_3(\bt)$, we have
5830: \begin{eqnarray*}
5831: A & = & \lm_1+\lm_2+\lm_3\,=\,(-t)^{e/2}(y+i\dl) \\
5832: B & = & \lm_1\lm_2+\lm_1\lm_3+\lm_2\lm_3\,=\,(-t)^{e/2}(y-i\dl) \\
5833: C & = & \lm_1\lm_2\lm_3\,=\,(-t)^e\,.
5834: \end{eqnarray*}
5835: Here $y$ is a real number we do not know, and we try to determine.
5836: Since $|\lm_k|=1$, we have
5837: as before $|A|\le 3$ and $|B|\le 3$. So the range
5838: for $y$ is $[-y_0,y_0]$ with $y_0=\sqrt{9-\dl^2}$.
5839: (If $|\dl|>3$ we are done as before.) Then
5840: \[
5841: \tr\psi_3\,\in\,[(-t)^{e/2}(-y_0+i\dl),\, (-t)^{e/2}(y_0+i\dl)]\,,
5842: \]
5843: with the interval understood lying in $\bC$.
5844: Let $\psi_-$ and $\psi_+$ be the endpoints of this interval.
5845: 
5846: Now one can restrict the interval $[-y_0,y_0]$ for $y$ using the
5847: Jones polynomial as follows.
5848: 
5849: Let $\rho\,:=(-t)^{-e/2}\tr\bar\psi_2\in [-2,2]$. The restriction
5850: to the given range follows because $J^{[3]}$ (see the proof of
5851: proposition \reference{p4.1}) is also definite when
5852: $J^{[4]}$ is. From \eqref{Vf} we have
5853: \[
5854: \tr \psi_3\,=\,\frac{1-t^2}{t(1-t^3)}\,\left[\,
5855: V(t)\,\left(-\frac{1}{\sqrt{t}}\right)^{e-3}-\frac{1-t^5}{1-t^2}
5856: -\frac{t^2}{1+t}(-t)^{e/2}\cdot \rho\,\right]\,=:\,\tl\psi(\rho)\,,
5857: \]
5858: so
5859: \[
5860: \tl\psi(\rho)=\frac{1-t^2}{t(1-t^3)}\left [
5861: V(t)\,\left(\frac{-1}{\sqrt{t}}\right)^{e-3}-
5862: \frac{1-t^5}{1-t^2}\right] -\frac{\rho}{1+2\Re t}(-t)^{e/2}\,.
5863: \]
5864: (Just for the purpose of defining $\tl\psi$, we should regard here
5865: $\rho$ as a formal parameter, rather than as a concrete value.)
5866: 
5867: Let $\tl\psi(\pm 2)=:\tl\psi_\pm$. Since $1+2\Re t>0$, we have
5868: $\Re((-t)^{-e/2}\tl\psi_+)>\Re((-t)^{-e/2}\tl\psi_-)$.
5869: 
5870: Then $[\tl\psi_-,\tl\psi_+]\subset \bC$ is an interval
5871: of the same slope as $[\psi_-,\psi_+]$, so we check if they overlap.
5872: 
5873: Let
5874: \[
5875: \tl y_{\pm}= (-t)^{-e/2}\tl\psi_\pm -i\delta.
5876: \]
5877: 
5878: Then for a consistent restriction on $\tr \psi_3$ the following holds:
5879: \begin{enumerate}
5880: \item $\tl y_{\pm}$ are real
5881: \item $ \tl y_+\tl y_-\le 0$ or $\min(\tl y_\pm^2+\delta^2)\le 9$.
5882: \end{enumerate}
5883: 
5884: Potentially these conditions may be violated, but in
5885: practice they seem always to hold. (We have not elaborated on why
5886: this is so, though it may be worth understanding.) Then at least,
5887: we consider 
5888: \begin{eqn}\label{qw0}
5889: y\in [-y_0,y_0]\cap [\tl y_-,\tl y_+]\,.
5890: \end{eqn}
5891: 
5892: We have now the cubic
5893: \[
5894: x^3 +ax^2 +bx +c \,:=\,x^3-A x^2+B x-C \,=\,0\,.
5895: \]
5896: One solution is obtained by Cardano's formula
5897: \begin{eqn}\label{qw}
5898: \lm_1\,=\,-\frac{a}{3}
5899: -\frac{\sqrt[3]{2}(-a^2+3b)}{3\Gm} + \frac{\Gm}{3\sqrt[3]{2}}\,,
5900: \end{eqn}
5901: where 
5902: \begin{eqn}\label{Gm}
5903: \Gm=\sqrt[3]{
5904:   -2a^3 + 9ab - 27c + 
5905:   \sqrt{
5906:     27\,\left( -a^2b^2 + 4b^3 + 4a^3c -18abc + 27c^2
5907:     \right)
5908:   }
5909: }\,.
5910: \end{eqn}
5911: We must have that $|\lm_1|=1$. Then we must check $|\lm_{2,3}|=1$.
5912: For this their exact determination is not necessary.
5913: We have $\lm_2+\lm_3=A-\lm_1$ and $\lm_2\lm_3=C/\lm_1$.
5914: In order $|\lm_{2,3}|=1$, we must have
5915: \[
5916: \lm_2+\lm_3=c\cdot \sqrt{\lm_2\lm_3}\,,
5917: \]
5918: with $c\in [-2,2]$, so $A-\lm_1=c\sqrt{C/\lm_1}$, which is
5919: equivalent to
5920: \begin{eqn}\label{qw2}
5921: \frac{(A-\lm_1)^2\lm_1}{C}\,\in\,[0,4]\,.
5922: \end{eqn}
5923: 
5924: \subsubsection{Applications and examples}
5925: 
5926: The image of the r.h.s. of \eqref{qw0} under \eqref{qw}
5927: will be some curve in $\bC$ that generically intersects
5928: $S^1$ in a finite number of points. This parametric
5929: equality can be examined numerically and allows to recover
5930: $A=\tr \psi_3$ up to finite indeterminacy. In particular,
5931: we have
5932: 
5933: \begin{prop}\label{4brt}
5934: Assume for some $t\in S^1$ with $|\arg t|\le \pi/2$ we
5935: cannot find $y$ as in \eqref{qw0}, such that $\lm_1$
5936: given by \eqref{qw} has norm 1 and \eqref{qw2} holds. Then
5937: there is no $\bt\in B_4$ with the given $e$ whose closure has
5938: the given Alexander and Jones polynomial. \qed
5939: \end{prop}
5940: 
5941: \begin{rem}
5942: To apply the test in practice, one chooses a small stepwidth $s$
5943: for $y$ in the interval \eqref{qw0}, and calculates the derivative
5944: of the r.h.s. of \eqref{qw} in $y$ to have an error bound on
5945: $||\lm_1|-1|$ in terms of $s$. When some of the radicands in \eqref{Gm}
5946: becomes close to $0$, some care is needed. One situation where
5947: such degeneracy occurs are the knots $15_{144634}$, $15_{144635}$,
5948: and $15_{145731}$. They have representations $\bt\in B_4$, whose
5949: Burau matrix is trivial for $t=e^{\pi i/3}$. This may be noteworthy
5950: on its own in relation to the problem whether $\psi_3$ is faithful. 
5951: (Clearly $\psi_3$ is unfaithful at any root of unity on the
5952: center of $B_4$, but here $\bt$ is not even a pure braid.)
5953: \end{rem}
5954: 
5955: \begin{exam}
5956: Applying proposition \reference{4brt} we can exclude $11_{387}$, one
5957: of the 7 prime 11 crossing knots with $b=5$ but $MWF\le 4$. Eleven
5958: prime knots with 12 crossings, and 63 of 13 crossings where braid
5959: index 4 is not prohibited by MWF can be ruled out. The correctness
5960: of these examples was later verified by the 2-cabled MWF. Up to 16
5961: crossings more than 4000 examples were obtained. (Let us note that
5962: from them only about 100 can be identified using the norm estimates.)
5963: \end{exam}
5964: 
5965: \begin{exam}
5966: The check of prime 14 crossing knots to which our criterion applied,
5967: revealed 6 knots with 2-cable MWF bound 8. One, $14_{22691}$,
5968: can be excluded from braid index 4 (as done with $14_{45759}$ 
5969: in \cite{posex_bcr}), by making the $v$-degrees of the
5970: polynomial of the 2-cable contradict the exponent sum of its
5971: possible 8-braid representation. However, for the other 5
5972: knots, $14_{28220}$, $14_{30960}$, $14_{41334}$, $14_{41703}$,
5973: and $14_{44371}$, the argument fails, and so our condition
5974: seems the only applicable one. (Clearly a 3-cable polynomial
5975: is not a computationally reasonable option, and even the 2-cable
5976: requires up to several hours, while our test lasts a few seconds.)
5977: \end{exam}
5978: 
5979: Still our criterion leaves open several interesting examples
5980: of (apparent) failure of the 2-cable MWF inequality, among
5981: them the knot $13_{9684}$ encountered with M.~Hirasawa. A more
5982: general, and important, possible application is as follows.
5983: 
5984: \begin{rem}\label{rJC}
5985: In relation to Jones's conjecture \reference{jcc}, we
5986: already quoted (in the proof of theorem \ref{thJC})
5987: the observation in \cite{posex_bcr} that counterexamples
5988: to the conjecture would make MWF and \em{all} of its cabled
5989: versions unsharp. In that sense, our 4-braid test may be
5990: the first possible approach toward identifying such a
5991: counterexample. Birman-Menasco claimed indeed a family
5992: of 6-string potential counterexamples, and K.~Kawamuro
5993: gave later a simpler family on 5 strings. Extensive
5994: checks with our test of Kawamuro's knots failed to turn up
5995: successful cases. This puzzled us a while, until Kawamuro
5996: reported recently that in fact H.~Matsuda falsified all
5997: Birman-Menasco (and hence also Kawamuro's) candidates.
