1: \documentclass[11pt]{report}
2:
3: %\textwidth14cm
4: %
5: \textwidth16.3cm
6: \textheight23.3cm
7: \addtolength{\evensidemargin}{-1.9cm}
8: \addtolength{\oddsidemargin}{-1.9cm}
9: \addtolength{\topmargin}{-.9in}
10: %
11: %\def\baselinestretch{1.4}
12: %\pagestyle{empty}
13: \def \dfll {\leaders \hbox to 1em {\hss.\hss}\hfill}
14: \def\beq {\begin{equation}}
15: \def\eeq {\end{equation}}
16: \newtheorem{theorem}{Theorem}
17: \newtheorem{propos}[theorem]{Proposition}
18: \newtheorem{cor}[theorem]{Corollary}
19: \newtheorem{conjecture}[theorem]{Conjecture}
20: \newtheorem{rem}[theorem]{Remark}
21: \newtheorem{defi}{Definition}
22: \newtheorem{qes}{Question}
23: \newtheorem{lemma}{Lemma}
24: \def\blm{\begin{lemma}}
25: \def\elm{\end{lemma}}
26: \def\bdf{\begin{defi}}
27: \def\edf{\end{defi}}
28: \def\btm{\begin{theorem}}
29: \def\etm{\end{theorem}}
30: \def\bpp{\begin{propos}}
31: \def\epp{\end{propos}}
32: \def\bQ {\begin{qes}}
33: \def\eQ {\end{qes}}
34: \def\btm{\begin{theorem}}
35: \def\etm{\end{theorem}}
36: \def\ben{\begin{enumerate}}
37: \def\een{\end{enumerate}}
38: \def\into{\hookrightarrow}
39: %\def\bar{\begin{array}}
40: %\def\ear{\end{array}}
41: %\def\ep {\hfill{\rule{2.3mm}{2.3mm}}\medskip}
42: \def\ep{\ \hfill{\rule {2.5mm}{2.5mm}}\smallskip}
43: %\pagestyle{myheadings}
44: %\markboth{Thomas Kerler}{Homology TQFT's via Hopf algebras}
45:
46: %\newcommand{\lbl}[1]{\label{#1}\mbox{[[{\bf\footnotesize#1}]]}}
47:
48: \newcommand{\lbl}[1]{\label{#1}}
49:
50: \def \isto {\widetilde{\longrightarrow}}
51:
52: \def \Cob {{\cal C}ob_3}
53:
54: \newcounter{cob}
55:
56: \def\Sp{{\rm Sp}}
57: \def\SL{{\rm SL}}
58:
59:
60: \def\opc{{\scriptstyle =}}
61:
62: \def\drulefill{$\mathord\opc\mkern-3mu\cleaders
63: \hbox{$\mkern-2mu\mathord\opc\mkern-2mu$}\hfill\mkern-2mu \mathord\opc$}
64:
65:
66: \newcommand{\dov}[1]{\vbox{\ialign{##\crcr\noalign
67: {\kern-3pt\nointerlineskip}\drulefill \crcr\noalign
68: {\kern0pt\nointerlineskip}
69: $\hfil\displaystyle{#1}\hfil$\crcr}}}
70:
71:
72:
73: \def\zer#1{\vbox{\ialign{##\crcr\noalign
74: {\kern-3pt\nointerlineskip}
75: {$\hfil\;\,\scriptscriptstyle\oslash\hfil\!$} \crcr\noalign
76: {\kern.5pt\nointerlineskip}
77: $\hfil\displaystyle{#1}\hfil$\crcr}}\!}
78:
79: \def\ste#1{\vbox{\ialign{##\crcr\noalign
80: {\kern-3pt\nointerlineskip}
81: {$\hfil\;\,\scriptscriptstyle\bowtie\hfil\!$} \crcr\noalign
82: {\kern.5pt\nointerlineskip}
83: $\hfil\displaystyle{#1}\hfil$\crcr}}\!}
84:
85:
86:
87: \newcommand{\TO}[2]{\stackrel {\mbox{#1}}{\hbox to #2pt{\rightarrowfill}}}
88:
89: %\def\twoheadrightarrow{\to\mkern-20mu\to}
90: \def\thrafill{$\mathsurround=0pt \mathord- \mkern-6mu
91: \cleaders\hbox{$\mkern-2mu
92: \mathord- \mkern-2mu$}\hfill \mkern-6mu\mathord\twoheadrightarrow$}
93: \newcommand {\onto} [1]{\hbox to #1pt{\thrafill}}
94: \newcommand {\ONTO} [2]{\stackrel{\mbox{#1}}{\onto {#2}}}
95:
96: \newcommand{\emptystuff}[1]{}
97:
98:
99: %\def\ltimes {\mbox{\raisebox{.4ex}{$\scriptstyle \rhd\mkern-4mu<$}}}
100: %\def\ltimes {\mbox{$\rhd\!\!\!<$}}
101: % \def\rtimes {\mbox{\raisebox{.4ex}{$\scriptstyle >\mkern-4mu\lhd$}}}
102:
103:
104: \usepackage{amsfonts,
105: amssymb}
106: %\usepackage{diagrams}
107:
108: \input{epsf.sty}
109:
110:
111: \newcommand{\Z}{{\mathbb Z}}
112: \newcommand{\D}{{\mathbb D}}
113: \newcommand{\Ll}{{\mathbb L}}
114: \newcommand{\R}{{\mathbb R}}
115: \newcommand{\Cc}{{\mathbb C}}
116: \newcommand{\N}{{\mathbb N}}
117: \newcommand{\Q}{{\mathbb Q}}
118: \newcommand{\E}{{\mathbb E}}
119: \newcommand{\Ss}{{\mathbb S}}
120: \newcommand{\Aa}{{\mathbb A}}
121: \newcommand{\Pp}{{\mathbb P}}
122:
123: \newcommand{\kk}{{\mathbb K}}
124:
125: \newcommand{\posit}[2]
126: {\raise -1.4ex\hbox{${\textstyle #1}\atop {\stackrel{\uparrow}{#2}}$}}
127:
128:
129: \newcommand{\head}[1]{
130: \smallskip
131:
132: \begin{center}{\large \sc #1}\end{center}
133:
134: \nopagebreak}
135:
136: \newcommand {\prg}[2]{\vspace*{.5cm}\subparagraph{ #1)\ #2\ }\lll{pg-#1}\ \newline}
137:
138: \def\ub{\underline}
139: \def \A {\mbox{$\cal N\,$}}
140: \def \F {I\mkern -6.2mu F}
141: \def \V {\mbox{$\cal V\,$}}
142: \def \C {\mbox{$\cal C\,$}}
143: \def \Am {\mbox{${\cal N}\!-\!mod\,$}}
144: \def\id{{1\mkern-5mu {\rm I}}}
145: \def \li {\mbox{$\cal L\,$}}
146: \def \lz {\langle}
147: \def \rz {\rangle}
148:
149: \def \th {\theta}
150: \def \tb {\bar{\theta}}
151: \def \tx {{\theta^{\#}}}
152:
153: \def \sempro {\mbox{$\bigcirc\mkern-14mu s\,\,$}}
154:
155: \def \hopf {\mbox{$^0\bigcirc\mkern-13mu\bigcirc^0\,$}}
156: \def \nhopf {\mbox{$^0\bigcirc\mkern-13mu\bigcirc\,$}}
157: \def \twhopf {\mbox{$^0\bigcirc\mkern-13mu\bigcirc^1\,$}}
158:
159: \def\B{{\cal B}}
160: \def\Bo{{\B}^{\circ}}
161:
162: \def\Hh{\mbox{\bm${\cal H}$\ubm}}
163:
164: \def\rcoa{\leftharpoondown}
165: \def\lcoa{\rightharpoonup}
166: \def\rac{\!\triangleleft\!}
167: \def\lac{\!\triangleright\!}
168:
169: \newcommand{\ext}[1] {\mbox{\raisebox{.4ex}{$\bigwedge^{\!#1}$}}\mkern-1mu}
170:
171: \def\Ii {\mbox{\raise .4 ex\hbox{$\int$}$\!\! I$}}
172:
173: \def \tgl{\mbox{\boldmath${\cal T}\! gl$\unboldmath}}
174: \def \stA{\mbox{\boldmath$s\!{\cal T}$\unboldmath$({\cal N})$}}
175:
176: \def \bm {\boldmath}
177: \def \ubm {\unboldmath}
178: \newcommand{\Qq}
179: {\mbox{\bm${\cal Q}$\ubm} \,}
180:
181:
182:
183: \def \vc {Vect({\bf C})}
184: %\newcommand{\lll}[1]{\label{#1}\raisebox{1ex}{\mbox{\scriptsize [#1]$\;$}}}
185: \def\lll{\label}
186:
187:
188: \begin{document}
189:
190: \vspace*{1cm}
191:
192:
193:
194: \begin{center}
195:
196:
197: \section*{Homology TQFT's and the Alexander-Reidemeister
198: Invariant of 3-Manifolds via Hopf Algebras and Skein Theory }
199:
200: \bigskip
201:
202: \medskip
203:
204:
205: {\large Thomas Kerler}\\
206: \medskip
207:
208: May 2001
209: \vspace*{1.2cm}
210:
211: \bigskip
212:
213:
214:
215: \end{center}
216: %\vspace*{1.7cm}
217:
218:
219: {\small \noindent{\bf Abstract :}
220: We develop an explicit skein theoretical algorithm to compute
221: the Alexander polynomial of a 3-manifold from a
222: surgery presentation employing the methods used in
223: the construction of quantum invariants of 3-manifolds.
224: As a prerequisite we establish and prove a rather unexpected
225: equivalence between the topological quantum field
226: theory constructed by Frohman and Nicas using the
227: homology of $U(1)$-representation varieties
228: on the one side and the combinatorially constructed Hennings-TQFT based on the
229: quasitriangular Hopf algebra ${\cal N}=\Z/2\ltimes \ext *\R^2$ on the
230: other side. We find that both TQFT's are $\SL(2,\R)$-equivariant functors and,
231: as such, are isomorphic. The
232: $\SL(2,\R)$-action in the Hennings construction comes from the natural
233: action on $\cal N$ and in the case of the Frohman-Nicas theory from the
234: Hard-Lefschetz decomposition of the $U(1)$-moduli spaces given that they
235: are naturally
236: K\"ahler. The irreducible components of this TQFT, corresponding to simple
237: representations of $\SL(2,\Z)$ and $\Sp(2g,\Z)$, thus yield a large family of
238: homological TQFT's by taking sums and products.
239: We give several examples of TQFT's and invariants
240: that appear to fit into this family, such as Milnor and Reidemeister Torsion,
241: Seiberg-Witten theories, Casson type theories for homology circles \'a la Donaldson,
242: higher rank gauge theories following Frohman and Nicas, and the $\Z/p\Z$ reductions of
243: Reshetikhin-Turaev theories over the cyclotomic integers $\Z[\zeta_p]$. We also
244: conjecture that the Hennings TQFT for quantum-${\mathfrak s}{\mathfrak l}_2$
245: is the product of the Reshetikhin-Turaev TQFT and such a homological TQFT.
246: \footnote{
247: 2000 Mathematics Subject Classification: Primary 57R56; Secondary 14D20, 16W30, 17B37, 18D35, 57M27.}
248:
249: }
250: \medskip
251:
252: \begin{center}
253: {\large \sc Contents}\
254: \smallskip
255:
256: \parbox[t]{12cm}{
257: 1. Introduction\dotfill\pageref{S1}
258:
259: 2. Topological Quantum Field Theory\dotfill\pageref{S2}
260:
261: 3. The Frohman-Nicas TQFT for $U(1)$\dotfill\pageref{S3}
262:
263: 4. The Mapping Class Groups and their Actions on Homology\dotfill\pageref{S4}
264:
265: 5. Hennings TQFT's\dotfill\pageref{S5}
266:
267: 6. The Algebra $\cal N$\dotfill\pageref{S6}
268:
269: 7. The Hennings TQFT for $\cal N$\dotfill\pageref{S7}
270:
271: 8. Skein theory for ${\cal V}_{\cal N}$\dotfill\pageref{S8}
272:
273: 9. Equivalence of ${\cal V}_{\cal N}^{(2)}$ and ${\cal V}^{FN}$\dotfill\pageref{S9}
274:
275: 10. Hard-Lefschetz Decomposition\dotfill\pageref{S10}
276:
277: 11. Alexander-Conway Calculus for 3-Manifolds\dotfill\pageref{S11}
278:
279: 12. Lefschetz compatible Hopf Algebra Structures on $H^*(J(\Sigma))$\dotfill\pageref{S12}
280:
281: 13. More Examples of Homological TQFT's and Open Questions\dotfill\pageref{S13}
282: }
283: \end{center}
284:
285:
286: \head{1. Introduction}\lbl{S1}
287:
288:
289:
290:
291: In recent years much energy has been put into finding new ways to
292: describe and compute classical invariants of
293: 3-manifolds using the tools and structures developed in the relatively new area
294: of quantum topology. In this paper we will establish another such relation between
295: quantum and classical invariants, which emerged in quite different guises in recent
296: research in 3-dimensional topology.
297:
298:
299: The classical invariant of a 3-manifold $M$ we are interested in here is its
300: Alexander polynomial $\Delta(M)\in\Z[H_1(M)]$. It is closely related and in most cases identical
301: to the Reidemeister Milnor Torsion $r(M)$, see \cite{Mil61} and \cite{Tur76}. More recently,
302: Meng and Taubes \cite{MengTaub} show that this invariant is also equal to the
303: Seiberg Witten invariant for 3-manifolds. Turaev \cite{Tur98} proves a refined version of this
304: theorem by comparing the behavior of both invariants under surgery.
305:
306:
307: On the side of the quantum invariants we consider the formalism used for the
308: Hennings invariant of 3-manifolds \cite{Hen96}. This invariant is motivated by
309: and follows the same principles as the Witten-Reshetikhin-Turaev invariant,
310: which is developed in \cite{Wi89}, \cite{ResTur91} and \cite{TurVi92}, in the sense
311: that it
312: assigns algebraic data to a surgery presentation for $M$. The innovation of the Hennings
313: approach is that it starts
314: directly from a possibly non-semisimple Hopf algebra ${\cal A}$ rather than its
315: semisimple representation theory. This formalism is refined by Kauffman and Radford in
316: \cite{KauRad95}. Also Kuperberg \cite{Kup96} gives a construction that assigns data
317: directly from a Hopf algebra to a Heegaard presentation of $M$.
318:
319:
320: In this article we discover and explain in detail the relation between the Hennings theory
321: for a certain
322: 8-dimensional Hopf algebra $\cal N$ and the (reduced) Alexander polynomial
323: $\Delta_{\varphi}(M)\in \Z[t,t^{-1}]$ for the cyclic covering given by an epimorphism
324: $\varphi:\pi_1(M)\to\Z$. As a consequence we have at our disposal
325: the entire combinatorial machinery of the Hennings formalism
326: in order to evaluate the Alexander polynomial from surgery
327: diagrams. Particularly, we are able to develop from this an efficient skein theoretical
328: algorithm.
329: The method of relating these two very differently defined theories is based
330: itself on a quite unexpected equivalence of more refined structures.
331:
332: More precisely, it turns out that underlying both invariants is the structure of a
333: topological quantum field theory (TQFT). The notion of a TQFT, which can be thought
334: of as a fiber functor on a category of cobordisms, was first cast into a mathematical
335: axiomatic framework by Atiyah \cite{Ati88}. Typically (or by definition)
336: all quantum invariants extend to TQFT's on 3-manifolds with boundaries.
337: In the case of the semisimple theories generalizing the Witten-Reshetikhin-Turaev
338: invariant these TQFT's are described in great detail in \cite{Tur94}. In our
339: context we need the non-semisimple version as it is worked out for the Hennings invariant
340: in \cite{Ker96} and in full generality in \cite{KerLub00}.
341:
342:
343: On the side of the classical invariants Frohman and Nicas \cite{FroNic92}
344: managed to give an interpretation of the Alexander polynomial of knot complements
345: in the setting of TQFT's. In particular, they
346: construct a TQFT ${\cal V}^{FN}$, which assigns to every surface $\Sigma$ as a vector
347: space the cohomology ring $H^*(J(\Sigma))$
348: of the $U(1)$-representation variety $J(\Sigma)=Hom(\pi_1(\Sigma),U(1))$.
349: The morphisms are constructed in the style of the Casson invariant from the intersection
350: numbers of representation varieties for a given Heegaard splitting of a cobordism. The
351: Alexander polynomial is thus given as the Lefschetz trace over ${\cal V}^{FN}(C_{\Sigma})$,
352: where $\Sigma$ is an arbitrary Seifert surface and $C_{\Sigma}$ is the 3-dimensional
353: cobordisms from $\Sigma$ to itself, obtained by cutting away a neighborhood of $\Sigma$.
354:
355:
356: The unexpected upshot is that this functor ${\cal V}^{FN}$ is isomorphic to the
357: Hennings TQFT ${\cal V}_{\cal N}$ for the non-semisimple Hopf algebra
358: ${\cal N}\cong \Z/2\ltimes \ext *\R^2$. The realization of the abelian gauge
359: field theory by a specific Hopf algebra
360: is not at all obvious since ${\cal V}^{FN}$ and ${\cal V}_{\cal N}$ are
361: defined in entirely different ways. In fact the isomorphism between these functors
362: on the vectors spaces mixes up the degrees of exteriors algebras in still
363: puzzling ways. For these reason the proof is rather explicit and computational.
364:
365: Nonetheless, it can be seen quite easily that it is not possible to realize
366: ${\cal V}^{FN}$ as a semisimple theory.
367: Particularly, ${\cal V}^{FN}$ represents Dehn twists by
368: matrices of the form $1+N$ where $N$ is nilpotent. Furthermore, the invariant vanishes
369: on $S^1\times S^2$. Yet, in the semisimple
370: theories from \cite{Tur94} Dehn twists are represented by
371: semisimple matrices $D$ with
372: $D^n=1$ and the invariant on $S^1\times S^2$ is never zero.
373:
374:
375:
376:
377: Once ${\cal V}^{FN}$ and thus the Alexander polynomial $\Delta_{\varphi}$
378: are translated into the language of the Hennings formalism for the
379: Hopf algebra ${\cal N}$ we are in the position to develop a skein theory
380: for the computation of $\Delta_{\varphi}$. The skein identities reflect
381: algebraic relations in ${\cal N}$. We derive from this a step by step recipe
382: for the computation of the Alexander polynomial.
383:
384:
385: Another intriguing feature of the two TQFT's is that both of them admit natural
386: equivariant $\SL(2,\R)$-actions that have very different origins but are,
387: nevertheless, intertwined by the isomorphism between them. In the case of ${\cal V}^{FN}$ the
388: $\SL(2,\R)$-action on $H^*(J(\Sigma))$
389: is given by the Hard Lefschetz decomposition of the cohomology ring
390: that arises from a K\"ahler structure on $J(\Sigma)$. For ${\cal V}_{\cal N}$
391: this action is derived from an $\SL(2,\R)$-actions on ${\cal N}$ as a Hopf algebra.
392: As a consequence $H^*(J(\Sigma))$
393: carries a nonstandard ring-structure induced by that of ${\cal N}^{\otimes g}$,
394: which, as opposed to the standard one,
395: is compatible with the Hard Lefschetz $\SL(2,\R)$-action.
396: \medskip
397:
398: Let us summarize the content and the
399: main results of this paper in better order and detail. In Section~2
400: we recall relevant
401: notions that characterize
402: topological quantum field theories, such as (non)semisimplicity. Section~3 reviews
403: the construction of the functor ${\cal V}^{FN}$
404: of Frohman and Nicas and its values on basic cobordisms. In Section~4 we
405: describe a convenient set of
406: generators of the mapping class groups as combinations of Dehn twists and tangles,
407: and determine
408: their actions on homology. Section~5 introduces the basic rules for the construction of
409: a Hennings TQFT as well as a method that allows us to construct TQFT's even from
410: {\em non-modular}
411: Hopf algebras or categories. In Section~6 we give the precise definition of $\cal N$ as
412: a quasi triangular Hopf algebra in the sense of Drinfel'd together with the
413: $\SL(2,\R)$-action on it. The vector spaces and the basic morphisms of the associated Hennings
414: TQFT are computed in Section~7 using standard tangle presentations.
415: We prove $\SL(2,\R)$-covariance and single out an index 2
416: subcategory of framed cobordisms that naturally
417: yields a {\em real} valued TQFT. For later applications
418: we also determine the {\em categorical}
419: Hopf algebra that is canonically associated to this TQFT.
420: The nilpotent braided structure of $\cal N$ is then used in Section~8 to develop
421: a skein theory for the evaluation of tangle diagrams. The pivotal equivalence
422: of TQFT's that relates this theory to the Alexander polynomial is given by a natural
423: isomorphism of functors as follows. This is
424: proven in Section~9 by explicit comparison of generating morphism.
425:
426: \begin{theorem}\lbl{thm-main}
427: There is an $\SL(2,\R)$-equivariant isomorphism
428: $$
429: \xi\;:\;\;\; {\cal V}_{\cal N}^{(2)}
430: \;\;\stackrel{\bullet\,\,\cong}{-\!\!\!-\!\!\!-\-\!\!\!\!\!\longrightarrow}
431: \;\;{\cal V}^{FN}\;\;,
432: $$
433: where both TQFT's are ``non-semisimple'',
434: $\Z/2\Z$-projective functors from the category $\Cob^{\bullet}$
435: of surfaces with one boundary component and relative cobordisms to the category
436: of real $\SL(2,\R)$-modules.
437: \end{theorem}
438:
439: The Hard Lefschetz $\SL(2,\R)$ action on the cohomology of the $U(1)$ moduli spaces and
440: its covariance with ${\cal V}^{FN}$ are described more precisely in Section~10. The fact
441: that $\xi$ is an $\SL(2,\R)$-equivariant transformation
442: is proven. Moreover, we describe the canonical decompositions of the TQFT and the
443: Alexander polynomial according to their dual $\SL(2,\R)$-representations. The summands
444: are irreducible TQFT's for which the mapping class groups are represented by fundamental
445: weight representations of the symplectic groups $\Sp(2g,\Z)$. In
446: Section~11 we use the equivalence from Section~9 and the skein theory for tangles
447: from Section~12 to lay out an explicit algorithm, based on a skein theory that extends
448: the Alexander-Conway calculus, for the computation of $\Delta_{\varphi}(M)$.
449:
450: \begin{theorem}\label{them-skeintheory}
451: Let $\cal L$ be a framed link and $\cal Z\subset \cal L$ a distinguished
452: component that has zero framing and algebraic linking number zero with all other components.
453: Let $M_{\cal L}$
454: be the 3-manifold obtained by surgery along ${\cal L}$ and $\varphi_{\cal Z}:\pi_1(M)\to\Z$
455: the linking number with $\cal Z$.
456:
457: Then $\Delta_{\varphi_{\cal Z}}(M_{\cal L})\in\Z[t,t^{-1}]$
458: can be computed systematically as follows:
459: \begin{itemize}
460: \item Use the skein relations from Proposition~\ref{propos-Yskein} to unknot the special
461: strand $\cal Z$.
462: \item Put the new configuration into a standard form as depicted in Figure~\ref{fig-stan},
463: yielding a tangle $\cal T$.
464: \item Use the skein relations from Theorem~\ref{thm-skein} and framing relations from
465: Figure~\ref{fig-frame} to decompose ${\cal T}^{\#}$ into elementary diagrams as
466: described in in Theorem~\ref{thm-solve}.
467: \item Translate the elementary tangle diagrams into Hopf algebra diagrams as in
468: (\ref{eq-symhopfeva}).
469: \item Go through the steps of Proposition~\ref{prop-HopfAlexeval} to assign
470: polynomials to each component of a diagram.
471: \item Take products over components and sums over elementary diagrams.
472: \end{itemize}
473:
474: \end{theorem}
475:
476: The calculus described here for the evaluation of tangle diagrams is precisely the
477: one used to compute the morphisms for the TQFT functors from Theorem~\ref{thm-main}
478: via tangle surgery presentations of cobordisms.
479:
480:
481: Another application of the equivalence established in
482: Theorem~\ref{thm-main} arises from the observation
483: that every TQFT ${\cal V}$ on $\Cob^{\bullet}$ naturally implies a braided
484: Hopf algebra structure ${\cal H}_{\cal V}$ on ${\cal N}_0:={\cal V}(\Sigma_{1,1})$.
485: Now, the cohomology ring $H^*(J(\Sigma_g, U(1)))\cong\ext *H_1(\Sigma_g)$ already has
486: a canonical structure ${\cal H}_{ext}$ of a $\Z/2$-graded Hopf algebra induced by
487: the group structure on $J(\Sigma_g, U(1))$. It is easy to see that
488: ${\cal H}_{ext}$ is {\em not} compatible with the Lefschetz $\SL(2,\R)$-action.
489: However, the braided Hopf algebra
490: structure ${\cal H}_{{\cal V}^{FN}}$ inherited from the TQFT's in
491: Theorem~\ref{thm-main} is naturally $\SL(2,\R)$-variant, and, furthermore,
492: equivalent to ${\cal H}_{ext}$:
493:
494: \begin{theorem}\lbl{thm-structure}
495: For any choice of an integral Lagrangian decomposition, $H_1(\Sigma_g,\Z)=\Lambda\oplus\Lambda^*$,
496: and volume forms, $\omega_{\Lambda}\in\ext g\Lambda$ and $\omega_{\Lambda^*}\in\ext g\Lambda^*$,
497: the space
498: $H^*(J(\Sigma_g))$ admits a canonical structure ${\cal H}_{\Lambda}$ of a $\Z/2$-graded
499: Hopf algebra. It coincides with the braided Hopf algebra structure induced
500: by ${\cal V}^{FN}$ and is isomorphic to the canonical structure ${\cal H}_{ext}$.
501:
502: In particular,
503: $(H^*(J(\Sigma_g)),{\cal H}_{\Lambda})$ is commutative and cocommutative in the graded
504: sense, with unit $\omega_{\Lambda^*}$, integral $\omega_{\Lambda}$, and primitive elements given by
505: $a\wedge\omega_{\Lambda^*}$ and $i^*_z\omega_{\Lambda^*}$ for $a\in H_1(\Sigma)$ and $z\in H^1(\Sigma)$.
506:
507: The structure ${\cal H}_{\Lambda}$ is, furthermore, compatible with the
508: Hard-Lefschetz $\SL(2,\R)$-action. Specifically, this action is the Howe dual to
509: the action of $\SL(g,\Z)$ on the Lagrangian subspace in the group of Hopf
510: automorphisms:
511: $$
512: \SL(2,\R)_{Lefsch.}\times \SL(\Lambda)\,\;\subset \;\, {\rm GL}(2g,\R)= Aut(H^*(J(\Sigma_g)),{\cal H}_{\Lambda})
513: $$
514: \end{theorem}
515:
516: In Section~13 we discuss the appearance of these TQFT's in other
517: contexts. To this end let us denote by
518: ${\cal V}^{(j)}$ the irreducible component of ${\cal V}^{FN}$ dual to the $j$-dimensional
519: $\SL(2,\R)$-representation. A detailed description of it is given in Theorem~\ref{cor-dec}.
520: Choose for a closed 3-manifold $M$ with Betti number
521: $\beta_1(M)\geq 1$ a surjection $\varphi:H_1(M)\onto{13}\Z$
522: (which would be canonical for homology circles as given by 0-surgeries on knots). A series
523: of invariants for the pair $(M,\varphi)$ can now be constructed by choosing any two-sided,
524: embedded surface $\Sigma \subset M$ that is dual to $\varphi$, and considering the cobordism
525: $C_{\Sigma}:\,\Sigma\to\Sigma$ obtained by removing an open tubular neighborhood of $\Sigma$
526: from $M$. The {\em $j$-th (fundamental) Alexander Character} is now defined to be the integer
527: \beq\label{eq-AlexMom}
528: \Delta^{(j)}_{\varphi}(M)\;\;=\;\;trace\Bigl({\cal V}^{(j)}(C_{\Sigma})\Bigr)\;,
529: \eeq
530: which is easily seen to depend only on $\varphi$ but not the choice of $\Sigma$.
531: Besides the Alexander Polynomial also two other invariants invariant $I^{SW}$ and
532: $I^{DC}$ depending this data have been constructed by Donaldson in \cite{Don99}
533: from a Seiberg-Witten Theory and an $SO(3)$-Casson-type gauge theory respectively.
534: Let us also denote by $\lambda_L$ the Lescop Invariant \cite{Les}. As specified in
535: the next theorem all of these invariants are in fact linear combinations
536: of the (fundamental) Alexander Characters.
537:
538: \begin{theorem}[{\rm\small mostly corollaries to} {\cite{FroNic92},
539: \cite{Don99},\cite{Les},\cite{KerKyoto}}]\label{thm-relations}
540: \begin{eqnarray}
541: \Delta_{\varphi}(M)\;\;&=&\;\quad\sum_{j\geq 1}\,[j]_{-t}\cdot \Delta^{(j)}_{\varphi}(M)\label{eq-AlexChar}\;
542: \\
543: &&\nonumber\\
544: I^{DC}_{\varphi}(M)\;\;&=&\;\;\;\sum_{j\geq 2}\,{{j+1}\choose 3}\cdot \Delta^{(j)}_{\varphi}(M)\label{eq-DCChar}\;
545: \\
546: &&\nonumber\\
547: I^{SW}_{d,\varphi}(M)\;\;&=&\;\;\sum_{j\geq d+2}\,\left[\!\!{\left[
548: \Bigl({\frac{j-d}2}\Bigr)^2\right]}\!\!\right]\cdot
549: \Delta^{(j)}_{\varphi}(M)\label{eq-SWChar}
550: \\
551: &&\nonumber\\
552: \lambda_{L}(M)\;\;&=&\;\;\sum_{j\geq 1} (-1)^{j-1}\frac {j(2j^2-3)}{12}\cdot
553: \Delta^{(j)}_{\varphi}(M)\;\label{eq-LesChar}
554: \end{eqnarray}
555: \end{theorem}
556: Here we denoted $[j]_q=\frac {q^j-q^{-j}}{q-q^{-1}}$ and by $\bigl[\![x]\!\bigr]$ the
557: largest integer $\leq x$. We further review in how far the higher $PSU(n)$ knot invariants
558: $I_{k,n,\varphi}^{FN}$
559: of Frohman and Nicas \cite{FroNic94} come out to be
560: polynomial expressions in the Alexander Characters.
561: As products of characters are associated to tensor products of TQFT's and their decompositions
562: into irreducible components it is natural to consider the corresponding higher,
563: irreducible Alexander Characters $\Delta^{(\gamma)}$. We conjecture that the
564: $I_{k,n,\varphi}^{FN}$ are linear combinations of the $\Delta^{(\gamma)}$ with
565: coefficients in ${\mathbb N}\cup\{0\}$ as it is the case for $I^{DC}$ and $I^{SW}$.
566:
567: %%%%%%%%%%%%%%%% BEGIN CUT %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
568:
569: \emptystuff{
570: From this sequence of TQFT's we are able to construct a large family of homological
571: TQFT functors ${\cal V}^{(P)}$ for
572: suitable polynomials $P$ in variables $x_0, x_1, \ldots$ by taking tensor products and direct sums
573: accordingly.
574:
575: One example of such a TQFT ${\cal V}^{(Y)}$ with $Y=x_0+x_3-x_5-x_8+x_{10}+\ldots\,$
576: taken over $\Z/5$ turns out describe the lowest order contribution
577: of the Reshetikhin Turaev invariant taken over the cyclotomic integers $\Z[\zeta_5]$ at least
578: in small genera and very likely in general \cite{Kerr=5}. A rather interesting application that
579: emerges from that is found in in joint work with Gilmer \cite{GilKer}. Namely, that non triviality
580: of the Alexander polynomial evaluated at a 5-th root of unity implies that a 3-manifold disconnects
581: if a surface with more than one component is removed from it.
582:
583: The polynomial homological TQFT's appear to be also isomorphic to ones
584: constructed by gauge theoretical means
585: using an approach similar to the Casson invariant or extending the methods
586: of Seiberg-Witten theory as described by Donaldson. Furthermore, Frohman and Nicas consider
587: generalizations to higher rank Lie groups. In all cases the Alexander polynomial appears
588: as the dominant invariant, suggesting the corresponding decomposition of the TQFT into
589: the basic functors ${\cal V}^{(j)}$.
590: }
591:
592: %%%%%%%%%%%%%%%% END CUT %%%%%%%%%%%%%%%%%%%%%%%%
593:
594: Moreover, we explain the irreducible $p$-modular reductions $\dov{\cal V}^{(j)}_p$
595: over ${\mathbb F}_p=\Z/p\Z$ of the ${\cal V}^{(j)}$ relate to the irreducible factors
596: of the $\Z[\zeta_p]\to{\mathbb F}_p$ of the
597: Reshetikhin Turaev TQFT's at a $p$-th root of unity $\zeta_p$.
598: We finally give evidence that the TQFT from Theorem~\ref{thm-main}
599: is essentially the missing tensor factor that relates the semisimple and the non-semisimple
600: TQFT constructions for $U_q({\mathfrak s}{\mathfrak l}_2)$
601: following Reshetikhin Turaev and Hennings respectively.
602:
603: \medskip
604:
605:
606: \paragraph{Acknowledgements:} I am indebted to Charlie Frohman for making me aware
607: of \cite{FroNic92} and explaining \cite{FroNic94} to me. I thank
608: Bernhard Kr\"otz for discussions about Howe pairs, and David Radford for helping me find
609: his example in \cite{Rad76}. Thanks also to Andrew Nicas and Hans Boden for their
610: interest and Pierre Deligne, Daniel Huybrechts, and Manfred Lehn
611: for discussions about Lefschetz decompositions in the higher rank case.
612: Finally, I want to thank Razvan Gelca, Pat Gilmer, Jozef Przytycki
613: for opportunities to speak about this paper, when it was still in its
614: early stages.
615: \newpage
616:
617: \head{2. Topological Quantum Field Theory}\lbl{S2}
618: We start with the definition of a TQFT as a functor as proposed by
619: Atiyah \cite{Ati88}, largely suppressing a more detailed discussion
620: of the tensor structures.
621:
622:
623: For every integer, $g\geq 0$, choose a compact, oriented
624: model surface, $\Sigma_g$, of genus $g\,$, and to a tuple of
625: integers $\underline g=(g_1,\ldots,g_n)$ associate the ordered union
626: $\Sigma_{\underline g}:=\Sigma_{g_1}\sqcup\ldots\sqcup\Sigma_{g_n}$.