5998: \end{rem}
5999: 
6000: \subsection{Mahler measures}
6001: 
6002: \subsubsection{3-braids}
6003: 
6004: The material in section \reference{SBurau} originated from
6005: a question of S.~Kamada whether the Alexander polynomial
6006: Mahler measure $M(\Dl)$ is bounded on closed 3-braids. 
6007: 
6008: \begin{defi}
6009: For a Laurent polynomial $p\in\bZ[t^{\pm 1}]$, and $n\in\bN_+$,
6010: we define the $n$-norm of $p$ by
6011: \[
6012: ||p||_n\,:=\,\sqrt[n]{\sum_{k=-\infty}^{\infty}\bigl([p]_k\bigr)^n}\,,
6013: \]
6014: and its Mahler measure by
6015: \[
6016: M(p)\,:=\,\prod_{|t|\ge 1,\,p(t)=0}\,|t|\,.
6017: \]
6018: This is extended to $p\in\bZ[t^{\pm 1/2}]$ in the obvious way.
6019: See \cite{SSW,CK}.
6020: \end{defi}
6021: 
6022: Kamada's question is related to controlling $|V(t)|$ and $|\Dl(t)|$
6023: for $t\in S^1$, since the 2-norm $||\,.\,||_2$ and the Mahler
6024: measure $M$ of polynomials have circle integral formulas (see
6025: \eqref{55'} and \eqref{Ji}). One
6026: can thus ask whether $||\Dl\cdot W_n||_2$ is bounded for proper
6027: $W_n\in \bZ[t,1/t]$, or (weaker) whether $M(\Dl)$ is bounded
6028: (and similarly $M(V)$ and $||V\cdot W_n||_2$) for braid
6029: index $\le n$.
6030: 
6031: For values of $t$ where $J_0$ is not definite, however, there
6032: seems little one can say on the range (on closed 3-braids)
6033: of $|V(t)|$ or $|\Dl(t)|$. Likely they are dense in $\bR_+$.
6034: % So bounded 2-norms seem unlikely, but even if $|\Dl(t)|$
6035: % is unbounded for any $t$ with indefinite $J_0$, that is
6036: % not sufficient.
6037: We can conclude boundedness properties in
6038: special cases where the indefinite $J_0$ values of $t$ are
6039: controllable. For example the following can be proved easily.
6040: (Note that for polynomials with integer coefficients the properties
6041: a set of polynomials to have bounded $2$-norms, or finitely many
6042: distinct $2$-norms, are equivalent, and also equivalent to the same
6043: two properties for the $1$-norm.)
6044: 
6045: \begin{prop}
6046: The set
6047: \[
6048: \{\,||(1-t^{2n})\Dl_{\hat\bt}(t)||_2\,:\,
6049: \bt\in B_{2n},\,\Dl_{\hat\bt}\in\bZ[t^{\pm n}]\,\}
6050: \]
6051: is finite for any $n\ge 2$.
6052: \end{prop}
6053: 
6054: \proof We have 
6055: \begin{eqn}\label{55'}
6056: ||X||_2^2\,=\,\int_{0}^{1}\,\big|\,X(e^{2\pi is})\,\big|^2\,ds\,.
6057: \end{eqn}
6058: If $\Dl\in\bZ[t^{\pm n}]$, then this integral for 
6059: \[
6060: X=\frac{(1-t^{2n})\Dl(t)}{1-t}=\det(1-\psi_{2n-1})
6061: \]
6062: is controlled from its part over $0<s<\ffrac{1}{2n}$, and
6063: this in turn is controlled by the unitarity of $\psi_{2n-1}$.
6064: To eliminate the denominator, observe that the norm in
6065: \eqref{55'} is just the one in $L^2(S^1)$, and so
6066: $||P\cdot Q||_2\le ||P||_2\cdot ||Q||_2$. \qed
6067: 
6068: However, in general a bound on the Mahler measure of
6069: arbitrary polynomials of closed $n$-braids does not exist
6070: even for $n=3$.
6071: 
6072: \begin{prop}
6073: The Mahler measure of $\Dl$ and $V$ is unbounded on closed
6074: 3-braids.
6075: \end{prop}
6076: 
6077: {\def\gm{g_{\min}'}
6078: \proof In the following $x\le O(y)$ means $\limsup |x/y|<\infty$
6079: (the limit following from the context), $x\ge O(y)$ means $\liminf
6080: |x/y|>0$ and $x=O(y)$ means both $x\le O(y)$ and $x\ge O(y)$.
6081: The quantities $C$ and $C'$ will stand for real constants, with
6082: $C$ being positive. They are understood to be independent on the
6083: indices of the terms in the formula they occur in, but
6084: may vary between the various inequalities.
6085: 
6086: We consider the links $L_n$ given by the closed 3-braids
6087: $(\sg_1\sg_2^{-1})^n$. Using the Burau matrix eigenvalues, we find
6088: \[
6089: V_n\,:=V(L_n)\,=\,t+\frac{1}{t}+e_-^n(t)+e_+^n(t)\,,
6090: \]
6091: where 
6092: \[
6093: e_{\pm}(t)\,=\,-\frac{t+1/t-1}{2}\,\pm
6094: \sqrt{\left( \frac{t+1/t-1}{2} \right)^2-1 }\,.
6095: \]
6096: Similarly
6097: \[
6098: \Dl_n\,:=\,\Dl(L_n)\,=\,\frac{1}{t+1/t+1}\left(2-e_-^n(t)-e_+^n(t)
6099: \right)\,.
6100: \]
6101: Now the Jensen integral gives
6102: \begin{eqn}\label{Ji}
6103: \log\,M(V_n)\,=\,\int_0^1\,\log\,|V_n(e^{2\pi is})|\,ds\,.
6104: \end{eqn}
6105: Since $V_n$ is reciprocal, it suffices to consider the integral
6106: over $s\in[0,1/2]$. Herein, with $t = e^{2\pi is}$, for
6107: $s\in (1/3,1/2]$, we have real $e_{\pm}(t)$, one with norm $>1$. So
6108: \[
6109: \int_{1/3}^{1/2}\,\log\,|V_n(e^{2\pi is})|\,ds\,\ge\,O(n)\,.
6110: \]
6111: We must argue about $s\in[0,1/3]$. In that case, $e_{\pm}(t)$
6112: are conjugate complex numbers of unit norm. This implies that
6113: the integrand is bounded above when $n\to\infty$, but our problem
6114: is to show that it does not decrease too quickly with $n$.
6115: The assignment (with $i=\sqrt{-1}$)
6116: \[
6117: t\,\longmapsto\,f(t)\,=\,-\frac{t+1/t-1}{2}\,+
6118: i\sqrt{ 1-\left( \frac{t+1/t-1}{2} \right)^2 }\,
6119: \]
6120: maps $e^{2\pi i[0,1/3]}\to e^{2\pi i[1/6,1/2]}$, so
6121: \[
6122: \frac{1}{2\pi}\,\arg t\,\stackrel{g}{\longmapsto}\,
6123: \frac{1}{2\pi}\,\arg f(t)
6124: \]
6125: is a map $g\,:\,[0,1/3]\to[1/6,1/2]$, which is
6126: bijective and monotonous, and $C^1$ except in $t=1/3$
6127: where $g'\to\infty$. In particular,
6128: \[
6129: \gm\,=\,\min\,\{\,g'(s)\,:\,s\in[0,1/3]\,\}\,>0\,.
6130: \]
6131: Now consider the functions 
6132: \[
6133: h(s)\,:=\,2\cos 2\pi ng(s)\quad\mbox{and}\quad
6134: m(s)\,:=\,-2\cos 2\pi s
6135: \]
6136: for $s\in[0,1/3]$. The singularities of the Jensen integral \eqref{Ji}
6137: correspond to the intersections of the graphs of $h$ and $m$.
6138: 
6139: If $n\gg \gm$, then all intersection points\footnote{For the rest
6140: of the proof $i$ is used as an index rather than the complex unit.}
6141: $s_i$ of the graphs of $h$ and $m$ have $|m'(s_i)|>|h'(s_i)|$. (Here
6142: we use also that $m\ne h$ when $s=0$.) In particular, between two
6143: critical points $x_i$ of $h$, it intersects $m$ at most once. So the
6144: number $n'$ of intersections $s_i$ satisfies
6145: \[
6146: \left|\,n'-\frac{2}{3}n\,\right|\,\le\,O(1)\,,\mbox{\quad
6147: and therefore\quad}\,n'=O(n)\,.
6148: \]
6149: 
6150: Moreover, 
6151: \begin{eqn}\label{20.5}
6152: |m'(s_i)|-|h'(s_i)|\,\ge\,(2-|m(s_i)|)\,\cdot C\,,
6153: \end{eqn}
6154: for a constant $C$ independent on $i$ and $n$. Now when $n$
6155: increases, the $s_i$ will concentrate around $s=\frac{1}{3}$,
6156: but since $\gm>0$, for $s$ small, the $x_i$ will be at distance
6157: $\ge O(\frac{1}{n})$. So
6158: \begin{eqn}\label{20.6}
6159: 2-|m(x_i)|\ge C\cdot \frac{i^2}{n^2}
6160: \end{eqn}
6161: for a constant $C$ independent on $i$ and $n$. Then
6162: if $s_1,\dots,s_{n'}$ are the $n'$ intersection points
6163: of $h$ and $m$, we can estimate for values of $s$
6164: between the critical points $x_i$ and $x_{i+1}$ of $h$
6165: \def\tsi{\bigl(2-|m(s_i)|\bigr)}
6166: \begin{eqn}\label{x5}
6167: |m(s)-h(s)|\,\ge\,C\,\cdot \tsi\,\cdot |s-s_i|^2
6168: \end{eqn}
6169: for a number $C$ independent on $i$ and $s$. Inequality
6170: \eqref{x5} follows because we can control the behaviour
6171: of $|m(s)-h(s)|$ from \eqref{20.5} for $s$ around $s_i$,
6172: and though $|m-h|$ decreases slightly around the endpoints
6173: of the interval $[x_i,x_{i+1}]$, this decrease can be
6174: controlled by \eqref{20.6}. So
6175: \begin{eqn}\label{x4}
6176: \int_{x_i}^{x_{i+1}}\,\log\,|\,m(s)-h(s)\,|\,ds\,\ge\,
6177: (x_{i+1}-x_i)\Bigl(\,\log\,\tsi\,+\log C\,\Bigr)\,+\,
6178: \,2\,\int_{x_i}^{x_{i+1}}\,\log |s-s_i|\,ds\,.
6179: \end{eqn}
6180: Since for small $i$ (when $2-|m(s_i)|$ is
6181: small), the $x_i$ will be at distance $\ge O(\frac{1}{n})$, using
6182: \eqref{20.6} and $m(s)\le 1$ when $0\le s\le \myfrac{1}{3}$ we have
6183: \begin{eqn}\label{x2}
6184: \sum_{i=1}^{n'}\,\Bigl|\,\log\,\tsi\,\Bigr|\,\le\,n'\cdot
6185: \left(\,C\cdot\left|\int_{0}^{1}\log |x|\,dx\right|+C'\,\right)\,=
6186: \,O(n)\,.
6187: \end{eqn}
6188: 
6189: Let $d_i:=x_{i+1}-x_i$. Because of $\gm>0$, we have 
6190: \begin{eqn}\label{63'}
6191: \max_{i}\,d_i\,\le\,O(1/n)\,.
6192: \end{eqn}
6193: So, similarly as in \eqref{x2}
6194: \begin{eqn}\label{x3}
6195: \sum_{i=1}^{n'}\,\Bigl|\,d_i\,\log\,\tsi\,\Bigr|\,
6196: \le O(1)\,.
6197: \end{eqn}
6198: Also, because of \eqref{63'} and $\int 1+\log x\ dx=x\cdot\log x+C$,
6199: for $n$ large
6200: \[
6201: \left|\,\int_{x_i}^{x_{i+1}}\,\bigl(\,1+\log |s-s_i|\,\bigr)\,ds\,
6202: \right|\,\le\,
6203: -C\cdot d_i\cdot \left[\,\log d_i\,+\,C'\,\right]\,\le\,
6204: \sqrt{d_i}\,\le\,O\left(\frac{1}{\sqrt{n}}\right)\,
6205: \]
6206: for some real number $C'$ independent on $i$ and $n$, and so
6207: \[
6208: \sum_{i=1}^{n'}\,\left|\,\int_{x_i}^{x_{i+1}}\,2\cdot
6209: \log |s-s_i|\,ds\,\right|\,\le\,O(\sqrt{n})\,.
6210: \]
6211: This, \eqref{x3} and \eqref{x4} imply that for $s\in[0,1/3]$ the
6212: (part of the) Jensen integral
6213: \eqref{Ji} has a lower bound that behaves like $-O(\sqrt{n})$, so the
6214: dominating part is for $s\in[1/3,1/2]$, and we are done for $V$.
6215: 
6216: For $\Dl$ the argument is similar. There $m(s)\equiv 2$, so
6217: instead of \eqref{x5} we have for $s\in[x_i,x_{i+1}]$ the estimate
6218: $|m(s)-h(s)|\ge |s-s_i|^2\cdot C\cdot \gm$, with the same further
6219: reasoning.  \qed
6220: }
6221: 
6222: \begin{rem}
6223: Dan Silver pointed out that for the Alexander polynomial a proof
6224: can be obtained (possibly more elegantly, but not immediately)
6225: from the ideas in \cite{SW}.
6226: \end{rem}
6227: 
6228: \subsubsection{Skein polynomial and generalized twisting}
6229: 
6230: By using the representation theory, one can extend the scope of
6231: the results in \cite{SSW,CK} on bounded Mahler measure
6232: to (parallel) multi-strand twisting.
6233: We give a version for the skein polynomial, since its special case,
6234: the Jones polynomial, was discussed in \cite{CK}. One can obtain
6235: this special case by setting $\lm=q$ in the below theorem. (The
6236: reversal of component orientation is not a serious problem for
6237: $V$, unlike for $P$.) We write as before $P_L(v,z)$ for the skein
6238: polynomial and use the skein rule $v^{-1}P_+-vP_-=zP_0$.
6239: 
6240: A \em{template} $T$ consists, as in \cite{SunThis}, of a number of
6241: strands, that create crossings and \em{slots}, only that for us
6242: strands are oriented and slots have the more general form
6243: (with an arbitrary number of in- and outputs)
6244: \[
6245: \diag{7mm}{3}{2}{
6246:   \picmultigraphics{5}{0.5 0}{\picvecline{0.5 0}{0.5 2}}
6247:   \picfilledbox{1.5 1}{2.8 1}{}
6248: }\,.
6249: \]
6250: A(n oriented) diagram $D$ is \em{associated} to $T$, if $D$ is
6251: obtained from $T$ by inserting into each slot of $T$ a braid of a
6252: certain number of full twists (on the proper number of strands).
6253: 
6254: The $1$-norm $||P||_1$ of a (Laurent) polynomial $P$ (even in
6255: several variables) is understood to be the sum of absolute values
6256: of all its (non-zero) coefficients (taken in all variables).
6257: We write $M(P)$ for its Mahler measure. It is known that for
6258: polynomials $P$ (of real coefficients), $M(P)\le ||P||_1$
6259: (see \S 2 of \cite{SSW}).
6260: 
6261: \begin{theorem}
6262: For each $n$ there is a polynomial $D_n(q)$ such that for
6263: each template of $n_1,\dots,n_k$-strand parallel twist
6264: slots and each diagram $D$ associated to $T$ we have
6265: \[
6266: \left|\left|\es\prod_{i=1}^k D_{n_i}(q)\es\cdot\es
6267: P_D(\lm,\sqrt{q}-1/\sqrt{q})\es\right|\right|_1\le C_T\,,
6268: \]
6269: where $C_T$ is a constant that depends only on $T$.
6270: Furthermore $D_n$ are made of a product of terms $1-q^{l}$
6271: for some $1\le l\le n$. In particular, 
6272: \[
6273: \{\,M(\,P_D(\lm,\sqrt{q}-1/\sqrt{q})\,)\,:\,\mbox{\ $D$ is
6274: associated to $T$\,}\,\}
6275: \]
6276: is bounded.
6277: \end{theorem}
6278: 
6279: \proof Write $D(t_1,\dots,t_k)=D_{t_1,\dots,t_k}$ for the diagram
6280: associated to $T$ by putting into the $i$-th slot $t_i$ full
6281: twists (on $n_i$ strands).
6282: 
6283: Let first $k=1$. Then the complement of the slot is a room with
6284: $n=n_1$ parallel in- and outputs. The skein module of such
6285: a room is generated by (positive permutation) braids, and
6286: thus it suffices to work with the $n$-strand braid
6287: group. Again, by \cite[Proposition 6.2]{Jones}
6288: \[
6289: P_L(\lm,\sqrt{q}-1/\sqrt{q})=X_L(q,\lm^2/q)\,,
6290: \]
6291: where
6292: % $X_L(q,\lm)$ is the invariant treated in \cite{Jones}.
6293: % By the work of \cite{Jones} we know that
6294: $X_L$ is a writhe-strand normalized (Definition 6.1 in \cite{Jones})
6295: weighted trace sum (equation 5.5), and the full twist has the
6296: effect of multiplying the traces by a power of $q$ (Lemma 9.3).
6297: 
6298: All terms in the demoninator of the (rational) generating series 
6299: \begin{eqn}\label{GS}
6300: \sum_{t=0}^{\infty}\,X(D_t)(q,\lm^2/q)\,x^t\,
6301: \end{eqn}
6302: are obtained from demoninators in the definition
6303: of $X_L(q,\lm)$ and quantities that differ by the
6304: number $t=t_1$ of twists. To identify these terms we 
6305: will work up to units in $\bZ[q^{\pm 1/2},\lm^{\pm 1/2}]$
6306: that do not depend on the number $t$ of twists but just 
6307: on $T$. The quantities depending on $t$ are only
6308: the power of $q$ in the traces and the term $\sqrt
6309: {\lm}^e$ in Definition 6.1, where $e$ is the exponent sum.
6310: Since the full twist has $e=n(n-1)$, we find the terms
6311: $1-\lm^{n(n-1)/2} q^jx$ for some values of $j\ge 0$. The
6312: exponent $n(n-1)/2$ becomes $n(n-1)$ and $j\ge -n(n-1)/2$
6313: when replacing $\lm^2/q$ for $\lm$. It suffices now
6314: to collect the demoninator terms occurring in the
6315: definition of $X_L$ (and replace $\lm^2/q$ for $\lm$).
6316: {}From Definition 6.1 we have $\lm^{(n-1)/2}$ (which is
6317: a unit, fixed for given $T$, and remains so when
6318: replacing $\lm^2/q$ for $\lm$) and some power of $1-q$.
6319: {}From the text below Figure 5.6 and (5.5) we have
6320: some power of $1-\lm q$ (that becomes $1-\lm^2$ under
6321: substitution) and (in $Q(q)$) products of terms $1-q^l$
6322: for some $1\le l\le n$ (with $l$ being a hook length of a
6323: box in a Young diagram of $n$ boxes). Thus the demoninator
6324: of \eqref{GS} is something we can take to be $D_n(q)$ times 
6325: a product of terms of the form $1-\lm^2$ and $1-\lm^{n
6326: (n-1)}q^{j}x$. 