627: A {\em cobordism}
628: is a collection, ${\bf M}=(M,\phi_{\#},\Sigma_{g_{\#}})$, of the following:
629:
630:
631: A compact, oriented 3-manifold, $M$, whose boundary is divided into
632: two components
633: $\partial M=-\partial_{in} M\sqcup\partial_{out} M$,
634: two standard surfaces $\Sigma_{\underline g_{in}}$ and
635: $\Sigma_{\underline g_{out}}$, and two orientation preserving
636: homeomorphisms $\phi_{in}:\Sigma_{\underline g_{in}}
637: \,\isto\, \partial_{in} M\,$ and $\phi_{out}:\Sigma_{\underline g_{out}}
638: \,\isto\, \partial_{out} M\,$.
639:
640: We say two cobordisms, ${\bf M}$ and ${\bf M}'$, are equivalent if
641: they have the same "in" and "out" standard surfaces, and
642: there is a homeomorphism $h:\,M\,\isto \, M'\,$, such that
643: $h\circ \phi_{\#}=\phi_{\#}'\,$.
644:
645: Let $\Cob$ be the category of cobordisms in dimension 2+1, which has
646: the standard surfaces as objects and equivalence classes of cobordisms
647: as morphism. The composition of morphisms is defined via gluing over
648: boundary components using the coordinate maps to the same standard
649: surfaces. In addition, $\Cob$ has a tensor
650: product given by disjoint unions of surfaces and cobordisms.
651:
652:
653: A {\em Topological Quantum Field Theory} (TQFT) is a functor,
654: ${\cal V}:\Cob\longrightarrow{\rm Vect}(\kk)$, from the category of cobordisms
655: to the category of vector spaces over a field $\kk\,$.
656: \medskip
657:
658: Let us recall next some generalizations of the definition given in
659: \cite{Ati88} that will be relevant for our purposes.
660: By $\Cob^{2fr}$ we denote the category of 2-framed cobordisms, where we
661: fixed some standard framings on the model surfaces $\Sigma_g\,$, see \cite{Ker99}. A
662: {\em 2-framed TQFT} is now a functor
663: ${\cal V}:\Cob^{2fr}\longrightarrow{\rm Vect}(\kk)$. The category of 2-framed
664: cobordisms can be understood as a central extensions
665: \beq\label{eq-cobext}
666: \,1\;\to\;\Z\;\longrightarrow\;\Cob^{2fr}\;\longrightarrow\;\Cob\;\to \;1\,
667: \eeq
668: of the
669: ordinary cobordism category, if restricted to connected cobordisms. Hence, an
670: irreducible 2-framed TQFT yields a {\em projective TQFT}
671: since $\Z$ is presented as a scalar.
672: See \cite{Ker99} for further descriptions of this extension in terms of signatures of
673: bounding 4-manifolds.
674:
675:
676: For a group, $G$, we introduce the notion of a {\em $G$-equivariant TQFT}.
677: It is a functor,
678: ${\cal V}:\Cob\longrightarrow G-{\rm mod}_{\kk}$, from the category of
679: cobordisms to the category of finite dimensional $G$-modules over a field $\kk$. This means
680: that the linear map associated to any cobordism commutes with
681: the action of $G$ on the vector spaces of the respective boundary
682: components.
683:
684: Recall also from \cite{Ker98} that a {\em half-projective} or {\em non-semisimple}
685: TQFT is one in which functoriality is weakened and replaced by
686: the composition law ${\cal V}(MN)=0^{\mu(M,N)}{\cal V}(M){\cal V}(N)\,$.
687: Here $\mu(M,N)=b(MN)-b(M)-b(N)\in\Z^{+,0}\,$, where $b(M)$ is the
688: number of components of $M$ minus half the number of components of
689: $\partial M$. Note that $0^0=1$.
690:
691: We often call a cobordism for which all (rational) homology comes from the homology
692: of the boundary {\em (rationally) homologically trivial} (r.h.t). More precisely, we mean by this
693: that $i_*:H_1(\partial M,\Q)\to H_1(M,\Q)$ is onto. Typical examples of r.h.t.
694: cobordisms are the ones in (\ref{eq-defpsi}) and (\ref{eq-hdlcob})
695: below and closed, rational homology spheres. Examples of cobordisms that are not
696: r.h.t. are any connected sums with closed manifolds $M$ with $\beta_1(M)\geq 1$.
697: We find the following vanishing property:
698: \begin{lemma}[\cite{Ker98}]\lbl{lm-vanish}
699: If ${\cal V}$ is a non-semisimple TQFT, then for any cobordism $M$,
700: $$
701: {\rm if}\quad {\cal V}(M)\, \neq \, 0\qquad{\rm then}\qquad
702: M\;\;\;\mbox{is r.h.t.}
703: $$
704: \end{lemma}
705:
706: We further introduce $\Cob^{\bullet}$, the category of cobordisms, for which
707: the surfaces are connected and have exactly one boundary component. As objects we
708: thus use model surfaces $\Sigma_{g,1}$, such that $\Sigma_{g+1,1}$ is obtained
709: from $\Sigma_{g,1}$ by gluing in a torus, $\Sigma_{1,2}$, with two boundary
710: components. Thus, we have a presentation
711: \begin{equation}\lbl{eq-modelsurf}
712: \Sigma_{g,1}\;=\;\underbrace{\Sigma_1\#\ldots\#\Sigma_1\#\Sigma_{1,1}}_g
713: \qquad\qquad
714: \mbox{with inclusions}\qquad\quad \Sigma_{g,1}\;\subset\;\Sigma_{g+1,1}\;.
715: \end{equation}
716: Instead of ordinary cobordisms we then consider {\em relative} ones. We finally introduce
717: categories of cobordisms with
718: combinations of these properties such as $\Cob^{2fr, \bullet}$, the category of
719: 2-framed, relative cobordisms.
720: \medskip
721:
722: For any homeomorphsim, $\psi\in Homeo^+(\Sigma_g)$, of a surface to itself
723: we define the cobordism
724: \beq\lbl{eq-defpsi}
725: {\bf I}_\psi\;\;=\;\;(\Sigma_g\times[0,1], id\sqcup \psi, \Sigma_g\sqcup \Sigma_g)\quad.
726: \eeq
727: The morphism $[{\bf I}_{\psi}]$
728: depends only on the isotopy class $\{\psi\}$ of $\psi$, and the resulting map
729: $\Gamma_g\to Aut(\Sigma_g):\{\psi\}\mapsto [{\bf I}_{\psi}]$ from the mapping class
730: group to the group of invertible
731: cobordisms on $\Sigma_g$ is an isomorphism, see \cite{KerLub00}. Consequently,
732: every TQFT defines a representation of the mapping class group
733: $\Gamma_g\to GL({\cal V}(\Sigma_g)):\,
734: \{\psi\}\mapsto {\cal V}([{\bf I}_{\psi}])\,$.
735:
736:
737: Moreover, let us introduce special cobordisms
738: \beq\lbl{eq-hdlcob}
739: {\bf H_{g}^+}\;:=\;=( H_{g}^+,id\sqcup id, \Sigma_g\sqcup \Sigma_{g+1})\;,
740: \eeq
741: where $H^+_g$ is obtained by adding a
742: full 1-handle to the cylinder $\Sigma_g\times[0,1]$
743: at two discs in $\Sigma_g\times 1$.
744: This is done in a way compatible with the choice of the model surfaces in
745: equation~(\ref{eq-modelsurf}). Another cobordism $H_{g}^-$ is built by gluing in
746: a 2-handle into the thickened surface $\Sigma_{g+1}\times[0,1]$ along a
747: curve $b_{g+1}$, which lies in the added torus from (\ref{eq-modelsurf})
748: and has geometric intersection number one with the meridian
749: of the 1-handle added by $H_g^+$. From this we obtain a cobordism
750: ${\bf H_g^-}=( H_g^-,\Sigma_{g+1}\sqcup\Sigma_g)$ in opposite
751: direction, with the property that ${\bf H_{g}^-}\circ {\bf H_{g}^+}$ is
752: equivalent to the identity.
753:
754:
755: Basic Morse theory implies a Heegaard decomposition as follows for any cobordism
756: \beq\lbl{eq-Heegaard}
757: {\bf M}\;\;\;\cong\;\;\;{\bf H_{g_2}^-}\circ {\bf H_{g_2+1}^-}\circ\ldots
758: \circ {\bf H_{N-1}^-}\circ {\bf I}_{\psi}\circ {\bf H_{N-1}^+}\circ\ldots\circ
759: {\bf H_{g_1+1}^+}
760: \circ {\bf H_{g_1}^+}\;,
761: \eeq
762: where $\psi\in Homeo^+(\Sigma_N)$. Hence, a TQFT is completely determined by
763: the induced representations of the mapping class groups and the maps
764: ${\cal V}([{\bf H}^+_g])$ and ${\cal V}([{\bf H}^-_g])$. Therefore, any two TQFT's
765: coinciding on the basic generators from (\ref{eq-defpsi}) and (\ref{eq-hdlcob})
766: have to be equal.
767:
768: \head{3. The Frohman-Nicas TQFT for $U(1)$}\lbl{S3}
769:
770: Let us review the basic steps in the construction of the topological
771: quantum field theory ${\cal V}^{FN}$ as given in \cite{FroNic92}
772: via intersection theory of $U(1)$-representation varieties:
773:
774: For a compact, connected manifold $X$ its $U(1)$-representation variety is
775: defined as
776: \beq\lbl{eq-defJX}
777: J(X)\quad:=\quad {\rm Hom}(\pi_1(X), U(1))\quad \cong \quad H^1(X,U(1))\;.
778: \eeq
779: Observe that $J(X)$ is a manifold of dimension $\beta_1(X)$. Specifically, it is
780: a torus if $H_1(X,\Z)$ is torsion free, and a discrete group if $\beta_1(X)=0$.
781:
782: The vector space associated to a surface $\Sigma_{\underline g}$ is given
783: by ${\cal V}^{FN}(\Sigma_{\underline g})
784: = H^*(J(\Sigma_{g_1})\times\ldots\times J(\Sigma_{g_N}) ,\R)$.
785:
786: We consider first cobordisms, $M$, between surfaces, $\partial_{in}M$
787: and $\partial_{out}M$, that are rationally homologically trivial in the
788: sense of Section~2.
789: In this case the map $j:\,J(M)\to J(\partial_{in}M)\times J(\partial_{out}M)$ is a half
790: dimensional immersion. Thus the top form $\pm[J(M)]$ defines (up to sign)
791: a middle dimensional
792: homology class in $H_*(J(\partial_{in}M),\R)\otimes H_*(J(\partial_{out}M),\R)$.
793: Using Poicar\'e Duality and the coordinate maps of the cobordism,
794: the latter space is isomorphic to the space of linear maps
795: from ${\cal V}^{FN}(\Sigma_{\underline g_{in}})$ to
796: ${\cal V}^{FN}(\Sigma_{\underline g_{out}})$. ${\cal V}^{FN}(M)$, for a
797: homologically trivial cobordism $M$, is now the linear map associated to
798: $j_*(\pm[J(M)])$.
799:
800:
801: In the general case Frohman and Nicas define ${\cal V}^{FN}(M)$ via
802: a Heegaard splitting of $M$ as in (\ref{eq-Heegaard}), and consider the intersection
803: number of representation
804: varieties of the elementary thick surfaces with handles separated by the Heegaard surface.
805: In the case where $H_1(\partial M,\R)\to H_1(M,\R)$ is not onto, i.e., $M$ is
806: not homologically trivial, these varieties
807: no longer transversely intersect so that ${\cal V}^{FN}(M)=0$.
808:
809: Regarding the composition structure ${\cal V}^{FN}$ has a couple of nonstandard
810: properties. For one, functoriality fails to hold when $M$ and $N$ are homologically
811: trivial but $M\circ N$ is not. Moreover, the orientation of the classes $\pm[J(M)]$ and cycles
812: cannot be chosen consistently with composition so that a sign-projectivity persists.
813: Recall, however, that a 2-framed TQFT is really defined on the $\Z$-extensions of cobordisms
814: given in (\ref{eq-cobext}).
815:
816: \begin{lemma}\lbl{lm-FNpart}
817: ${\cal V}^{FN}$ is a non-semisimple, $\Z/2\Z$-projective TQFT in the sense of Section~2.
818: \end{lemma}
819:
820: The mechanism by which the universal $\Z$-extension is factored into a $\Z/2\Z$-extension
821: is explained further for the quantum theory in Lemma~\ref{lm-mod2prop} and
822: Proposition~\ref{pp-ATQFT} of Section~7. At least indirectly, we have thus related the
823: orientation ambiguities in \cite{FroNic92} to the usual framing ambiguities of quantum
824: theories.
825:
826: Now, in the $U(1)$ case $J(X)$ has a group structure itself, which induces a
827: coalgebra structure on the cohomology ring so that $H^*(J(X))$ is endowed
828: with a canonical Hopf algebra structure ${\cal H}_{ext}$. If $H_1(X)$ is torsion free
829: then $H^*(J(X))$ is connected and we obtain a natural isomorphism
830: $H^*(J(X))\cong\ext * H_1(X)$ of $\Z/2$-graded Hopf algebras, and
831: $H_1(X)$ is the space of primitive elements. Hence, we can write for the vector spaces:
832: \begin{equation}
833: {\cal V}^{FN}(\Sigma_{\underline g})\;=\;\ext * H_1(\Sigma_{\underline g})\;.
834: \end{equation}
835: The representation of the mapping class group $\Gamma_{\underline g}$
836: on this space is given by the obvious action
837: \beq\lbl{eq-FNmapcg}
838: {\cal V}^{FN}([{\bf I_{\psi}}])\;\;=\;\;\ext * [\psi]\qquad
839: \forall \{\psi\}\in\Gamma_g\;.
840: \eeq
841: Here, $[\psi]\in \Sp(H_1(\Sigma_{\underline g}))$ is the natural, induced action on homology.
842: For a connected surface $\Sigma_g$ we have the associated short exact sequence
843: \beq\lbl{eq-TGSseq}
844: 1\;\to\;{\cal J}_g\;\longrightarrow\;\Gamma_g\;\stackrel{\psi\mapsto[\psi]}
845: {-\!\!\!-\!\!\!-\!\!\!-\!\!\!\longrightarrow}\;\Sp(2g,\Z)\;\to\;1\;,
846: \eeq
847: where ${\cal J}_g$ is the Torelli group.
848:
849: Let
850: ${\bf H_{g}^+}$ be the cobordism as defined in (\ref{eq-hdlcob}), and
851: let $[a_{g+1}]$ be a generator of
852: $ker(H_1(\Sigma_{g+1},\Z)\to H_1(H_{g+},\Z))$ seen as an element
853: of $H_1(\Sigma_{g+1},\R)$. It is represented by the meridian $a_{g+1}$ of
854: the added handle.
855: In a slight variation of the Frohman Nicas formalism we see that the associated
856: linear map is given as
857: \beq\lbl{eq-FNmcg}
858: {\cal V}^{FN}({\bf H_{g}^+}):\ext * H_1(\Sigma_g)\longrightarrow\ext * H_1(\Sigma_{g+1})\;:
859: \quad \alpha\;\mapsto\;i_*(\alpha)\wedge [a_{g+1}]\;.
860: \eeq
861: Here we use the fact that $H_1(\Sigma_{g,1})=H_1(\Sigma_g)$ so that the
862: inclusion of surfaces in (\ref{eq-modelsurf}) implies also an inclusion
863: $i_*:H_1(\Sigma_g)\subset H_1(\Sigma_{g+1})$.
864:
865:
866:
867: Let ${\bf H_{g}^-}$ be the cobordism obtained by gluing a 2-handle along $b_{g+1}$
868: as defined above. We note that
869: $H_1(\Sigma_{g+1})=H_1(\Sigma_{g})\oplus\lz[a_{g+1}],[b_{g+1}]\rz$ so that
870: $\ext * H_1(\Sigma_{g+1})$ is the direct sum of spaces
871: $V_1\oplus V_a\oplus V_b\oplus V_{a\wedge b}$ where
872: $V_{x}=[x_{g+1}]\wedge \ext *H_1(\Sigma_{g})$. The linear map associated in \cite{FroNic92}
873: to ${\bf H_{g}^-}$ acts on $V_a$ as
874: \beq\lbl{eq-FNgcm}
875: {\cal V}^{FN}({\bf H_{g}^-}): V_a\longrightarrow\ext * H_1(\Sigma_{g})\;:
876: \quad i_*(\alpha)\wedge [a_{g+1}]\;\mapsto\;\alpha
877: \eeq
878: and is zero on all other summands.
879:
880:
881: % Furthermore, let us assume that we have chose an inner product
882: % $\lz .,.\rz$ on $H_1(\Sigma_g)$
883: % such that $i_*$ is an isometrie, and $\|[a_{g}]\|=1$.
884: % Together with the symplectic form on $H_1(\Sigma_g)$ this
885: % yields an almost complex structure $J$ with $J^2=-1$.
886: % If we set $[b_{g+1}]:=J[a_{g+1}]$
887: % we see that $a_g\cdot b_g=\pm 1$ and intersection
888: % numbers with cycles from $i_*(H_1(\Sigma_g))$ are all zero.
889:
890: % We assume that the attaching curve $b_{g+1}$ for the 2-handle in the
891: % construction of $H_{g-}$ is a representative of $[b_{g+1}]$.
892: % The associated linear map comes out to be
893: % $$
894: % {\cal V}^{FN}({\bf H_{g}^-})\quad=\quad {\cal V}^{FN}({\bf H_{g}^+})^T
895: % $$
896: % where $.^T$ denotes the transpose with respect to $\lz .,.\rz$.
897:
898:
899: \head{4. The Mapping Class Groups and their Actions on Homology}\lbl{S4}
900: The mapping class group
901: $\Gamma_{g,1}=\pi_0(Homeo^+(\Sigma_{g,1}))$ on a model surface
902: $\Sigma_{g,1}$ is generated by the right handed
903: Dehn twists along oriented curves $a_j$, $b_j$,
904: and $c_j$, as depicted in
905: Figure~\ref{fig-modsurf}. We denote them by capital letters
906: $A_j,\,B_j,\,C_j\, \in\Gamma_{g,1}$ respectively.
907: In fact we only need the $A_2$ of the $A_j$'s to generate
908: $\Gamma_{g,1}$. A presentation of $\Gamma_{g,1}$ in
909: these generators is given by Wajnryb \cite{Waj83}.
910: \begin{figure}[ht]
911: \begin{center}
912: \leavevmode
913: \epsfxsize=7cm
914: \epsfbox{mcg-gen.eps}
915: \end{center}
916: \caption{Curves on $\Sigma_{g,1}$}\lbl{fig-modsurf}
917: \end{figure}
918: For our purposes we prefer the set $\{A_j, D_j, S_j\}$ of generators
919: defined as follows:
920: \begin{equation}\lbl{eq-newgen}
921: D_j\;:=\;A_j^{-1}A_{j+1}^{-1}C_j\qquad\mbox{and}\qquad S_j\;:=\;A_jB_jA_j\qquad\quad
922: \mbox{for}\;\;j=1,\ldots,g \;.
923: \end{equation}
924: In \cite{MatPol94} a tangle presentation of $\Gamma_{g,1}$ is given using the results
925: in \cite{Waj83}. The same presentation results from the tangle presentation of
926: $\Cob^{2fr,\bullet}$ in \cite[Proposition 14]{Ker99}, which extends to the central
927: extension $1\to\Z\to\Gamma_{g,1}^{2fr}\to\Gamma_{g,1}\to 1$ that stems from the
928: 2-framing of cobordisms. The framed tangles associated to our preferred generators
929: are given in Figures~\ref{fig-A-tgl}, \ref{fig-D-tgl}, and \ref{fig-S-tgl}.
930: We use an empty circle to indicate a right handed $2\pi$-twist on the framing
931: of a strand as in Figure~\ref{fig-A-tgl}, and a full circle for a left handed one as in
932: Figure~\ref{fig-Rigg-tgl}. Note, that the extra 1-framed circle in
933: Figure~\ref{fig-S-tgl} does not change the 3-cobordism in $\Cob^{\bullet}$ but
934: shifts its 2-framing in $\Cob^{2fr, \bullet}$ by one.
935:
936: \begin{figure}[ht]
937: \begin{center}
938: \leavevmode
939: \epsfysize=2.2cm
940: \epsfbox{A-tgl.eps}
941: \end{center}
942: \caption{Tangle for $A_j$}\lbl{fig-A-tgl}
943: \end{figure}
944:
945: \begin{figure}[ht]
946: \begin{center}
947: \leavevmode
948: \epsfysize=2.2cm
949: \epsfbox{D-tgl.eps}
950: \end{center}
951: \caption{Tangle for $D_j$}\lbl{fig-D-tgl}
952: \end{figure}
953:
954:
955: \begin{figure}[ht]
956: \begin{center}
957: \leavevmode
958: \epsfysize=2.2cm
959: \epsfbox{S-tgl.eps}
960: \end{center}
961: \caption{Tangle for $S_j$}\lbl{fig-S-tgl}
962: \end{figure}
963: $\Gamma_{g,1}^{2fr}$ can then be thought of as the sub-group of tangles generated
964: by these diagrams, modulo isotopies, 2-handle slides, the $\sigma$-move and the
965: Hopf link move, see \cite{Ker99}.
966: \medskip
967:
968: For later purposes we give the explicit action of these generators on
969: $H_1(\Sigma_g,\Z)=H_1(\Sigma_{g,1},\Z)$ in the sense of (\ref{eq-TGSseq}).
970: Suppose $p,\,f\,\subset\Sigma_{g,1}$
971: are two transverse, oriented curves. We denote by $P$ the Dehn twist along $p$,
972: by $[P]\in \Sp(2g,\Z)$ its action on homology,
973: and by $[p]$ and $[f]$ the respective homology classes. We have
974: \begin{equation}\lbl{eq-DehnHom}
975: [P].[f]\quad=\quad[f]\,+\,([p]\cdot[f])[p]\;.
976: \end{equation}
977: Here $([p]\cdot[f])\in\Z$ is the algebraic intersection number of $p$ with $f$,
978: counting $+1$ for a crossing if the tangent vectors of $p,f$ form an oriented
979: basis and $-1$ if the basis has opposite orientation.
980:
981: A basis for $H_1(\Sigma_g)$ is given by
982: $\{[a_1],\ldots, [a_g],[b_1],\ldots, [b_g]\}$,
983: and intersection numbers can be read
984: off Figure~\ref{fig-modsurf}.
985: For example $a_j$ intersects $b_j$ in only one point, where
986: $[a_j]\cdot[b_j]= +1$ since $b_j$ follows $a_j$ counter clockwise at the
987: crossing. Hence
988: \begin{equation}\lbl{eq-A-hom}
989: [A_j].[b_j]=[b_j]+[a_j]\qquad \mbox{and}\quad [A_j].[x]=[x]\quad
990: \mbox{for all other basis vectors.}
991: \end{equation}
992: Similarly, we have that $[C_j]$ only acts on $[b_j]$ and $[b_{j+1}]$
993: with $[C_j].[b_j]=[b_j]+[c_j]$ and $[C_j].[b_{j+1}]=[b_{j+1}]-[c_j]$.
994: Substituting $[c_j]=[a_j]-[a_{j+1}]$, and using the definition of $D_j$ in
995: (\ref{eq-newgen}) and (\ref{eq-A-hom}) we compute
996: \begin{equation}\lbl{eq-D-hom}
997: [D_j].[b_j]=[b_j]-[a_{j+1}]\quad \mbox{and}\quad [D_j].[b_{j+1}]=[b_{j+1}]-[a_{j}] \;,
998: \end{equation}
999: and, again, $[D_j].[x]=[x]$ for all other basis vectors $[x]$ of $H_1(\Sigma_1,\Z)$.
1000: Finally, we find $[B_j].[a_j]=[a_j]-[b_j]$ so that
1001: \begin{equation}\lbl{eq-S-hom}
1002: [S_j].[a_j]=-[b_j] \quad \mbox{and}\quad [S_j].[b_j]=[a_{j}]
1003: \end{equation}
1004: and $[S_j].[x]=[x]$ elsewise.
1005:
1006: The above action can be identified with specific generators of
1007: the Lie algebra ${\mathfrak s}{\mathfrak p}(2g,\R)$ as follows:
1008: $$
1009: [A_j]\;
1010: =\;I_{2g}\;+\;E_{j,-j}\;=\;I_{2g}\;+\;e_{2\epsilon_j}\;=\;\exp(e_{2\epsilon_j})
1011: $$
1012: \begin{equation}\lbl{eq-sproots}
1013: [B_j]\;=\;I_{2g}\;-\;E_{-j,j}\;=\;I_{2g}\;-\;f_{2\epsilon_j}\;=\;\exp(-f_{2\epsilon_j})
1014: \end{equation}
1015: $$
1016: [D_j]\;=\;I_{2g}\;-\;E_{j,-(j+1)}-E_{j+1,-j}\;=
1017: \;I_{2g}\;-\;e_{\epsilon_j+\epsilon_{j+1}}\;=\;\exp(-e_{\epsilon_j+\epsilon_{j+1}})\\
1018: $$
1019: The conventions and notations for the weights $\epsilon_j$ and the
1020: matrices $E_{i,j}$ are taken from \cite[Chapter 2.3]{GooWal98}. Hence,
1021: the natural representation on ${\rm Sp}(2g,\Z)$ clearly lifts to the
1022: fundamental representation of ${\rm Sp}(2g,\R)$.
1023:
1024:
1025: Finally, there is an
1026: ${\rm Sp}(2g,\Z)$-invariant 2-form, which is unique up to signs
1027: and given in our basis as:
1028: \begin{equation}\lbl{eq-defsymform}
1029: \omega_g\quad:=\quad \sum_{j=1}^g[a_j]\wedge[b_j]\qquad\in\;
1030: \ext 2 H_1(\Sigma_g)\,=\,H^2(J(\Sigma_g))\;.
1031: \end{equation}
1032: It is identical to twice the K\"ahler metric form in
1033: $H^2(J(\Sigma_g))$, see Section~10 and \cite{GriHar78}.
1034:
1035:
1036:
1037: \head{5. Hennings TQFT's}\lbl{S5}
1038: In \cite{Hen96} Hennings describes a calculus
1039: that allows us to compute an invariant, ${\cal V}_{\cal A}^H(M)$,
1040: for a closed 3-manifold, $M$, starting from a
1041: surgery presentation, $M=S^3_{\cal L}$, by a framed link, ${\cal L}\subset S^3$, and
1042: a quasitriangular Hopf algebra $\cal A$. It is obtained by inserting and moving elements
1043: of $\cal A$ along the strands of a projection of $\cal L$ and evaluating them against
1044: integrals. This procedure was refined by Kauffman and
1045: Radford \cite{KauRad95} permitting
1046: unoriented links and simplifying the evaluation and proofs substantially.
1047: ${\cal V}^H_{\cal A}$ turns out to be a special case of the invariant given by
1048: Lyubashenko \cite{Lub95}, which is constructed from general abelian categories.
1049: In \cite[Theorem 14]{Ker96} we generalize the Hennings procedure to
1050: tangles and cobordisms and thus construct a topological
1051: quantum field theory ${\cal V}^H_{\cal A}$ for any modular Hopf algebra $\cal A$.
1052: In turn ${\cal V}^H_{\cal A}$ is derived as a special case of
1053: the general TQFT construction by Lyubashenko and
1054: the author in \cite{KerLub00}.
1055:
1056: The TQFT in \cite{Ker96} was formulated as
1057: a contravariant functor, ${\cal V}_{\cal A}^{*}:\,Cob_3^{\bullet}\to Vect(\kk)$,
1058: where ${\cal V}_{\cal A}^{*}(\Sigma_{g,1})={\cal A}^{\otimes g}$. In this section we
1059: will give the rules for construction for the covariant version, defined by
1060: ${\cal V}_{\cal A}(M)=(f^{\otimes g})^{-1}({\cal V}^*_{\cal A}(M))^*f^{\otimes g}\,$,
1061: where $f:{\cal A}\to{\cal A}^*:\, x\mapsto \mu(S(x)\_\_)$. We generalize
1062: \cite{Ker96} further by allowing Hopf algebras, $\cal A$, that are not modular,
1063: at the expense of reducing the vector space by a canonical projection.
1064:
1065:
1066: Let $M$ be a 2-framed cobordism between two model surfaces,
1067: $\Sigma_{g_1}$ and $\Sigma_{g_2}$.
1068: As in \cite{Ker99} we associate to the homeomorphism class of $M$ an equivalence
1069: class of framed tangle diagrams. The projection of a representative tangle, $T_M$, in
1070: $\R\times [0,1]$ has $2g_1$ endpoints $\,1^-<1^+<2^-<\ldots<g_1^-<g_1^+\,$
1071: in the top line $\R\times 1$ and $2g_2$ endpoints $\,1^-<1^+<2^-<\ldots<g_2^-<g_2^+\,$
1072: in the bottom line $\R\times 0$. Besides closed components ($\cong S^1$) the tangle can
1073: have components with boundary ($\cong[0,1]$). An interval component, $J$, of the tangle
1074: can either run between points $j^-$ and $j^+$ at the top line or between
1075: $j^-$ and $j^+$ at the bottom line.
1076: As a forth possibility we admit pairs of components, $I$ and $J$, of
1077: which each starts at the top line and ends at the bottom line and cobords a pair
1078: $\{j^-,j^+\}$ to a pair $\{k^-,k^+\}$. The equivalences of tangles are generated
1079: by isotopies, 2-handle slides (second Kirby move) over closed components, the
1080: addition and removal of an isolated Hopf link, in which one component has 0-framing, and
1081: additional boundary moves, called $\sigma$- and $\tau$-Moves, see \cite{Ker99}.
1082: For later purposes we also depict here the $\sigma$-Move:
1083:
1084: \beq\label{eq-boundmove}
1085: \epsfbox{boundmove.eps}
1086: \eeq
1087:
1088:
1089: The next ingredient is a unimodular, ribbon Hopf algebra, ${\cal A}$,
1090: in the sense of \cite{Rad94}, over a
1091: perfect field $\kk$ with $char(\kk)=0$.
1092: In particular, ${\cal A}$ is a {\em quasitriangular} Hopf algebra
1093: as introduced by Drinfel'd \cite{Dri90}. This means there exists an element
1094: ${\cal R}=\sum_je_j\otimes f_j\,\in{\cal A}^{\otimes 2}$, called the {\em R-matrix},
1095: which fulfills several natural conditions. As in \cite{Dri90} we define the element
1096: $u=\sum_jS(f_j)e_j$, which implements the square of the antipode $S$ by $S^2(x)=uxu^{-1}$.
1097: A {\em ribbon} Hopf algebra is now a quasitriangular Hopf algebra with a group like
1098: element, $G$, such that $G$ also implements $S^2$ and $G^2=uS(u)^{-1}$.
1099: From this we define the ribbon element $v:=u^{-1}G$, which is central in
1100: ${\cal A}$. Furthermore, it satisfies the equation
1101: \begin{equation}\lbl{eq-M-matrix}
1102: \,{\cal M}\;\;=\;\;{\cal R}^{\dagger}{\cal R}\;\;=\;\;\Delta(v^{-1})v\otimes v\;\;,
1103: \end{equation}
1104: where $(a\otimes b)^{\dagger}=b\otimes a$ is the transposition of tensor factors.
1105:
1106: Now, any finite dimensional Hopf algebra contains a {\em right integral},
1107: which is an element $\mu \in {\cal A}^*$ characterized by the equation:
1108: \beq\lbl{eq-int-def}
1109: ( \mu\otimes id_{\cal A})(\Delta(x))\;=\;1\cdot\mu(x)
1110: \eeq
1111: Its existence and uniqueness (up to scalar multiplication) has been proven
1112: in \cite{LarSwe69}. The adjective ``unimodular''
1113: implies that
1114: \beq\lbl{eq-muSinvar}
1115: \mu(xy)\;\;=\;\;\mu(S^2(y)x)\qquad\qquad\mbox{and}\qquad\qquad\mu(S(x))\;\;=\;\;\mu(G^2x)\;,
1116: \eeq
1117: see \cite{Rad94}. For the remainder of this article we will also
1118: assume the following normalizations:
1119: \beq\lbl{eq-balnorm}
1120: \mu\otimes\mu({\cal M})\;\;=\;\;1\qquad\mbox{and}\qquad\;\;\mu(v)\mu(v^{-1})=1
1121: \eeq
1122:
1123:
1124: The next step in the Hennings procedure is to replace the tangle projection $T_M$
1125: with distinguished over and under crossings by a formal linear combination of copies of
1126: the projection $T_M$ in which we do not distinguish between over and under crossings
1127: but decorate
1128: segments of the resulting planar curve with elements of $\,{\cal A}\,$. Specifically,
1129: we replace an over crossing by an indefinite crossing and insert at the two incoming
1130: pieces the elements occurring in the $R$-matrix, and similarly for an under crossing,
1131: as indicated in the following diagrams.
1132:
1133:
1134: \beq\lbl{fig-Dec-cross}
1135: \setlength{\unitlength}{0.25mm}
1136: \begin{picture}(470,50)
1137: \put(20,5){\line(1,1){40}}
1138: \put(20,45){\line(1,-1){17}}
1139: \put(60,5){\line(-1,1){17}}
1140:
1141: \put(85,30){}
1142: \put(80,25){\vector(1,0){26}}
1143:
1144: \put(125,20){$\displaystyle \sum_j$}
1145:
1146: \put(165,50){\line(1,-1){50}}
1147: \put(165,0){\line(1,1){50}}
1148: \put(180,44){$\scriptscriptstyle e_j$}
1149: \put(175,40){\circle*{10}}
1150: \put(205,40){\circle*{10}}
1151: \put(214,38){$\scriptscriptstyle f_j$}
1152:
1153:
1154: %\put(275,50){\line(1,-1){50}}
1155: %\put(275,0){\line(1,1){50}}
1156: %\put(285,40){\circle*{10}}
1157: %\put(260,44){$\scriptscriptstyle e_j$}
1158: %\put(315,40){\circle*{10}}
1159: %\put(294,38){$\scriptscriptstyle f_j$}
1160:
1161: \put(280,5){\line(1,1){17}}
1162: \put(320,45){\line(-1,-1){17}}
1163: \put(320,5){\line(-1,1){40}}
1164:
1165:
1166: %\put(308,30){$\scriptscriptstyle {\cal D}ec_{\cal A}$}
1167: \put(333,25){\vector(1,0){27}}
1168:
1169: \put(375,20){$\displaystyle \sum_j$}
1170:
1171:
1172: \put(425,50){\line(1,-1){50}}
1173: \put(425,0){\line(1,1){50}}
1174: \put(435,10){\circle*{10}}
1175: \put(405,15){$\scriptscriptstyle S(e_j)$}
1176: \put(465,10){\circle*{10}}
1177: \put(482,13){$\scriptscriptstyle f_j$}
1178: \end{picture}
1179: \eeq
1180:
1181:
1182: The elements on the segments of
1183: the planar diagram can then be moved along the connected components according to
1184: the following rules.