6327: 
6328: Now if $k>1$, we first observe that $X(D_{t_1,\dots,t_k})(q,\lm)$
6329: satisfy a linear recurrence (with coefficients in $\bZ[
6330: q^{\pm 1/2},\lm^{\pm 1/2}]$) in $t_i$ for any fixed value
6331: of $t_j, j\ne i$. Moreover that recurrence itself does not
6332: depend on $t_j, j\ne i$ for fixed $i$, only its initial values
6333: do depend. Then one inductively argues over $k$ that the
6334: generating series 
6335: \[
6336: \sum_{t_1,\dots,t_k\ge 0} X(D_{t_1,\dots,t_k})(q,\lm^2/q)\es
6337: x_1^{t_1} \dots x_k^{t_k}
6338: \]
6339: has a demoninator which is the product of the $D_{n_i}$ and terms
6340: $1-\lm^2$ and $1-\lm^{n_i(n_i-1)}q^{j}x_i$ for $1\le i\le k$ and $j\ge 
6341: -n_i(n_i-1)/2$. The cases when $t_i<0$ are analogous. From this
6342: one easily concludes the claim up to the factors $1-\lm^2$.
6343: Since for fixed $T$ the power $p_T$ of $1-\lm^2$ is fixed,
6344: and the $\lm$-span of $P_D$ is bounded by the Morton-Williams-%
6345: Franks inequality (see proposition 15.1 of \cite{Jones}),
6346: one can get disposed of the $1-\lm^2$
6347: factors by linear combinations of the coefficients of
6348: $(1-\lm^2)^{p_T}P_D\cdot \prod D_{n_i}$ in the powers
6349: of $\lm$. \qed
6350: 
6351: 
6352: \subsection{Knots with unsharp MWF inequality}
6353: 
6354: This brief part was motivated by the paper of Kawamuro
6355: \cite{Kawamuro}, which I came across during the continuous work
6356: on my paper. Kawamuro was interested in finding infinitely many
6357: knots for which the MWF inequality is strict, in particular such
6358: satisfying Jones' conjecture in \S\ref{SJS}. The point to make
6359: is that we have at least two ways allowing us to obtain her, and
6360: further such knots much more easily. (In particular it is not
6361: necessary to appeal to the heavy geometric machinery of
6362: Neumann-Rudolph-Giroux.)
6363: 
6364: We already saw an infinite family of examples with unsharp MWF
6365: (and satisfying Jones' conjecture) in the proof of theorem
6366: \ref{thJC}. This family is, of course, little insightful,
6367: since it is settled already by Birman-Menasco's 3-braid
6368: work. The 4-braid examples they proposed, but could not
6369: decide about, deserve more attention. As given in figure 4
6370: of \cite{Kawamuro}, consider for $(x,y,z,w)\in\bZ^4$ the
6371: 4-braids $\bt_{x,y,z,w}=[212^213^x2^y-12^z3^w]$, and let
6372: $K_{x,y,z,w}=\hat\bt_{x,y,z,w}$. Birman-Menasco observe that
6373: $MWF(K_{x,y,z,w})\le 3$ (see lemma 2.9 in \cite{Kawamuro}),
6374: but suspect that many of the $K_{x,y,z,w}$ have braid index 4.
6375: One can exhibit an infinite family of $K_{x,y,z,w}$ of $b=4$,
6376: and thus recover theorem 2.8 of \cite{Kawamuro}, like this.
6377: 
6378: \begin{prop}
6379: There is a natural number $N$ and 6 tuples $(\ap_1,\dots,
6380: \ap_4)\in \bZ^4$ of different mod-2-reductions in $(\bZ/2)
6381: ^4$, such that if $K_\gm:=K_{\gm_1,\dots,\gm_4}$ is a knot,
6382: then $b(K_\gm)=4$ whenever $(\ap_i)\equiv (\gm_i)\bmod
6383: 2$ and $\gcd\Bigl(\ffrac{\gm_i-\ap_i}{2}\Bigr)\ge N$.
6384: \end{prop}
6385: 
6386: \proof Since (complex) roots of unity are dense on the complex
6387: unit circle, choose a root of unity $t$, for which the 3-braid
6388: test in corollary \ref{pp1} applies for $9_{42}$ or $9_{49}$ (see
6389: example \ref{4249}). Then, using the formula for $\Dl$ in lemma 3.1
6390: of \cite{SSW}, we see that for each such $t$ there is an $n\in2\bZ
6391: \sm\{0\}$, such that $\Dl_{\hat\bt}(t)$ is preserved when $\bt$
6392: is replaced by $\sg_i^n\bt$.
6393: 
6394: Moreover, it is easy to check that for all 6 of the 16 possible
6395: vectors of parities of $(x,y,z,w)$, for which $K_{x,y,z,w}$ is a
6396: knot, there is a representation of (at least one of) $9_{42}$ or
6397: $9_{49}$ of that parity combination. (The representations given
6398: after definition 2.7 of \cite{Kawamuro} show 3 of the parity
6399: types.) Now, take such a representation $\bt_{\ap_1,\dots,\ap_4}$
6400: and vary the parameters $\ap_i$ by multiples of $n$. Then, since
6401: we can choose (for proper $t$) $n$ to be any sufficiently
6402: large even natural number, we obtain in the claim. \qed
6403: 
6404: Of course, this elementary argument could be further concretified
6405: and strengthened. Presumably, one can do much better using
6406: J.~Murakami's test theorem \reference{TMF} (since it is an equality).
6407: Note that our proof does not confirm Jones' conjecture on any
6408: $K_{x,y,z,w}$, so it may trigger the question: would it help
6409: to find a counterexample? This seems very optimistic, though,
6410: as we will soon see from an alternative approach to Kawamuro's
6411: theorem, using the work in \S\ref{SJS}. First we can settle
6412: (also with regard to Jones' conjecture) the concrete examples
6413: $K_{-1,-2,m,2}$ for $m\ge 2$ even, obtained in her proof.
6414: 
6415: \begin{prop}\label{pZ}
6416: We have $b(K_{-1,-2,m,2})=4$ for $m\ne -1,0$.
6417: \end{prop}
6418: 
6419: \proof We give just a brief explanation. We try to calculate the
6420: skein polynomial of the 2-cable $\bt^{[2]}_{-1,-2,m,2}$ of $\bt_
6421: {-1,-2,m,2}$, obtained by replacing $\sg_i$ by $\sg_{2i}\sg_{2i+1}
6422: \sg_{2i-1}\sg_{2i}$, and we may consider just the coefficient
6423: of $v^{4[\bt_{-1,-2,m,2}]+7}$. Again $\pi_Y([2312])$ has at
6424: most 7 different eigenvalues $\dl_i$, for all $Y\vdash 8$ taken
6425: together. But now we have no full twist, and so do not need to
6426: deal with the various $Y$ (and their weights $W_Y$ etc.) one by
6427: one. We can sum $W_Y\cdot c(i,Y)$ over $Y\vdash 8$, and are left
6428: with 7 different $q$-rational expressions $c_i$ to determine. 
6429: Again this can be done from 7 explicit polynomials, which can be
6430: obtained using Morton's program. (If we focus on one parity of $m$
6431: only, we can again reduce the unknowns $c_i$ and test polynomials 
6432: to 5, but need to calculate polynomials from slightly more
6433: complicated braids.)
6434: 
6435: We calculated in fact 11 polynomials, for $|m|\le 5$, and
6436: used those 9 for $|m|\le 4$ in the determination of $c_i$
6437: to have extra safety. The resulting linear
6438: equation system for $c_i$ (with matrix $\{\dl_i^m\}$) is now
6439: not a serious problem to MATHEMATICA, which gives the solution
6440: (for generic $q$) in just a few seconds. Clearing denominators,
6441: and looking already at the extremal $q$-degrees, we see that for
6442: $|m|\ge 6$ among the 7 terms $c_i\dl_i^m$ the one for $i=2$ has
6443: the lowest minimal or highest maximal degree (as a unique term,
6444: except for $m=6$, where we must use also that the leading
6445: coefficient of $c_6\dl_6^m$ has the same sign). With an
6446: explicit check for the other $m$, we conclude the claim. \qed
6447: 
6448: This method of proof has also the advantage of easily leading
6449: to a qualitative improvement of Kawamuro's theorem, which is
6450: closer to what one should expect.
6451: 
6452: \begin{prop}
6453: The tuples $(x,y,z,w)$ for which $K_{x,y,z,w}$ has braid index 4 (and
6454: satisfies Jones' conjecture) are generic in $\bZ^4$, in the sense that
6455: \begin{eqn}\label{bla}
6456: \lim_{n\to\infty}\es
6457:   \frac{
6458:    \big|\,
6459:     \{\,(x,y,z,w)\in \bZ^4\,:\,b(K_{x,y,z,w})=4\,\}\,\cap\, {[-n,n]^4} 
6460:    \,\big|
6461:   \rule[-0.6em]{\z@}{1.3em}}
6462:   {\big|\,[-n,n]^4\,\big|\rule[0.1em]{\z@}{1.0em}}
6463: \es=\,1\,.
6464: \end{eqn}
6465: \end{prop}
6466: 
6467: \proof Extending the argument for proposition \ref{pZ}, we see that 
6468: \[
6469: \tl P(x,y,z,w):=
6470:   [P(\hat\bt^{[2]}_{x,y,z,w})]_{v^{4[\bt_{x,y,z,w}]+7}}(
6471:   \sqrt{q}-1/\sqrt{q})\cdot \sqrt{q}^{\,4[\bt_{x,y,z,w}]+7}
6472:   \,=\,\sum_{i_x,i_y,i_z,i_w=1}^7\,c_{i_x,i_y,i_z,i_w}\dl_{i_x}^x
6473:   \dl_{i_y}^y\dl_{i_z}^z\dl_{i_w}^w\,,
6474: \]
6475: where $c_{i_x,i_y,i_z,i_w}$ are again some rational expressions
6476: in $q$. They are obviously not all zero, and the polynomials $\tl
6477: P(x,y,z,w)$ have the following property: if one fixes three of the
6478: parameters (say $x,y,z$), and $\tl P(x,y,z,w)=0$ for 7 different
6479: values of the remaining parameter (here $w$), then it vanishes
6480: for all values of that parameter (at the given fixed 3 others).
6481: It is then not hard to deduce \eqref{bla} from this (see e.g.