1185:
1186: \beq\lbl{eq-elem-mv}
1187: \setlength{\unitlength}{0.25mm}
1188: \begin{picture}(440,85)(0,-15)
1189: \put(20,5){\line(0,1){60}}
1190: \put(20,20){\circle*{10}}
1191: \put(27,21){$x$}
1192: \put(20,50){\circle*{10}}
1193: \put(27,51){$y$}
1194:
1195: \put(37,35){$=$}
1196:
1197: \put(60,5){\line(0,1){60}}
1198:
1199: \put(60,35){\circle*{10}}
1200: \put(67,36){$xy$}
1201:
1202:
1203: \put(135,35){\oval(40,40)[b]}
1204: \put(115,35){\line(0,1){20}}
1205: \put(155,35){\line(0,1){20}}
1206: \put(155,42){\circle*{10}}
1207: \put(162,43){$S(x)$}
1208:
1209: \put(190,33){$=$}
1210:
1211: \put(235,35){\oval(40,40)[b]}
1212: \put(215,35){\line(0,1){20}}
1213: \put(255,35){\line(0,1){20}}
1214: \put(215,42){\circle*{10}}
1215: \put(222,43){$x$}
1216:
1217:
1218: \put(300,55){\line(1,-1){40}}
1219: \put(300,15){\line(1,1){40}}
1220: \put(330,25){\circle*{10}}
1221: \put(340,25){$x$}
1222:
1223: \put(358,35){$=$}
1224:
1225: \put(380,55){\line(1,-1){40}}
1226: \put(380,15){\line(1,1){40}}
1227: \put(390,45){\circle*{10}}
1228: \put(397,48){$x$}
1229:
1230: \end{picture}
1231: \eeq
1232:
1233:
1234: Finally, every diagram can be untangled using the local moves given below,
1235: and the usual planar third Reidemeister move. In particular, undoing a closed
1236: curve in the diagram
1237: yields an extra overall factor $G^d$, where $G$ is the group like element defined
1238: above and $d$ the Whitney number of the curve.
1239:
1240: \beq\lbl{fig-st-isom}
1241: \setlength{\unitlength}{0.25mm}
1242: \begin{picture}(400,100)(0,-15)
1243: \put(30,30){\oval(40,40)[b]}
1244: \put(10,30){\line(2,5){6}}
1245: \put(50,30){\line(-2,5){6}}
1246: \put(16,45){\line(1,1){34}}
1247: \put(44,45){\line(-1,1){34}}
1248:
1249: \put(72,43){$=$}
1250:
1251: \put(120,30){\oval(40,40)[b]}
1252: \put(100,30){\line(0,1){46}}
1253: \put(140,30){\line(0,1){46}}
1254: \put(140,55){\circle*{10}}
1255: \put(147,56){$G$}
1256:
1257: %\put(60,-15){\em Rm1}
1258:
1259: \put(220,40){\line(0,1){20}}
1260: \put(250,40){\line(0,1){20}}
1261: \put(220,60){\line(2,5){6}}
1262: \put(250,60){\line(-2,5){6}}
1263: \put(220,40){\line(2,-5){6}}
1264: \put(250,40){\line(-2,-5){6}}
1265: \put(250,99){\line(-1,-1){24}}
1266: \put(220,99){\line(1,-1){24}}
1267: \put(250,1){\line(-1,1){24}}
1268: \put(220,1){\line(1,1){24}}
1269:
1270: \put(270,50){$=$}
1271:
1272: \put(290,4){\line(0,1){92}}
1273: \put(320,4){\line(0,1){92}}
1274:
1275: %\put(263,-15){\em Rm2}
1276:
1277: \end{picture}
1278: \eeq
1279:
1280: The assignments that result from this for the left and right ribbon $2\pi$-twists are
1281: summarized in Figure~\ref{fig-Rigg-tgl}. Note, that in the assignment on the
1282: right hand side the full circle on the left side stands for a left handed twist
1283: for the framing, while the fat dot on the right hand side indicates a decoration of the strand
1284: by the element $v^{-1}$.
1285:
1286:
1287: \begin{figure}[ht]
1288: \begin{center}
1289: \leavevmode
1290: \epsfxsize=13cm
1291: \epsfbox{Rigg-tgl.eps}
1292: \end{center}
1293: \caption{Twist Assignments}\lbl{fig-Rigg-tgl}
1294: \end{figure}
1295:
1296:
1297:
1298:
1299:
1300:
1301:
1302: It is clear that after application of these types of manipulations to any decorated diagram
1303: we eventually obtain a set of disjoint, planar curves which can be one of four types.
1304: For each of these types we describe next the evaluation rule that leads to the definition
1305: of a linear map ${\cal V}^{\#}(T_M)$:
1306:
1307: Components of the first type are closed circles decorated with one element $a_i\in{\cal A}\,$ on the
1308: right side. To this we associate the number $\mu(a_i)\in\kk\,$.
1309:
1310: Next, we may have an arc at the bottom line of the diagram connecting points
1311: $p_k'$ and $q_k'$ with one decoration
1312: $\,b_k\in{\cal A}\,$ at the left strand. To this to we associate
1313: the vector $\,b_k\in{\cal A}^{(k)}\,$ in the $k$-th copy of
1314: the tensor product $\,{\cal A}^{\otimes g_2}\,$.
1315:
1316: Thirdly, for an arc at the top line between points $p_j$ and $q_j$ with decoration
1317: $c_j\in{\cal A}\,$ on the right we assign the linear form $l_{c_j}:{\cal A}^{(j)}\to\kk\,$
1318: given by $l_{c_j}(x)=\mu(S(x)c_j)$
1319: on the $j$-th copy of
1320: the tensor product $\,{\cal A}^{\otimes g_1}\,$.
1321:
1322: Finally, we may have pairs of straight strands that connect a pair $\,\{p_j,q_j\}\,$ to the pair
1323: $\,\{p_k',q_k'\}\,$, carrying decorations, $a$ and $b$. In case the strands are parallel,
1324: that is, one connects $p_j$ to $p'_k$ and the other $q_j$ to $q'_k$, we assign a linear
1325: map $T_{a,b}\,:{\cal A}^{(j)}\to {\cal A}^{(k)}\,$ between the $j$-th copy of
1326: $\,{\cal A}^{\otimes g_1}\,$ to the $k$-th copy of $\,{\cal A}^{\otimes g_2}\,$,
1327: by $T_{a,b}(x)=axS(b)\,$.
1328:
1329: If the connecting strands cross over we apply in addition the endomorphism $K(x)=G^{-1}S(x)$ on the
1330: $k$-th copy ${\cal A}^{(k)}$ for a crossing right at the bottom line. It is quite useful to summarize
1331: these rules also pictorially as follows:
1332: \beq\lbl{pic-Rclos}
1333: \setlength{\unitlength}{0.25mm}
1334: \begin{picture}(250,70)(0,-5)
1335: \put(35,35){\circle{50}}
1336: \put(60,35){\circle*{8}}
1337: \put(65,37){$a_i$}
1338:
1339: \put(130,37){\vector(1,0){20}}
1340:
1341: \put(200,35)
1342: {$\;\mu(a_i)\;$}
1343: \end{picture}
1344: \eeq
1345: \beq\lbl{pic-Rbot}
1346: \setlength{\unitlength}{0.25mm}
1347: \begin{picture}(400,80)(0,-15)
1348: \put(5,10){\rule{15mm}{1mm}}
1349: \put(12,-10){$p_k'$}
1350: \put(52,-10){$q_k'$}
1351: \put(35,35){\oval(40,40)[t]}
1352: \put(15,35){\line(0,-1){25}}
1353: \put(55,35){\line(0,-1){25}}
1354: \put(15,24){\circle*{10}}
1355: \put(21,25){$b_k$}
1356:
1357: \put(130,30){\vector(1,0){20}}
1358:
1359: \put(200,25)
1360: {$b\,:\kk\longrightarrow {\cal A}^{(k)} \,\,\,:\;\,\,\;1\,\mapsto\,\,b_k\,$}
1361: \end{picture}
1362: \eeq
1363: \beq\lbl{pic-Rtop}
1364: \setlength{\unitlength}{0.25mm}
1365: \begin{picture}(500,93)(0,-5)
1366: \put(5,60){\rule{15mm}{1mm}}
1367: \put(12,80){$p_j$}
1368: \put(52,80){$q_j$}
1369: \put(35,35){\oval(40,40)[b]}
1370: \put(15,35){\line(0,1){25}}
1371: \put(55,35){\line(0,1){25}}
1372: \put(55,42){\circle*{10}}
1373: \put(62,43){$c_j$}
1374:
1375: \put(130,53){\vector(1,0){20}}
1376:
1377: \put(200,40)
1378: {$l_{c_j}\,:{\cal A}^{(j)}\longrightarrow \kk\,\,\,:\;\,\,\;x\,\mapsto\,\,\mu(S(x)c_j)\,$}
1379: \end{picture}
1380: \eeq
1381: \beq\lbl{pic-Rthru}
1382: \setlength{\unitlength}{0.25mm}
1383: \begin{picture}(500,112)(0,-15)
1384: \put(20,5){\line(0,1){60}}
1385: \put(20,35){\circle*{10}}
1386: \put(27,36){$a$}
1387:
1388: \put(12,80){$p_j$}
1389: \put(52,80){$q_j$}
1390:
1391: \put(12,-10){$p_k'$}
1392: \put(52,-10){$q_k'$}
1393: \put(60,5){\line(0,1){60}}
1394:
1395: \put(60,35){\circle*{10}}
1396: \put(67,36){$b$}
1397:
1398: \put(130,33){\vector(1,0){20}}
1399:
1400:
1401: \put(200,27){$T_{a,b}\,:{\cal A}^{(j)}\longrightarrow {\cal A}^{(k)}\,\,\,:\;\,\,\;x\,\mapsto\,\,
1402: a x S(b)\,$}
1403: \end{picture}
1404: \eeq
1405:
1406: \setlength{\unitlength}{0.25mm}
1407: \begin{picture}(400,80)(0,-15)
1408: \put(5,10){\rule{15mm}{1mm}}
1409: \put(12,-10){$p_k'$}
1410: \put(52,-10){$q_k'$}
1411: \put(15,48){\line(1,-1){37}}
1412: \put(55,48){\line(-1,-1){37}}
1413:
1414:
1415: \put(130,30){\vector(1,0){20}}
1416:
1417: \put(200,25){$K\,:{\cal A}^{(k)}\longrightarrow {\cal A}^{(k)}
1418: \,\,\,:\;\,\,\;x\,\mapsto\,\, G^{-1}S(x)\,$}
1419:
1420: \end{picture}
1421:
1422:
1423:
1424:
1425: From these rules for evaluating diagrams
1426: we obtain a linear map $\,{\cal A}^{\otimes g_1}\to{\cal A}^{\otimes g_2}\,$
1427: for any decorated planar tangle. For a given tangle $T_M$ we denote by $\,{\cal V}^{\#}(T_M)\,$ the
1428: sum of all of these maps associated to the sum of decorated diagrams for $T_M$.
1429: Thus, if we consider, for simplicity, a tangle $T_M$ without
1430: components of the fourth type, and denote by $a^{\nu}_i\,$,
1431: $b^{\nu}_j\,$ and $c^{\nu}_k\,$ the respective elements of the $\nu$-th summand of
1432: the same untangled curve of $T_M$, this linear map can be expressed as
1433: $$
1434: {\cal V}^{\#}(T_M)\;\;:=\;\;\sum_{\nu\, }\mu(a_1^{\nu})\ldots \mu(a_N^{\nu})\,
1435: b_1^{\nu}\otimes\ldots\otimes b_{g_2}^{\nu}\, l_{a_1^{\nu}}\otimes\ldots\otimes
1436: l_{a_{g_1}^{\nu}}\;\;.
1437: $$
1438: For tangles with strand pairs that connect top and bottom pairs we insert the operators
1439: $T_{a,b}\,$ in the respective positions.
1440: \medskip
1441:
1442:
1443: \begin{lemma}\lbl{lm-Vprop}
1444: The linear maps ${\cal V}^{\#}(T_M)$ are well defined, (covariantly)
1445: functorial under the composition of tangles, and they commute with
1446: the adjoint action of ${\cal A}$ on ${\cal A}^{\otimes g}$. They are also invariant under isotopies
1447: and the following moves:
1448: \begin{enumerate}
1449: \item 2-handle slides of any type of strand over a closed component of $\,T_M\,$
1450: \item Adding/removing an isolated Hopf link for which one component has 0-framing
1451: and the other framing 0 or 1.
1452: \end{enumerate}
1453: \end{lemma}
1454:
1455: {\em Proof:} The fact that the construction procedure for a given diagram is unambiguous
1456: is almost straight forward, except that one has to pay attention to the positioning of the
1457: resulting elements. Details for closed links can be found in \cite{Kau94}.
1458: Functoriality is easily checked from the rules of construction. The fact that the maps
1459: are ${\cal A}$-equivariant follows from the fact that it is a special
1460: case of the categorical construction in \cite{KerLub00}, and the fact that
1461: $f:{\cal A}\to{\cal A}^*$ intertwines the adjoint with the coadjoint action.
1462: Invariance under isotopies follows, as in \cite{Hen96} or \cite{KauRad95}, from the
1463: properties of the $R$-matrix of a quasitriangular Hopf algebra. In the same articles
1464: the 2-handle slide is directly related to the defining equation (\ref{eq-int-def})
1465: of the right integral, see also \cite{Lub95} for the categorical version of the
1466: argument. Invariance under the Hopf link moves is a direct consequence of the normalizations
1467: in (\ref{eq-balnorm}), since they imply that the Hennings invariants on the Hopf links are all one.
1468: \ep
1469: \medskip
1470:
1471: In order to describe the reduction procedure that allows us to define a TQFT also
1472: for non-modular Hopf algebras
1473: we introduce the operators associated to the diagrams in Figure~\ref{fig-Smat},
1474: the left being isotopic to the one in Figure~\ref{fig-S-tgl}.
1475: \begin{figure}[ht]
1476: \begin{center}
1477: \leavevmode
1478: \epsfysize=2.2cm
1479: \epsfbox{Smat.eps}
1480: \caption{$S^{\pm}$-Transformations}\lbl{fig-Smat}
1481: \end{center}
1482: \end{figure}
1483: The double crossing is replaced by the elements $m^{+}_j, n^{+}_j$ from
1484: ${\cal M}\,=\,\sum_jm_j^+\otimes n_j^+\,$, as defined in (\ref{eq-M-matrix}).
1485: The transformation $S^+:{\cal A}\to{\cal A}$ is readily worked out to be
1486: \begin{equation}\lbl{eq-S-matrix}
1487: S^+(x)\;=\;\sum_j\mu(S(x) m_j^+) n_j^+\qquad.
1488: \end{equation}
1489: The formula for $S^-$ follows analogously, substituting ${\cal M}$ for
1490: ${\cal M}^{-1}=\sum_jm_j^-\otimes n_j^-$. We consider next the result $\Pi$
1491: of stacking the two tangles in Figure~\ref{fig-Smat} on top of each other:
1492: \begin{lemma}\lbl{lm-Pi}
1493: Let $\Pi:=S^+\circ S^-=S^-\circ S^+$, and denote
1494: $\Pi^{(j)}=1\otimes\ldots1\otimes \Pi\otimes 1\ldots\otimes 1$, with $\Pi$
1495: occurring in the $j$-th tensor position.
1496: \begin{enumerate}
1497: \item $\Pi$ is an idempotent that commutes with the adjoint action of $\cal A$.
1498: \item ${\cal V}^{\#}(T_M)\Pi^{(j)}={\cal V}^{\#}(T_M)$ if the $j$-th top index pair in
1499: $T_M$ is attached to a top ribbon in $T_M$. (Analogously for bottom ribbons).
1500: \item $\Pi^{(k)}{\cal V}^{\#}(T_M)={\cal V}^{\#}(T_M)\Pi^{(j)}$ if $T_M$ has
1501: a through pair
1502: connecting the $j$-th top pair to the $k$-th bottom pair.
1503: \end{enumerate}
1504: \end{lemma}
1505:
1506: {\em Proof:} For {\em 1.} note that the picture for $\Pi$ consists of two arcs
1507: that are connected by a circle. Stacking $\Pi$ on top of itself we obtain the picture
1508: for $\Pi^2$ by functoriality in
1509: Lemma~\ref{lm-Vprop}. The resulting tangle is the chain of circles $C_j$ and
1510: arcs $A_{t/b}$ depicted on the left of Figure~\ref{fig-pi-idem}. By
1511: {\em 1.} of Lemma~\ref{lm-Vprop} we may use 2-handle slides to manipulate
1512: this picture. We first slide $C_1$ over $C_3$, and then $A_b$ over $C_2$.
1513: The result is the tangle for $\Pi$ and a separate Hopf link. The value of
1514: the latter, however, is 1 by (\ref{eq-balnorm}). Hence, $\Pi^2=\Pi$.
1515: \begin{figure}[ht]
1516: \begin{center}
1517: \leavevmode
1518: \epsfysize=2.2cm
1519: \epsfbox{pi-idem.eps}
1520: \caption{$\Pi$ is idempotent}\lbl{fig-pi-idem}
1521: \end{center}
1522: \end{figure}
1523:
1524: Equivariance with respect to the action of ${\cal A}$ is immediate from
1525: Lemma~\ref{lm-Vprop}.
1526:
1527: For {\em 2.} we repeat an argument from \cite{KerLub00}. Suppose $\tau$ is a top
1528: component and $\eta$ any band connecting two intervals $I_i$ in $\tau$ in an
1529: orientation preserving way. To this we associated the surgered diagram in which
1530: the component $\tau$ is replaced by the union $\tau_{\eta}$ of three components.
1531: They are obtained by cutting away the intervals $I_i$ from $\tau$ and inserting
1532: the other two edges of $\eta$ at the endpoints $\partial I_i$ as indicated in
1533: Figure~\ref{fig-toppath}. Furthermore, we insert a 0-framed annulus $A$ around $\eta$.
1534: \begin{figure}[ht]
1535: \begin{center}
1536: \leavevmode
1537: \epsfysize=2.2cm
1538: \epsfbox{toppath.eps}
1539: \caption{$\eta$-Surgery}\lbl{fig-toppath}
1540: \end{center}
1541: \end{figure}
1542: Sliding any other component over $A$ at an arbitrary point along $\eta$
1543: has the effect of just moving it through $\eta$ at this point. Moreover, we
1544: can slide a $\pm 1$-framed annulus $K$ over $A$ so that it surround the two
1545: parallel strands in $\tau_{\eta}\,$, and then slide the two strands over $K$. The
1546: effect is the same as putting a $2\pi$-twist into $\eta$. These two operation
1547: allow us to move any band $\eta$ to any other band $\eta'$ such that $\tau_{\eta}$
1548: and $\tau_{\eta'}$ are related by a sequence of two handle slides.
1549:
1550: Now, adding the picture of $\Pi$ to the top-component $\tau$ of a tangle $T_M$
1551: is the same as surgering $\tau$ along a straight band parallel and close to
1552: the interval between the attaching points of $\tau$ at the top line. We replace
1553: this $\eta$ by a small planar arc at $\tau$ separate from the rest of the
1554: tangle. Surgery along this corresponds to linking a Hopf link to $\tau$,
1555: as $C_2\cup C_3$ is linked to $A_b$ in the middle of Figure~\ref{fig-pi-idem},
1556: and consequently can be removed by the same argument.
1557: The proofs for the formulas for bottom and through strands are entirely analogous.
1558: \ep
1559:
1560: Set $\Pi^{\#}=\Pi^{\otimes g}$, when acting on ${\cal A}^{\otimes g}$. It follows
1561: now easily from Lemma~\ref{lm-Pi} that
1562: ${\cal V}^{\#}(T_M)\Pi^{\#}=\Pi^{\#}{\cal V}^{\#}(T_M)$ for all $T_M$. Thus
1563: each ${\cal V}^{\#}(T_M)$ maps the image of $\Pi^{\#}$ to itself so that we can
1564: define the restriction
1565: \begin{equation}\lbl{eq-redtqft}
1566: {\cal V}(T_M):={\cal V}^{\#}(T_M)\Bigl |_{im(\Pi^{\#})}:\;\;
1567: {\cal V}_{\cal A}(\Sigma_{g_1,1})\;\longrightarrow\;{\cal V}_{\cal A}(\Sigma_{g_2,1})\;\;,
1568: \end{equation}
1569: where the vector spaces are given as
1570: \beq\lbl{eq-A0def}
1571: {\cal V}_{\cal A}(\Sigma_{g,1})=\Pi^{\#}({\cal V}^{\#}(\Sigma_{g}))={\cal A}_0^{\otimes g}
1572: \qquad\mbox{with} \qquad {\cal A}_0=\Pi({\cal A})\;\;\;.
1573: \eeq
1574: \begin{theorem}\lbl{thm-henn}
1575: The assignment $\cal V$ as given in (\ref{eq-redtqft}) yields a
1576: well defined, 2-framed, relative, ${\cal A}-equivariant$
1577: topological quantum field theory
1578: $$
1579: {\cal V}_{\cal A}\;:\quad \Cob^{2fr,\bullet}\;\longrightarrow\;{\cal A}{\rm -mod}_{\kk}\quad\subset
1580: \quad Vect(\kk)\quad.
1581: $$
1582: Using the invariance functor ${\rm Inv}={\rm Hom}(1,\_):{\cal A}{\rm -mod}\to
1583: Vect(\kk)$ we obtain an ordinary 2-framed TQFT for closed surfaces as
1584: $$
1585: {\cal V}_{\cal A}^0:={\rm Inv}\circ{\cal V}_{\cal A}\;:
1586: \quad \Cob^{2fr}\;\longrightarrow\;
1587: \quad Vect(\kk)\quad.
1588: $$
1589: \end{theorem}
1590:
1591: {\em Proof:}
1592: We recall from \cite[Proposition 12]{Ker99} that two presentations, $T_M$ and $T'_M$,
1593: of a framed, relative cobordism $M\in \Cob^{2fr,\bullet}$ are related by the
1594: moves described in Lemma~\ref{lm-Vprop} and the so called $\sigma$-moves, which
1595: consist of adding the picture of $\Pi$ to a pair of points at the top or bottom
1596: line of the diagram. From ${\cal V}(T_M)\Pi^{(j)}={\cal V}^{\#}(T_M)\Pi^{\#}\Pi^{(j)}
1597: ={\cal V}^{\#}(T_M)\Pi^{\#}$ we see that ${\cal V}(T_M)$ is invariant under this
1598: move. Hence, ${\cal V}(T_M)$ only depends on the cobordism represented by $T_M$
1599: and we can write ${\cal V}_{\cal A}(M):={\cal V}(T_M)$.
1600:
1601: Due to the equivariance of $\Pi$ also ${\cal A}_0$ from (\ref{eq-A0def}) is invariant
1602: under the adjoint action of ${\cal A}$, and the restricted maps commute with
1603: the action of ${\cal A}$ as well. Functoriality of ${\cal V}$ follows from
1604: functoriality of ${\cal V}^{\#}$ and the fact that $\Pi^{\#}$ commutes with
1605: ${\cal V}^{\#}$.
1606:
1607: Since each ${\cal V}(M)$ commutes with the action of ${\cal A}$ they also
1608: map the ${\cal A}$-invariant subspaces
1609: ${\cal V}^0(\Sigma_{g}):={\rm Inv}({\cal V}(\Sigma_{g,1}))$ to themselves.
1610: This implements the additional $\tau$-move \cite{Ker99} needed to represent
1611: cobordisms between closed surfaces.
1612: \ep
1613:
1614:
1615: \head{6. The Algebra $\cal N$}\lbl{S6}
1616: The Hopf algebra $\cal N$ we will define in this section is the same
1617: as the algebra $A_2$ described by Radford in Example~1 of Section~4.1
1618: in \cite{Rad76}. The quasitriangular structure that we endow $\cal N$
1619: with is essentially distilled from the one of $U_{-1}({\mathfrak s}{\mathfrak l}_2)$.
1620:
1621: Let $\E\cong\R^2$ be the Euclidean plane, and consider
1622: the 8-dimensional algebra
1623: \beq\lbl{eq-Adef}
1624: {\cal N}\;:=\;\Z/2\ltimes \ext * \E\;.
1625: \eeq
1626: The generator of $\Z/2$ is denoted by $K$, with $K^2=1$, and we write $x^K=KxK$
1627: for any $x\in{\cal N}$. We thus have relations $w'w=-ww'$ and $w^K:=KwK=-w$
1628: for all $w,w'\in\E$.
1629:
1630: \blm\lbl{lm-Ahopf}
1631: $\cal N$ is a Hopf algebra with coproducts
1632: \beq\lbl{eq-coprod}
1633: \Delta(K)=K\otimes K\quad\mbox{ and}\qquad \Delta(w)\,=\,w\otimes 1\,+\,K\otimes w\;\;
1634: \forall w\in\E\;.
1635: \eeq
1636: \elm
1637: {\em Proof:} The fact that $\Delta:{\cal N}\to{\cal N}^{\otimes 2}$ is a
1638: coassociative homomorphism
1639: is readily verified. The antipode is given by
1640: \beq\lbl{eq-antipode}
1641: S(K)\,=\,K\qquad\mbox{and}\qquad S(w)=-Kw,\;\;\forall w\in\E\;.
1642: \eeq
1643: \ep
1644:
1645: We note the following formulas for the adjoint action and antipode:
1646: \beq\lbl{eq-somefor}
1647: ad(w)(x)=wx-x^Kw\qquad, \qquad S^2(x)=x^K\qquad\forall x\in{\cal N}, w\in\E
1648: \eeq
1649:
1650: Let us pick a non-zero element $\rho\in\ext 2\E\subset{\cal N}$, and for this define
1651: a form $\mu_0\in{\cal N}^*$ as follows:
1652: \beq\lbl{eq-Ainteg}
1653: \mu_0(\rho)\;=\;1\;\,,\qquad\quad \mu_0(K\rho)=0\;,\qquad\mbox{and}
1654: \eeq
1655: $$
1656: \mu_0(K^{\delta}x)=0\;,\qquad\forall x\in\ext j\E\;\;,\;\;\mbox{whenever}
1657: \;\;j,\delta\in\{0,1\}\;\;.
1658: $$
1659: \blm\lbl{lm-Ainteg}\ \ \
1660: $\mu_0$ is a right (and left) integral on ${\cal N}$. Moreover,
1661: \beq\lbl{eq-intIN}
1662: \lambda_0:= (1+K)\rho\qquad\mbox{with}\quad\mu_0(\lambda_0)=1
1663: \eeq
1664: is a two sided integral in ${\cal N}$.
1665: \elm
1666:
1667: {\em Proof:} Straightforward verification of (\ref{eq-int-def}). The defining
1668: equation for a two sided integral in ${\cal N}$ is
1669: $x\lambda_0=\lambda_0 x=\epsilon(x)\lambda_0$, which is also readily found.
1670: \ep
1671:
1672: Next, we fix a basis $\{\th,\tb\}$ for $\E$.
1673: We define an $R$-matrix, ${\cal R}\in{\cal N}\otimes{\cal N}$, by the formula
1674: \beq\lbl{eq-ARmat}
1675: {\cal R}\quad:=\quad \Bigl(1\otimes 1\,+\,\th\otimes K\tb\Bigr)\cdot {\cal Z}\;,
1676: \qquad\mbox{where}\quad {\cal Z}:=\frac 12 \sum_{i,j=0}^1(-1)^{ij}K^i\otimes K^j
1677: %\quad =\quad \cdot {\cal Z}\Bigl(1\otimes 1\,+\,\th K\otimes \tb\Bigr)
1678: \eeq
1679:
1680: \blm\lbl{lm-Rmat}
1681: The element ${\cal R}$ makes ${\cal N}$ into a quasitriangular Hopf algebra.
1682:
1683: Moreover, ${\cal N}$ is a ribbon Hopf algebra with unique balancing element $G=K$.
1684:
1685: \elm
1686: {\em Proof:} Quasitriangularity follows from a
1687: straightforward verification of the axioms in \cite{Dri90}.
1688: We compute the special element $u^{-1}=\sum_jf_jS^2(e_j)=K(1+\tb\th)$ for which
1689: $uS(u)^{-1}=uu^{-1}=1$ so that $G=K$ is a valid and unique choice. The ribbon element
1690: is then given by
1691: \beq\lbl{lm-Aribb}
1692: v\;:=\;1+\rho\qquad\mbox{with}\qquad \rho:=\tb\th
1693: \eeq
1694: \ep
1695:
1696: % For the computations involved in the proof and later purposes we note the following
1697: % formulas for ${\cal Z}$:
1698: % \beq\lbl{eq-AZrel}
1699: % \begin{array}{lll}
1700: % a)\;{\cal Z}^2=1\;\;\; & c)\;{\cal Z}(1\otimes w)=(K\otimes w){\cal Z} \;\;\;& e)\;
1701: % 1\otimes \Delta({\cal Z})={\cal Z}_{13}{\cal Z}_{12}\\
1702: % b)\;{\cal Z}^{\dagger}={\cal Z} & d)\;{\cal Z}(w\otimes 1)=(w\otimes K){\cal Z} & f)\;
1703: % \Delta\otimes 1({\cal Z})={\cal Z}_{13}{\cal Z}_{23}\\
1704: % \end{array}
1705: % \eeq
1706: % Here $w\in\E$ and the convention for ${\cal Z}_{ij}\in{\cal N}^{\otimes 3}$ is as in \cite% {Dri90}.
1707:
1708:
1709: For the monodromy matrix, as defined in (\ref{eq-M-matrix}), we obtain:
1710: \beq\lbl{eq-Amono}
1711: {\cal M} \;= \; 1\,+\,K\tb\otimes \th \,+\,\th K\otimes \tb-
1712: \rho\otimes\rho\;\;.
1713: \eeq
1714: Setting $T=K\tb\otimes \th \,+\,\th K\otimes \tb$ we compute $T^2=-2\rho\otimes\rho\,$ and
1715: $T^3=0\,$
1716: so that ${\cal M}=\exp(T)$. Hence we can also compute $p$-th powers of the monodromy matrix:
1717: \beq\lbl{eq-qAmono}
1718: {\cal M}^p\;=\;\exp(pT)\;=\;1\,+\,pT\,+\,\frac {p^2}2 T^2\;.
1719: \eeq
1720:
1721: With $\mu_0$ as defined in (\ref{eq-Ainteg}), and for $\rho$ as in (\ref{lm-Aribb})
1722: we find $\mu_0\otimes\mu_0({\cal M})=
1723: \mu_0(v)\mu_0(v^{-1})=-1$. Hence, in order to fulfill (\ref{eq-balnorm}) we need to use the
1724: renormalized integrals
1725: \beq\lbl{eq-renormint}
1726: \mu\;=\;i\mu_0\;\;, \qquad\;\;\lambda=\frac 1 i\lambda_0\;\;,
1727: \qquad\;\; \mbox{with}\;\;\;i=\sqrt{-1}\;\;.
1728: \eeq
1729: For these choices we compute the $S^{\pm}$-transformations assigned to (\ref{eq-S-matrix}) as follows:
1730: \beq\lbl{eq-ASmat}
1731: \begin{array}{lll}
1732: \frac 1 iS^{\pm}(w)=\mp w\;\;\;\forall w\in\E &\qquad \frac 1 i S^{\pm}(\rho)=1\\
1733: \frac 1 iS^{\pm}(Kx)=\,0\;\;\;\forall x\in\ext * \E &\qquad \frac 1 i S^{\pm}(1)=-\rho\;.\\
1734: \end{array}
1735: \eeq
1736: This implies that the projector $\Pi$ from Lemma~\ref{lm-Pi} has
1737: kernel $ker(\Pi)=\{Kw\, :\, w\in\ext *\E\}$ and image
1738: \beq\lbl{eq-APi}
1739: {\cal N}_0\;\;=\;\;im(\Pi)\;\;\;=\;\;\ext * \E\;.
1740: \eeq
1741: From (\ref{eq-somefor}) we see that ${\cal N}_0$ acts trivially
1742: on itself so that the action of ${\cal N}$ factors through the obvious
1743: $\Z/2\Z={\cal N}/{\cal N}_0$-action.
1744:
1745:
1746: Finally, we note that $\SL(2,\R)$ acts on $\E$ and, hence, also on $\cal N$,
1747: assuming $K$ is $\SL(2,\R)$-invariant.
1748: \begin{lemma}\lbl{lm-ASLinv}
1749: $\SL(2,\R)$ acts on $\cal N$
1750: by Hopf algebra automorphisms.
1751:
1752: The ribbon element $v$, the monodromy $\cal M$, and the two integrals
1753: are invariant under this action.
1754: \end{lemma}
1755:
1756: {\em Proof:} The fact that $\SL(2,\R)$ yields algebra automorphsims is
1757: obvious by construction. Linearity of coproduct and antipode in $w$
1758: in (\ref{eq-coprod}) and (\ref{eq-antipode}) imply that this is, in fact,
1759: a Hopf algebra homomorphism. $v$ and $\lambda$ are invariant
1760: since $\SL(2,\R)$ acts trivially on $\E\wedge\E\,$. Invariance of $\cal M$
1761: follows then from (\ref{eq-M-matrix}).
1762: \ep
1763:
1764: Note, that $\cal R$ itself is {\em not} $\SL(2,\R)$-invariant.
1765:
1766:
1767: \bigskip
1768:
1769: \head{7. The Hennings TQFT for $\cal N$}\lbl{S7}
1770:
1771:
1772: From (\ref{eq-APi}) and (\ref{eq-redtqft}) we see that the vector spaces of the
1773: Hennings TQFT for the algebra from (\ref{eq-Adef}) are given as
1774: \beq\lbl{eq-AVvect}
1775: {\cal V}_{\cal N}(\Sigma_g)\;\;:=\;\;\Bigl(\ext *\E\Bigr)^{\otimes g}
1776: \qquad\mbox{with}\qquad {\rm dim}({\cal V}_{\cal N}(\Sigma_g))=4^g\;.
1777: \eeq
1778: We now compute the action of the mapping class group generators from the
1779: tangles in Figures~\ref{fig-A-tgl}, \ref{fig-D-tgl}, and \ref{fig-S-tgl}.
1780: From the extended Hennings rules it is clear that
1781: the pictures for both $A_j$ and $S_j$ result in actions only on the $j$-th
1782: factor in the tensor product in (\ref{eq-AVvect}). For $A_j$ we use the presentation
1783: from Figure~\ref{fig-A-tgl} and the rules from Figure~\ref{fig-Rigg-tgl} and
1784: (\ref{pic-Rthru}) to obtain the linear map $\Aa(x):=x\cdot v$.
1785:
1786: The extra 1-framed circle in Figure~\ref{fig-S-tgl} results in an extra factor
1787: $\mu(v)=i$, since an empty circle corresponds to an insertion of $v$. The action on
1788: the $j$-th factor is thus given by an application of $\Ss:=iS^{+}\bigl |_{{\cal N}_0}$
1789: so that
1790: \beq\lbl{eq-SSform}
1791: \Ss(\rho)=-1\;,\qquad\quad \Ss(1)=\rho\;,\qquad{\rm and}\qquad \Ss(w)=w\;,\quad\forall w\in\E\;\;.
1792: \eeq
1793:
1794: Similarly, $D_j$ acts only on the $j$-th and the $(j+1)$-st
1795: factors of ${\cal N}_0^{\otimes g}$. From (\ref{pic-Rthru}) and the formula
1796: for ${\cal M}^{-1}$ in (\ref{eq-Amono}) we compute for the action on these two factors
1797: \beq\lbl{eq-D-2act}
1798: \D\;:\;{\cal N}_0^{\otimes 2}\,\to\,{\cal N}_0^{\otimes 2}, \qquad
1799: x\otimes y \;\mapsto\; x\otimes y + x\th\otimes \tb y -
1800: x\tb\otimes \th y - x\rho \otimes \rho y\;.
1801: \eeq
1802: The generators of the mapping class group $\Gamma_g$ are thus represented as follows:
1803: \beq\lbl{eq-AVgen}
1804: \begin{array}{ll}
1805: {\cal V}_{\cal N}({\bf I}_{A_j})\;\;=\;\;I^{\otimes j-1}\otimes \Aa\otimes I^{\otimes g-j}\;,
1806: &\qquad \quad {\cal V}_{\cal N}({\bf I}_{S_j})\;\;=\;\;I^{\otimes j-1}\otimes \Ss\otimes I^{\otimes g-j}\\
1807: &\\
1808: \qquad\qquad\mbox{and}&
1809: {\cal V}_{\cal N}({\bf I}_{D_j})\;\;=\;\;I^{\otimes j-1}\otimes \D\otimes I^{\otimes g-j-1}\;.\\
1810: \end{array}
1811: \eeq
1812: Let us also compute the linear maps associated to the cobordisms ${\bf H}_g^{\pm}$
1813: from (\ref{eq-hdlcob}). Their tangle presentations follow from \cite{Ker99} and
1814: have the forms given in Figure~\ref{fig-tglhdl}.
1815:
1816: \begin{figure}[ht]
1817: \begin{center}
1818: \leavevmode
1819: \epsfysize=3.2cm
1820: \epsfbox{HG-tgls.eps}
1821: \caption{Tangles for Handle additions}\lbl{fig-tglhdl}
1822: \end{center}
1823: \end{figure}
1824:
1825: We included $\pm 1$-framed circles to adjust the 2-framings of ${\bf H}_g^{\pm}$.
1826: A 0-framed circle around a
1827: strand has the effect of inserting $\lambda= S^+(1)=\frac 1 i\rho$. In this
1828: normalization we find with $\rho=i\Pi\lambda$ and (\ref{pic-Rbot}) that
1829: \beq\lbl{eq-Ahdl+map}
1830: {\cal V}_{\cal N}({\bf H}_g^{+})\;:\;\;\alpha\;\mapsto\;\;\alpha\otimes \rho
1831: \qquad \quad \forall \alpha \in{\cal N}_0^{\otimes g}\;.
1832: \eeq
1833: Similarly, we obtain from (\ref{pic-Rtop}) that
1834: \beq\lbl{eq-Ahdl-map}
1835: {\cal V}_{\cal N}({\bf H}_g^{-})\;:\;\;\alpha\otimes x\;\mapsto\;\;\mu_0(x)\alpha
1836: \qquad \quad \forall \alpha \in{\cal N}_0^{\otimes g}, \; x\in {\cal N}_0\;,
1837: \eeq
1838: where $\mu_0$ is as in (\ref{eq-Ainteg}). We note the following:
1839:
1840: \begin{lemma}\lbl{lm-ASL2inv}
1841: The generators in (\ref{eq-AVgen}), (\ref{eq-Ahdl+map}), and (\ref{eq-Ahdl-map})
1842: intertwine the $\SL(2,\R)$-action on ${\cal N}_0^{\otimes g}$.
1843: \end{lemma}
1844:
1845: {\em Proof:} The fact that $\Aa$ and $\D$ commute with the $\SL(2,\R)$-action follows from
1846: invariance of $v$ and $\cal M$. From (\ref{eq-ASmat}) we see that $\Ss$ is scalar on the
1847: non-invariant part, and thus commutes as well. Finally, $\rho$ and $\mu_0$ are clearly invariant.
1848:
1849: \ep
1850:
1851: For $g\geq 0$ set $\chi_g:=S_g\circ\ldots\circ S_1$, $h^+_g:= {\bf H}_{g-1}^+\circ \ldots
1852: \circ {\bf H}_{0}^+$, and $h^-_g:={\bf H}_{0}^-\circ \ldots
1853: \circ {\bf H}_{g-1}^-$. We define a standard closure of a 2-framed 3-cobordism as the closed
1854: 3-manifold
1855: \beq\lbl{eq-closM}
1856: \lz M\rz\;\;:=\;\;h^-_{g_2}\circ\chi_{g_2}\circ M\circ \chi^{-1}_{g_1}\circ h^+_{g_1}\;\cup\,D^3\;.
1857: \eeq
1858: If $M$ is represented by a tangle $T$ we obtain, similarly, a link $\lz T\rz$. We introduce
1859: the following function from the class of 2-framed cobordisms into $\Z/2$:
1860: \beq\lbl{eq-number}
1861: \varphi(M)\quad:=\quad \beta_1(\lz M\rz)\;+\;{\rm sign}(\lz T\rz)\;\;{\rm mod}\; 2\;,
1862: \eeq
1863: where $\beta_j$ denotes the $j$-th Betti number.
1864: We further denote by $\Cob^{22fr,*}\subset \Cob^{2fr,*}$ the subset of all cobordisms $M$
1865: with $\varphi(M)=0$, which we will call {\em evenly} 2-framed.
1866:
1867: \begin{lemma}\lbl{lm-mod2prop}
1868: \begin{enumerate}
1869: \item $\varphi(M)\;=\;|\lz T\rz|\;{\rm mod}\, 2$, where $|\lz T\rz|\,:=$ $\#$ components of $\lz T\rz$.
1870: \item $\varphi(M)\;=\;\#$ components of $ T$ not connected to the bottom line.
1871: \item ${\cal V}_{\cal N}(M)$ is real if $\varphi(M)=0$ and imaginary for $\varphi(M)=1$.
1872: \item $\Cob^{22fr,*}$ is a subcategory.
1873: \end{enumerate}
1874: \end{lemma}
1875:
1876:
1877: {\em Proof:} Let $W$ be the 4-manifold given by adding 2-handles to $D^4$ along
1878: $\lz T\rz\subset S^3$ so that $\lz M\rz=\partial W$, and let $L_T$ be the linking matrix of
1879: $\lz T\rz$. We have $\beta_2(W)= |\lz T\rz |=d_++d_-+d_0$, where $d_+$, $d_-$, and $d_0$ are
1880: the number of eigenvalues of $L_T$ that are $>0$, $<0$, and $=0$ respectively.
1881: From the exact sequence
1882: $0\to H_2(\lz M\rz )\to H_2(W) \stackrel {L_T}{\longrightarrow} H^2(W)\to H_1(\lz M\rz)\to 0$
1883: we find that $\beta_1(\lz M\rz)=d_0$, which implies {\em 1.} using ${\rm sign}(W)=d_+-d_-$.
1884: {\em 2.} follows immediately from the respective tangle compositions.
1885:
1886: The possible components
1887: not connected to the bottom line are strands connecting point pairs at the top line
1888: or closed components. From the rules (\ref{pic-Rclos}) through (\ref{pic-Rthru}) we see
1889: that these are just the types of components that involve an evaluation against $\mu=i\mu_0$.
1890: All other parts of the Hennings procedure involve only real maps. Finally, {\em 4.} follows
1891: from counting tangle components under composition.
1892: \ep
1893:
1894: \begin{propos}\lbl{pp-ATQFT}
1895: The Hennings procedure yields a relative, 2-framed, $\SL(2,\R)$-equivariant, half-projective TQFT
1896: $$
1897: {\cal V}_{\cal N}\;:\;\;\Cob^{2fr,\bullet}\;\longrightarrow\;\SL(2,\R)-mod_{\Cc}\;,
1898: $$
1899: which is $\Z/4$-projective on $\Cob^{\bullet}$. We have a restriction
1900: $$
1901: {\cal V}_{\cal N}^{(2)}\;:\;\;\Cob^{22fr,\bullet}\;\longrightarrow\;\SL(2,\R)-mod_{\R}\;,
1902: $$
1903: which is $\Z/2$-projective on $\Cob^{\bullet}$.
1904: \end{propos}
1905:
1906:
1907: {\em Proof:} From Lemma~\ref{lm-ASL2inv} we know that the generators of $\Gamma_g$ are
1908: represented $\SL(2,\R)$-equivariantly, hence also $\Gamma_g$ itself. The decomposition in
1909: (\ref{eq-Heegaard}) and equivariace of the maps in (\ref{eq-Ahdl+map}) and
1910: (\ref{eq-Ahdl-map}) implies the same for general cobordisms. That this TQFT is
1911: half-projective follows from the fact that $\cal N$ is non-semisimple, or, equivalently,
1912: that ${\cal V}_{\cal N}(S^1\times S^2)=\mu(1)=\varepsilon(\lambda)=0$, see \cite{Ker98}.
1913: The projective phase of the TQFT is determined by the value $\mu(v)=i$ on the 1-framed
1914: circle.
1915:
1916: Lemma~\ref{lm-mod2prop}, {\em 3.} implies that ${\cal V}_{\cal N}^{(2)}$ maps into
1917: the {\em real} $\SL(2,\R)$-equivariant maps and modules. This reduces the
1918: ambiguity of multiplication with $i$ to a sign ambiguity.
1919: \ep
1920:
1921: \medskip
1922:
1923: An important point of view in the TQFT constructions in \cite{KerLub00}
1924: is the existence of a categorical Hopf algebra, which can be understood as
1925: the TQFT image of a topological Hopf algebra given as an object in $\Cob^{\bullet}$.
1926:
1927: To be more precise, in
1928: \cite{Yet97} and \cite{Ker96} $\Cob^{\bullet}$ is described as a braided tensor
1929: category, and it is found
1930: that the object $\Sigma_{1,1}\in \Cob^{\bullet}$ is naturally identified as a
1931: braided Hopf algebra in this category in the sense of \cite{Maj93} and \cite{LubMaj94}.
1932: Particularly, $\Sigma_{2,1}$ is identified with $\Sigma_{1,1}\otimes\Sigma_{1,1}$
1933: since the tensor product on $\Cob^{\bullet}$ is defined by sewing two
1934: surfaces together along a pair of pants. The
1935: multiplication and comultiplication are thus given by
1936: elementary cobordisms ${\bf M}:\Sigma_{2,1}\to\Sigma_{1,1}$ and
1937: ${\bf \Delta} :\Sigma_{1,1}\to \Sigma_{2,1}$. Their tangle diagrams are worked out
1938: explicitly in \cite{BKLT00}, and depicted in Figure~\ref{fig-MC-tgl}
1939: with minor modifications in the conventions:
1940:
1941: \begin{figure}[ht]
1942: \begin{center}
1943: \leavevmode
1944: \epsfysize=2.2cm
1945: \epsfbox{MC-tgl.eps}
1946: \caption{Tangles for Mulitplications}\lbl{fig-MC-tgl}
1947: \end{center}
1948: \end{figure}
1949:
1950: Here ${\bf c}:\Sigma_{2,1}\to\Sigma_{2,1}$ is the braid isomorphism.
1951: The braided antipode is given by the tangle
1952: $\Gamma=(S^+)^2$, with $S^+$ as in Figure~\ref{fig-Smat}.
1953:
1954: \begin{lemma}
1955: \lbl{lm-multHeeg}
1956: The cobordisms {\bf M} and ${\bf \Delta}$ have the following
1957: Heegaard decompositions.
1958: $$
1959: {\bf M}\;=\;{\bf H}_2^-\circ {\bf I}_{D_1 \circ S_2} \qquad\mbox{and}
1960: \qquad {\bf \Delta}\;=\;
1961: {\bf I}_{S_1\circ D_1^{-1} \circ S_1^{-1}\circ S_2^{-1}}
1962: \circ{\bf H}_2^+
1963: $$
1964: \end{lemma}
1965:
1966: {\em Proof:} Verification by composition of the associated tangles.\ep
1967:
1968: The explicit formulae for the linear maps associated to the generators
1969: of the mapping class group and the handle attachments in Section~7 allow us now to
1970: compute the braided Hopf algebra structure induced on
1971: ${\cal N}_0={\cal V}_{\cal N}(\Sigma_{1,1})$. We write
1972: $M_0:={\cal V}_{\cal N}({\bf M})$, $\Delta_0:={\cal V}_{\cal N}({\bf \Delta})$,
1973: $S_0:={\cal V}_{\cal N}(S_1^2)$, and $c_0:={\cal V}_{\cal N}({\bf c})$ for
1974: the braided multiplication, comultiplication, antipode and braid isomorphism
1975: respectively.
1976:
1977: \begin{lemma}\lbl{lm-super-Hopf}
1978: The induced braided Hopf algebra structure on ${\cal N}_0$ is the canonical $\Z/2$-graded
1979: Hopf algebra with:
1980: $$
1981: M_0(x\otimes y)= xy\qquad\quad c_0(x\otimes y)=(-1)^{d(x)d(y)}y\otimes x\qquad\quad
1982: \forall x,y\in {\cal N}_0
1983: $$
1984: $$
1985: \mbox{and} \qquad \Delta_0(w)=w\otimes 1 + 1\otimes w \qquad\quad
1986: \Gamma_0(w)=-w \qquad\forall w\in\E\quad.
1987: $$
1988: In particular, ${\cal N}_0$ is commutative and cocommutative in the graded and braided sense,
1989: ${\cal N}_0\cong{\cal N}_0^*$ is self dual, $\SL(2,\R)$ still acts by Hopf automorphisms
1990: on ${\cal N}_0$, and $S_0$ is an involutory homomorphism on ${\cal N}_0$.
1991: \end{lemma}
1992:
1993: {\em Proof:} For {\bf M} and ${\bf \Delta }$ insert the morphism associated to
1994: the generators in Lemma~\ref{lm-multHeeg}. The braid isomorphism is given via the Hennings
1995: rules by acting with the operator $ad\otimes ad({\cal R})$ on
1996: ${\cal N}_0^{\otimes 2}$ and then permuting the factors. It is easy to see that
1997: $ad\otimes ad({\cal Z})$ acts on $x\otimes y$ by multiplying $(-1)^{d(x)d(y)}$,
1998: where $d(x)$ is the $\Z/2$-degeree of $x$ in ${\cal N}_0$. Moreover, we we know
1999: that the adjoint action of ${\cal N}_0$ on itself is trivial so that
2000: the term $\th\otimes K\tb$ in the second factor of ${\cal R}$ in (\ref{eq-ARmat})
2001: does not contribute.
2002: \ep
2003:
2004:
2005: \smallskip
2006:
2007:
2008: \medskip
2009:
2010:
2011: \head{8. Skein theory for ${\cal V}_{\cal N}$}\lbl{S8}
2012:
2013: The skein theory of the Hennings calculus over ${\cal N}$ is mostly a consequence
2014: on the form
2015: $v=1+\rho$ of the ribbon element
2016: as in (\ref{lm-Aribb}). In the Hennings procedure we substitute a strand
2017: with decoration $\frac 1 i\rho$ by a dotted strand (with possibly more decorations)
2018: as shown on the left of Figure~\ref{fig-rhostrand}. Observe from (\ref{eq-Amono}) that
2019: $$
2020: {\cal M}^{\pm 1}(1\otimes \rho)=(1\otimes \rho)\qquad\mbox{and}
2021: \qquad {\cal M}^{\pm 1}(\rho\otimes 1)=(\rho\otimes 1).
2022: $$
2023: This means that for a dotted strand we do not have to distinguish between over
2024: and undercrossing with other strands as indicated on the right of Figure~\ref{fig-rhostrand}.
2025: As a result such a strand can be
2026: disentangled from the rest of the diagram.
2027:
2028: \begin{figure}[ht]
2029: \begin{center}
2030: \leavevmode
2031: \epsfbox{rhostrand.eps}
2032: \end{center}
2033: \caption{Transparent $\rho$-decorated strand}\lbl{fig-rhostrand}
2034: \end{figure}
2035:
2036: The next additional ingredient in the calculus are symbols for 1-handles. They are used
2037: in the bridged link calculus as described in \cite{Ker99} and \cite{KerLub00}. We indicate
2038: a pair of 1-surgery balls by pairs of coupons. The defining relation is the modification
2039: move depicted on the left of Figure~\ref{fig-modif}. The move indicated on the right
2040: of Figure~\ref{fig-modif} and its reflections is a standard consequence of the boundary
2041: move from (\ref{eq-boundmove}).
2042:
2043: \begin{figure}[h]
2044: \begin{center}
2045: \leavevmode
2046: \epsfbox{modif.eps}
2047: \end{center}
2048: \caption{Coupons for 1-handles}\lbl{fig-modif}
2049: \end{figure}
2050:
2051: Since $v^{k}=1+k\rho$ for $k\in\Z$ we find that the framing of any component can be changed
2052: at the expense of introducing dotted lines. This translates to the diagrams in
2053: Figure~\ref{fig-frame}.
2054:
2055: \begin{figure}[h]
2056: \begin{center}
2057: \leavevmode
2058: \epsfbox{frame.eps}
2059: \end{center}
2060: \caption{Framing shift}\lbl{fig-frame}
2061: \end{figure}
2062:
2063: The skein relation is now obtained by applying Figure~\ref{fig-frame} to
2064: the Fenn Rourke move as in Figure~\ref{fig-FR}, see also\cite{MatPol94}.
2065:
2066: \begin{figure}[h]
2067: \begin{center}
2068: \leavevmode
2069: \epsfysize=2.6cm
2070: \epsfbox{FRmove.eps}
2071: \end{center}
2072: \caption{Fenn Rourke Move}\lbl{fig-FR}
2073: \end{figure}
2074:
2075: \begin{lemma}\label{lm-skein} For two strands belonging to two different components of a tangle
2076: diagram we have the relation
2077: \begin{center}
2078: \leavevmode
2079: \epsfbox{skein.eps}
2080: \end{center}
2081: For strands belonging to the same component of the tangle the relation is
2082: \begin{center}
2083: \leavevmode
2084: \epsfbox{skeinsame.eps}
2085: \end{center}
2086: \end{lemma}
2087:
2088: At this point it is convenient to extend the tangle presentations to general diagrams,
2089: dropping the condition that a strand starting at a point $j^-$ has to end at a point
2090: $j^+$ (or the corresponding condition for through strands). From such a general tangle
2091: diagram we can get to an admissible one by applying boundary moves (\ref{eq-boundmove})
2092: at all intervals $[j^-,j^+]$. (This is in fact the original definition used in
2093: \cite{Ker99}.) We shall allow the occurrence of coupons but restrict ourselves to the cases
2094: where exactly two strands enter (or exit) a coupon as in Lemma~\ref{lm-skein}.
2095:
2096: We also introduce two notions of components:
2097: The first is that of a {\em diagram component} $\cal X$ of a generalized tangle diagram.
2098: It is given by a concatenation of curve segments, coupons that have two strands going in
2099: on one side, and intervals $[j^-,j^+]$ connecting a strand ending in $j^-$ with the one
2100: ending in $j^+$.
2101:
2102: The second is a {\em strand component}, which is also a collection of curves that can be
2103: joined in two ways. As before curves that end in two sides of the same interval
2104: $[j^-,j^+]$ belong to the same strand component, as well as curves exiting and entering a coupon
2105: pair that would be connected under application of Figure~\ref{fig-modif}.
2106:
2107:
2108:
2109: We have the following rules for manipulating the coupons:
2110:
2111: \begin{lemma} \label{lm-1handlemoves}
2112: In the following equivalences the labels $A, B,\ldots$ indicate which coupons form a pair.
2113: \begin{enumerate}
2114: \item 1-handles can be slid over other 1-handles, through a boundary interval, and hence
2115: anywhere along a strand component.
2116: \beq\label{eq-1handlemove}
2117: \epsfbox{1handlemoves.eps}
2118: \eeq
2119: \item If in a diagram the coupons of a pair belong to different diagram components the entire
2120: diagram does not contribute, i.e., is evaluated as zero. Hence only diagrams contribute in which
2121: the diagram components coincide with strand components.
2122: \beq\label{eq-pairvanish}
2123: \epsfbox{pairvanish.eps}
2124: \eeq
2125: \item Direct 1-handle cancellation: If coupons with the same label are adjacent on the same
2126: side of a strand they can be canceled:
2127: \beq\label{eq-1handlecanc}
2128: \epsfbox{1handlecanc.eps}
2129: \eeq
2130: \item Opposite 1-handle cancellation: If coupons with the same label are adjacent on opposite
2131: sides of a strand the strand is replaced by a dotted strand and the evaluation gains a
2132: factor of 4.
2133: \beq\label{eq-1handleoppcanc}
2134: \epsfbox{1handleoppcanc.eps}
2135: \eeq
2136: \item If a generalized tangle diagram contains a coupon configuration as indicated the entire
2137: diagram is evaluated as zero.
2138: \beq\label{eq-coupontorus}
2139: \epsfbox{coupontorus.eps}
2140: \eeq
2141: \end{enumerate}
2142: \end{lemma}
2143:
2144:
2145: {\em Proof:} The slide of $B$ over the pair $A$ in (\ref{eq-1handlemove}) translates to
2146: a simple isotopy if we apply the move in
2147: Figure~\ref{fig-modif} to the $A$-pair. Similarly, the slide
2148: through a boundary interval is given by an isotopy conjugated by a $\sigma$-move as in
2149: (\ref{eq-boundmove}).
2150:
2151:
2152: For {\em b)}
2153: let $\cal X$ be a diagram component that contains coupons $A_1,\ldots, A_n$ whose partner
2154: lie on different diagram components. Performing boundary moves we can make $\cal X$ to be a true inner
2155: component. Furthermore, we can eliminate the other coupons on $\cal X$ that occur in pairs
2156: by undoing the modification from Figure~\ref{fig-modif}. The component $\cal X$ is now a
2157: closed curve interrupted only by coupons $A_1,\ldots, A_n$. We undo the modification also
2158: for these and the corresponding annuli added in the move
2159: bound discs that we denote by $D_1,\ldots, D_n$. Note, that the arcs of
2160: $\cal X$ all end in only one side of a disc $D_j$ since the strands emerging from the other
2161: side belong to a different component. We can thus surger the discs along the arcs, as shown in
2162: (\ref{eq-pairvanish}), so that we
2163: obtain a torus $T$ with $n$ holes
2164: $\partial T=\partial D_1\sqcup\ldots\sqcup\partial D_n$
2165: which misses all other parts of the tangle. After surgery along the annuli the
2166: torus $T$ can be capped off so that we have found a non-separating surface
2167: inside the represented cobordism. Since we are dealing with a non-semisimple
2168: TQFT this implies that the associated linear map is zero.
2169:
2170: The direct cancellation in (\ref{eq-1handlecanc}) follows by applying Figure~\ref{fig-modif}.
2171: In the resulting configuration in the middle of (\ref{eq-1handlecanc}) the Hopf link can
2172: be slid off and removed.
2173:
2174: The opposite cancellation in (\ref{eq-1handleoppcanc}) and the remodification from
2175: Figure~\ref{fig-modif} give the tangle in the middle. Now consider in general a
2176: straight strand that is entangled with an annulus with $2p$ positive crossings as
2177: in (\ref{eq-annuluswinding}).
2178: \beq\label{eq-annuluswinding}
2179: \epsfbox{annuluswinding.eps}
2180: \eeq
2181: Using the formula in (\ref{eq-qAmono}) we find by applying the Hennings procedure and
2182: evaluating the elements on the annulus against the integral
2183: that the resulting element on the open strand
2184: is
2185: $$
2186: \mu\otimes id({\cal M}^p)\;\;=\;\;\frac {p^2} i \,\rho
2187: $$
2188: which with Figure~\ref{fig-rhostrand} implies the claim.
2189:
2190: Finally, we also
2191: reexpress the coupons in
2192: in (\ref{eq-coupontorus}) by a tangle. As before non-semisimplicity
2193: of the TQFT implies that a diagram containing such a subdiagram is always zero. For example the
2194: 0-framed annulus clearly bounds a surface disjoint from the rest of the link so that the
2195: cobordism contains a non separating surface.
2196:
2197: \ep
2198:
2199:
2200:
2201: We now combine the previous two lemmas in the following skein relations without coupons.
2202:
2203: \begin{theorem} \label{thm-skein}
2204: For generalized tangle diagrams we have the following skein relations:
2205:
2206: For crossings of strands of different components:
2207:
2208: \beq\label{eq-SKEIN}
2209: \epsfbox{SKEIN.eps}
2210: \eeq
2211:
2212: For crossing of strands of the same component we need to introduce an orientation on the component.
2213:
2214: \beq\label{eq-SKEINsame}
2215: \epsfbox{SKEINsame.eps}
2216: \eeq
2217:
2218: \end{theorem}
2219:
2220: {\em Proof:} The proof is given by moving the coupons in the skein relations of
2221: Lemma~\ref{lm-skein} through the components using Lemma~\ref{lm-1handlemoves}.
2222: \ep
2223:
2224: Note, that relation (\ref{eq-SKEINsame}) implies the relation for the Kauffman
2225: polynomial for $z=\frac 12$. However, the framing relations are quite different.
2226:
2227: Let $\widehat {\cal B}_g$ be the group of tangles in $2g$ strands generated
2228: by the braidings $\bf c$ of double strands and the
2229: braided antipodes $\Gamma$ as in Figure~\ref{fig-MC-tgl}
2230: acting in different positions. It is thus the image of the abelian extension $B_g\ltimes {\Z/2}^g $
2231: of the braid group.
2232:
2233: Moreover, let us introduce a few elementary generalized tangles $M_k: k\to 0$, $\varepsilon:1\to 0$
2234: and $X_n:0\to 2n$ as depicted below.
2235:
2236: \beq\label{eq-planelem}
2237: \epsfbox{planelem.eps}
2238: \eeq
2239:
2240: \begin{theorem} \label{thm-solve}
2241: Every tangle $T:G\to 0$ with $2G$ starting (top) points and {\em no} endpoints
2242: can be resolved via the skein relations in Theorem~\ref{thm-skein} into a combination
2243: of tangles of the form
2244: $$
2245: T\;=\;(M_{k_1}\otimes\ldots M_{k_r}\otimes \epsilon^{\otimes N})\circ B\;,
2246: $$
2247: with $B\in \widehat {\cal B}_G$ and $\sum_{i=1}^rk_i=G-N$.
2248: \end{theorem}
2249:
2250: {\em Proof:} We consider generalized tangles without coupons. We proceed by induction on
2251: the number $m$ of connected components of $T$. We only count components that involve solid lines,
2252: those with dotted lines reduce to a collection of $\varepsilon$-diagrams at the intervals belonging to
2253: that component or closed dotted circles that do not contribute.
2254: Suppose now $T$ has only one component, which we equip with some orientation.
2255: Applying $\Gamma$'s to the intervals we can arrange it that the strands enter an interval
2256: $[j^-,j^+]$ at the left point $j^-$ and leave at the right one $j^+$. Furthermore, we can find a
2257: permutation of intervals so that the strand exiting $j^+$ enters at $(j+1)^-$, except for
2258: $G^+$, which is connected to $1^-$. Hence, by multiplying an element of $\widehat {\cal B}_G$
2259: to $T$ we can assume that the endpoints of the intervals are connected to each other by strands
2260: as they are for $M_G$.
2261:
2262: Next we note that the skein relation (\ref{eq-SKEINsame}) from Theorem~\ref{thm-skein} does not
2263: change this connectivity property for the solid lines and any diagram with dotted lines
2264: collapses to $\varepsilon$-diagrams.
2265:
2266: For diagrams where equally labeled coupons are on the same components
2267: there are three planar moves that allow us to
2268: manipulate the arrangement of coupons. They are the 1-handle slide and the
2269: 1-handle cancellation depicted below, and the boundary flip as in Figure~\ref{fig-modif}.
2270: In fact it is easy to see that
2271: we have the skein relation $T=M_G+iw(T)\varepsilon^{\otimes G}$, where $w(T)$ is the generalization
2272: of the writhe number of the diagram as defined, for example, in \cite{Lik97}. In case $G=0$ the diagram
2273: $M_0$ is a closed solid circle which therefore makes the entire diagram zero.
2274:
2275: Assume now $T$ has $m$ components and the claim is true for all diagrams with $m-1$ components.
2276: Pick one component $C$ and apply an element of $\widehat {\cal B}_G$ such that the intervals included in
2277: this component are all to the left of the other intervals. Note,
2278: that the set of intervals that belongs to
2279: $C$ may also be empty. Next apply the skein relations (\ref{eq-SKEIN})
2280: from Theorem~\ref{thm-skein} to untangle $C$ from the other components. In each step of changing
2281: crossings of a strand of $C$ with the strand of another component $D$ we can choose the relation
2282: for which the tangle that belongs to the first local diagram on the right side of the equation has
2283: one component less since $C$ and $D$ are connected. The other diagrams on the right side also
2284: have one less component since we do not count dotted lines. Hence, by induction, the error of changing
2285: a crossing between $C$ and another component can be resolved into elementary diagrams as claimed.
2286: After $C$ is untangled we have expressed $T$, modulo elementary diagrams,
2287: in the form $C\otimes T'$ (juxtaposition) where $T'$
2288: has $m-1$ components. Again each factor can be resolved independently by induction, and, hence,
2289: the whole
2290: diagram since $\otimes$-products of elementary diagrams are again elementary.
2291:
2292: \ep
2293:
2294:
2295: Next note that every tangle $R:\,g_1\to g_2$ is in fact of the form
2296: \beq\label{eq-RT}
2297: R\;=\; (T\otimes id_{g_2})\circ (id_{g_1}\otimes X_{g_2})
2298: \eeq
2299: for some $T:\,g_1+g_2\to 0$. Thus, in order to evaluate a general tangle diagram it suffices
2300: by Theorem~\ref{thm-solve} to specify the evaluations of the elementary tangles in (\ref{eq-planelem}). To this end we define the tensor
2301: \beq\label{eq-defAtens}
2302: A\;=\;\frac 1 i S\otimes 1\Delta(\rho)\;=\;\frac 1 i \Bigl( \rho\otimes 1\,+\,1\otimes\rho\,-\,\tb\otimes\th\,+\,\th\otimes\tb\Bigr)\;\;\in\,{\cal N}_0^{\otimes 2}\;.
2303: \eeq
2304:
2305: \begin{cor} \label{cor-eval}
2306: Every diagram can be resolved into a sum of composites of diagrams in
2307: (\ref{eq-planelem}). The linear maps associated to them are
2308: \beq\label{eq-X1}
2309: {\cal V}_{\cal N}(X_1)\,:\,\Cc\to {\cal N}_0^{\otimes 2}\,:\;1\,\mapsto
2310: A\,=\,\sum_{\nu}x_{\nu}\otimes y_{\nu}
2311: \eeq
2312: \begin{eqnarray}\label{eq-Xn}
2313: {\cal V}_{\cal N}(X_n)\;&=&\;
2314: (1^{\otimes (n-1)}\otimes {\cal V}_{\cal N}(X_1)\otimes1^{\otimes (n-1)})\circ
2315: {\cal V}_{\cal N}(X_{n-1})\,
2316: \;
2317: :\, \Cc\to {\cal N}_0^{\otimes 2n}\,
2318: \\
2319: &:&\;
2320: 1\,\mapsto
2321: A_{\{n\}}\,
2322: =\,\sum_{\nu_1,\ldots,\nu_n}x_{\nu_1}\otimes x_{\nu_2}\otimes \ldots\otimes x_{\nu_n}\otimes y_{\nu_n}
2323: \otimes \ldots\otimes y_{\nu_2} \otimes y_{\nu_1}\nonumber
2324: \end{eqnarray}
2325: \beq\label{eq-Mn}
2326: {\cal V}_{\cal N}(M_n):\;\; {\cal N}_0^{\otimes n}\to \Cc\;\;:\;\;\;\;a_1\otimes\ldots\otimes a_n\;\mapsto\;\mu(a_1\cdot\ldots\cdot a_n)
2327: \eeq
2328:
2329: Dotted circles can be removed and diagrams with solid circles do not contribute.
2330: \end{cor}
2331:
2332: {\em Proof:} The formulae follow easily from the pictures in Figure~\ref{fig-MC-tgl} to which we
2333: assigned linear maps in Lemma~\ref{lm-super-Hopf}. Particularly, we find that the upside down
2334: reflection of the multiplication tangle {\bf M} is mapped to the S-conjugate coproduct
2335: \beq\lbl{eq-Scoprod}
2336: \widetilde \Delta= i \Ss^{-1}\otimes\Ss^{-1}\Delta_0\Ss\,:\;{\cal N}_0\otimes{\cal N}_0\to
2337: {\cal N}_0\;.
2338: \eeq
2339: The tangle $X_1$ is obtained by capping this off with an arc at the top, which corresponds to
2340: the insertion of the unit. Hence, $A=\widetilde \Delta(1)$. The diagrams $M_p$ are easily
2341: identified as composites $M^p=(M\otimes 1^{\otimes(p-1)})\circ M^{p-1}$ capped off with
2342: an arc at the bottom, which is hence assigned to the $p$-fold multiplication followed
2343: by an evaluation against the integral $\mu\in{\cal N}^*$.