6482: the proof of lemma 11.1 of \cite{gen2}). \qed
6483: 
6484: 
6485: 
6486: \noindent{\bf Acknowledgement.} I would like to thank to M. Hirasawa,
6487: M. Ishiwata and K.~Murasugi for their contribution, and S.~Kamada,
6488: K.~Kawamuro, T.~Nakamura and D.~Silver for some helpful remarks.
6489: The calculations were performed largely using the programs of
6490: \cite{MorSho2} and \cite{KnotScape}, and MATHEMATI\-CA\TM{}
6491: \cite{Wolfram}. I also wish to thank to my JSPS host Prof. T.~Kohno
6492: at University of Tokyo for his support.
6493: 
6494: % \vspace{1cm}\mbox{}
6495: % \newpage
6496: 
6497: \begin{appendix}
6498: 
6499: \renewcommand{\thesection}{\Alph{section}}
6500: {\def\@seccntformat#1{Appendix \csname the#1\endcsname . \quad}
6501: 
6502: \section{Postliminaries}
6503: }
6504: % \def\@seccntformat#1{\csname the#1\endcsname . \quad}
6505: 
6506: \subsection{Fibered Dean knots (Hirasawa-Murasugi)\label{HM}}
6507: 
6508: Here we present some material due to Hirasawa and Murasugi,
6509: who studied fibering of generalized Dean knots. An overview is given
6510: in \cite{Hirasawa}. As this is work in progress, a longer exposition
6511: may appear subsequently elsewhere.
6512: 
6513: \begin{defi}
6514: The Dean knot $K(p,q|r,rs)$ is given by the closed $p$-braid
6515: \[
6516: (\sg_{p-1}\sg_{p-2}\cdots \sg_1)^{q}(\sg_1\sg_2\cdots \sg_{r-1})^{rs}\,,
6517: \]
6518: with $p>r>1$ and $q,s$ non-zero integers such that $(q,p)=1$.
6519: \end{defi}
6520: 
6521: Hirasawa and Murasugi proposed a conjecture on these knots
6522: and obtained so far the following partial results.
6523:  
6524: \begin{conjecture}(Hirasawa-Murasugi)
6525: A Dean's knot $K(p,q|r,rs)$ is a fibred knot if and
6526: only if its Alexander polynomial is monic, that is,
6527: $\Mc\Dl=\pm 1$.
6528: \end{conjecture}
6529: 
6530: \begin{figure}[th]
6531: \[
6532: \begin{array}{ccc}
6533: \epsfs{6cm}{fig2a} & 
6534: \epsfs{5cm}{fig2b} & 
6535: \epsfs{5cm}{fig2c}\\
6536:  (a) & (b) & (c) \\
6537: \end{array}
6538: \]
6539: \caption{\label{fig2}}
6540: \end{figure}
6541: 
6542: \begin{prop}(Hirasawa-Murasugi)
6543: This conjecture has been proven for the following cases.
6544: \def\theenumi{\alph{enumi}}
6545: \def\labelenumi{(\theenumi)}
6546: \begin{enumerate}
6547: \item     $q = kp + 1$, and $r$ and $s$ are arbitrary,
6548: \item     $q =kp - 1$, and  $r$ and $s$ are arbitrary,
6549: \item     $r =  p - 1$, and  $q$ and $s$ are arbitrary.
6550: \end{enumerate}
6551: \end{prop}
6552: 
6553: The last case implies in particular that the conjecture is
6554: true for $p=3$. Below follows a part of the argument in this
6555: case that settles lemma \reference{lki}. (Other parts of their
6556: proof are very similar to some of our previous arguments.
6557: It seems, for example, that Hirasawa and Murasugi were to
6558: some extent aware of Theorem \reference{Thq}.)
6559: 
6560: \proof [of lemma \reference{lki}] We consider the braids $[(123)^k-2]$,
6561: with $k>0$ fixed. By isotopies and Hopf plumbings, we modify our
6562: surface 
6563: \begin{eqnarray*}
6564: % & (123)^k-2\to 1223(123)^{k-1}-2\to -21223(123)^{k-1}= & \\
6565: % & 12 -123(123)^{k-1}\to 23(123)^{k-1}12 -1= 2(312)^k-1\to
6566: % (2312)^k-1\,.
6567: &\rx{-1mm} [(123)^k -2] \to [(1223)^k - 2]=[12(2312)^{k-1}23 - 2] \to
6568: [(2312)^{k-1}23 - 212] = [(2312)^{k-1}2312 -1]=[(2312)^k - 1].
6569: \end{eqnarray*}
6570: Let $F$ be the surface of the band representation $\bt=[(2312)^k-1]$.
6571: We will show that $F$ is a fiber surface.
6572: 
6573: Consider the subsurface $F_0$ of $F$ that spans $[2 3 1 2]$
6574: in the natural manner.
6575: We deform the $k$ copies of $F_0$ to the (isotopic sub)surfaces
6576: $F'_0$ by a series of diagrams, see figure \reference{fig2}.
6577: (Here strands are numbered from right to left and words composed
6578: downward.)
6579: 
6580: In figure \reference{fig2}, for the move (a) $\to$ (b), we slide
6581: $B$ and $C$, respectively, along $D$ and $A$, then delete $D$
6582: (by deplumbing a Hopf band), and then slide $C$ back along
6583: $A$. For the move (b) $\to$ (c), we slide $C$ along $B$,
6584: then slide $B$ along $C$. The $k$ bands $A$ can be subsequently
6585: removed by Murasugi desumming a $(2,k)$-torus link fiber surface.
6586: Thus the surface $F$, spanned by $\hat\bt$, turns after
6587: (de)summing Hopf bands into a surface $F'$ consisting of $k$
6588: copies of $F'_0$ and one negative band $N$. See figure \ref{fig3} (a).
6589:  
6590: 
6591: {
6592: % \tm
6593: \def\@captype{figure}
6594: \long\def\@makecaption#1#2{%
6595:    % \tm
6596:    \vskip 10pt
6597:    {\let\label\@gobble
6598:    \let\ignorespaces\@empty
6599:    \xdef\@tempt{#2}%
6600:    %\typeout{`#2'}%
6601:    }%
6602:    \ea\@ifempty\ea{\@tempt}{%
6603:    \setbox\@tempboxa\hbox{%
6604:       \fignr#1#2}%
6605:       }{%
6606:    \setbox\@tempboxa\hbox{%
6607:       {\fignr#1:}\capt\ #2}%
6608:       }%
6609:    \ifdim \wd\@tempboxa >\captionwidth {%
6610:       \rightskip=\@captionmargin\leftskip=\@captionmargin
6611:       \unhbox\@tempboxa\par}%
6612:    \else
6613:       \hbox to\captionwidth{\hfil\box\@tempboxa\hfil}%
6614:    \fi}%
6615: % %
6616: % \def\fignr{\small\sffamily\bfseries}%
6617: % \def\capt{\small\sffamily}%
6618: % 
6619: % 
6620: % \newdimen\@captionmargin\@captionmargin2cm\relax
6621: % \newdimen\captionwidth\captionwidth\hsize\relax
6622: 
6623: \
6624: \captionwidth0.3\textwidth\relax
6625: \[
6626: \begin{array}{cc@{\qquad}c}
6627: \epsfs{5cm}{fig3a} & \epsfs{5cm}{fig3b} & \epsfs{5cm}{surf} \\
6628: (a) & (b) \\[-2mm]
6629: \multicolumn{2}{c}{ \vbox{\caption{\label{fig3}}} } &
6630: \raisebox{1em}{\vbox{\caption{\label{fig4}}}}
6631: \end{array}
6632: \]
6633: 
6634: }
6635: 
6636: In figure \reference{fig3}, we perform the move (a) $\to$ (b), sliding
6637: $B$ along $N$. The last surface $F'$, in figure \reference{fig3} (b),
6638: is Murasugi sum of a fibre surface spanning the $(2,2,-2)$-pretzel
6639: link and $\tl F$, where $\tl F$ consists 
6640: of $k - 1$ copies of $F'_0$ and the band $N$. By induction on
6641: $k$, we see that $F'$ is a fibre surface, and hence $\hat\bt$
6642: is a fibred link. \qed
6643:  
6644: \subsection{$A$-decomposition (joint with Hirasawa-Ishiwata)\label{HI}}
6645: 
6646: For the proof of theorem \reference{thuq} we introduce the
6647: $A$-decomposition due to T.~Kobayashi \cite{Kobayashi}.
6648: 
6649: A \em{sutured manifold} in the sense of Gabai \cite{Gabai3} can be
6650: understood as a pair $(L,H)$ consisting of a closed 3-dimensional
6651: submanifold $H$ of $\bR^3$ with boundary $S=\prt H$ a connected
6652: surface, and a set of oriented loops $L\subset H$. We require that
6653: one can orient the connected components of $S\sm L$ so that the
6654: induced orientations on $L$ coincide from both sides of $L$ (in
6655: particular a connected component of $S\sm L$ never bounds to itself
6656: along a loop of $L$), and are given by the orientation of $L$.
6657: 
6658: Let $F$ be a connected Seifert surface of a(n oriented) link $L=\prt
6659: F$. We embed $F$ as $F\times\{0.5\}$ into the \em{bicolar} $H=
6660: F\times I$ (with $I=[0,1]$). Then $(L,H)$ becomes a sutured
6661: manifold. We call it \em{canonical sutured manifold} $C(F)$ of $F$.
6662: 
6663: We describe some basic operations on sutured manifolds $(L,H)$.
6664: 
6665: A \em{decomposition disk} $D$ is a disk with $P=\prt D\subset \prt
6666: H$, properly embedded in the complement of $H$ (i.e. $D\cap H=P$).
6667: We require that $D$ is not parallel to $S=\prt H$, and
6668: satisfies $P\cap F\ne\vn$. We assume also that the intersection of
6669: $P$ and $F$ is transversal, so that it is a collection of points.
6670: 
6671: Since $L=\prt F$ is separating on $S$, the intersection $D\cap L= P
6672: \cap F$ is an even number of points, and the orientation of $L$ at
6673: the intersection points is alternating (with respect to the
6674: orientation of the loop $P$). See figure \reference{fig5} (a).
6675: 
6676: \begin{figure}[th]
6677: \[
6678: \begin{array}{c@{\kern1.4cm}c}
6679: \diag{1cm}{3}{3}{
6680:   \piccircle{1.5 d}{0.6}{$D$}
6681:   \pictranslate{1.5 1.5}{
6682:     { \piclinedash{0.05 0.05}{0.02}
6683:       \picmultigraphics[rt]{4}{90}{
6684:         \picline{0.0 0.6}{-0.3 1.4}
6685:       }
6686:     }
6687:     \picputtext{0.4 0.9}{$L$}
6688:     \picmultigraphics[rt]{2}{180}{
6689:       \picvecline{0.3 1.4}{0.0 0.6 0.3 1.4 0.5 conv}
6690:       \picline{0.0 0.6}{0.0 0.6 0.3 1.4 0.5 conv}
6691:       \picrotate{90}{
6692:         \picvecline{0.0 0.6}{0.0 0.6 0.3 1.4 0.5 conv}
6693:         \picline{0.3 1.4}{0.0 0.6 0.3 1.4 0.5 conv}
6694:       }
6695:     }
6696:   }
6697: }
6698: &
6699: \diag{1cm}{2.5}{2.0}{
6700:   \picline{1 0}{1 2}
6701:   \picline{0 0.5}{1 0.5}
6702:   \picvecline{2 0.5}{1 0.5}
6703:   \picvecline{0 1.5}{2 1.5}
6704:   \picputtext{1 2.3}{$D$}
6705:   \picputtext{2.3 0.5}{$L$}
6706:   \picputtext{2.3 1.5}{$L$}
6707:   \picputtext{0.8 1.0}{$a$}
6708: }\quad\lra\quad
6709: \diag{1cm}{2}{2.0}{
6710:   \picline{1 0}{1 2}
6711:   \picputtext{1 2.3}{$D$}
6712:   \pictranslate{1 1}{
6713:     \picmultigraphics[rt]{2}{180}{
6714:       \picline{1 -0.5}{0.2 -0.5}
6715:       \piclineto{0.2 0.5}
6716:       \picveclineto{1 0.5}
6717:     }
6718:   }
6719: }   
6720: \\
6721: \ry{1.7em}(a) & \raisebox{0.8em}{$(b)$} \\
6722: \end{array}
6723: \]
6724: \caption{\label{fig5}}
6725: \end{figure}
6726: 
6727: Then $L\cap P$ separates $P$ into a collection of intervals or
6728: \em{arcs}. Let $a$ be such an arc. An \em{$A$-operation} on $D$ along
6729: $a$ is a transformation of $(L,H)$ into a sutured manifold $(L',H)$,
6730: where $L'$ is obtained by splicing $L$ along $a$. See figure
6731: \ref{fig5} (b).
6732: 
6733: A \em{product decomposition} along $D$ is a similar operation, due
6734: to Gabai \cite{Gabai3}, and can be described as an $A$-operation
6735: if $|L\cap D|=2$ (in which case which of the two arcs is chosen
6736: is irrelevant), followed by a subsequent gluing of a $D^2\times I$ into
6737: $H$ along a neighborhood $N(P)\simeq S^1\times I$ of $P$ on $S$.
6738: 
6739: \begin{defi}
6740: We define a sutured manifold $(L,H)$ to be \em{$A$-decomposable} as
6741: follows:
6742: \def\theenumi{\arabic{enumi}}
6743: \def\labelenumi{\theenumi)}
6744: \begin{enumerate}
6745: \item Assume $H$ is a standardly embedded handlebody (i.e. so that
6746: $\ol{S^3\sm H}$ is also one). If $L$ is a collection of trivial loops
6747: on $\prt H$, and all loops bound \em{disjoint} disks in $\prt H$,
6748: then $(L,H)$ is $A$-decomposable.
6749: %
6750: \item If $(L',H')$ is obtained from $(L,H)$ by a product decomposition
6751: (along some decomposition disk $D$),
6752: and $(L',H')$ is $A$-decomposable, then so is $(L,H)$.
6753: %
6754: \item Let $D$ be a decomposition disk of $H$ with $|L\cap D|=2n$ and
6755: choose among the $2n$ arcs on $P=\prt D$ a collection of $n$ cyclically
6756: consecutive arcs $a_1,\dots,a_n$. (Consecutive is to mean that, taken
6757: with their boundary in $L\cap P$, their union is a single interval
6758: in $P$, and not several such intervals.) Let $(L_i,H)$ be obtained
6759: from $(L,H)$ by $A$-decomposition on $D$ along $a_i$ for $i=1,\dots,n$.
6760: Then if all $(L_i,H)$ are $A$-decomposable, so is $(L,H)$.
6761: \end{enumerate}
6762: \end{defi}
6763: 
6764: \begin{theorem}(T.~Kobayashi \cite{Kobayashi}, O.~Kakimizu
6765: \cite{Kakimizu}) 
6766: \def\theenumi{\arabic{enumi}}
6767: \def\labelenumi{\theenumi)}
6768: \begin{enumerate}
6769: \item A fiber surface is a unique incompressible surface.
6770: \item The property a surface to be a unique incompressible surface 
6771:   is invariant under Hopf (de)plumbing.
6772: \item If $C(F)$ is $A$-decomposable, then $F$ is a
6773:   unique incompressible surface for $L=\prt F$.
6774: \end{enumerate}
6775: \end{theorem}
6776: 
6777: \begin{figure}%[th]
6778: \[
6779: \epsfsv{4.5cm}{dec1} \stackrel{a_1}{\lra} \epsfsv{4.7cm}{dec2}
6780: \]
6781: \[
6782: \epsfsv{4.5cm}{dec3} \qquad \epsfsv{4.5cm}{dec4} 
6783: \]
6784: \[
6785: \epsfsv{4.5cm}{dec5} \qquad \epsfsv{4.5cm}{dec6}
6786: \]
6787: \[
6788: \epsfsv{4.2cm}{dec7} \qquad \epsfsv{4.2cm}{dec8}
6789: \]
6790: \[
6791: \epsfsv{3.5cm}{dec9} \qquad \epsfsv{3.7cm}{dec10}
6792: \]
6793: \[
6794: \epsfsv{3.5cm}{dec11} \qquad \epsfsv{3.2cm}{dec12}
6795: \]
6796: \caption{\label{fig6} This sequence of diagrams describes the
6797: $A$-decomposition along arc $a_1$ of the canonical sutured
6798: manifold corresponding to the surface in figure \reference{fig4}.
6799: The $A$-operation is the change between the first and the second
6800: diagram. The following diagrams (left to right in each row)
6801: display isotopies and product decompositions, leading finally to
6802: a trivial curve on a torus. The $A$-operation along $a_2$
6803: must be performed too, but the decomposition is similar.}
6804: \end{figure}
6805: 
6806: Now we complete the proof of theorem \reference{thuq}.
6807: 
6808: \proof[of theorem \reference{thuq}] 
6809: Let $L$ be a 3-braid link. If $L$ is split, the splitting sphere
6810: also splits any incompressible surface for $L$. Since
6811: $L$ is a split union of a 2-braid link and an unknot
6812: (incl.~2 and 3-component unlinks), the claim is easy.
6813: Excluding split links, corollary \reference{cbd} shows we
6814: need to consider only connected Seifert surfaces. It also
6815: suffices to deal with the non-fibered links only. These are
6816: equivalent under Hopf (de)plumbing to some of $(123)^k$ ($k\ne
6817: 0$). For such links, we can modify the transformation of
6818: Hirasawa and Murasugi from \S\reference{HM}. Then we can turn
6819: by Hopf (de)plumbing the surfaces into those like in figure
6820: \reference{fig3} (a), consisting of $k$ copies of $F_0'$, but
6821: now \em{without} the lower band $N$. (Figure \reference{fig4}
6822: shows the case $k=4$.) Then we use $A$-decomposition, as
6823: shown in figure \reference{fig6}. One applies $A$-operations
6824: along the arcs $a_{1,2}$. We show only the result for $a_1$,
6825: the case of $a_2$ and other $k$ is analogous. \qed
6826: 
6827: \end{appendix}
6828: 
6829: % \vspace{1cm}\mbox{}
6830: 
6831: % \tm
6832: {\small
6833: \begin{thebibliography}{BLM}
6834: \bibitem[Al]{Alexander} J. W.~Alexander, \em{A lemma on systems of
6835:         knotted curves}, Proc. Nat. Acad. Sci. U.S.A.
6836:         {\bf 9} (1923), 93--95.
6837: \bibitem[Al2]{Alexander2} \bysame, \em{Topological invariants
6838:         of knots and links}, Trans.\ Amer.\ Math.\ Soc. {\bf 30} (1928),
6839:         275--306.
6840: \bibitem[Ar]{Artin} E.~Artin, \em{Theorie der Z\"opfe}, 
6841:         Abh. Math. Sem. Hamburgischen Univ. {\bf 4} (1926), 47--72.
6842: \bibitem[Ar2]{Artin2} \bysame, \em{Theory of braids}, 
6843:         Ann. of Math. {\bf 48(2)} (1947), 101--126.
6844: \bibitem[Be]{Bennequin} D.~Bennequin, \em{Entrelacements et \'equations
6845: 	de Pfaff}, Soc.\ Math.\ de France, Ast\'erisque {\bf 107-108}
6846: 	(1983), 87--161.
6847: \bibitem[Bi]{Bigelow} S.~Bigelow, \em{Representations of braid
6848: 	groups}, Proceedings of the International Congress of
6849: 	Mathematicians, Vol. {\bf II} (Beijing, 2002),  37--45.
6850: \bibitem[B]{Birman} J.\ S.\ Birman, \em{On the Jones polynomial of
6851: 	closed $3$-braids}, Invent. Math. {\bf 81(2)} (1985), 287--294.
6852: \bibitem[BKL]{BKL} \bysame, K.\ Ko and S.\ J.\ Lee, \em{A new
6853: 	approach to the word and conjugacy problems in the braid
6854: 	groups}, Adv. Math. {\bf 139(2)} (1998), 322--353.