2344: \ep
2345: \medskip
2346:
2347: Let us consider a few examples. One useful case is when the braid $B\in\widehat{\cal B}_n$ can
2348: be chosen trivially. Hence the contribution to the linear map for a tangle $R:g_1\to g_2$ is
2349: given by a union of planar diagrams as depicted in (\ref{eq-plancomp}):
2350: \beq\label{eq-plancomp}
2351: \epsfbox{plancomp.eps}
2352: \eeq
2353: Define the map
2354: \beq\label{eq-defCpq}
2355: C_p^q\,=\,\widetilde\Delta^{q-1}\circ M_0^{p-1}\;:
2356: \quad {\cal N}_0^{\otimes p}\longrightarrow {\cal N}_0^{\otimes q}\;,
2357: \eeq
2358: where the exponents denote the usual multiple products and coproducts.
2359: The linear map associated to a planar diagram is now the tensor product
2360: of maps associated to the individual components of the diagram. For example,
2361: if we want to evaluate the linear map on a homogeneous vector $x_1\otimes\ldots\otimes x_{g_1}$
2362: an the diagram has
2363: a component with solid lines as in (\ref{eq-plancomp}) containing
2364: top intervals $[i_1^-,i_1^+], \ldots, [i_p^-,i_p^+]$ and bottom intervals
2365: $[j_1^-,j_1^+], \ldots, [j_q^-,j_q^+]$ we compute the vector
2366: $C_p^q(x_{i_1}\otimes\ldots \otimes x_{i_p})\in {\cal N}_0^{\otimes q}$ and insert the
2367: entries in order into the positions $j_1,\ldots, j_q$ in ${\cal N}_0^{\otimes g_2}$.
2368:
2369:
2370:
2371: With these rules the computation of the maps associated to the generators of the
2372: mapping class group are readily carried out. For
2373: example we can evaluate the diagram for the $S$-transformation from
2374: Figure~\ref{fig-S-tgl}. We resolve the right most crossing by taking the
2375: skein relation in the first row in Proposition~\ref{thm-skein} but with every
2376: diagram rotated clockwise by $\frac \pi 2$. The result is
2377: $$
2378: \Ss\;=\;id - \rho\otimes\mu_0 - 1\otimes \epsilon - 1\otimes \mu_0 + \rho\otimes \epsilon
2379: $$
2380: This yields exactly the formula from (\ref{eq-SSform}).
2381:
2382:
2383: As another example we may consider the $C_1$ waist cycle in $\Sigma_2$. The diagram consists
2384: of four parallel strands with a 1-framed annulus around the second and third. We apply
2385: Figure~\ref{fig-frame} and then Figure~\ref{fig-modif} to this annulus. The resulting coupons
2386: can be canceled. We find
2387: $${\cal V}_{\cal N}({\bf I}_{C_1})=id - iC_1^1\,.
2388: $$
2389: This implies the formula for the $D$-transformation from (\ref{eq-D-2act}).
2390: \smallskip
2391:
2392:
2393: Finally, let us show how to use the skein calculus to find
2394: the precise formula for the invariant of a 2-framed
2395: closed 3-manifold presented by a link ${\cal L}\subset S^3$. It is
2396: basically given by the order of the first integral homology. More precisely,
2397: let
2398: \beq\label{eq-defeta}
2399: \eta(M)\;:=\;\left\{\begin{array}{cl}
2400: \Bigl| H_1(M,\Z)\Bigr| & \mbox{for}\; \beta_1(M)=0\\
2401: 0 & \mbox{for}\;\beta_1(M)>0\end{array}\right.
2402: \eeq
2403:
2404: \begin{lemma}\label{lm-defeta} For a given framed link ${\cal L}\subset S^3$ and
2405: $\eta$ as in (\ref{eq-defeta}) we have
2406: $$
2407: {\cal V}_{\cal N}(M_{\cal L})\;=\; i^{|{\cal L}|}det({\cal L}\cdot{\cal L})
2408: =\;\pm i^{|{\cal L}|} \eta(M)
2409: $$
2410: \end{lemma}
2411:
2412: {\em Proof:}
2413: By 2-handle slides we can move ${\cal L}$ into a link ${\cal L}^{\delta}$
2414: so that the linking form ${\cal L}^{\delta}\cdot {\cal L}^{\delta}$ is
2415: diagonal and equivalent to the original one ${\cal L}\cdot{\cal L}$. Suppose
2416: $f_j$ is the framing number of the $j$-th component ${\cal L}_j^{\delta}$.
2417: From Figure~\ref{fig-frame} we see that
2418: $$
2419: {\cal V}_{\cal N}({\cal L}^{\delta})=
2420: {\cal V}_{\cal N}({\cal L}^{\delta, -f_j}) + if_j
2421: {\cal V}_{\cal N}({\cal L}^{\delta}-{\cal L}_j^{\delta})
2422: $$
2423: Here, ${\cal L}^{\delta, -f_j}$ is the link in which the framing of the $j$-th
2424: component is shifted to zero. As a result the manifold represented by this
2425: link has non-trivial rational homology. Since ${\cal V}_{\cal N}$ is a
2426: non-semisimple theory this implies that
2427: ${\cal V}_{\cal N}({\cal L}^{\delta, -f_j})=0$. Iterating the above identity
2428: we find
2429: ${\cal V}_{\cal N}({\cal L}^{\delta})=\prod_{j=1}^{|{\cal L}|}(if_j)
2430: {\cal V}_{\cal N}(\emptyset)$. Clearly, $\prod_{j=1}^{|{\cal L}|}(f_j)$
2431: is the determinant of the linking form of ${\cal L}^{\delta}$ and
2432: hence also the one of ${\cal L}$.
2433:
2434: \ep
2435:
2436:
2437: \head{9. Equivalence of ${\cal V}_{\cal N}^{(2)}$ and ${\cal V}^{FN}$}\lbl{S9}
2438:
2439: In this section we compare the two topological quantum field theories ${\cal V}^{FN}$
2440: described in Section~3 and ${\cal V}_{\cal N}^{(2)}$ constructed in Section~7. We already
2441: found a number of general properties that are shared by both theories:
2442:
2443: By Lemma~\ref{lm-FNpart} and Proposition~\ref{pp-ATQFT} both theories are $\Z/2$-projective
2444: on $\Cob^{\bullet}$ and non-semisimple, fulfilling the property of Lemma~\ref{lm-vanish}.
2445: The $\Z/2$-projectivity is due to ambiguities of even 2-framings in the case of
2446: ${\cal V}_{\cal N}^{(2)}$ and ambiguities of orientations in the case of ${\cal V}^{FN}$.
2447: The non-semisimple half-projective property results in the case of ${\cal V}^{FN}$ from
2448: representation varieties that are transversely disjoint, and in the case of
2449: ${\cal V}_{\cal N}^{(2)}$ from the nilpotency
2450: of the integral $\lambda\in\cal N$.
2451: Further common features are the dimensions of vector spaces $(=4^g)$,
2452: actions of $\SL(2,\R)$, see Section~9, and the fact that ${\cal J}_g$ lies in the kernel
2453: of the mapping class group representations.
2454:
2455: We construct now an explicit isomorphism between ${\cal V}^{FN}$ and ${\cal V}_{\cal N}^{(2)}$.
2456: Let ${\cal Q}=\ext *\lz a,b\rz$ be the exterior algebra over $\R^2$ with basis $a,b\in\R^2$.
2457: We obtain a canonical isomorphism, which is defined on monomial elements as follows:
2458: \begin{equation}\lbl{eq-defi}
2459: i_*\;:\quad {\cal Q}^{\otimes g}\;\isto\;\ext * H_1(\Sigma_g)\;\;:\qquad
2460: q_1\otimes \ldots\otimes q_g\mapsto i_1(q_1)\wedge\ldots\wedge i_g(q_g) \;,
2461: \end{equation}
2462: where $i_j:{\cal Q}\,\isto\,\ext *\lz [a_j], [b_j]\rz$ is the canonical map sending $a$ and $b$
2463: to $[a_j]$ and $[b_j]$ respectively. Next,
2464: we define an isomorphism between ${\cal Q}$ and ${\cal N}_0$, seen as linear spaces,
2465: by the following assignment of basis vectors:
2466: %$\,\,\phi_j\,:\;\mbox{$\bigwedge^*$} (\lz \th,\tb\rz)\;
2467: %\longrightarrow\;\mbox{$\bigwedge^*$} (\lz a_j,b_j\rz)\;$
2468: \begin{equation}\lbl{eq-phidef}
2469: \begin{array}{lcc}
2470: &\quad \phi(1)=b \quad
2471: &\quad\phi(\tb\th)=a\quad\\
2472: \phi\;:\quad{\cal N}_0\;\isto\;{\cal Q}\qquad\;\;\mbox{with}\;\;&&\\
2473: &\qquad \phi(\th)=a\wedge b \quad &\quad \phi(\tb)=1\quad \\
2474: \end{array}\;.
2475: \end{equation}
2476: Note, that this map has odd $\Z/2$-degree and is, in particular,
2477: not an algebra homomorphism. From (\ref{eq-phidef}) we infer
2478: directly the following identities:
2479: \beq\lbl{eq-}
2480: \phi(\th x)=-\phi(x)\wedge a \qquad\qquad \phi(x \th )=a\wedge \phi(x)\\
2481: \eeq
2482: \beq\lbl{eq-AS1phi}
2483: \phi(\Aa x)\;=\;[A_1]\phi(x)\qquad\qquad \phi(\Ss x)\;=\;[S_1]\phi(x)\\
2484: \eeq
2485: Here, $\Aa$ and $\Ss$ are as in (\ref{eq-AVgen}), and $[A_1]$ and $[S_1]$ are the
2486: maps on $H_1(\Sigma_1)$ as in (\ref{eq-A-hom}) and (\ref{eq-S-hom}).
2487:
2488:
2489: Moreover, let us introduce a sign-operator $(-1)^\Lambda$ on ${\cal Q}^{\otimes g}$
2490: defined on monomials by
2491: \begin{equation}\lbl{eq-Lambda}
2492: (-1)^{\Lambda_g}(q_1\otimes \ldots\otimes q_g)\quad=\quad
2493: (-1)^{\lambda_g(d_1,\ldots,d_g)}q_1\otimes \ldots\otimes q_g\;.
2494: \end{equation}
2495: The function $\lambda_N$ is defined in the $N$-fold product of
2496: $\Z/2$'s as follows:
2497: \begin{equation}\lbl{eq-lambdadef}
2498: \lambda_N:\;(\Z/2)^N\,\to\,\Z/2\qquad\mbox{with}\quad
2499: \lambda_N(d_1,\ldots,d_N)\;=\;\sum_{i<j}d_i(1-d_j)\;,
2500: \end{equation}
2501: where $d_j={\rm deg}(q_j)\,{\rm mod}\, 2\,$. Consider now the following isomorphism
2502: of vector spaces.
2503: \begin{equation}\lbl{eq-defxi}
2504: \xi_g:=\;i_*\circ(-1)^{\Lambda_g}\circ\phi^{\otimes g}\quad:
2505: \qquad{\cal N}_0^{\otimes g}\,\;\isto\,\;\ext *H_1
2506: \end{equation}
2507: Given a linear map, $F:{\cal N}^{\otimes g_1}\to {\cal N}^{\otimes g_2}$, we write
2508: $(F)^{\xi}:=\xi_{g_2}\circ F\circ \xi_{g_1}^{-1}$ for the respective map on homology.
2509: Moreover, we denote by ${\bf L}^{(k)}_x$ the operator on ${\cal N}^{\otimes g}$
2510: that multiplies the $k$-th factor in the tensor product by $x$ from the left,
2511: and by ${\bf R}^{(k)}_x$ the respective operator for multiplication from the right. We
2512: compute:
2513: \beq\lbl{eq-LRdef}
2514: \begin{array}{cl}
2515: \quad &({\bf L}^{(k)}_{\th})^{\xi}(\alpha\wedge u_k\wedge\beta)\;\;=\;\;(-1)^{g-k+s+1}\,\,
2516: \alpha\wedge a_k\wedge u_k\wedge\beta\;,\\
2517: \\
2518: \mbox{and}\qquad &({\bf R}^{(k)}_{\th})^{\xi}(\alpha\wedge u_k\wedge\beta)\;\;=\;\;(-1)^{g-k+s}\,\,
2519: \alpha\wedge u_k\wedge a_k \wedge\beta\;,
2520: \end{array}
2521: \eeq
2522: where $s=\sum_{j=1}^g d_j$ is the total degree of $\alpha\wedge u_k\wedge\beta$,
2523: $\alpha\in\ext *\lz a_1,\ldots, b_{k-1}\rz$, and $\beta\in\ext *\lz a_{k+1},\ldots, b_{g}\rz$.
2524:
2525: \begin{lemma}\lbl{lm-commut}
2526: For every standard generator $G\in \{A_j, D_j, S_j\}$, we have
2527: $$
2528: ({\cal V}_{\cal N}({\bf I}_{G}))^{\xi}\;=\;\ext * [G]\;\;,
2529: $$
2530: where $[G]$ denotes as before the action on homology.
2531: \end{lemma}
2532:
2533: {\em Proof:} For the $A_j$ and $S_j$ this follows readily from (\ref{eq-AS1phi}), and
2534: the fact that $[A_j]$ and $[S_j]$ do not change the degrees $d_j$ and hence commute with
2535: $(-1)^{\Lambda_g}$.
2536:
2537: The operator in (\ref{eq-D-2act}) decomposes into $\D=\D^0+\D^1$, where
2538: $\D^0=id - {\bf R}_{\rho}\otimes {\bf L}_{\rho}$ and
2539: $\D^1={\bf R}_{\th}\otimes {\bf L}_{\tb}-{\bf R}_{\tb}\otimes {\bf L}_{\th}$.
2540: Now $\D^0$ does not change the $\Z/2$-degree of both factors, and $\D^1$ flips the
2541: degree of both factors. One readily verifies that
2542: $$
2543: \lambda_g(\ldots, 1-d_j, 1-d_{j+1},\ldots)-\lambda_g(\ldots, d_j, d_{j+1},\ldots)=d_j + d_{j+1}\qquad
2544: {\rm mod}\,2
2545: $$
2546: $$
2547: \mbox{so that}\qquad
2548: {\cal V}_{\cal N}({\bf I}_{D_j})^{\xi}\;\;=\;\;({\cal V}_{\cal N}^0({\bf I}_{D_j}))^{\zeta}\,+\,
2549: (-1)^{d_j+d_{j+1}}({\cal V}_{\cal N}^1({\bf I}_{D_j}))^{\zeta}\quad
2550: %\mbox{with}\;\;\zeta_g=i_*\circ \phi^{\otimes g}
2551: $$
2552: $$\qquad\quad=\;\;(I^{\otimes j-1}\otimes (\D^0)^{\phi^{\otimes 2}}\otimes I^{\otimes g-j-1})^{i_*}\;+\;(-1)^{d_j+d_{j+1}}(I^{\otimes j-1}\otimes (\D^1)^{\phi^{\otimes 2}}
2553: \otimes I^{\otimes g-j-1})^{i_*}
2554: $$
2555:
2556: Here, $\zeta_g=i_*\circ \phi^{\otimes g}$ and
2557: ${\cal V}_{\cal N}^i({\bf I}_{D_j})$ is the operator with $\D^i$ in $j$-th position.
2558: Since $\zeta_g=\zeta^{\otimes g}_1$ the $\zeta$-conjugate maps only act on the generators
2559: $\{a_j, b_j, a_{j+1}, b_{j+1}\}$ the action is the same for all positions $j\,$.
2560: Observe that also $[D_j]$ acts only on the homology generators $\{a_j, b_j, a_{j+1}, b_{j+1}\}$.
2561: It is, therefore, enough to prove the relation for $g=2$ and ${\cal V}_{\cal N}({\bf I}_{D_1})
2562: =\D$.
2563:
2564: Now, from (\ref{eq-D-2act}) it is obvious
2565: that ${\cal V}_{\cal N}({\bf I}_{D_j})$ commutes
2566: with ${\bf L}^{(j)}_{\th}$ and ${\bf R}^{(j+1)}_{\th}$. Moreover, it is easy to see that
2567: $\ext *[D_j]$, as given in (\ref{eq-D-hom}), commutes with
2568: $({\bf L}^{(j)}_{\th})^{\xi}$ and $({\bf R}^{(j+1)}_{\th})^{\xi}$ from (\ref{eq-LRdef}).
2569: Specifically, we use that $\ext *[D_j]$ does not change the total degree, and acts trivially on
2570: $a_j$ and $a_{j+1}$. It thus suffices to check
2571: \beq\lbl{eq-12Drel}
2572: \ext 2[D_1]\circ\zeta_2(x_1\otimes x_2)=\zeta_2\circ \D^0(x_1\otimes x_2) +(-1)^{d_1+d_2}
2573: \zeta_2\circ\D^1 (x_1\otimes x_2)
2574: \eeq
2575: with $d_i={\rm deg}(\phi(x_i))$, and only for $x_i\in\{1,\tb\}$. For example, for $x_1=x_2=1$,
2576: with $d_1+d_2=0$, we find from (\ref{eq-D-2act}) and (\ref{eq-D-hom}) that
2577: $$
2578: \begin{array}{lcl}
2579: \zeta_2 \circ \D(1\otimes 1)&=&\zeta_2(
2580: 1\otimes 1 + \th\otimes \tb -
2581: \tb\otimes \th - \rho \otimes \rho)\;\\
2582: &=& b_1\wedge b_2 + a_1\wedge b_1 - a_2\wedge b_2 - a_1\wedge a_2\\
2583: &=& (b_1-a_2)\wedge (b_2-a_1) \;=\;\ext 2[D_1](b_1\wedge b_2)\;=\;\ext 2[D_1](\zeta_2(1\otimes 1))\\
2584: \end{array}
2585: $$
2586: We also compute for the case $x_1=\tb$ and $x_2=1$, with $d_1+d_2=1$:
2587: $$
2588: \begin{array}{lcl}
2589: \zeta_2 \circ (\D^0-\D^1)(\tb\otimes 1)&=&\zeta_2(\tb\otimes 1 -\tb\th\otimes\tb)
2590: \;=\;b_2 - a_1\\
2591: &=& \ext 2[D_1](b_2)\;\;\;=\;\;\;\ext 2[D_1](\zeta_2(\tb\otimes 1))
2592: \end{array}\;.
2593: $$
2594: The other two cases follow similarly.
2595: \ep
2596:
2597: As the $\{A_j, D_j, S_j\}$ generate $\Gamma_g$ we conclude from Lemma~\ref{lm-commut} and
2598: (\ref{eq-FNmapcg}) that $({\cal V}_{\cal N}({\bf I}_{\psi}))^{\xi}
2599: ={\cal V}^{FN}({\bf I}_{\psi})$ for all $\psi\in\Gamma_g\,$.
2600:
2601: Let us also consider the maps associated by both functors to the
2602: handle additions ${\bf H}_g^{\pm}$. We note that
2603: $$
2604: \lambda_{g+1}(d_1,\ldots,d_g,1)=\lambda_g(d_1,\ldots,d_g)
2605: $$
2606: so that we find from (\ref{eq-Ahdl+map}), (\ref{eq-FNmcg}) and (\ref{eq-phidef}) that
2607: $({\cal V}_{\cal N}({\bf H}_g^+))^{\xi}={\cal V}^{FN}({\bf H}_g^+)$. Similarly,
2608: (\ref{eq-Ahdl-map}), (\ref{eq-FNgcm}) and (\ref{eq-Ainteg}) imply
2609: $({\cal V}_{\cal N}({\bf H}_g^-))^{\xi}={\cal V}^{FN}({\bf H}_g^-)$.
2610: Using the Heegaard
2611: decomposition (\ref{eq-Heegaard}) we finally infer equivalence:
2612: \begin{propos}\lbl{pp-Vequiv}
2613: The maps $\xi_g$ defined in (\ref{eq-defxi}) give rise to an isomorphism
2614: $$
2615: \xi\;:\; {\cal V}_{\cal N}\;\;\;\;\stackrel{\bullet\,\cong}
2616: {-\!\!\!-\!\!\!-\!\!\!\longrightarrow}\;
2617: \;\;\;{\cal V}^{FN}\;\;
2618: $$
2619: of relative, non-semisimple, $\Z/2$-projective functors from $\Cob^{\bullet}$ to $Vect(\kk)$.
2620: \end{propos}
2621:
2622:
2623:
2624: \head{10. Hard-Lefschetz decomposition and Invariants}\lbl{S10}
2625:
2626: The tangent bundle over the moduli space $J(\Sigma_g)$ is trivial with fiber
2627: $H^*(\Sigma_g,\R)$ so that its cohomology ring is naturally $\ext *H_1(\Sigma_g,\R)$.
2628: There is an almost complex structure on $J(\Sigma_g)$ given by a map ${\sf J}$ with
2629: ${\sf J}^2=-1$ in the cohomology. It is given by ${\sf J}.[a_j]=-[b_j]$ and
2630: ${\sf J}.[b_j]=[a_j]$.
2631: With the K\"ahler form
2632: $\omega_g\in H^2(J(\Sigma_g))$ defined in (\ref{eq-defsymform})
2633: it is also a K\"ahler manifold. The dual K\"ahler metric provides us with a Hodge star
2634: $\star:\ext jH_1(\Sigma_g)\to\ext {2g-j}H_1(\Sigma_g)$ for a given volume form
2635: $\Omega\in\ext {2g} H_1(\Sigma_g)$ by the equation
2636: $\alpha\wedge\star\beta=\lz\alpha,\beta\rz\Omega$.
2637: Specifically, the $2g$ generators $\{[a_1],\ldots, [b_g]\}$ of $H_1(\Sigma_g)$,
2638: with volume form $\Omega=[a_1]\wedge\ldots\wedge [b_g]$ the Hodge star is given by
2639: $\star(a_1^{1-\epsilon_1}\wedge\ldots\wedge b_g^{1-\epsilon_{2g}})
2640: =(-1)^{\lambda_{2g}(\epsilon_1,\ldots,\epsilon_{2g})}a_1^{\epsilon_1}\wedge\ldots\wedge
2641: b_g^{\epsilon_{2g}}$, where $\lambda_{2g}$ is as in (\ref{eq-lambdadef}).
2642:
2643:
2644: As a K\"ahler manifold $H^*(J(\Sigma_g))$ admits an $\SL(2,\R)$-action,
2645: see for example \cite{GriHar78}, given for the standard generators
2646: $E, F, H\,\in\,{\mathfrak s}{\mathfrak l}_2(\R)$ by
2647: \beq\lbl{eq-defsl2act}
2648: H\alpha:=(j-g)\alpha\quad\; \forall \alpha\in \ext jH_1(\Sigma_g)\;,
2649: \qquad\quad E\alpha:=\alpha\wedge\omega_g\;,\qquad\quad F:=\star\circ E\circ\star^{-1}
2650: \eeq
2651: \begin{lemma}\lbl{lm-FN-equiv}
2652: The functor ${\cal V}^{FN}$ is $\SL(2,\R)$-equivariant with respect to the action
2653: in (\ref{eq-defsl2act}).
2654: \end{lemma}
2655:
2656: {\em Proof:} Commutation with $H$ follows from counting degrees. Since $\omega_g$ is invariant
2657: under the $\Sp(2g,\R)$-action, $E$ commutes with the maps in (\ref{eq-FNmapcg}),
2658: and since $\omega_g\wedge [a_{g+1}]=[a_{g+1}]\wedge \omega_{g+1}$ also with the ones
2659: in (\ref{eq-FNmcg}) and (\ref{eq-FNgcm}). Finally, as all maps ${\cal V}^{FN}(M)$
2660: are isometries with respect to $\lz.,.\rz$ they also commute with $F$. \ep
2661:
2662: In order to finish the proof of Theorem~\ref{thm-main} we still need to show that
2663: the $\xi_g$ are $\SL(2,\R)$-equivariant as well. The fact that $H$ commutes with $\xi_g$
2664: is again a matter of counting degrees. We have $E=\sum (E_1^{(i)})^{i_*}$,
2665: where $E_1^{(i)}$ acts
2666: on the $i$-th factor of ${\cal Q}^{\otimes g}$
2667: by $q\mapsto E_1(q)= q\wedge a\wedge b$. Since $E$ does not change
2668: degrees we find that $E^{\xi}=\sum (E^{(i)})^{\phi^{(i)}}$, where
2669: $(E^{(i)})^{\phi^{(i)}}$ acts on the $i$-th factor by $E_1^{\phi}$.
2670: We find $E_1^{\phi}(\tb)=\th$, and
2671: $E_1^{\phi}(1)=E_1^{\phi}(\th)=E_1^{\phi}(\tb\th)=0$, which yields precisely
2672: the desired action of $E$
2673: on ${\cal N}_0$. The conjugate action of $\star$ on ${\cal N}_0^g$ is as follows:
2674: \beq\lbl{eq-xistar}
2675: \star^{\xi}\;:\;\; x_1\otimes\ldots x_g\;\;\mapsto \; (-1)^{\sum_{i<j}d_id_j}(\star x_1)
2676: \otimes\ldots\otimes (\star x_g)\qquad\;\forall x_j\in{\cal N}_0\;,
2677: \eeq
2678: where $\star \th=\tb$, $\star\tb=\th$, $\star\tb\th=1$, and $\star 1=-\tb\th$.
2679: From this we see that $F^{\xi}$ acts on each factor by
2680: $F_1^{\phi}(\th)=\tb$, and
2681: $F_1^{\phi}(1)=F_1^{\phi}(\tb)=F_1^{\phi}(\tb\th)=0$, as required.
2682:
2683: With Lemma~\ref{lm-FN-equiv} and equivariance of $\xi_g$
2684: we have thus completed the proof of Theorem~\ref{thm-main}.
2685: Henceforth, we will use the simpler notation ${\cal V}={\cal V}^{FN}={\cal V}_{\cal N}$
2686: \ep
2687:
2688: \medskip
2689:
2690: The $\SL(2,\R)$-action implies a Hard-Lefschetz decomposition \cite{GriHar78} as follows
2691: \beq\lbl{eq-Lefsch}
2692: H^*(J(\Sigma_g))\quad\cong\quad \bigoplus_{j=1}^{g+1} V_j\otimes W_{g,j}\;.
2693: \eeq
2694: Here, $V_j$ is the irreducible ${\mathfrak s}{\mathfrak l}_2$-module
2695: with ${\rm dim}(V_j)=j\,$, and
2696: \beq\lbl{eq-W}
2697: W_{g,j}\;:=\;\{u\in\ext {g-j+1}H_1(\Sigma_g)\,:\;\omega_g\wedge u=0\}
2698: \eeq
2699: is the space of {\em isotropic} vectors of degree $(g-j+1)$, or, equivalently, the
2700: space of ${\mathfrak s}{\mathfrak l}_2$-highest weight vectors of weight $(j-1)$. On each of these
2701: spaces we have an action of the mapping class groups from (\ref{eq-FNmapcg})
2702: factoring through $\Sp(2g,\R)$.
2703: \begin{theorem}[\cite{GooWal98} Chapter 5.1.8]\lbl{thm-GW}
2704: Each $W_{g,j}$ is an irreducible $\Sp(2g,\R)$-module with fundamental highest weight
2705: $\varpi_{g-j+1}$ and dimension
2706: $$
2707: {\rm dim}(W_{g,j})={2g\choose g-j+1}\,-\, {2g\choose g-j-1}
2708: $$
2709: In particular, the pair of subgroups
2710: $$
2711: \SL(2,\R)\,\times\,\Sp(2g,\R)\;\;\subset\;\;{\rm GL}(H^*(J(\Sigma_g)))
2712: $$
2713: forms a Howe pair, that is, the two subgroups are exact commutants of each other.
2714: \end{theorem}
2715: The fundamental weights are given as in \cite{GooWal98} by
2716: $\varpi_{k}=\epsilon_1+\ldots +\epsilon_{k}$ with $\epsilon_j$ as
2717: in (\ref{eq-sproots}).
2718:
2719:
2720:
2721:
2722:
2723: In the decomposition into irreducible TQFT's the one for $j=1$ associated to the
2724: trivial $\SL(2,\Cc)$ representation plays a special role for invariants of closed
2725: manifolds.
2726:
2727: For any invariant, $\tau$, of closed 3-manifolds there is a standard ``reconstruction''
2728: of TQFT vector spaces
2729: as follows. We take the formal $\kk$-linear span ${\mathfrak C}_g^+$ of
2730: cobordisms $M:\emptyset\to\Sigma_g$
2731: and ${\mathfrak C}_g^-$ of cobordisms $N:\Sigma_g\to\emptyset$. We obtain a pairing
2732: ${\mathfrak C}_g^-\times {\mathfrak C}_g^+\to\kk:\,(N,M)\to \tau(N\circ M)$. If
2733: ${\mathfrak N}^+_g\subset {\mathfrak C}_g^+$ is the null space of this pairing
2734: we define ${\cal V}^{\tau-rec}(\Sigma_g)={\mathfrak C}_g^+/{\mathfrak N}^+_g$.
2735: For generic $\tau$ these vector spaces are infinite dimensional. The exception is when $\tau$
2736: stems from a TQFT. In this case
2737: ${\cal V}^{\tau-rec}(\Sigma_g)^*={\mathfrak C}_g^-/{\mathfrak N}^-_g$, and the linear map
2738: ${\cal V}^{\tau-rec}(P)$ associated to a cobordism $P$ is reconstructed from its
2739: matrix elements $\tau(N\circ P\circ M)$. This construction, which basically imitates the
2740: GNS construction of operator algebras, is folklore since the emergence of TQFT's and
2741: appears, for example, in \cite{Tur94}.
2742:
2743: \begin{theorem}\lbl{cor-dec}
2744: \ben
2745: \item
2746: The TQFT functor from Theorem~\ref{thm-main} decomposes into a direct sum
2747: $$
2748: {\cal V}\;=\;\;\bigoplus \R^{j}\otimes {\cal V}^{(j)}\;\;\;=\;\;\;
2749: {\cal V}^{(1)}\,\oplus\, \R^2\otimes {\cal V}^{(2)}
2750: \,\oplus\, \R^3\otimes{\cal V}^{(3)}\,\ldots \,
2751: $$
2752: of irreducible TQFT's with multiplicities.
2753: \item
2754: The associated vector space for each TQFT is ${\cal V}^{(j)}(\Sigma_g)=W_{g,j}$
2755: so that ${\cal V}^{(j)}(\Sigma_g)=0$ whenever $j>g+1$.
2756: In particular, for any closed 3-manifold $M$ and $j>1$ we have ${\cal V}^{(j)}(M)=0$ so that
2757: ${\cal V}(M)={\cal V}^{(1)}(M)$.
2758: \item
2759: The vector spaces associated to the invariant $\pm\eta$ from (\ref{eq-defeta})
2760: are finite dimensional. The
2761: reconstructed $\Z/2$-projective TQFT is ${\cal V}^{\eta-rec}={\cal V}^{(1)}$ with dimensions
2762: ${\rm dim}({\cal V}^{\eta-rec}(\Sigma_g))=
2763: {\rm dim}(W_{g,1})=\frac 2{g+2}{2g+1\choose g}\,$.
2764: \een
2765: \end{theorem}
2766:
2767:
2768: {\em Proof:} The fact that the TQFT's decompose in the prescribed manner follows
2769: from the $\SL(2,\R)$-covariance. Irreducibility of each ${\cal V}^{(j)}$, meaning there
2770: are no proper sub-TQFT's, results from the fact that each $\Sp(2g,\Z)$ representation
2771: is irreducible so that in a sub-TQFT the vector spaces for each $g$ are either
2772: ${\cal V}^{(j)}(\Sigma_g)$ or 0. Since the handle maps yield non-zero maps between
2773: these vector spaces if one space is non-zero none of them can be. The reconstructed TQFT
2774: must be a quotient TQFT of ${\cal V}^{(1)}$, which is, however, irreducible.
2775: Hence, they are equal.
2776:
2777: \ep
2778:
2779: %%%%%%%%%%%%%%%%% BEGIN CUT %%%%%%%%%%%%%%%%%%%%%%%%%%%
2780:
2781: \emptystuff{
2782: From the irreducible TQFT's in Theorem~\ref{cor-dec} we can construct a much larger class of TQFT's,
2783: which appear to be related to higher rank gauge theories,
2784: as follows. Let ${\cal P}^+
2785: \subset {\Z}^{0,+}[x_1,x_2,\ldots]$ be the set of formal power series
2786: $$
2787: P(x_1,x_2,\ldots)=\sum_{k=1}^{\infty}\,\sum_{n_1, \ldots , n_k=1}^{\infty}c_{n_1,n_2,\ldots,n_k}
2788: x_1^{n_1}x_2^{n_2}\ldots x_k^{n_k}\;,
2789: $$
2790: such that all $c_{n_1,n_2,\ldots,n_k}\in {\Z}^{0,+}$ are non-negative integers, and for fixed $k$
2791: only finitely many $c_{n_1,n_2,\ldots,n_k}$ are non-zero. To every such $P$ we associate a
2792: TQFT by the formula
2793: \beq\label{eq-defpolyTQFT}
2794: {\cal V}^{(P)}\;=\;\bigoplus_{k=1}^{\infty}\bigoplus_{n_1, \ldots , n_k=1}^{\infty}
2795: \, \R^{c_{n_1,n_2,\ldots,n_k}}\otimes ({\cal V}^{(1)})^{\otimes n_1}\otimes
2796: ({\cal V}^{(2)})^{\otimes n_2}\otimes \ldots\otimes ({\cal V}^{(k)})^{\otimes n_k}
2797: \;.
2798: \eeq
2799: For example ${\cal V}^{FN}={\cal V}^{(F)}$, where $F(x_1,x_2,\ldots)=\sum_j j x_j$.
2800: The restriction on the coefficients together with the second part of Theorem~\ref{cor-dec}
2801: implies that all vector spaces are finite dimensional.
2802: \begin{lemma}\label{lm-polyTQFT}
2803: The TQFT functor ${\cal V}^{(P)}$ is well defined for every $P\in {\cal P}^+$ .
2804: \end{lemma}
2805: }
2806:
2807: %%%%%%%%%%%%%%%%%%% END CUT %%%%%%%%%%%%%%%%%%%%%%
2808:
2809:
2810: Let us finally give an alternative proof of Lemma~\ref{lm-defeta} using the language in
2811: which the Frohman Nicas invariant is constructed.
2812:
2813: We present $M$ by a Heegaard splitting
2814: $M_{\psi}=h^-_g\circ{\bf I}_{\psi}\circ h^+_g$, as defined in
2815: (\ref{eq-Heegaard}) and (\ref{eq-closM}).
2816: The invariant is given as
2817: the matrix coefficient of $\ext g[\psi]$ for the basis vector
2818: ${\cal V}(h^+_g)=[a_1]\wedge[a_2]\wedge\ldots\wedge[a_g]$. If we denote by
2819: $[\psi]_{aa}$ the $g\times g$-block of $[\psi]$ acting on the Lagrangian
2820: subspace spanned by the $[a_i]$'s this number is just ${\rm det}([\psi]_{aa})$.
2821: At the same time, the Mayer-Vietoris sequence for $M_{\psi}$ shows that
2822: $[\psi]_{aa}$ is a presentation matrix for the group $H_1(M_{\psi},\Z)$
2823: so that the order of $H_1(M_{\psi},\Z)$ is, indeed, given by $\pm{\rm det}([\psi]_{aa})$.
2824:
2825: \ep
2826:
2827:
2828: \head{11. Alexander-Conway Calculus for 3-Manifolds}\lbl{S11}
2829:
2830: Let $M$ be a 3-manifold with an epimorphism
2831: $\varphi:H_1(M,\Z)\twoheadrightarrow \Z$. We recall the definition of
2832: the {\em (reduced) Alexander polynomial}
2833: $\Delta_{\varphi}(M)$, as it is given in the case of
2834: knot and link complements for example in \cite{BurZie}.
2835:
2836:
2837: Let $\widetilde M\to M$
2838: be the cyclic cover associated to $\varphi$ and view $H_1(\widetilde M)$ as
2839: a $\Z[t,t^{-1}]$-module with $t$ acting by Decktransformation. Let
2840: $E_1\subset \Z[t,t^{-1}]$ be the first elementary ideal generated by the
2841: $n\times n$ minors of an $n\times m$ presentation matrix $A(t)$ of $H_1(\widetilde M)$.
2842: Then $\Delta_{\varphi}(M)$ is the generator of the smallest principal idea containing
2843: $E_1$, or, equivalently, the g.c.d. of the $n\times n$ minors of a presentation
2844: matrix. Particularly, if $A(t)$ is a square matrix $\Delta_{\varphi}(M)=det(A(t))$ and
2845: if $n>m$, i.e., there are more rows than columns, $\Delta_{\varphi}(M)=0$.
2846:
2847:
2848: Another important invariant of a 3-manifold is its Reidemeister Torsion, which is
2849: obtained as the torsion of a chain complex over $\Q [t,t^{-1}]$ obtained from a
2850: cell decomposition of $\widetilde M$.
2851: The Alexander polynomial turns out to be
2852: almost the same as the Reidemeister Torsion of a 3-manifold.
2853: The relation described in the next theorem was first proven for homology circles by
2854: Milnor and in the general case by Turaev.
2855: \begin{theorem}[\cite{Mil61}\cite{Tur76}]\label{thm-ReidAlex}
2856: Let $M$ be a compact, oriented 3-manifolds,
2857: $\varphi: H_1(M)\to\Z$ an epimorphism as above,
2858: $r_{\varphi}(M)$ its Reidemeister Torsion, and $\Delta_{\varphi}(M)$ its Alexander polynomial.
2859: \ben
2860: \item If $\partial M\neq \emptyset$ then
2861: $\displaystyle r_{\varphi}(M)\,=\,\frac 1 {(t-1)}\Delta_{\varphi}(M)$
2862: \item If $\partial M = \emptyset$ then
2863: $\displaystyle r_{\varphi}(M)\,=\,\frac 1 {(t-1)^2}\Delta_{\varphi}(M)$
2864: \een
2865: \end{theorem}
2866:
2867:
2868: For a 3-manifold given by surgery along a framed link we will now give a
2869: procedure to compute the Alexander polynomial (and thus also Reidemeister Torsion).
2870:
2871:
2872: Let ${\cal Z}\sqcup{\cal L}\subset S^3$ be a framed link consisting of a
2873: framed link ${\cal L}$ and
2874: a curve ${\cal Z}$ which has
2875: trivial linking number of all components of ${\cal L}$, i.e.,
2876: with ${\cal L}\cdot{\cal Z}=0$.
2877: We denote by $M_{{\cal Z},{\cal L}}^{\bullet}$ the manifold obtained by cutting
2878: out a tubular neighborhood of ${\cal Z}$ and doing surgery along ${\cal L}$.
2879: Hence, $\partial M_{{\cal Z},{\cal L}}^{\bullet}=S^1\times S^1$, with canonical
2880: meridian and longitude (given by 0-framing).
2881: Also let $M_{{\cal Z},{\cal L}}$ be the closed manifold obtained
2882: by doing 0-surgery along ${\cal Z}$ so that
2883: $M_{{\cal Z},{\cal L}}=M_{{\cal Z},{\cal L}}^{\bullet}\cup D^2\times S^1$.
2884: The special component $\cal Z$ defines an epimorphism
2885: $\varphi_{\cal Z}: H_1(M^{(\bullet)})\to\Z$, for example via intersection
2886: numbers with a Seifert surface. We write
2887: $\Delta_{{\cal Z},{\cal L}}=\Delta_{\varphi_{\cal Z}}(M_{{\cal Z},{\cal L}})=
2888: \Delta_{\varphi_{\cal Z}}(M_{{\cal Z},{\cal L}}^{\bullet})$ for the associated
2889: reduced Alexander polynomial, which is the same in both cases.
2890:
2891: Consider a general Seifert surface $\Sigma^{\bullet}\subset S^3$ with
2892: $\partial \Sigma^{\bullet} ={\cal Z}$ and
2893: $\Sigma^{\bullet}\cap{\cal L}=\emptyset$. By removing a neighborhood
2894: of the surface we obtain a relative cobordism
2895: $C_{\Sigma}^{\bullet}=M_{{\cal Z},{\cal L}}^{\bullet}-\Sigma^{\bullet}
2896: \times(-\epsilon,\epsilon)$ from $\Sigma^{\bullet}$ to itself.
2897: Similarly, $C_{\Sigma}=M_{{\cal Z},{\cal L}}-\Sigma
2898: \times(-\epsilon,\epsilon)$, where $\Sigma$ is the closed capped off
2899: surface $\Sigma^{\bullet}\cup D^2$. The cobordism $C_{\Sigma}$ is obtained
2900: from $C_{\Sigma}^{\bullet}$ by gluing in a full cylinder $D^2\times[0,1]$.
2901:
2902:
2903: Denote by
2904: $\psi^{(\bullet)}_{\pm}:\Sigma^{\bullet}_{\pm}\hookrightarrow C_{\Sigma}$
2905: the inclusion maps of the bounding surfaces, and by
2906: $$
2907: A_{\pm}=H_1(\psi^{(\bullet)}_{\pm}):\,H_1(\Sigma)\to
2908: H_1(C_{\Sigma}^{(\bullet)})\to H_1^{\it free}(C_{\Sigma}^{(\bullet)})\;,
2909: $$
2910: the maps on the free part of homology, where the free part is
2911: $G^{\it free}=\frac G {Tors(G)}\,$. As
2912: $H_1(\widetilde M)\cong
2913: H_1^{\it free}(\widetilde M)\oplus\,Tors(H_1(M))\otimes\Z[t,t^{-1}]$
2914: we will consider the first elementary ideal for the free part, which differs only by
2915: a factor of $|Tors(H_1(M))|$.
2916:
2917:
2918: Suppose first that $C$ does not have interior homology. This means the $A_{\pm}$
2919: can be presented as square matrices, and $A_+-tA_-$ is a presentation matrix.
2920: Consequently
2921: $\Delta_{{\cal Z},{\cal L}}=\pm t^p det(A_+-tA_-)$. By some linear algebra
2922: \cite{FroNic92}
2923: this is the same as the Lefschetz polynomial
2924: $$
2925: det(A_+-tA_-)\;=\;\sum_{k=0}^{2g}(-t)^{2g-k}
2926: trace\Bigl((\ext k A_+)\circ \star^{-1}\circ
2927: (\ext {2g-k}A_-^*)\circ\star\Bigr)
2928: $$
2929: In \cite{FroNic92} it is also shown that the expression inside the trace is
2930: the same as ${\cal V}^{FN}(C_{\Sigma}^{\bullet})_k$ or
2931: ${\cal V}^{FN}(C_{\Sigma})_k$ depending on context. Hence, we have (multiplying by a unit $(-t)^{-g}$)
2932: that
2933: %$$
2934: \begin{eqnarray}
2935: \Delta_{{\cal Z},{\cal L}} & = &\sum_{k=0}^{2g}(-t)^{g-k} trace({\cal V}^{FN}(C_{\Sigma})_k)
2936: \label{eq-leftrace}
2937: \\
2938: &=& trace((-t)^{-H}{\cal V}^{FN}(C_{\Sigma}))
2939: \label{eq-leftrace2}
2940: \\
2941: &=& \sum_{j=1}\;[j]_{-t}\,trace({\cal V}^{(j)}(C_{\Sigma}))\;=\;
2942: \sum_{j=1}\;[j]_{-t}\,\Delta_{{\cal Z},{\cal L}}^{(j)}\;,
2943: \label{eq-leftrace3}
2944: \end{eqnarray}
2945: where $[n]_q=\frac{q^n-q^{-n}}{q-q^{-1}}$. In (\ref{eq-leftrace2}) we used the generator $H$ of
2946: the $\SL(2,\R)$-Lefschetz action. Formula (\ref{eq-leftrace3}) is a consequence of the
2947: Hard-Lefschetz decomposition from (\ref{eq-Lefsch}). We call the invariant
2948: $\Delta_{{\cal Z},{\cal L}}^{(j)}$ the {\em $j$-th Alexander Character} of the
2949: Alexander polynomial.
2950:
2951:
2952: In case $C$ does have interior rational homology the dimension of
2953: $H_1^{\it free}(C_{\Sigma}^{(\bullet)})$ is bigger than $H_1(\Sigma)$ so that
2954: $H_1(\widetilde M)$ has $\Z[t,t^{-1}]$ as a direct summand. Consequently, the Alexander
2955: polynomial vanishes. At the same time ${\cal V}^{FN}(C_{\Sigma})$ is zero since it is
2956: a non-semisimple TQFT. Hence, (\ref{eq-leftrace3}) holds for all cases.
2957:
2958:
2959: Suppose that in our presentation ${\cal Y}\subset S^3$ is the unknot. In this case we
2960: can isotop the diagram ${\cal L}\sqcup {\cal Y}\subset S^3$ into the form shown on the
2961: right side of Figure~\ref{fig-stan}. Specifically, we arrange
2962: it that the strands of one link component alternate orientations
2963: as we go from left to right. By application of the connecting annulus
2964: moves, see for example \cite{Ker99}, we can modify the link further such that the resulting
2965: tangle ${\cal T}$ in the indicated box is admissible without through pairs
2966: as described in the beginning of
2967: Section~5 or, again, \cite{Ker99}. There is a canonical Seifert surface $\Sigma_{\cal T}$
2968: associated to a diagram as in Figure~\ref{fig-stan} obtained by surgering the disc
2969: bounded by ${\cal Z}$ along the framed
2970: components of ${\cal L}$ emerging at the bottom side. By construction ${\cal T}$ is
2971: then a tangle presentation of $C_{\Sigma_{\cal T}}$.
2972:
2973:
2974: \begin{figure}[ht]
2975: \begin{center}
2976: \leavevmode
2977: \epsfbox{stan.eps}
2978: \end{center}
2979: \caption{Standard Presentation }\lbl{fig-stan}
2980: \end{figure}
2981:
2982: For the evaluation of this diagram it
2983: is convenient to introduce an extension of ${\cal N}$ over $\Z[t,t^{-1}]$,
2984: given by $\Z[\gamma^{\pm 1}]\ltimes {\cal N}$. The extra generator $\gamma$ is
2985: group like with $S(\gamma)=\gamma^{-1}$ and it acts on ${\cal N}$ by
2986: $\gamma x\gamma^{-1}=t^Hx=t^{deg(x)}x$ for $x\in{\cal N}$ and $deg(x)$ the degree
2987: for homogenous elements.
2988:
2989: In order to evaluate the diagram we apply the Hennings substitutions for crossing
2990: (\ref{fig-Dec-cross}) and rules (\ref{eq-elem-mv}) through (\ref{pic-Rclos})
2991: to the ${\cal T}$ part to obtain a combination of ${\cal N}$-decorated
2992: arcs as in (\ref{pic-Rbot}) and (\ref{pic-Rtop}). Furthermore, we
2993: remove the circle ${\cal Y}$ at the expense of introducing a $\gamma$-decoration
2994: on each strand. The Hennings procedure is continued with the extended algebra over
2995: $\Z[t,t^{-1}]$. It is easy to see that the elements that have to be evaluated against
2996: the integral all lie in $\Z[t,t^{-1}]\otimes{\cal N}$ and that $\mu$ is cyclic also
2997: with respect to $\gamma$. Hence, the evaluation is well defined.
2998:
2999: \begin{lemma}
3000: The evaluation procedure for a diagram as in Figure~\ref{fig-stan} yields the
3001: Alexander polynomial.
3002: \end{lemma}
3003:
3004: {\em Proof:} The standard evaluation of ${\cal T}$ yields a sum of diagrams with top and bottom
3005: arcs, where the $j$-th bottom arc is decorated by $b_j$ and the $j$-th top arc by $c_j$
3006: as in (\ref{pic-Rbot}) and (\ref{pic-Rtop}). Hence, ${\cal V}_{\cal N}(C_{\Sigma})$ is the sum
3007: over all diagrams of linear maps
3008: $\bigotimes_j^g (b_j\otimes \mu(S(\_)c_j))$.
3009: The extended evaluation yields closed curves, each of
3010: which is decorated with four elements $b_j$, $c_j$, $\gamma$, and $\gamma^{-1}$. Using the
3011: antipodal sliding rule from (\ref{pic-Rclos}) we collect them at one side of a circle so that the
3012: evaluation becomes
3013: $$
3014: \mu(S^{-1}(b_j)\gamma c_j\gamma^{-1})\;=\;(-1)^{deg(b_j)}t^{deg(c_j)}\mu(S(b_j)c_j)
3015: \;=\; (-t)^{-deg(b_j)}trace(b_j\otimes \mu(S(\_)c_j))\;.
3016: $$
3017: Note here, that $S^2(b_j)=(-1)^{deg(b_j)}$ and that the evaluation is non zero only if
3018: $deg(c_j)+deg(b_j)=0$. The sum (over all decorations) of the products (over $j$)
3019: of these individual traces is thus just the trace of $(-t)^{-H}{\cal V}_{\cal N}(C_{\Sigma})$.
3020: Since this is (up to sign) identical with $(-t)^{-H}{\cal V}^{FN}(C_{\Sigma})$ it follows
3021: from (\ref{eq-leftrace2}) that the evaluation gives the Alexander polynomial.
3022:
3023: \ep
3024:
3025:
3026: The evaluation of a standard diagram can be described also more explicitly without the use
3027: of the $\Z[\gamma]$ extension. Let ${\cal T}^{\#}: 2g\to 0$ be the diagram consisting of
3028: the tangle ${\cal T}: g\to g$ and the lower arcs. That is,
3029: ${\cal T}=(1^g\otimes {\cal T}^{\#})\circ (X_g\otimes 1^g)$ and
3030: ${\cal T}^{\#}= (X_g^{\dagger})\circ(1^g \otimes {\cal T})$, where $X_g^{\dagger}$ is the
3031: upside down reflection of $X_g\,$. We define $A^{\gamma}\in
3032: {\cal N}_0^{\otimes 2}\otimes\Z[t,t^{-1}]$ as
3033: \beq\label{eq-defAg}
3034: A^{\gamma}\;=\;({\gamma}\otimes 1)A({\gamma^{-1}}\otimes 1)\;=\;
3035: \frac 1 i \Bigl( \rho\otimes 1\,+\,1\otimes\rho\,-\,t^{-1}\tb\otimes\th\,+\,t
3036: \th\otimes\tb\Bigr)\;\;\;.
3037: \eeq
3038: Moreover, we define
3039: $A_{\{g\}}^{\gamma}\in{\cal N}_0^{\otimes 2g}\otimes\Z[t,t^{-1}]$ from $A^{\gamma}$
3040: as $A_{\{g\}}$ in (\ref{eq-Xn}) is defined from $A$ in (\ref{eq-defAtens}) and (\ref{eq-X1}), or,
3041: equivalently, by
3042: $$
3043: A_g^{\gamma}\;=\;({\gamma}^{\otimes g}\otimes 1^{\otimes g})\circ A_{\{g\}}\circ
3044: ({(\gamma^{-1})}^{\otimes g}\otimes 1^{\otimes g})\;.
3045: $$
3046: This tensor is assigned to the upper arcs and the $\gamma$ elements in the standard diagram.
3047: Hence, by the extended Hennings evaluation procedure the Alexander polynomial is given by the
3048: composition
3049: $$
3050: \Delta_{{\cal Z},{\cal L}}\;=\;{\cal V}^{FN}({\cal T}^{\#}) (A^{\gamma}_g)\;,
3051: $$
3052: where we think of ${\cal V}^{FN}({\cal T}^{\#}):{\cal N}_0^{\otimes 2g}\to \Cc$ as being
3053: naturally extended to a $\Z[t,t^{-1}]$-map from ${\cal N}_0^{\otimes 2g}\otimes\Z[t,t^{-1}]
3054: \,\to\, \Cc[t,t^{-1}]\,$.
3055:
3056:
3057: For further evaluation we use Theorem~\ref{thm-solve} to write
3058: ${\cal V}^{FN}({\cal T}^{\#})=\sum_{\nu}{\cal V}^{FN}(E_{\nu})$
3059: as a combination of elementary tangles
3060: $E_{\nu}=(M_{k_1}\otimes\ldots M_{k_r}\otimes \epsilon^{\otimes N})\circ B\,$ so that
3061: the Alexander polynomial is the sum of polynomials $E_{\nu}(A^{\gamma}_g)$.
3062: For the computation of these elementary polynomials it is convenient to use
3063: the following
3064: graphical notation. As shown in (\ref{eq-symhopfeva}) we indicate the morphism $M_k$
3065: by a tree with $k$ incoming branches. The morphism $X_1$ is drawn as an arc and
3066: $X_g$ as $g$ concentric arcs.
3067: \beq\label{eq-symhopfeva}
3068: \epsfbox{symhopfeval.eps}
3069: \eeq
3070: For $E=(M_1^{\otimes 3}\otimes M_2 \otimes M_4\otimes \epsilon)\circ B$ we obtain
3071: the composite shown on the right of (\ref{eq-symhopfeva}). Using relations
3072: $(\mu\otimes 1)A^{\gamma}=( 1\otimes \mu)A^{\gamma}= 1$,
3073: $(\varepsilon\otimes 1)A^{\gamma}=( 1\otimes\varepsilon)A^{\gamma}= \frac 1 i\rho$,
3074: and $\mu(x\frac 1i \rho)=\varepsilon(x)$ we find the graphical relations
3075: depicted in (\ref{eq-elimeval}).
3076: \beq\label{eq-elimeval}
3077: \epsfbox{elimeval.eps}
3078: \eeq
3079: Now, to each of the arcs
3080: the tensor $A^{\gamma}$ is associated containing the four terms $\rho\otimes 1$,
3081: $1\otimes \rho$, $\tb\otimes\th$, and $\th\otimes\tb$ with coefficients of the form
3082: $\pm it^m$. We represent the elementary
3083: polynomial thus as a sum over all combinations of these terms, i.e., $4^g$ terms
3084: for $A^{\gamma}_{\{g\}}$. We indicate a combination in a diagram by drawing a line
3085: with a down arrow for $\tb$, a line with an up arrow for $\th$, a line with arrows for
3086: $\rho$ and a dashed line for $1$. Hence, (\ref{eq-defAg}) becomes the first line in
3087: (\ref{eq-termsarr}).
3088: \beq\label{eq-termsarr}
3089: \epsfbox{termsarr.eps}
3090: \eeq
3091: The tensors associated to the $M_k$ are non zero only in two cases. Namely, if one
3092: element is $\th$, another $\tb$ and all other $1$, or if one element is $\rho$
3093: and all others 1. In diagrams we obtain the evaluation rules as depicted. All
3094: other configurations are evaluated to zero.
3095:
3096: For an elementary diagram let $N_x(=g)$ be the number of arcs at the top,
3097: $N_0$ the number of $\varepsilon$'s, and $N_k$ the number of $M_k$'s at
3098: the bottom of the diagram for $k\geq 1$. Let us also call an elementary
3099: diagram {\em reduced} if $N_0=N_1=0$. We can now give the recipe for evaluating
3100: elementary diagrams:
3101:
3102: \begin{propos}\label{prop-HopfAlexeval}
3103: \
3104:
3105: \ben
3106: \item We have the relations \ \ \ \
3107: $\displaystyle
3108: 2N_x\,=\, N_0\,+\,\sum_{k\geq 1} kN_k\;,$ \ \ and \ \
3109: $\displaystyle N_x\;=\;\sum_{k\geq 1} N_k\;.$
3110: \item Every elementary diagram is zero or equivalent to a reduced one by application
3111: of the moves in (\ref{eq-elimeval}).
3112: \item A reduced diagram is non zero only of $N_j=0$ for $j\geq 3$. That is, if
3113: the diagram is of the form $D=M_2^{\otimes g}\circ B\circ X_g$.
3114: \item A contributing reduced diagram $D=P_1\sqcup \ldots\sqcup P_n$ is the union of closed
3115: paths $P_j$, and the
3116: polynomial $\Delta_D=\prod_j\Delta_{P_j}$ assigned to $D$ is the product
3117: of the polynomials assigned to the the components $P_j$.
3118: \item The polynomial associated to a connected component is
3119: $$
3120: \Delta_P\;=\;2\,-(-1)^b\,(t^p+t^{-p})\;,
3121: $$
3122: where $p$ is the algebraic intersection number of the closed
3123: path $P$ with a radial line segment $\Xi$ as in (\ref{eq-symhopfeva}), and $b$
3124: is the total number of half twists (or antipode insertions) in $B$.
3125: \een
3126: \end{propos}
3127:
3128: {\em Proof:} {\em 1.}
3129: In a diagram as in (\ref{eq-symhopfeva}) the number of strands entering from
3130: the top is $2N_x$, two for each arc, and the number of strands entering from the bottom
3131: is $N_0+\sum_{k\geq 1}kN_k$. Obviously,
3132: both numbers have to be equal. For an admissible configuration
3133: of a contributing diagram we can also call weighted edges, where the dashed ones are
3134: weighted 0, the ones with one arrow as 1, and those with double arrows as 2. The
3135: top part of the diagram shows that the total weight has to be $2N_x$ since every
3136: admissible arc has weight 2. Also every tree has weight 2 and the $\epsilon$'s have
3137: weight 0 so that the total weight must also be given by $\sum_{k\geq 1} 2N_k$.
3138:
3139: {\em 2.} This is clear since every non-reduced one allows the application of a
3140: move that reduces the number of edges.
3141:
3142: {\em 3.} If we subtract twice the second identity in {\em i)} from the first we
3143: find $0=N_0-N_1+N_3+2N_4+3N_5+\ldots\,$. In the reduced case with $N_0=N_1=0$ this
3144: implies $0=N_3=N_4=N_5=\ldots$ since these are all non negative integers.
3145:
3146: {\em 4.} Any graph where all vertices have valency 2 is the union of closed
3147: paths. Since we have a symmetric commutativity constraint we can untangle components
3148: from each other and move them apart. The evaluation of disjoint unions of diagrams
3149: is given by their products.
3150:
3151: {\em 5.} There are four configurations that contribute to $\Delta_P$ for a
3152: closed path. Two if them are given by dashed lines alternating with double arrow
3153: lines. This corresponds to paring factors $\frac 1 i\rho$ with integrals $\mu$ in
3154: two different ways each evaluated as 1. Thus these two cases contribute the 2 in the
3155: expression. The other two configurations are given by two orientations of $P$ with
3156: single arrows everywhere. For one given orientation we get from
3157: (\ref{eq-termsarr}) a factor $\frac 1 i t$ if
3158: $P$ crosses $\Xi$ left to right and a factor $\frac 1 i (-t^{-1})$ if $P$ crosses
3159: right to left. Thus the arcs yield a tensor $\pm (\frac 1 i)^g t^b(x_1\otimes\ldots\otimes x_{2g})$,
3160: where each $x_i$ is either $\th$ or $\tb$. Application of $B$ yields a tensor
3161: $\pm (\frac 1 i)^g t^b(y_1\otimes\ldots\otimes y_{g})$ where each $y_j$ is either $\th\otimes\tb$
3162: or $\tb\otimes\th$ depending on which way the path runs through the $M_2$ piece. The
3163: pairwise multiplication thus yields the tensor $\pm t^b(\frac 1i\rho)^{\otimes g}$ and
3164: evaluation against $\mu$ the factor $\pm t^b$. For the opposite orientation the tensor
3165: for the arcs is obtained by exchanging $t$ for $t^{-1}$ and multiplying a factor $(-1)^g$.
3166: The factor picked up by
3167: application of $B$ is unchanged, and in the evaluation against the $\mu$ we pick up a factor
3168: $(-1)^g$ because the orders of $\th$ and $\tb$ are exchanged canceling the one from the top.
3169: Hence the contribution for the opposite orientation is the same with $t$ and $t^{-1}$ exchanged.
3170: Thus $\Delta_P=2\pm(t^b+t^{-b})$. The sign can be determined by evaluating the polynomial
3171: at $t=1$. This is identical with the usual Hennings invariant of the 3-manifold given by
3172: surgery along a link associated to the connected diagram $P$ as follows.
3173:
3174: First choose
3175: over and under crossing for $P$ pushing it slightly outside the plane of projection into
3176: a knot $P^*$. This knot is thickened to a band $N(P^*)$, which is parallel to the
3177: plane of projection except for half twists that are
3178: introduced at the points where $B\subset P$ has antipodes
3179: inserted.
3180:
3181: Consider the link $\partial N(P^*)$ given by the edges of the band. Generically this
3182: link consists of parallel strands that double cross as in Figure~\ref{fig-MC-tgl} at
3183: simple crossings of $P^*$ and has $\Gamma$-diagram also as in Figure~\ref{fig-MC-tgl}
3184: for every half twist. We further modify this link at some generic point in the band
3185: by replacing the parallel strands
3186: by a configuration with a connecting annulus as in the $\sigma$-Move of (\ref{eq-boundmove}).
3187: We obtain a two component link ${\cal L}_P={\cal A}_P\sqcup{\cal C}_P$, where
3188: ${\cal A}_P$ is the 0-framed annulus. The other part ${\cal C}_P$ bounds the disc obtained
3189: by removing the small piece from the band where we applied the $\sigma$-Move and thus
3190: carries a natural framing. We have by construction that
3191: $\Delta_P(1)=\pm\eta(M_{{\cal L}_P})$ with $\eta$ as in (\ref{eq-defeta}). For self intersection
3192: numbers we clearly have ${\cal A}_P\cdot {\cal A}_P=0$ and
3193: ${\cal C}_P\cdot {\cal C}_P=0$. For an even number of twists in the band $N(P^*)$ we
3194: obtain also ${\cal A}_P\cdot {\cal C}_P=0$ and for an odd number of twists we have
3195: ${\cal A}_P\cdot {\cal C}_P=\pm 2$. Hence $\eta(M_{{\cal L}_P})=0$ in the first case
3196: and $\eta(M_{{\cal L}_P})=4$ in the second.
3197: \ep
3198:
3199:
3200: Note, that the form of the $\Delta_P$ implies again the symmetry $\Delta(t)=\Delta(t^{-1})$
3201: of the Alexander polynomial. In order to instill some confidence in our procedure let us
3202: recalculate the familiar formula for the left-handed trefoil in this setting. Using the Fenn Rourke
3203: move from Figure~\ref{fig-FR} we present the trefoil as an unknotted curve $\cal Z$
3204: in a surgery diagram of Borromean rings as in (\ref{eq-trefoil}).
3205: \beq\label{eq-trefoil}
3206: \epsfbox{trefoil.eps}
3207: \eeq
3208: The standard form is obtained by moving ${\cal C}_1$ to the right off $\cal Z$ and
3209: letting ${\cal C}_2$ follow at the ends. The tangle ${\cal T}^{\#}$ is then as depicted on
3210: the left of (\ref{eq-trefoiltgl}) below. Using the framing moves from Figure~\ref{fig-frame} we
3211: expand it into elementary diagrams as on the right of (\ref{eq-trefoiltgl}).
3212: \beq\label{eq-trefoiltgl}
3213: \epsfbox{trefoiltgl.eps}
3214: \eeq
3215: The translation into Hopf algebra diagrams and subsequently polynomials is indicated
3216: next in (\ref{eq-trefoilhopf}).
3217: \beq\label{eq-trefoilhopf}
3218: \epsfbox{trefoilhopf.eps}
3219: \eeq
3220: Thus the polynomial comes out to be $t+t^{-1}-1$ as it had to be.
3221: The same calculation carries through
3222: if we change the framings $f_j$ of the components ${\cal C}_j$ in (\ref{eq-trefoil}). The
3223: difference is the sign of the first summand, that is
3224: $\Delta_{\cal Z}=f_1 f_2(t+t^{-1}-2)+1$. Thus, if we flip both framings we obtain the right-handed
3225: trefoil with the same polynomial. If we flip only one framing so that $f_1=-f_2$ we obtain one
3226: of two figure-eight knots with polynomial $-t-t^{-1}+3$. Many other Alexander polynomials with
3227: multiple twists as for example $(p,q,r)$-pretzel knots can be computed quite conveniently
3228: in this fashion using Fenn Rourke moves and the nilpotency of the ribbon element $v^k=1+k\rho$.
3229: Thus, our method proves to be
3230: quite useful in the calculation of the Alexander Polynomial for knots although its primary
3231: application is the generalization to 3-manifolds.
3232:
3233:
3234: We describe next a more systematic way to unknot the special strand $\cal Z$ in a general diagram
3235: more akin the traditional skein theory. The additional
3236: relations that allow us to put any diagram ${\cal L}\sqcup {\cal Z}$ into a standard form
3237: are as follows.
3238:
3239: \begin{propos}\label{propos-Yskein}
3240: We have the following two skein relations for the special strand ${\cal Z}$
3241: \beq\label{eq-Yskein}
3242: \epsfbox{Yskein.eps}
3243: \eeq
3244: and
3245: \beq\label{eq-Ycoupon}
3246: \epsfbox{Ycoupon.eps}
3247: \eeq
3248: as well as the slide and cancellations moves analogous to (\ref{eq-1handlemove}),
3249: and a vanishing property as in (\ref{eq-pairvanish}). % Lemma~\ref{lm-1handlemoves}
3250: % (\ref{eq-1handlecanc}) (\ref{eq-1handleoppcanc}).
3251:
3252: These equivalences allow us to express the Alexander polynomial of
3253: any diagram ${\cal Y}\sqcup {\cal L}\subset S^3$ as a combination of the
3254: evaluations of diagrams in standard form.
3255: \end{propos}
3256:
3257: {\em Proof:} As before we change a self crossing of ${\cal Y}$ by sliding a 1-framed annulus
3258: ${\cal A}$ over the crossing. Note, that we do not have to keep track of the framing of
3259: ${\cal Y}$ as it is unchanged and by convention zero.
3260: Using the orientation of ${\cal Y}$ we can do this
3261: such that the linking numbers of ${\cal Y}$ and ${\cal A}$ remain zero. It is easy to see
3262: that we can bring a diagram into the standard position as in Figure~\ref{fig-stan} without
3263: ever sliding a strand over the new component ${\cal A}$. The evaluation is obtained as the
3264: weighted trace over the linear map associated by ${\cal V}_{\cal N}$ to the
3265: cobordism represented by the tangle,
3266: which contains ${\cal A}$. Inserting
3267: the relation from Figure~\ref{fig-frame} we see that this linear map, and hence
3268: the associated polynomial, is the
3269: combination of the one for which ${\cal A}$ has been removed and the one for which the framing
3270: of ${\cal A}$ has been shifted by one. In both cases the unknotting procedure can be reversed
3271: so that we obtain the original pictures with ${\cal A}$ removed or its framing shifted by one.
3272: The situation in which ${\cal A}$ is removed corresponds to the opposite crossing. In the other
3273: contribution we have a 0-framed annulus around the crossing which can be rewritten as an index-1
3274: surgery represented by a pair of coupons. This yields (\ref{eq-Yskein}).
3275:
3276: The coupon combination in (\ref{eq-Ycoupon}) can be reexpressed by a tangle as in
3277: (\ref{eq-coupontorus}), can be isotoped into the position shown in (\ref{eq-Yext}).
3278: \beq\label{eq-Yext}
3279: \epsfbox{Yext.eps}
3280: \eeq
3281: The extra tangle piece ${\cal Q}$ maps to the identity on a torus block. More precisely,
3282: ${\cal V}_{\cal N}({\cal Q}\sqcup{\cal T})=id_{{\cal N}_0}\otimes
3283: {\cal V}_{\cal N}({\cal T})$. The weighted traces thus differ by a factor
3284: $trace_{{\cal N}_0}((-t)^{-H})=-t+2-t^{-1}=-(t^{\frac 12}-t^{-\frac 12})^2$.
3285: \ep
3286:
3287: For ordinary link and knot complements there are well known skein relations
3288: that uniquely characterize the Alexander-Conway polynomial of the knot,
3289: see for example \cite{BurZie} Chapter 12.C.
3290:
3291: \begin{cor}
3292: For ordinary knot complements (that is if ${\cal L}=\emptyset$) the relations
3293: Proposition~\ref{propos-Yskein} reduce to the ordinary Alexander-Conway skein
3294: relations.
3295: \end{cor}
3296:
3297: {\em Proof:} It is clear that with Proposition~\ref{propos-Yskein} we can resolve every diagram
3298: into disjoint circles in the plane with coupons on them in exactly the same way as
3299: for the Alexander-Conway polynomial. The difference is that wherever we pick up a
3300: factor $(t^{\frac 12}-t^{-\frac 12})$ from the smoothening in the traditional calculus
3301: we obtain a factor $\frac 1 i$ and a pair of coupons in our case, but all other
3302: numbers are the same.
3303:
3304:
3305: Suppose now after resolving the crossings
3306: we have more than one circle. Since the strand $\cal Z$ has to run
3307: though all of these components we must have coupons that are paired but on different
3308: circles. By (\ref{eq-pairvanish}) of Lemma~\ref{lm-1handlemoves} it follows that such a
3309: configuration must vanish. In the Alexander-Conway calculus we also have the rule
3310: that the link invariant for the unlinked union of an unknot with a non-trivial link
3311: is zero. Hence we only need to compare the contributions that come from single circles.
3312: If in the process of applying the skein relations we carried out $N$ smoothenings of
3313: crossings the circle will carry $2N$ coupons.
3314:
3315: Next we claim that it is not possible to slide two paired coupons in adjacent position.
3316: To this end note that the coupons in the resolution of
3317: Proposition~\ref{propos-Yskein} stay all on one side of the
3318: special strand. I.e., in the depicted orientation of $\cal Z$
3319: the coupons are always on the left of $\cal Z$. Thus,
3320: if they become adjacent
3321: we would have a situation as in (\ref{eq-1handlecanc}) of Lemma~\ref{lm-1handlemoves}.
3322: This is not possible since then $\cal Z$ would have at least two components.
3323: Thus the number $2N$ of coupons will remain the same under handle slides.
3324:
3325: We next observe that a circle with edges that are labeled in pairs and subject to
3326: handle slides also occurs in the classification of compact, oriented surfaces via
3327: their triangulations as in \cite{massey} Chapter 1. It is shown there that any
3328: such configuration is under application of handle slides and cancellation
3329: moves as in (\ref{eq-1handlecanc}) equivalent to a sequence of blocks as in
3330: (\ref{eq-Ycoupon}). As before we may assume that all coupons lie on one side of
3331: the circle. In fact, as $\cal Z$ is connected we see from \cite{massey} that
3332: we can move to the configuration in standard block form without the use of
3333: cancellations.
3334:
3335: Thus, we have $\frac N 2$ 4-coupon (torus) blocks as in (\ref{eq-Ycoupon})
3336: contributing a factor of
3337: $(-(t^{\frac 12}-t^{-\frac 12})^2)^{\frac N 2}=(i)^N(t^{\frac 12}-t^{-\frac 12})^N$.
3338: Recall that in each resolution we also had a factor $\frac 1 i$ so that the
3339: total factor for the circle is just $(t^{\frac 12}-t^{-\frac 12})^N$ and $N$ is the
3340: number of smoothenings. But $(t^{\frac 12}-t^{-\frac 12})$ is precisely the factor
3341: assigned to each smoothening by the usual Alexander-Conway calculus.
3342:
3343: \ep
3344:
3345: Although we now have a systematic procedure for computing the Alexander polynomial
3346: of a 3-manifold it is often times efficient to use the skein relations leading up
3347: to it directly. We illustrate this by computing $\Delta_{{\cal C}_{k,l},{\cal Z}}$,
3348: where ${\cal C}_{k,l}$ is the component depicted in (\ref{eq-exampleknot})
3349: \beq\label{eq-exampleknot}
3350: \epsfbox{exampleknot.eps}
3351: \eeq
3352: The two middle strands are twisted with each other $k$ times generating $2k$ crossings,
3353: and we have $q$ full circles on the upper strand indicating shifts in the framing by -1.
3354: The definition for $k<0$ or $q<0$ is given by choosing the opposite twistings.
3355:
3356: \begin{lemma} The Alexander Polynomial of $M_{{\cal C}_{k,l},{\cal Z}}$ is given by
3357: the ordinary Alexander polynomial of the knot as follows:
3358: $$
3359: \Delta_{{\cal C}_{k,l},{\cal Z}}\;=\;i(k(t+t^{-1})-q)\Delta_{{\cal Z}}
3360: $$
3361: \end{lemma}
3362:
3363:
3364: {\em Proof:} We combine every twist with two circles so that we have $k$ twist
3365: configurations as in Figure~\ref{fig-FR} and $l=q-2k$ remaining circles. Applying
3366: the Fenn Rourke move to each of these we obtain a configuration in which we have a
3367: parallel instead of twisted pair of strands in the middle surrounded by $k$
3368: annuli with an empty circle on them. In addition, we have $k$ separate annuli with
3369: full circles. Denote by $\Delta_{k,l}$ the associated Alexander Polynomial. For $k>0$
3370: we choose one of the first annuli and apply the framing shift relation
3371: (\ref{fig-frame}) to the empty circle on it. In the second contribution we omit the
3372: dotted line so that we obtain the same configuration with one less annulus around
3373: the double strands. The factor $i$ in (\ref{fig-frame}) is canceled against one
3374: of the separate annuli with a full circle so that the second contribution is exactly
3375: $\Delta_{l,k-1}$. In the first contribution we have a 0-framed annulus which, by
3376: Figure~\ref{fig-modif}, can be turned into a pair of coupons. The other $k-1$ coupons
3377: can thus be slid off and canceled against $k-1$ annuli with full circles. Moreover,
3378: the remaining $l$ full circles on the upper strand can be removed since inserting a
3379: dotted line leaves two isolated coupons, which yields zero. The resulting
3380: configuration is the knot $\cal Z$ with a tangle piece ${\cal Q}$ as in
3381: (\ref{eq-Yext}), contributing an
3382: extra factor $-(t^{\frac 12}-t^{-\frac 12})^2$, and an
3383: extra annulus with full circle with a factor $-i$. We thus obtain the recursion
3384: relation $P_{k,l}\,=\,i(t^{\frac 12}-t^{-\frac 12})^2\,+\,P_{k-1,l}$ so that
3385: $P_{k,l}\,=\,ik(t^{\frac 12}-t^{-\frac 12})^2\Delta_{\cal Z}\,+\,P_{0,l}$.
3386: But the configuration
3387: for $k=0$ is the separate union of $\cal Z$ and an annulus with $l$ full circles.
3388: The latter yields a factor $-il$ so that
3389: $P_{k,l}\,=\,i(k(t^{\frac 12}-t^{-\frac 12})^2-l)\Delta_{\cal Z}$, which computes
3390: to the desired formula.
3391:
3392: \ep
3393:
3394:
3395:
3396: \head{12. Lefschetz compatible Hopf algebra structures on $H^*(J(\Sigma))$}\lbl{S12}
3397:
3398:
3399: It is easy to see that the natural ring structure on the cohomology
3400: $H^*(J(\Sigma))\cong\ext * H_1(\Sigma)$ is not compatible with the $\SL(2,\R)$
3401: Lefschetz action as described in Section~10. For example $E(x\wedge y)=x\wedge y\wedge \omega$
3402: but $(Ex)\wedge y +x\wedge (Ey) =2x\wedge y\wedge \omega$. The isomorphism with
3403: ${\cal N}_0^{\otimes g}$ however induces another multiplication structure compatible
3404: with the $\SL(2,\R)$ action. In this section we will describe it explicitly.
3405:
3406: The $\Z/2$-graded Hopf algebra structure on ${\cal N}_0$ given in
3407: Lemma~\ref{lm-super-Hopf} extends to a $\Z/2$-graded Hopf algebra
3408: structure ${\cal H}_{\cal N}$ on ${\cal N}_0^{\otimes g}$ with
3409: $$
3410: (x_1\otimes \ldots\otimes x_g)(y_1\otimes \ldots\otimes y_g)
3411: \;\;=\;\;(-1)^{\sum_{i<j} d(x_j)d(y_i)}\,x_1y_1\otimes \ldots\otimes x_gy_g\;.
3412: $$
3413: The formula for ${\bf \Delta}$ is the dual analog.
3414: The precise form of ${\cal H}_{\cal N}$ is given as follows:
3415:
3416:
3417: \begin{lemma}\lbl{lm-StrucN} For a choice of basis of $\R^g$ there
3418: is a natural isomorphism of Hopf algebras
3419: $$
3420: \varrho\;:\;\;\ext *(\E\otimes \R^g)\;\;\isto\;\;{\cal N}_0^{\otimes g}\;\;
3421: $$
3422:
3423: so that \ \ \ ${\rm Aut}({\cal N}_0^{\otimes g},
3424: {\cal H}_{\cal N})\;\cong\;{\rm GL}(\E\otimes \R^g)$.
3425: \end{lemma}
3426:
3427: {\em Proof:} Let $\{e_j\}$ be a basis of $\R^g$. The generating set of primitive vectors
3428: of $\ext *(\E\otimes \R^g)$ is given by $\E\otimes \R^g$. On this subspace we set
3429: $\varrho(w\otimes e_j)=1\otimes\ldots 1\otimes w \otimes 1\ldots\otimes 1$, with $w$ in
3430: $j$-th position. We easily see that the vectors in $\varrho(\E\otimes \R^g)$ form
3431: again a generating set of anticommuting, primitive vectors of ${\cal N}_0^{\otimes g}$
3432: so that $\varrho$ extends to a Hopf algebra epimorphism. Equality of dimensions thus implies
3433: that $\varrho$ is an isomorphism.\ep
3434:
3435: The canonical $\SL(2,\R)$-action on ${\cal N}_0^{\otimes g}$ is still compatible
3436: with ${\cal H}_{\cal N}$ since it preserves the degrees and factors. Under the isomorphism
3437: in Lemma~\ref{lm-StrucN} it is readily identified as the $\SL(2,\R)$-action on the $\E$-factor.
3438: The remaining action on the $\R^g$-part can be understood geometrically. Specifically,
3439: $\Sp(2g,\Z)$ acts on ${\cal N}_0^{\otimes g}$ since the ${\cal V}$-representation of the
3440: mapping class group factors through a the symplectic group with representation
3441: ${\cal V}^{Sp}:\,\Sp(2g,\Z)\to {\rm GL}({\cal N}_0^{\otimes g})\,:\,[\psi]\mapsto
3442: {\cal V}^{Sp}([\psi]):={\cal V}({\bf I}_{\psi})$.
3443: For a given decomposition into Lagrangian subspaces we denote the standard inclusion
3444: \beq\lbl{eq-defkappa}
3445: \kappa\,:\;
3446: \SL(g,\Z)\,\into {\rm }GL(g,\Z)\,\into \,\Sp(2g,\Z)\;:
3447: \;\;A\,\mapsto\, \kappa(A):=\,A\oplus (A^{-1})^T\;.
3448: \eeq
3449:
3450:
3451: \begin{lemma}\lbl{lm-slgz-invar}
3452: The action of $\SL(g,\Z)$ on ${\cal N}_0^{\otimes g}$ induced by ${\cal V}^{Sp}\circ\kappa$
3453: is compatible with ${\cal H}_{\cal N}$, and under the isomorphism $\varrho$ from
3454: Lemma~\ref{lm-StrucN} it is identical with the $\SL(g,\Z)$-action on $\R^g$ for the given basis.
3455: In particular, it commutes with the $\SL(2,\R)$-action so that we have the following
3456: natural inclusion of the Howe pairs:
3457: $$
3458: \SL(2,\R)\times \SL(g,\Z) \;\;\subset\;\;
3459: {\rm GL}(\E\otimes \R^g)={\rm Aut}({\cal N}_0^{\otimes g},{\cal H}_{\cal N} )\;\;.
3460: $$
3461: \end{lemma}
3462:
3463: {\em Proof:} Consider the elements
3464: $P_j:=S_j\circ D_j^{-1}\circ S_j^{-1}$ and
3465: $Q_j:=S_{j+1}\circ D_j^{-1}\circ S_{j+1}^{-1}$ of $\Gamma_{g,1}$.
3466: From (\ref{eq-D-hom}) and (\ref{eq-S-hom}) we compute the homological
3467: action as $[R_j]=\kappa(I_g+E_{j+1, j})$ and $[Q_j]=\kappa(I_g+E_{j, j+1})$,
3468: with conventions again as in \cite{GooWal98}.
3469: The matrices $I_g+E_{j+1, j}$ and $I_g+E_{j, j+1}$ generate $\SL(g,\Z)$, and hence
3470: $[P_j]$ and $[Q_j]$ generate $\kappa(\SL(g,\Z))\subset \Sp(2g,\Z)\,$.
3471: The actions of ${\cal V}({\bf I}_{P_j})$ and
3472: ${\cal V}({\bf I}_{Q_j})$ on ${\cal N}_0^{\otimes g}$ are given by
3473: placing the maps $\Pp:=(\Ss\otimes 1)\D^{-1}(\Ss^{-1}\otimes 1)$ and
3474: $\Q:=(1 \otimes\Ss )\D^{-1}(1 \otimes\Ss^{-1})$ in the $j$-th and $j+1$-st tensor
3475: positions. In order to show that the actions of $P_j$ and $Q_j$ on ${\cal N}_0^{\otimes g}$
3476: yield Hopf algebra automorphisms it thus suffices to prove this for the maps
3477: $\Pp$ and $\Q$ in the case $g=2$. From the tangle presentations we find
3478: identities ${\bf I}_{Q_1}= ({\bf M}\otimes 1)\circ (1\otimes {\bf \Delta})$
3479: and ${\bf I}_{P_1}= (1\otimes {\bf M})\circ({\bf \Delta}\otimes 1)$.
3480: It follows that $\Pp(x\otimes y)=\Delta_0(x)(1\otimes y)$ and
3481: $\Q(x\otimes y)=(x\otimes 1)\Delta_0(y)$. The fact that these are Hopf automorphisms
3482: on ${\cal N}_0\otimes {\cal N}_0$ can be verified by direct computations. For the
3483: multiplication this amounts to verification of equations such as
3484: $\Delta(w)1\otimes v=-1\otimes v\Delta(w), \forall v,w\in\E\,$, and for the
3485: comultiplication we use the
3486: fact that ${\cal N}_0$ is self dual.
3487:
3488: From the above identities we have that
3489: ${\cal V}({\bf I}_{Q_1})=(M_0\otimes 1)\circ (1\otimes \Delta_0)$ so that
3490: ${\cal V}({\bf I}_{Q_j})$ is given on a monomial by taking the coproduct of the element in
3491: $(j+1)$-st position, multiplying the first factor of that to the element in $j$-th position
3492: and placing the second factor into $(j+1)$-st position. We readily infer for every $w\in\E$ that
3493: ${\cal V}({\bf I}_{Q_j})(\varrho(w\otimes e_k))\,=\,
3494: \varrho(w\otimes e_k+\delta_{j+1,k}w\otimes e_j)\,=\,\varrho(w\otimes (I_g+E_{j+1,j})e_k)$.
3495: The analogous relation holds for $[P_j]$ so that
3496: $$
3497: {\cal V}^{Sp}(\kappa(A))(w\otimes x)\quad=\quad w\otimes (Ax)\;\;\;
3498: \forall A\in \SL(g,\Z).
3499: $$
3500: This is precisely the claim made in Lemma~\ref{lm-slgz-invar}.
3501: \ep
3502:
3503:
3504: The structure ${\cal H}_{\cal N}$ is mapped by the isomorphism $\xi_g$ from (\ref{eq-defxi})
3505: to a $\Z/2$-graded Hopf algebra structure ${\cal H}_{\Lambda}$ on $H^*(J(\Sigma_g))$.
3506: A-priori the isomorphism $\xi_g$ and thus also ${\cal H}_{\Lambda}$ depend on the choice of
3507: a basis of $H_1(\Sigma_g)$. However, the $\SL(g,\Z)$-invariance determined in Lemma~\ref{lm-slgz-invar}
3508: translates to the $\SL(g,\Z)$-invariance of ${\cal H}_{\Lambda}$, where
3509: $\kappa(\SL(g,\Z))\subset \Sp(2g,\Z)$ acts in the canonical way on $H^*(J(\Sigma_g))$.
3510: Hence, ${\cal H}_{\Lambda}$ only depends on the oriented subspaces
3511: $\Lambda=\lz [a_1],\ldots,[a_g]\rz\subset H_1(\Sigma_g,\Z)$
3512: and $\Lambda^*=\lz [b_1],\ldots,[b_g]\rz\subset H_1(\Sigma_g,\Z)$,
3513: but not the specific choice of basis within them.
3514: The orientations can be given by volume forms $\omega_{\Lambda}:=[a_1]\wedge\ldots\wedge[a_g]$
3515: and $\omega_{\Lambda^*}:=[b_1]\wedge\ldots\wedge[b_g]$.
3516: The primitive elements $\varrho(\th\otimes e_j)$ and $\varrho(\tb\otimes e_j)$
3517: of ${\cal N}_g^{\otimes g}$ are mapped by $\xi_g$ to
3518: \beq\lbl{eq-primit}
3519: \pm[a_j]\wedge\omega_{\Lambda^*}\in\ext {g+1}H_1(\Sigma_g)\qquad\mbox{and}\qquad
3520: \pm i^*_{z_j}(\omega_{\Lambda^*})\in\ext {g-1}H_1(\Sigma_g)
3521: \eeq
3522: respectively, where $[a_j]\in H_1(\Sigma_g)$ and $z_j\in H^1(\Sigma_g)$, with $z_j([b_j])=1$ and
3523: $z_j([x])=0$ on all other basis vectors. We also have $\xi_g(1)=\omega_{\Lambda^*}$ and
3524: $\xi_g(\rho^{\otimes g})=\omega_{\Lambda}$.
3525:
3526: This completes the proof of Theorem~\ref{thm-structure}. \ep
3527: \medskip
3528:
3529: In the remainder of this section we give a more explicit description of the
3530: structure ${\cal H}_{\Lambda}$ on $H^*(J(\Sigma_g))$, and relate it to an involution,
3531: $\tau$,
3532: on $H^*(J(\Sigma_g))$, which acts as identity on the $\Lambda$-factor and, modulo signs,
3533: as a Hodge star on the opposite $\Lambda^*$-factor.
3534:
3535: The product $\diamond$ on $(H^*(J(\Sigma_g)),{\cal H}_{\Lambda})$
3536: is given on a genus one block, $\ext * \lz[a],[b]\rz$, as follows:
3537: \beq\lbl{eq-diamond}
3538: \begin{array}{c}
3539: \mbox{Table for}\\
3540: \\
3541: \quad u\,\diamond\, t\;:=\;\phi(\phi^{-1}(u)\phi^{-1}(t))\\
3542: \end{array}
3543: \quad\quad
3544: \begin{tabular}{ | c || c|c|c|c|}
3545: \hline
3546: ${}_u \,\backslash \,{}^t$\ \
3547: &\ \ 1\ \ &\ \ $[a]$\ \ &\ \ $[b]$\ \ &\ \ $[a]\wedge [b]$ \ \ \\
3548: \hline\hline
3549: 1 & 0 & 0 & 1 & $[a]$ \\
3550: \hline
3551: $[a]$ & 0 & 0 & $a$ & 0 \\
3552: \hline
3553: $[b]$ & 1 & $[a]$ & $[b]$ & $[a]\wedge [b]$ \\
3554: \hline
3555: $[a]\wedge [b]$ & $-[a]$ & 0 & $[a]\wedge [b]$ & 0\\
3556: \hline
3557: \end{tabular}
3558: \qquad\qquad
3559: \eeq
3560: It extends to $\ext * H_1(\Sigma_g)$ via the formula
3561: \beq\lbl{eq-totprod}
3562: (u_1\wedge\ldots\wedge u_g)\diamond (t_1\wedge\ldots\wedge t_g)
3563: \quad=\quad (-1)^{\sum_{i<j}d_il_j}\,\,(u_1\diamond t_1)
3564: \wedge \ldots\wedge(u_g\diamond t_g)\;,\quad
3565: \eeq
3566: where $u_i, t_i\in\ext *\lz [a_i],[b_i]\rz$, $d_i=1-deg(u_i)$ and $l_j=1-deg(t_j)$.
3567: In particular, we have $\,u\diamond t=(-1)^{d l}\,t\diamond u\,$,
3568: with $d=\sum_id_i=g-deg(u)$ and $l=\sum_il_i=g-deg(t)$, which reflects the
3569: $\Z/2$-commutativity of $H^*(J(\Sigma_g))$.
3570:
3571:
3572: The product structure and another proof of Lemma~\ref{lm-slgz-invar} can be also
3573: found from an involution, $\tau$, defined as follows:
3574:
3575: Every cohomology
3576: class $x\in H^*(J(\Sigma_g))$ is uniquely written as $x=\alpha\wedge\beta$, where
3577: $\alpha\in\ext *\Lambda$ and $\beta\in \ext *\Lambda^*$. For $x$ in this form the map $\tau$
3578: is uniquely determined by the relations
3579: \beq\lbl{eq-deftau}
3580: \tau(\alpha\wedge\beta)=\alpha\wedge\tau(\beta)\quad\mbox{and}\quad
3581: \tau(b_1^{\epsilon_1}\wedge\ldots\wedge b_g^{\epsilon_g})=
3582: b_1^{1-\epsilon_1}\wedge\ldots\wedge b_g^{1-\epsilon_g}\;.
3583: \eeq
3584: From the formulae in (\ref{eq-diamond}) and (\ref{eq-totprod}) we find that $\tau^2=1\,$, and
3585: \beq\lbl{eq-tauprop}
3586: \tau(u\diamond t)\;\;=\;\;\tau(t)\wedge\tau(u)\;,\quad\quad\;
3587: \eeq
3588: and that $\tau$ maps $\ext *\Lambda$ as well as $\ext *\Lambda^*$ to itself. It is clear from
3589: (\ref{eq-deftau}) and (\ref{eq-tauprop}) that $\SL(g,\Z)$-variance
3590: of $\diamond$ on $H^*(J(\Sigma_g))$ is equivalent to $\SL(g,\Z)$-variance
3591: of $\diamond$ on $\ext * \Lambda^*$. Now, for any $A\in \SL(\Lambda^*)$ the following
3592: identity holds:
3593: \beq\lbl{eq-tauAid}
3594: \tau\circ(\ext * A)\circ \tau\;\;=\;\;\ext* \iota(A)\;,
3595: \eeq
3596: where $\iota$ is the involution on $\SL(\Lambda^*)$ defined by
3597: $$
3598: \iota(A)\;:=\;D\circ(A^{-1})^T\circ D\,,\qquad\mbox{with}\;\; D[b_j]=(-1)^j[b_j]\;.
3599: $$
3600: This can be proven either by considering again generators of $\SL(\Lambda^*)$,
3601: or by applying the generalized Leibniz formula for the expansion of the determinant
3602: of a $g\times g$-matrix into products of determinants of $k\times k$ and
3603: $(g-k)\times(g-k)$-submatrices. See also Lemma~5.2 in \cite{FroNic94}. (\ref{eq-tauprop})
3604: together with (\ref{eq-tauAid}) implies now that $\diamond$ depends only on the
3605: decomposition $H_1(\Sigma_g,\Z)=\Lambda\oplus\Lambda^*$.
3606:
3607: In summary, we have the following isomorphism of $\Z/2$-graded Hopf algebras:
3608: $$
3609: \tau':=\ext *D\circ\tau\,:\,(H^*(J(\Sigma_g)),{\cal H}_{\Lambda})\;\isto\;
3610: (H^*(J(\Sigma_g)),{\cal H}_{ext})\;\;,
3611: $$
3612: The Howe pair
3613: $\SL(2,\R)\times \SL(g,\R)\subset GL(H_1(\Sigma_g))=
3614: {\rm Aut}(H^*(J(\Sigma_g)),{\cal H}_{ext})$, with
3615: $H_1(\Sigma_g)=\E\otimes \Lambda$, is conjugated by $\tau'$ to the pair
3616: $\SL(2,\R)_{Lefsch.}\times \kappa(\SL(g,\R))\subset
3617: {\rm Aut}(H^*(J(\Sigma_g)),{\cal H}_{\Lambda})$.
3618: \medskip
3619:
3620: \newpage
3621:
3622:
3623: \head{13. More Examples of Homological TQFT's and Open Questions}\lbl{S13}
3624:
3625:
3626: %%%%%%%%%%%%%%%%% BEGIN CUT %%%%%%%%%%%%%%%%%%%%%%%%%%%
3627:
3628: \emptystuff{
3629: \paragraph{A. Homology TQFT's over $\Z/r$ and cut numbers.}\
3630:
3631:
3632: Although the TQFT's of Reshetikhin and Turaev are semisimple and
3633: non trivial on the Torelli groups they contain homological TQFT's in an
3634: indirect manner. Specifically, if we consider the TQFT for $U_q({\mathfrak s}{\mathfrak l}_2)$
3635: for $q$ a primitive $r$-th root of unity and $r$ is an odd prime
3636: Gilmer \cite{Gil01} shows that it can be defined essentially as a
3637: theory ${\cal W}^*_r$
3638: over the ring over cyclotomic integers $\Z[q]$. This generalizes
3639: the integrality results in \cite{MasRob97} and \cite{Mur94} for the
3640: invariants of closed manifolds.
3641:
3642: Of particular interest are expansions in $(q-1)$ which on the level of
3643: invariants of closed manifolds lead to the Ohtsuki invariants in $\Z/r$
3644: \cite{Oh96} which in 0-th order coincides with the invariant from (\ref{eq-defeta})
3645: and in next order is identical to the Casson invariant \cite{Mur94}.
3646:
3647: A candidate for a useful homological TQFT is the lowest order of the TQFT
3648: over the cyclotomic integers. It is given by the extending the
3649: trace function $\Z[q]\to\Z/r$ to a transformation
3650: ${\cal W}^*_r\,\to\,{\cal W}_r$, where ${\cal W}_r$ is the respective TQFT
3651: defined over the finite field $\Z/r$.
3652:
3653:
3654: In \cite{Kerr=5} we consider the first non-trivial prime $r=5$ and
3655: find an explicit basis for ${\cal W}^*_5$ and hence a description of
3656: ${\cal W}_5$. We find that the Torelli group is not entirely in the
3657: kernel of ${\cal W}_5$ but factors through the Johnson homomorphism.
3658: It does, however, contain a sub-TQFT $ {\cal U}_5\subset {\cal W}_5$
3659: which is homological, meaning does not see the Torelli group,
3660: such that also the quotient TQFT
3661: ${\cal Q}_5={\cal W}_5/ {\cal U}_5$ is homological.
3662:
3663: Explicit computations strongly suggest that these homological TQFT's are
3664: also of the general form (\ref{eq-defpolyTQFT}). More precisely,
3665: we define the following (linear)
3666: polynomials in $\Z[x_0,x_1,\ldots]$.
3667: $$
3668: Q\,=\,x_0-x_8+x_{10}-x_{18}+x_{20}-x_{28}+\ldots
3669: $$
3670: $$\mbox{and}
3671: \qquad\qquad
3672: U\,=\,x_3-x_5+x_{13}-x_{15}+x_{23}-x_{25}+\ldots \;\;.
3673: \qquad\qquad \qquad
3674: $$
3675: Explicit verification for low genera and comparison of dimensions leads
3676: us to the following.
3677: \begin{conjecture}[see\cite{Kerr=5}]\label{conj-5}
3678: We have isomorphisms of TQFT's defined
3679: over $\Z/5$:
3680: $$
3681: {\cal Q}_5\,\,\cong\,\,{\cal V}^{(Q)}
3682: \qquad\qquad \mbox{and}\qquad \qquad
3683: {\cal U}_5\,\,\cong\,\,{\cal V}^{(U)}
3684: $$
3685: \end{conjecture}
3686: Note that the polynomials $Q$ and $U$ also have negative integers so that
3687: we need to make sense of subtracting vector spaces or TQFT's. To this end note that
3688: the $\Sp(2g,\Z)$-representations $W_{g,j}$ are irreducible over $\Z$ but
3689: become reducible if we take them, for a given basis, over $\Z/5$. For instance the
3690: $\Z/5$-reduction of $W_{6,3}$ contains a subrepresentation isomorphic to
3691: the $\Z/5$-reduction of $W_{6,5}$. This explains the meaning of the
3692: difference $x_3-x_5$ in the expression for $U$.
3693:
3694: A fascinating topological application is the determination of so called
3695: {\em cut numbers}, which is investigated in joint work with Gilmer \cite{GilKer}.
3696: Let us denote by $cut(M)$ the maximal number rank $n$ of a (non-abelian) free group
3697: $F_n=\Z*\ldots*\Z$ such that there is an epimorphism $\phi:\pi_1(M)\to F_n$.
3698: This is also the maximal number of surfaces that can be removed from $M$ without
3699: disconnecting the manifold. For a given epimorphism $\varphi:H_1(M)\to \Z$ we
3700: also define the relative cut number $cut(M,\varphi)$ as the maximal $n$ such that
3701: there is an epimorhpism $\phi:\pi_1(M)\to F_n$ which factors through $\varphi$,
3702: meaning there is a map $\tau:F_m\to\Z$ such that $\varphi=\tau\circ\phi$. This
3703: counts non separating surfaces with the constraint that one represents $\varphi$, \cite{GilKer}.
3704:
3705: Clearly we have $cut(M)\geq cut(M,\varphi)\geq 1$ if defined. Aside from these
3706: constraints the
3707: absolute and relative cut number are independent. For example
3708: let $M=S^1\times \Sigma_g$ with canonical projection $\pi:M\to S^1$. Then
3709: $cut(M,\pi_*)=1$ but $cut(M)\geq g$.
3710:
3711:
3712: As we remarked in the beginning of Section~11 an additional non-separating
3713: surface in the cut cobordism $C$ used to define the Alexander polynomial
3714: implies ${\cal V}^{FN}(C)=0$ by non-semisimplicity. Thus for a 3-manifold $M$
3715: with epimorphism $\varphi:H_1(M)\to\Z$ as before we have the implication:
3716: \beq\label{eq-cutrel}
3717: \mbox{If}\qquad\quad cut(M,\varphi) >1\qquad\quad{\rm then}
3718: \qquad\quad \Delta_{M,\varphi}\,=\,0.
3719: \eeq
3720: In \cite{GilKer} we manage to obtain a criterium on the bare cut number
3721: {\em independent} of a choice of $\varphi$:
3722: \beq\label{eq-cut}
3723: \mbox{If}\qquad\quad cut(M) >1\qquad\quad{\rm then}
3724: \qquad\quad trace({\cal W}_5(C))\,\equiv\,0\;{\rm mod}\,5.
3725: \eeq
3726: Note that the expression on the right only depends on the homological
3727: functors $\cal Q$ and $\cal U$. It turns out that under the assumption of
3728: Conjecture~\ref{conj-5} the respective traces are easily computed form the Alexander polynomial.
3729: In fact, under this assumption,
3730: the trace expression
3731: in (\ref{eq-cut}), which is just the sum of the traces for $\cal Q$ and $\cal U$,
3732: comes out to be equal to
3733: the unique coefficient $T_{M,\varphi}$
3734: of the Alexander polynomial when written as follows.
3735: $$
3736: \Delta_{M,\varphi}(q)\;=\;T_{M,\varphi}\;+\;B(q+q^{-1})\;\;\;\in\;\;\;\Z[q]\qquad\mbox{with}\quad
3737: T_{M,\varphi}, B\in \Z\;\;.
3738: $$
3739: The contrapositive of (\ref{eq-cut}) under the assumption of the conjecture thus becomes
3740: \beq\label{eq-contracut}
3741: \mbox{If}\qquad\quad T_{M,\varphi}\neq 0 \,\, {\rm mod}\,5\qquad\quad{\rm then}
3742: \qquad\quad cut(M)=1\;.\qquad\mbox{(for any choice of $\varphi$)}
3743: \eeq
3744: See \cite{GilKer} for more details and applications.
3745: }
3746:
3747: %%%%%%%%%%%%%%%%%%%%%%%%%% END CUT %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3748:
3749:
3750: \paragraph{A. Relations to Gauge Theories and the TQFT-Ring \Qq Generated by ${\cal V}$.}\
3751:
3752: We begin by collecting the ingredients that imply Theorem~\ref{thm-relations}.
3753: The first identity (\ref{eq-AlexChar}) has already been computed in
3754: (\ref{eq-leftrace3}).
3755:
3756: The invariants $I^{DC}$ and $I^{SW}_d$ are obtained by Donaldson in \cite{Don99}
3757: from TQFT's ${\cal V}^{DC}$ and ${\cal V}^{SW}_d$ respectively. For both TQFT's the
3758: vector spaces associated to a surface $\Sigma$ are the homologies of natural
3759: moduli spaces. In the case of ${\cal V}^{DC}$ this is the moduli space ${\cal M}(\Sigma)$
3760: of flat connections on a non-trivial $SO(3)$ bundle. For ${\cal V}^{SW}_d$
3761: the moduli space of solutions to certain vortex equations is considered, which turn
3762: identified with the symmetric products of the surface. The action of the mapping
3763: class group on the resulting homologies also factors through the symplectic groups (with
3764: the familiar ${\mathbb F}_2$-ambiguity). Donaldson thus derives the following
3765: isomorphisms between $\Sp(2g,\Z)$-modules.
3766: \beq\label{eq-Donaldson}
3767: \begin{array}{rcccl}
3768: \displaystyle {\cal V}^{DC}(\Sigma_g)&\;\;=\;\;&H_*\bigl({\cal M}(\Sigma_g)\bigr)
3769: &\;\;\cong\;\;&\displaystyle
3770: \bigoplus_{j=0}^g\,{\mathbb Q}^{j^2}\otimes \ext {g-j}H_1(\Sigma_g)\;\;\mbox{and}\\
3771: %&&&\\
3772: \displaystyle {\cal V}^{SW}_d(\Sigma_g)&\;\;=\;\;&H_*\bigl(Sym^k(\Sigma_g)\bigr)
3773: &\;\;\cong\;\;&\displaystyle
3774: \bigoplus_{j=1}^{g-d}\,{\mathbb Q}^{j}\otimes \ext {g-d-j}H_1(\Sigma_g)\;,\\
3775: \end{array}
3776: \eeq
3777: where $k=g-1-d$ is the degree the holomorphic line bundle of which the vortex solutions
3778: are sections. The $\Sp(2g,\Z)$-representations can be further identified with the
3779: irreducible parts, which takes in our notation the form
3780: \beq\label{eq-extexpand}
3781: \ext {g-j}H_1(\Sigma_g)\;\;=\;\;{\cal V}^{(j+1)}(\Sigma)\;\oplus\;
3782: {\cal V}^{(j+3)}(\Sigma)\;\oplus\;{\cal V}^{(j+5)}(\Sigma)\;\oplus\;\ldots
3783: \eeq
3784: Inserting (\ref{eq-extexpand}) into the isomorphisms in (\ref{eq-Donaldson}) we
3785: find that the ${\cal V}^{DC}(\Sigma_g)$ and ${\cal V}^{SW}(\Sigma_g)$ are direct
3786: sums of the ${\cal V}^{(j)}(\Sigma_g)$ with $g$-independent
3787: multiplicities given by precisely the
3788: non-negative coefficients in (\ref{eq-DCChar}) and (\ref{eq-SWChar}). In Chapter~5 of
3789: \cite{Don99}
3790: Donaldson exploits this fact to show that the decomposition thus extends to the
3791: entire TQFT's, meaning that cobordisms act trivially on the mulitiplicity spaces and
3792: have block-wise actions on the ${\cal V}^{(j)}$ components equivalent to those in
3793: the ${\cal V}^{FN}$ case. Summarily, we have the following isomorphisms
3794: {\em of TQFT's}.