6855: \bibitem[BM]{BirMen} \bysame\ and W.~W.~Menasco, \em{Studying knots
6856: 	via braids III: Classifying knots which are closed 3 braids},
6857: 	Pacific J.~Math. {\bf 161}(1993), 25--113.
6858: \bibitem[BM2]{BirMen2} \bysame\ and \bysame\,, \em{Studying links
6859: 	via closed braids II: On a theorem of Bennequin},
6860: 	Topology Appl. {\bf 40(1)} (1991), 71--82.
6861: \bibitem[BM3]{BirMen3} \bysame\, and \bysame\,, \em{Studying knots
6862: 	via braids VI: A non-finiteness theorem},
6863: 	Pacific J.~Math. {\bf 156} (1992), 265--285.
6864: \bibitem[BW]{BW} \bysame\ and H.\ Wenzl, \em{Braids, link polynomials
6865: 	and a new algebra}, Trans. Amer. Math. Soc. {\bf 313(1)}
6866: 	(1989), 249--273.
6867: \bibitem[Bl]{Bleiler} S.~A.~Bleiler, \em{Realizing concordant
6868: 	polynomials with prime knots}, Pacific J. Math. {\bf 100(2)}
6869: 	(1982), 249--257.
6870: \bibitem[BLM]{BLM} R.~D.~Brandt, W. B. R. Lickorish and K. Millett,
6871: 	\em{A polynomial invariant for unoriented knots and links},
6872: 	Inv. Math. {\bf 74} (1986), 563--573.
6873: \bibitem[CK]{CK} A.~Champanerkar and I.~Kofman, \em{On the Mahler
6874: 	measure of Jones polynomials under twisting},
6875: 	Algebr. Geom. Topol. {\bf 5} (2005), 1--22.
6876: \bibitem[Cr]{Cromwell} P.~R.~Cromwell, {\em Homogeneous links},
6877: 	J. London Math. Soc. (series 2) {\bf 39} (1989), 535--552.
6878: \bibitem[DL]{DasLin} O.~Dasbach and X.-S.~Lin, {\em On the Head and
6879: 	the Tail of the Colored Jones Polynomial}, Compositio Math.
6880: 	{\bf 142(5)} (2006), 1332--1342.
6881: \bibitem[EKT]{EKT} S.\ Eliahou, L.\ H. Kauffman and M.\ Thistlethwaite,
6882: 	{\em Infinite families of links with trivial Jones polynomial},
6883: 	Topology {\bf 42(1)} (2003), 155--169.
6884: \bibitem[Fi]{Fied} T.~Fiedler, {\em A small state sum for knots},
6885: 	Topology {\bf 32~(2)} (1993), 281--294.
6886: \bibitem[Fi2]{Fied2} \bysame, {\em Gauss sum invariants for knots
6887: 	and links}, Kluwer Academic Publishers, Mathematics and Its
6888: 	Applications Vol {\bf 532} (2001).
6889: \bibitem[FK]{FK} \bysame\, and V.\ Kurlin, \em{A one-parameter
6890: 	approach to knot theory}, preprint \web|math.GT/0606381|.
6891: \bibitem[FW]{WilFr} J.\ Franks and R.\ F.\ Williams, \em{Braids and the
6892: 	Jones-Conway polynomial}, Trans.\ Amer.\ Math.\ Soc. {\bf 303}
6893: 	(1987), 97--108.
6894: \bibitem[F\&]{HOMFLY} P. Freyd, J. Hoste, W. B. R. Lickorish,
6895: 	K. Millett, A. Ocneanu and D. Yetter, {\it A new polynomial
6896: 	invariant of knots and links}, Bull. Amer. Math. Soc.
6897: 	{\bf 12} (1985), 239--246.
6898: \bibitem[Fu]{Fujii} H.~Fujii, \em{Geometric indices and the Alexander
6899: 	polynomial of a knot},
6900: 	Proc. Amer. Math. Soc. {\bf 124(9)} (1996), 2923--2933.
6901: \bibitem[Ga]{Gabai} D.~Gabai, \em{The Murasugi sum is a natural
6902: 	geometric operation}, Low-dimensional topology (San Francisco,
6903: 	Calif., 1981), Contemp. Math. {\bf 20}, 131--143, Amer. Math.
6904: 	Soc., Providence, RI, 1983.
6905: \bibitem[Ga2]{Gabai2} \bysame, \em{The Murasugi sum is a natural
6906: 	geometric operation II}, Combinatorial methods in
6907: 	topology and algebraic geometry (Rochester, N.Y., 1982),
6908: 	93--100, Contemp. Math. {\bf 44}, Amer. Math. Soc.,
6909: 	Providence, RI, 1985.
6910: \bibitem[Ga3]{Gabai3} \bysame, \em{Detecting fibred links in $S^3$},
6911:  	Comment. Math. Helv.  {\bf 61(4)} (1986), 519--555.
6912: \bibitem[Gr]{Garside} F.~Garside, \em{The braid group and other groups},
6913: 	Q.~J.~Math.~Oxford {\bf 20} (1969), 235--264.
6914: \bibitem[HM]{Hirasawa} M.~Hirasawa and K.~Murasugi, \em{Double-torus
6915: 	fibered knots and pre-fiber surfaces}, Musubime to Teijigen
6916: 	Topology (Dec. 1999), 43--49.
6917: \bibitem[HT]{KnotScape} J.~Hoste and M.~Thistlethwaite, {\em
6918: 	KnotScape}, a knot polynomial calculation and table access
6919: 	program, available at \webb:|http://www.math.utk.edu/\~morwen|.
6920: \bibitem[HTW]{HTW} \bysame, \bysame\, and J.~Weeks, \em{The first
6921: 	1,701,936 knots}, Math. Intell. {\bf 20 (4)} (1998), 33--48.
6922: \bibitem[I]{Ishikawa} M.~Ishikawa, {\em On the Thurston-Bennequin
6923: 	invariant of graph divide links}, Math. Proc. Cambridge
6924: 	Philos. Soc. {\bf 139(3)} (2005), 487--495.
6925: \bibitem[J]{Jones} V.~F.~R.~Jones, {\em Hecke algebra representations 
6926: 	of braid groups and link polynomials}, Ann. of Math.
6927: 	{\bf 126} (1987), 335--388.
6928: \bibitem[J2]{Jones2} \bysame, {\em A polynomial
6929: 	invariant of knots and links via von Neumann algebras},
6930: 	Bull. Amer. Math. Soc. {\bf 12} (1985), 103--111.
6931: \bibitem[Kk]{Kakimizu} O.~Kakimizu, \em{Classification of the
6932: 	incompressible spanning surfaces for prime knots of $\leq 10$
6933: 	crossings}, Hiroshima Math. J. {\bf 35} (2005), 47--92.
6934: \bibitem[K]{Kanenobu} T.~Kanenobu, \em{Relations between the
6935: 	Jones and Q polynomials of 2-bridge and 3-braid links},
6936: 	Math. Ann. {\bf 285} (1989), 115--124.
6937: \bibitem[K2]{Kanenobu2} \bysame, \em{Examples on polynomial invariants
6938: 	of knots and links II}, Osaka J. Math. {\bf 26(3)} (1989),
6939: 	465--482.
6940: \bibitem[K3]{Kanenobu3} \bysame, \em{Examples on polynomial invariants
6941: 	of knots and links}, Math. Ann. {\bf 275} (1986), 555--572.
6942: \bibitem[K4]{Kanenobu4} \bysame, \em{An evaluation of the first
6943: 	derivative of the $Q$ polynomial of a link},
6944: 	Kobe J. Math. {\bf 5(2)} (1988), 179--184.
6945: \bibitem[Ka]{Kauffman} L.\ H.\ Kauffman, {\em An invariant of regular
6946: 	isotopy}, Trans. Amer. Math. Soc. {\bf 318} (1990), 417--471.
6947: \bibitem[Ka2]{Kauffman2} \bysame, {\em State models and
6948: 	the Jones polynomial}, Topology {\bf 26} (1987), 395--407.
6949: \bibitem[Kw]{Kawamuro} K.~Kawamuro, \em{The algebraic crossing number
6950: 	and the braid index of knots and links},
6951: 	Algebr. Geom. Topol. {\bf 6} (2006), 2313--2350. 
6952: \bibitem[Ki]{Kidwell} M.~E.~Kidwell, \em{On the degree of the
6953: 	Brandt-Lickorish-Millett-Ho polynomial of a link},
6954: 	Proc. Amer. Math. Soc. {\bf 100(4)} (1987), 755--762.
6955: \bibitem[Kn]{Kneissler} J. A. Kneissler, \em{Woven braids and their
6956: 	closures}, J. Knot Theory Ramifications {\bf 8(2)} (1999),
6957: 	201--214.
6958: \bibitem[Ko]{Kobayashi} T.~Kobayashi, \em{Uniqueness of minimal
6959: 	genus Seifert surfaces for links},
6960: 	Topology Appl. {\bf 33(3)} (1989), 265--279.
6961: \bibitem[Kr]{Kreimer} D.\ Kreimer, \em{Knots and Feynman diagrams},
6962: 	Cambridge Lecture Notes in Physics {\bf 13},
6963: 	Cambridge University Press, Cambridge, 2000.
6964: \bibitem[LM]{LickMil} W. B. R. Lickorish and K.~C.~Millett, {\em A
6965: 	polynomial invariant for oriented links}, Topology
6966: 	{\bf 26 (1)} (1987), 107--141.
6967: \bibitem[LT]{LickThis} \bysame\, and M.~B.~Thistlethwaite,
6968: 	\em{Some links with non-trivial polynomials and their crossing
6969: 	numbers}, Comment. Math. Helv. {\bf 63} (1988), 527--539.
6970: \bibitem[MT]{MenThis} W.~W.~Menasco and M.~B.~Thistlethwaite, \em{%
6971:   	The Tait flyping conjecture}, Bull. Amer. Math. Soc. {\bf
6972:     	25 (2)} (1991), 403--412.