3795: \beq\label{eq-TQFTidentities}
3796: \begin{array}{rcl}
3797: \displaystyle {\cal V}^{DC}
3798: &\;\;\cong\;\;&\displaystyle
3799: \bigoplus_{j\geq 2}\,{\mathbb Q}^{{j+1}\choose 3}\otimes {\cal V}^{(j)}\;\\
3800: %&&&\\
3801: \displaystyle {\cal V}^{SW}_d
3802: &\;\;\cong\;\;&\displaystyle
3803: \bigoplus_{j\geq d+2}\,{\mathbb Q}^{\left[\!\!{\left[
3804: \Bigl({\frac{j-d}2}\Bigr)^2\right]}\!\!\right]}\otimes {\cal V}^{(j)}\;.\\
3805: \end{array}
3806: \eeq
3807: Identities (\ref{eq-DCChar}) and (\ref{eq-SWChar}) are now immediate. For the
3808: last equation (\ref{eq-LesChar}) in Theorem~\ref{thm-relations}
3809: we refer to \cite{KerKyoto}.
3810:
3811: \ep
3812:
3813: In an effort to find new knot invariants Frohman and Nicas generalized their
3814: approach in \cite{FroNic94} to higher rank Lie algebras. They construct a TQFT
3815: ${\cal V}^{PSU(n)}_k$,
3816: whose vector spaces are
3817: given as intersection homology groups of certain restricted moduli spaces
3818: of $PSU(n)$-representations and derive from these by similar trace formulae
3819: invariants $\lambda_{n,k}$ depending on the
3820: rank $n$ and weight $k$. In \cite{Fro93} Frohman finds a recursive
3821: procedure to compute the invariants $\lambda_{n,k}$ and shows that they
3822: are determined by the polynomial expressions in the
3823: coefficients of the Alexander polynomial. Consequently, they are also
3824: polynomial in the Alexander Characters so that we can write
3825: \beq\label{eq-laRnk}
3826: \lambda_{n,k}\;=\; R_{n,k}(\Delta^{(1)}, \Delta^{(2)}, \ldots)\;,
3827: \eeq
3828: with $R_{n,k}\in{\mathbb Z}[x_1,x_2,\ldots]$. A general, closed
3829: formula and some integrality issues for the $R_{n,k}$ are still unresolved
3830: though, see also \cite{BodNic00}. This relation in (\ref{eq-laRnk}) is more
3831: general than those expressed in Theorem~\ref{thm-relations} as it is no longer
3832: linear.
3833:
3834: More precisely, define the space of invariants
3835: $\Qq^{[0]}\;=\;\left\{{n_1\Delta^{(1)}\,+\,n_2\Delta^{(2)}\,+\,\ldots\,|\,
3836: n_i\in{\mathbb Z}^{+,0}}\right\}$. Then is clear that invariant that descends
3837: from a TQFT that is homological must be in $\Qq^{[0]}$ where the $n_i\geq 0$
3838: are the multiplicities of the irreducible summands. Thus $I^{DC},\,I^{SW}_d\,
3839: \in\,\Qq^{[0]}$, but we also have ${\lambda}_L\not\in\Qq^{[0]}$ since some of
3840: the coefficients are negative. ${\lambda}_L$ is, nevertheless, related to the
3841: quantum TQFT's, but the derivations use slightly more subtle $p$-modular
3842: interpretations, see \cite{KerKyoto}.
3843:
3844: \bigskip
3845:
3846: Similar, to sums we can derive the
3847: invariant given by the product
3848: of two Alexander Characters, say $\Delta^{(i)}\cdot \Delta^{(j)}$,
3849: from the tensor product of the corresponding TQFT's, namely
3850: ${\cal V}^{(i)}\otimes {\cal V}^{(j)}$. Thus is the coefficients of all
3851: the $R_{n,k}$ were non-negative integers we could easily produce a
3852: homological TQFT by taking corresponding direct sums and tensor products
3853: of the ${\cal V}^{(i)}$ in order to reproduce $\lambda_{n,k}$. This
3854: invariant, indeed, descends from the TQFT ${\cal V}^{PSU(n)}_k$, however,
3855: the coefficients of the $R_{n,k}$. The point to observe here is that, for
3856: example, ${\cal V}^{(i)}\otimes {\cal V}^{(j)}$ is generally not an
3857: irreducible TQFT and can be decomposed.
3858:
3859: Denote by ${\cal V}^{(\vec \lambda)}$ the irreducible TQFT's obtained as summands of
3860: quotients of multiple tensor products of the ${\cal V}^{(i)}$. The superscript label,
3861: $\vec\lambda\in\vec\Lambda$, may be roughly thought of as a semi-infinite branching
3862: path for $\Sp(2)\subset \Sp(4)\subset \Sp(6)\subset\ldots\,$. The space of TQFT's
3863: \beq
3864: \Qq^{[+]}\;=\;
3865: \Biggl\{{\,\bigoplus_{{\vec\lambda}\in{\vec\Lambda}} \,{\mathbb Q}^{n_{\vec\lambda}}
3866: \otimes {\cal V}^{(\vec\lambda)}\,\,\Bigg |\,\,\,n_{\vec\lambda}\in {\mathbb N}\cup
3867: \{0\}}\Biggr\}\;
3868: \eeq
3869: thus has a natural ring structure with operations $\oplus$ and $\otimes$ and can be
3870: thought of as a type of Grothendieck $K_0$-ring for a homological subquotient of $\Cob$.
3871: We denote the corresponding set of higher Alexander Characters abusively in the same
3872: way, since it possesses the same ring structure under usual addition and multiplication.
3873: Clearly, $\Qq^{[0]}\subset\Qq^{[+]}$. The following conjecture together with an understanding
3874: of the ring structure of $\Qq^{[+]}$ should shed light
3875: onto the general structure of the polynomials $R_{n,k}$.
3876: \begin{conjecture}
3877: $$
3878: {\cal V}^{PSU(n)}_k\;\;\;\in\;\;\;\Qq^{[+]}
3879: $$
3880: \end{conjecture}
3881:
3882: \emptystuff{
3883:
3884:
3885:
3886: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% BEGIN CUT %%%%%%%%%%%%%%%%%%%%%%%%%%
3887:
3888:
3889:
3890:
3891:
3892: \bigskip
3893:
3894: \bigskip
3895:
3896:
3897: Using the coefficients of
3898: the Conway polynomial $\nabla (z)=\sum_j c_j z^{2j}$ with
3899: $z=t^{\frac 12}-t^{-\frac 12}$ this yields polynomials
3900: $\lambda_{n,k}=q_{n,k}(c_1,c_2,\ldots)$, which appear to
3901: have non-negative
3902: integer coefficients, that is $q_{n,k}\in{\cal P}^+$ as defined in
3903: Section~10. See also \cite{BodNic00} for more explicit formulae.
3904: Changing the basis of the polynomial ring from $z^{2j}$ to
3905: the $[j]_{-t}$ we are similarly able to express the higher
3906: rank invariants in terms of the Alexander Characters $\Delta^{(j)}$ defined
3907: in (\ref{eq-leftrace3}). We can thus write
3908: $$
3909: \lambda_{n,k}\;=\; R_{n,k}(\Delta^{(1)}, \Delta^{(2)}, \ldots)
3910: $$
3911: for some polynomial $R_{n,k}$ with integral coefficients.
3912: Note further that if $C$ is the cobordism on a Seifert surface of a knot
3913: and $P\in{\cal P}^+$ then
3914: $trace({\cal V}^{(P)}(C))=P(\Delta^{(1)}, \Delta^{(2)}, \ldots)$
3915:
3916: The natural
3917: question tied to these observations is whether the TQFT's constructed
3918: in \cite{FroNic94} for general gauge groups are related or equivalent
3919: to a TQFT of the polynomial form ${\cal V}^{(R_{n,k})}$ as defined in
3920: (\ref{eq-defpolyTQFT}). If the coefficients of the $R_{n,k}$ are not
3921: all non-negative we may have to consider two theories
3922: ${\cal V}^{(R_{n,k}^{\pm})}$ with $R_{n,k}=R_{n,k}^+-R_{n,k}^-$ and
3923: $R_{n,k}^{\pm}\in{\cal P}^+$ and make sense of their difference.
3924:
3925:
3926:
3927:
3928:
3929:
3930: In \cite{Don99} Donaldson describes a slightly different TQFT, ${\cal V}^{DF}$,
3931: modeled on moduli spaces ${\cal M}_g$ of connections on a non-trivial $SO(3)$ bundle.
3932: This TQFT leads up to a Casson type invariant for homology circles $Y$, which is
3933: determined by the coefficients of the Alexander polynomial $\Delta_Y$.
3934: The vector spaces are given as
3935: \beq\label{eq-Donaldson}
3936: {\cal V}^{DF}(\Sigma_g)\;\;=\;\;H_*\bigl({\cal M}_g\bigr)
3937: \;\;\cong\;\;\bigoplus_{j=0}^g\,\R^{j^2}\otimes \ext {g-j}H_1(\Sigma_g)\;.
3938: \eeq
3939: The morphisms ${\cal V}^{DF}(M)$ are similarly constructed via the intersection
3940: theory of representation varieties, using also a dimension
3941: reduction of the Floer-cohomology on $\hat M\times S^1$.
3942:
3943: Now from Corollary~5.1.9 in \cite{GooWal98} we see that
3944: $\ext {g-j}H_1(\Sigma_g)\,\cong\,W_{g,j+1}\oplus W_{g,j+3}\oplus W_{g,j+5}\ldots$
3945: as $\Sp(2g,\Z)$ modules. Inserting this decomposition into (\ref{eq-Donaldson}) we
3946: obtain the multiplicities stated in the following conjecture.
3947: \begin{conjecture} Let $\displaystyle D=\sum_{k\geq 1}
3948: {\scriptstyle { k+1 \choose 3}} x_k\;\in\;{\cal P}^+$. Then
3949: $$
3950: {\cal V}^{DF}\;\cong\;{\cal V}^{(D)}\;.
3951: $$
3952: \end{conjecture}
3953: Note that on the level of vector spaces and invariants we do in fact have equality.
3954:
3955:
3956: The theories in \cite{FroNic94} and \cite{Don99} are all inherently
3957: $\Z/2$-projective, and have the vanishing properties from Lemma~\ref{lm-vanish}.
3958: This indicates that they also belong into the class of
3959: half-projective or non-semisimple TQFT's.
3960:
3961:
3962: Another conjecture that is independent of a particular gauge theory may be stated
3963: as follows.
3964: \begin{conjecture} Suppose ${\cal V}$ is a non-semisimple TQFT in which the kernel
3965: of the mapping class group representations are precisely the Torelli groups. The
3966: ${\cal V}$ is isomorphic to a sub-TQFT of some ${\cal V}^{(P)}$ for some
3967: $P\in{\cal P}^+$.
3968: \end{conjecture}
3969: To say that the Torelli group is {\em precisely} the kernel implies that we
3970: have faithful $\Sp(2g,\Z)$-representations so that by Margulis' rigidity
3971: these lift to algebraic $\Sp(2g,\R)$-representations. Classifying homological
3972: TQFT's as described will thus involve exercises in branching rules as given
3973: for example in Section~8.3.4 of \cite{GooWal98}.
3974:
3975:
3976: Another approach to describing TQFT's behind the higher rank Frohman Nicas
3977: $PSU(n)$-theories or the Donaldson $SO(3)$-construction is to try to extract
3978: a categorical Hopf algebra ${\cal A}_{\cal V}$ for the given TQFT by evaluation
3979: of the cobordisms in Figure~\ref{fig-MC-tgl}. The problem with this approach,
3980: however, is that in the non-abelian TQFT's we do not seem to have a nicely
3981: defined tensor structure arising from gluing two one holed surfaces over a
3982: pair of pants. Specifically we need an isomorphism
3983: ${\cal V} (\Sigma_{2,1})=
3984: {\cal V} (\Sigma_{1,1})\otimes{\cal V} (\Sigma_{1,1})$ which is generally
3985: not true because of gauge constraints over the pair of pants.
3986:
3987: It is also not clear whether higher rank theories such as those in \cite{FroNic94}
3988: exhibit symmetries similar to the $\SL(2,\R)$-equivariance
3989: that yields a type of Lefschetz
3990: decomposition. Particularly, the non-abelian moduli spaces have
3991: no canonical K\"ahler structure. They do, however, admit useful Poisson structures
3992: \cite{FocRos97}.
3993: }
3994:
3995:
3996: %%%%%%%%%%%%%%%%%%%%%% END CUT %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3997:
3998:
3999: \paragraph{B. Homology TQFT's from the Reshetikhin-Turaev Theory:}\
4000:
4001:
4002: Recall that the TQFT ${\cal V}^{(j)}$ is in fact a functor to
4003: the category of free $\Z$-modules rather than just the category
4004: of vector spaces over ${\mathbb Q}$. Now, for any prime $p\geq 3$,
4005: by taking all lattices modulo $p$
4006: this in turn maps to the category of vector spaces over the finite
4007: field ${\mathbb F}_p=\Z/p\Z$. The resulting TQFT ${\cal V}^{(j)}_p$
4008: over ${\mathbb F}_p$
4009: is now no longer irreducible, but it has a {\em unique}
4010: irreducible subquotient, which we denote by $\dov {\cal V}^{(j)}_p$,
4011: see \cite{KerRes}.
4012:
4013: Another way of generating TQFT's over ${\mathbb F}_p$ is to consider
4014: the Reshetikhin Turaev Theory for quantum-$SO(3)$ at a primitive
4015: $p$-th root of unity $\zeta_p$. As shown in \cite{Gil01} this can
4016: be regarded as a TQFT over the ring of cyclotomic integers
4017: $\Z[\zeta_p]$. The TQFT obtained from the ring reduction
4018: $\Z[\zeta_p]\onto{12}{\mathbb F}_p\,:\,\zeta_p\mapsto 1$ is denoted
4019: ${\cal V}^{RT}_p$. The example $p=5$, which is in some sense a
4020: fundamental case, is analyzed in \cite{Kerr=5}. We obtain an exact,
4021: but non-split sequence of TQFT's as follows
4022: \beq\label{eq-ext}
4023: 0\;\;\to\;\;\dov {\cal V}^{(4)}_5
4024: \;\;\longrightarrow\;\;
4025: {\cal V}^{RT}_5
4026: \;\;\longrightarrow\;\;
4027: \dov {\cal V}^{(1)}_5
4028: \;\;\to\;\;0
4029: \eeq
4030: As an extension of the mapping class group $\Gamma_g$ (\ref{eq-ext})
4031: involves a Johnson-Morita subquotient of $\Gamma_g$. The precise modular
4032: structure of the ${\cal V}^{(j)}_p$ and $\dov {\cal V}^{(j)}_p$
4033: TQFT's is unraveled in \cite{KerRes}. There we find resolutions
4034: of the $\dov {\cal V}^{(j)}_p$ in terms of the ${\cal V}^{(j)}_p$,
4035: which lead to important identities between the $p$-modular versions
4036: of the invariants from Theorem~\ref{thm-relations} and the
4037: Reshetikhin Turaev Invariants.
4038:
4039: It is easy to see that the irreducible factors of ${\cal V}^{RT}_p$ for
4040: $p\geq 7$ can no longer be reductions of the ${\cal V}^{(j)}$. There is,
4041: however, evidence that suggests that the irreducible factors are
4042: reductions of summands in the symmetric powers of the fundamental ones.
4043: That is, TQFT's of the form
4044: \beq
4045: {\cal V}^{\vec\lambda}
4046: \;\subseteq\;S^{\frac {p-3}2}{\cal V}^{FN}\,\;\in\,\;\Qq^{[+]}\,.
4047: \eeq
4048: This is closely related to the conjecture that the Lescop invariant
4049: for a closed 3-manifold $M$ with $\beta_1(M)\geq 1$ relates to the
4050: Reshetikhin Turaev Invariant as follows.
4051: \beq
4052: {\cal V}^{RT}_{\zeta_p}(M)\;\;=\;\;C_p\cdot
4053: \Bigl((\zeta_p-1)\lambda_L(M)\Bigr)^{\frac {p-3}2}\;\;+\;\;{\cal O}
4054: \bigl((\zeta_p-1)^{\frac {p-1}2}\bigr)\;\;.
4055: \eeq
4056: This has been verified for $p=5$ in \cite{KerKyoto}.
4057:
4058:
4059: \paragraph{C. Relation of Reshetikhin-Turaev and Hennings Theory:}\
4060:
4061: Given a quasitriangular Hopf algebra, ${\cal A}$, we have described in
4062: Section~5 a procedure to construct a topological quantum field theory,
4063: ${\cal V}_{\cal A}^{H}$. In \cite{ResTur91} and \cite{Tur94} Reshetikhin and
4064: Turaev give another procedure to construct a TQFT, ${\cal V}^{RT}_{\cal S}$,
4065: from a {\em semisimple} modular category, ${\cal S}$. A more general
4066: construction in \cite{KerLub00} allows us to construct a TQFT,
4067: ${\cal V}_{\cal C}^{KL}$, also for modular categories, ${\cal C}$,
4068: that are not semisimple, and we show in \cite{Ker96} that
4069: ${\cal V}_{\cal A}^{H}= {\cal V}_{{\cal A}-mod}^{KL}$ and
4070: ${\cal V}^{RT}_{\cal S}={\cal V}_{\cal S}^{KL}$ for semisimple
4071: ${\cal S}$. For a non-semisimple, quasitriangular algebra, ${\cal A}$,
4072: the semisimple category used in \cite{ResTur91} and \cite{Tur94}
4073: is given as the semisimple trace-quotient
4074: ${\cal S}({\cal A})=\overline{{\cal A}-mod}$ of the representation
4075: category of $\cal A$. The relation between ${\cal V}_{\cal A}^{H}$
4076: and ${\cal V}^{RT}_{{\cal S}({\cal A})}$ is generally unknown.
4077: We make the following conjecture in the case of quantum ${\mathfrak s}{\mathfrak l}_2$.
4078:
4079: \begin{conjecture}
4080: Let ${\cal A}=U_q({\mathfrak s}{\mathfrak l}_2)^{red}$, with $q$ an odd $p$-th
4081: root of unity, and relations $E^p=F^p=0$ and $K^{2p}=1$ for the standard generators. Then there is monomorphic, natural transformation
4082: \beq\label{eq-HRT}
4083: {\cal V}^{FN}\otimes
4084: {\cal V}^{RT}_{{\cal S}({\cal A})}\;\;\into\;\;{\cal V}_{\cal A}^H\;.
4085: \eeq
4086: \end{conjecture}
4087:
4088: In the genus one case we have shown in \cite{Ker94} and \cite{Ker96}
4089: that the mapping class group representations and invariants of
4090: lens spaces of both theories in ({\ref{eq-HRT}) are in fact equal.
4091: The above inclusion of TQFT functors can also be phrased in the form
4092: ${\cal V}_{{\cal C}^{\#}}^{KL}\into{\cal V}^{KL}_{{\cal C}}$,
4093: where ${\cal C}:=U_q({\mathfrak s}{\mathfrak l}_2)^{red}-mod$ and
4094: ${\cal C}^{\#}:=({\cal N}-mod)\otimes \overline{\cal C}$.
4095: The categories ${\cal C}$ and ${\cal C}^{\#}$ are in fact rather
4096: similar as linear abelian categories. From \cite{Ker98} it follows
4097: that there an isomorphism of {\em abelian} categories
4098: \beq
4099: %\mbox{\bm${\cal H}$\ubm}
4100: \Hh\;\;:\;\;
4101: {\cal C}^{\#}\;\oplus\;2\cdot {\rm Vect}({\mathbb C}) \;\;\;
4102: \stackrel{\cong}{\longrightarrow}\;\;\;{\cal C}\;,
4103: \eeq
4104: where the two extra ${\rm Vect}({\mathbb C})$'s account for the two $p$-dimensional,
4105: irreducible Steinberg modules. This, however, is {\em not} a
4106: monoidal functor. Instead we have a natural set of monomorphisms of
4107: the form $\Hh(X)\otimes\Hh(Y)\,\into\,\Hh(X\otimes Y)$. As a result
4108: the braidings, integrals, and coends, that enter in a crucial way
4109: the construction of the TQFT's \cite{KerLub00}
4110: can no longer be na\"\i vely
4111: identified. Strategies of proof would include a basis
4112: of ${\cal A}$ as worked out in \cite{Ker94} and the use of the special
4113: central, nilpotent element ${\sf Q}$ defined in \cite{Ker96}.
4114:
4115: %%%%%%%%%%%%%%%%%%% BEGIN CUT %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4116:
4117: \emptystuff{
4118:
4119:
4120: Specifically, we know the following:
4121: \begin{theorem}[\cite{Ker89}]\ Let ${\cal A}=U_q({\mathfrak s}{\mathfrak l}_2)^{red}$ and
4122: $\cal N$ as in Section~6.
4123: \begin{enumerate}
4124: \item
4125: For any generic Casimir value, $c\in ({\mathfrak z}({\cal A}))^*$,
4126: the corresponding subcategory ${\cal C}_c\subset{\cal A}-mod$
4127: of representations is isomorphic to ${\cal N}-mod$.
4128: \item The representations with non-generic Casimir values are
4129: sums of the two irreducible Steinberg modules of dimension $r$
4130: and quantum dimension 0.
4131: \item An indecomposable representation of ${\cal N}$ is either
4132: one of the two 4-dim
4133: projective representations in ${\cal N}={\cal N}^+\oplus{\cal N}^-$,
4134: or an indecomposable representation of one of the two Kronecker quivers
4135: $\bullet \raisebox{-.5ex}{\hbox{$\stackrel{\mbox{$\longrightarrow$}}{\longrightarrow}$}}\bullet$ and
4136: $\bullet \raisebox{-.5ex}{\hbox{$\stackrel{\mbox{$\longleftarrow$}}{\longleftarrow}$}}\bullet$,
4137: where the $\bullet$'s stand for an eigenspaces of $K$.
4138: \end{enumerate}
4139: \end{theorem}
4140: The generic Casimir values are in a two to one correspondence with the
4141: admissible irreducible representations, and we have
4142: ${\cal C}=\bigoplus_c{\cal C}_c$ and
4143: ${\cal C}^{\#}=\bigoplus_j {\cal N}-mod$, where $j$ runs over
4144: irreducible representations. Thus we have a close correspondence between
4145: the modules in both categories. They differ, however, more strongly
4146: as tensor categories. Strategies of proof would include a basis
4147: of ${\cal A}$ as worked out in \cite{Ker94} and the use of the special
4148: central, nilpotent element ${\sf Q}$ defined in \cite{Ker96}.
4149:
4150: }
4151:
4152: %%%%%%%%%%%%%%%%%%%%%%%%% BEGIN CUT %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4153:
4154: \emptystuff{
4155:
4156: \paragraph{C. Universality of ${\cal V}$ and Casson Type Gauge Theories.}\
4157:
4158: In order to find new knot invariants Frohman and Nicas generalized their
4159: approach in \cite{FroNic94} to higher rank Lie algebras. They construct a TQFT
4160: whose vector spaces are
4161: given as intersection homology groups of certain restricted moduli spaces
4162: of $PSU(n)$-representations and derive from these by similar trace formulae
4163: invariants $\lambda_{n,k}$ depending on the
4164: rank $n$ and weight $k$. In \cite{Fro93} Frohman finds a recursive
4165: procedure to compute the invariants $\lambda_{n,k}$ and shows that they
4166: are determined by the polynomial expressions in the
4167: coefficients of the Alexander polynomial. Using the coefficients of
4168: the Conway polynomial $\nabla (z)=\sum_j c_j z^{2j}$ with
4169: $z=t^{\frac 12}-t^{-\frac 12}$ this yields polynomials
4170: $\lambda_{n,k}=q_{n,k}(c_1,c_2,\ldots)$, which appear to
4171: have non-negative
4172: integer coefficients, that is $q_{n,k}\in{\cal P}^+$ as defined in
4173: Section~10. See also \cite{BodNic00} for more explicit formulae.
4174: Changing the basis of the polynomial ring from $z^{2j}$ to
4175: the $[j]_{-t}$ we are similarly able to express the higher
4176: rank invariants in terms of the Alexander Characters $\Delta^{(j)}$ defined
4177: in (\ref{eq-leftrace3}). We can thus write
4178: $$
4179: \lambda_{n,k}\;=\; R_{n,k}(\Delta^{(1)}, \Delta^{(2)}, \ldots)
4180: $$
4181: for some polynomial $R_{n,k}$ with integral coefficients.
4182: Note further that if $C$ is the cobordism on a Seifert surface of a knot
4183: and $P\in{\cal P}^+$ then
4184: $trace({\cal V}^{(P)}(C))=P(\Delta^{(1)}, \Delta^{(2)}, \ldots)$
4185:
4186: The natural
4187: question tied to these observations is whether the TQFT's constructed
4188: in \cite{FroNic94} for general gauge groups are related or equivalent
4189: to a TQFT of the polynomial form ${\cal V}^{(R_{n,k})}$ as defined in
4190: (\ref{eq-defpolyTQFT}). If the coefficients of the $R_{n,k}$ are not
4191: all non-negative we may have to consider two theories
4192: ${\cal V}^{(R_{n,k}^{\pm})}$ with $R_{n,k}=R_{n,k}^+-R_{n,k}^-$ and
4193: $R_{n,k}^{\pm}\in{\cal P}^+$ and make sense of their difference.
4194:
4195:
4196:
4197:
4198:
4199:
4200: In \cite{Don99} Donaldson describes a slightly different TQFT, ${\cal V}^{DF}$,
4201: modeled on moduli spaces ${\cal M}_g$ of connections on a non-trivial $SO(3)$ bundle.
4202: This TQFT leads up to a Casson type invariant for homology circles $Y$, which is
4203: determined by the coefficients of the Alexander polynomial $\Delta_Y$.
4204: The vector spaces are given as
4205: \beq\label{eq-Donaldson}
4206: {\cal V}^{DF}(\Sigma_g)\;\;=\;\;H_*\bigl({\cal M}_g\bigr)
4207: \;\;\cong\;\;\bigoplus_{j=0}^g\,\R^{j^2}\otimes \ext {g-j}H_1(\Sigma_g)\;.
4208: \eeq
4209: The morphisms ${\cal V}^{DF}(M)$ are similarly constructed via the intersection
4210: theory of representation varieties, using also a dimension
4211: reduction of the Floer-cohomology on $\hat M\times S^1$.
4212:
4213: Now from Corollary~5.1.9 in \cite{GooWal98} we see that
4214: $\ext {g-j}H_1(\Sigma_g)\,\cong\,W_{g,j+1}\oplus W_{g,j+3}\oplus W_{g,j+5}\ldots$
4215: as $\Sp(2g,\Z)$ modules. Inserting this decomposition into (\ref{eq-Donaldson}) we
4216: obtain the multiplicities stated in the following conjecture.
4217: \begin{conjecture} Let $\displaystyle D=\sum_{k\geq 1}
4218: {\scriptstyle { k+1 \choose 3}} x_k\;\in\;{\cal P}^+$. Then
4219: $$
4220: {\cal V}^{DF}\;\cong\;{\cal V}^{(D)}\;.
4221: $$
4222: \end{conjecture}
4223: Note that on the level of vector spaces and invariants we do in fact have equality.
4224:
4225:
4226: The theories in \cite{FroNic94} and \cite{Don99} are all inherently
4227: $\Z/2$-projective, and have the vanishing properties from Lemma~\ref{lm-vanish}.
4228: This indicates that they also belong into the class of
4229: half-projective or non-semisimple TQFT's.
4230:
4231:
4232: Another conjecture that is independent of a particular gauge theory may be stated
4233: as follows.
4234: \begin{conjecture} Suppose ${\cal V}$ is a non-semisimple TQFT in which the kernel
4235: of the mapping class group representations are precisely the Torelli groups. The
4236: ${\cal V}$ is isomorphic to a sub-TQFT of some ${\cal V}^{(P)}$ for some
4237: $P\in{\cal P}^+$.
4238: \end{conjecture}
4239: To say that the Torelli group is {\em precisely} the kernel implies that we
4240: have faithful $\Sp(2g,\Z)$-representations so that by Margulis' rigidity
4241: these lift to algebraic $\Sp(2g,\R)$-representations. Classifying homological
4242: TQFT's as described will thus involve exercises in branching rules as given
4243: for example in Section~8.3.4 of \cite{GooWal98}.
4244:
4245:
4246: Another approach to describing TQFT's behind the higher rank Frohman Nicas
4247: $PSU(n)$-theories or the Donaldson $SO(3)$-construction is to try to extract
4248: a categorical Hopf algebra ${\cal A}_{\cal V}$ for the given TQFT by evaluation
4249: of the cobordisms in Figure~\ref{fig-MC-tgl}. The problem with this approach,
4250: however, is that in the non-abelian TQFT's we do not seem to have a nicely
4251: defined tensor structure arising from gluing two one holed surfaces over a
4252: pair of pants. Specifically we need an isomorphism
4253: ${\cal V} (\Sigma_{2,1})=
4254: {\cal V} (\Sigma_{1,1})\otimes{\cal V} (\Sigma_{1,1})$ which is generally
4255: not true because of gauge constraints over the pair of pants.
4256:
4257: It is also not clear whether higher rank theories such as those in \cite{FroNic94}
4258: exhibit symmetries similar to the $\SL(2,\R)$-equivariance
4259: that yields a type of Lefschetz
4260: decomposition. Particularly, the non-abelian moduli spaces have
4261: no canonical K\"ahler structure. They do, however, admit useful Poisson structures
4262: \cite{FocRos97}.
4263:
4264:
4265: \paragraph{D. Milnor Torsion and Seiberg Witten invariants}\
4266:
4267:
4268: The Milnor Torsion of a 3-manifold $M$ is defined from a cell complex of a
4269: simplicial representation of the cyclic covering space $\widetilde M$.
4270: The relation between the Alexander polynomial and Milnor Torsion as stated
4271: in Theorem~\ref{thm-ReidAlex} suggests that there should be a quantum topological
4272: description of the invariants obtained from a simplicial complex. In fact
4273: the Turaev Viro and Kuperberg invariant as examples of quantum invariants
4274: that start from a cell decomposition of $M$. Given the non-semisimple nature
4275: of our theory the Kuperberg invariant \cite{Kup96} is a more natural candidate.
4276: The basic Hopf algebra is likely to be similar to the Borel subalgebra
4277: generated by $K$ and $\th$ but not $\tb$ following a conjecture that the
4278: Hennings and Kuperberg are related by Drinfel'd double construction. The
4279: difficulties, however, consist in describing a cell decomposition of the
4280: cyclic covering space $\widetilde M$ from a Heegaard diagram for $M$.
4281: The main problem being that the Kuperberg theory has no easy extension
4282: to a TQFT.
4283:
4284: Nevertheless, we propose as a problem to find a direct description of
4285: Milnor-Reidemeister torsion via the picture developed by Kuperberg for
4286: the construction of 3-manifold invariants.
4287:
4288: In \cite{Tur98} Turaev shows that an extensions of
4289: the Milnor torsion is equal to the Seiberg Witten invariant
4290: for 3-manifolds equipped with $Spin^C$-structures or, equivalently, Euler
4291: structures and with $\beta_1(M)>0$. A weaker version of such an equivalence
4292: without additional structures was shown by Meng and Taubes
4293: \cite{MengTaub}.
4294: The proof in \cite{Tur98} uses the fact that
4295: both invariants follows the same recursion formulae under surgery.
4296: It should be interesting to relate these formulae to the skein theory
4297: developed here and find ways of including the additional structures
4298: in our context.
4299:
4300:
4301: In general our procedure is also limited to either the reduced Torsion
4302: or Alexander polynomial if $\beta_1(M)\geq 2$. Additional generators
4303: of homology can be represented as additional 0-framed surgery links with
4304: zero linking numbers. It is, however, not as obvious in this case how
4305: to generate a TQFT picture that would allow us to describe the full
4306: invariants with values in $\Z[H_1^{(free)}(M)]$.
4307:
4308:
4309: \paragraph{E. Relations to quantum field theories:}\
4310:
4311: Let us mention here only briefly interpretations of the homological TQFT's in
4312: a physics context. For small genera Rozansky and Saleur
4313: \cite{RosSal93} find the same vector spaces and representations of the
4314: mapping class groups from the $U(1,1)$ Wess-Zumino-Witten theory.
4315:
4316: The exterior product spaces may also be interpreted as fermionic Fock spaces.
4317: Ideas of constructing such ferminoc topological $U(1)$-theories in general
4318: have been suggested for example by Louis Crane.
4319: }
4320:
4321:
4322: %%%%%%%%%%%%%%%%%%%%%%%%%% END CUT %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4323:
4324: {\small
4325:
4326:
4327: \begin{thebibliography}{99}
4328:
4329: %\bibitem[Dr]{Dri:QG}
4330: %Drinfeld, V.G.: {Quantum groups},
4331: %Proceedings of the ICM, Amer. Math. Soc.,
4332: %Providence, Rhode Island, {\bf 1} (1987), 798--820.
4333: \bibitem{Ati88}
4334: M. Atiyah: {Topological quantum field theories}, Inst. Hautes
4335: \'Etudes Sci. Publ. Math. \textbf{68} (1988), 175--186.
4336:
4337: \bibitem{BodNic00} H. Boden, A. Nicas:
4338: Universal formulae for ${\rm SU}(n)$ Casson invariants of knots.
4339: {\em Trans. Amer. Math. Soc.} \textbf{352} (2000), no.~7, 3149--3187.
4340:
4341: \bibitem{BKLT00} Yu. Bespalov , T. Kerler, V. Lyubashenko, V. Turaev:
4342: Integrals for braided Hopf algebras. {\em J. Pure Appl. Algebra}
4343: \textbf{148} (2000) no.~2, 113--164.
4344:
4345: \bibitem{BurZie} G. Burde, H. Zieschang: Knots.
4346: {\em de Gruyter Stud. Math.} \textbf{5} Walter de Gruyter 1985.
4347:
4348: \bibitem{Don99}
4349: S. K. Donaldson:
4350: Topological field theories and formulae of Casson and Meng-Taubes.
4351: Proceedings of the Kirbyfest, {\em Geom. Topol.} {\bf 2} (1999) 87--102
4352: (electronic).
4353: \bibitem{Dri90}
4354: V.G. Drinfeld: \emph{On almost cocommutative {H}opf algebras},
4355: {\em Leningrad Math. J.} \textbf{1} (1990) no.~2, 321--342.
4356: \bibitem{Fro93} C. Frohman:
4357: Unitary representations of knot groups.
4358: {\em Topology} \textbf{32} (1993), no.~1, 121--144.
4359:
4360: \bibitem{FroNic92} C. Frohman, A. Nicas: The Alexander Polynomial via topological quantum field
4361: theory, {\em Differential Geometry, Global Analysis, and Topology, Canadian Math. Soc. Conf.
4362: Proc.} Vol. 12, Amer. Math. Soc