6973: \bibitem[Mo]{Morton} H.~R.~Morton, \em{Seifert circles and knot
6974: 	polynomials}, Proc. Camb. Phil. Soc. {\bf 99} (1986), 107--109.
6975: \bibitem[Mo2]{MortonPb} \bysame\ (ed.), {\em Problems},  in
6976: 	``Braids'', Santa Cruz, 1986 (J.~S.~Birman and A.~L.~Libgober,
6977: 	eds.), Contemp. Math. {\bf 78}, 557--574.
6978: \bibitem[MS]{MorSho}  \bysame\ and H. B. Short, \em{The
6979: 	$2$-variable polynomial of cable knots}, Math. Proc.
6980: 	Cambridge Philos. Soc. {\bf 101(2)} (1987), 267--278.
6981: \bibitem[MS2]{MorSho2} \bysame\ and \bysame, {\tt br9z.p}, a Pascal
6982: 	program for calculation of the skein polynomial from braids,
6983: 	\web|http://www.liv.ac.uk/~su14/knotprogs.html|.
6984: \bibitem[Mr]{Murakami} J.\ Murakami, \em{The Kauffman polynomial of
6985: 	links and representation theory},
6986: 	Osaka J. Math. {\bf 24(4)} (1987), 745--758.
6987: \bibitem[Mu]{Murasugi} K.~Murasugi, \em{On the braid index of
6988: 	alternating links}, Trans. Amer. Math. Soc. {\bf 326 (1)}
6989: 	(1991), 237--260.
6990: \bibitem[Mu2]{Murasugi2} \bysame, \em{On closed 3-braids},
6991:   	Memoirs AMS {\bf 151} (1974), AMS, Providence.
6992: \bibitem[Mu3]{Murasugi3} \bysame\,, \em{Jones polynomial and classical
6993: 	conjectures in knot theory}, Topology {\bf 26} (1987), 187--194.
6994: \bibitem[MP]{MurPrz} \bysame\ and J.~Przytycki, \em{The skein
6995: 	polynomial of a planar star product of two
6996: 	links}, Math. Proc. Cambridge Philos. Soc. {\bf 106(2)}
6997: 	(1989), 273--276.
6998: \bibitem[MP2]{MurPrz2} \bysame\ and \bysame, \em{An index of a graph
6999: 	with applications to knot theory}, Mem. Amer. Math. Soc.
7000: 	{\bf 106 (508)} (1993).
7001: \bibitem[Na]{Nakamura} T.~Nakamura, \em{Notes on the braid index of
7002: 	closed positive braids}, Topology Appl. {\bf 135(1-3)}
7003: 	(2004), 13--31.
7004: \bibitem[Na2]{Nakamura2} \bysame, \em{Braidzel surfaces and the
7005: 	Alexander polynomial}, Proceedings of the Workshop
7006: 	``Intelligence of Low Dimensional Topology'', Osaka
7007: 	City University (2004), 25--34.
7008: \bibitem[Ni]{Ni} Y.~Ni, \em{Closed 3-braids are nearly fibred},
7009: 	preprint \web|math.GT/0510243|.
7010: \bibitem[Oh]{Ohyama} Y.~Ohyama, \em{On the minimal crossing number
7011: 	and the braid index of links}, Canad. J. Math. {\bf 45(1)}
7012: 	(1993), 117--131.
7013: \bibitem[Or]{Orevkov} S.~Orevkov, \em{Quasipositivity problem for
7014: 	3-braids}, Turkish Journal of Math. {\bf 28} (2004),
7015: 	89--93; also available at
7016: 	\web|http://picard.ups-tlse.fr/homepage/orevkov.html|.
7017: \bibitem[PV]{VirPol} M.~Polyak and O.~Viro,
7018: 	{\em Gauss diagram formulas for Vassiliev invariants},
7019: 	Int.\ Math.\ Res.\  Notes {\bf 11} (1994), 445--454.
7020: \bibitem[PV2]{VirPol2} \bysame\, and \bysame, {\em On the Casson
7021: 	knot invariant}, Knots in Hellas '98, Vol. {\bf 3} (Delphi),
7022: 	J. Knot Theory Ramifications {\bf 10(5)} (2001), 711--738.
7023: \bibitem[Ro]{Rolfsen} D.~Rolfsen, {\em Knots and links}, Publish
7024: 	or Perish, 1976.
7025: \bibitem[Ru]{Rudolph} L.\ Rudolph, \em{Braided surfaces and
7026: 	Seifert ribbons for closed braids}, Comment. Math. Helv.
7027: 	{\bf 58} (1983), 1--37.
7028: \bibitem[Ru2]{Rudolph2} \bysame, \em{Quasipositivity as an
7029: 	obstruction to sliceness}, Bull. Amer. Math. Soc. (N.S.)
7030: 	{\bf 29(1)} (1993), 51--59.
7031: \bibitem[Sc]{Sc} O.~Schreier, \em{\"Uber die Gruppen $A^aB^b=1$},
7032: 	Abh. Math. Sem. Univ. Hamburg {\bf 3} (1924), 167--169.
7033: \bibitem[SSW]{SSW} D.~Silver, A.~Stoimenow and S.~G.~Williams, \em{%
7034: 	Euclidean Mahler measure and Twisted Links},
7035: 	\web|math.GT/0412513|,
7036: 	Algebr. Geom. Topol. {\bf 6} (2006), 581--602.
7037: \bibitem[SW]{SW} \bysame\, and S.~Williams, \em{Coloring link diagrams
7038: 	with a continuous palette}, Topology {\bf 39} (2000),
7039: 	1225--1237.
7040: \bibitem[Sq]{Squier} C.\ Squier, \em{The Burau representation is
7041:   	unitary}, Proc.\ Amer.\ Math.~Soc. {\bf 90} (1984), 199--202.
7042: \bibitem[St]{posqp} A.~Stoimenow, {\em On polynomials and surfaces of
7043: 	variously positive links}, \web|math.GT/0202226|,
7044: 	Jour. Europ. Math. Soc. {\bf 7(4)} (2005), 477--509.
7045: \bibitem[St2]{3br} \bysame, {\em The skein polynomial of closed
7046: 	3-braids}, J. Reine Angew. Math. {\bf 564} (2003), 167--180.
7047: \bibitem[St3]{posex_bcr} \bysame, {\em On the crossing number of
7048:    	positive knots and braids and braid index criteria of Jones
7049: 	and Morton-Williams-Franks},
7050:         Trans.\ Amer.\ Math.\ Soc. {\bf 354(10)} (2002), 3927--3954.
7051: \bibitem[St4]{ntriv} \bysame, {\em Coefficients and non-triviality
7052: 	of the Jones polynomial}, preprint {\tt math.GT/0606255}.
7053: \bibitem[St5]{gener} \bysame, {\em Diagram genus, generators and
7054: 	applications}, preprint.
7055: \bibitem[St6]{bind} \bysame, {\em The braid index and the growth of
7056: 	Vassiliev invariants}, J.~Of Knot Theory and Its Ram.
7057: 	{\bf 8(6)} (1999), 799--813.
7058: \bibitem[St7]{pos} \bysame, {\em Positive knots, closed braids,
7059: 	and the Jones polynomial}, {\tt math.GT/9805078}, Ann. Scuola
7060: 	Norm. Sup. Pisa Cl. Sci. {\bf 2(2)} (2003), 237--285.
7061: \bibitem[St8]{gen2}\bysame, {\em Knots of genus two}, preprint
7062: 	\web|math.GT/0303012|.
7063: \bibitem[SV]{SV} \bysame\ and A.\ Vdovina, {\em Counting alternating
7064:   	knots by genus}, Math.\ Ann. {\bf 333} (2005), 1--27.
7065: \bibitem[ST]{SunThis} C.~Sundberg and M.~B.~Thistlethwaite, {\em The
7066: 	rate of growth of the number of prime alternating links and
7067: 	tangles}, Pacific Journal of Math. {\bf 182 (2)} (1998),
7068: 	329--358.
7069: \bibitem[Th]{Thistle} M.~B.~Thistlethwaite, {\em On the Kauffman
7070: 	polynomial of an adequate link}, Invent.\ Math. {\bf 93(2)}
7071: 	(1988), 285--296.
7072: \bibitem[Th2]{Thistle2} \bysame, {\em A spanning tree expansion for the
7073: 	Jones polynomial}, Topology {\bf 26} (1987), 297--309.
7074: \bibitem[Tr]{Traczyk} P.~Traczyk, \em{$3$-braids with proportional
7075:   	Jones polynomials}, Kobe J. Math. {\bf 15(2)} (1998), 187--190.
7076: \bibitem[Vo]{Vogel} P.~Vogel, \em{Representation of links by braids:
7077: 	A new algorithm}, Comment. Math. Helv. {\bf 65} (1990),
7078: 	104--113.
7079: \bibitem[Wi]{Williams} R. F. Williams, \em{Lorenz knots are prime},
7080: 	Ergodic Theory Dynam. Systems  {\bf 4(1)} (1984), 147--163. 
7081: \bibitem[Wo]{Wolfram} S. Wolfram, {\em Mathematica --- a system
7082: 	for doing mathematics by computer}, Addison-Wesley, 1989.
7083: \bibitem[Xu]{Xu} Peijun Xu, {\em The genus of closed $3$-braids},
7084: 	J. Knot Theory Ramifications {\bf 1(3)} (1992), 303--326.
7085: \bibitem[Y]{Yamada} S.~Yamada, \em{The minimal number of Seifert
7086: 	circles equals the braid index}, Invent. Math. {\bf 88}
7087: 	(1987), 347--356.
7088: \bibitem[Yo]{Yokota} Y.\ Yokota, {\em Polynomial invariants of
7089: 	positive links}, Topology {\bf 31(4)} (1992), 805--811.
7090: \end{thebibliography}
7091: }
7092: 
7093: \end{document}
7094:  
7095: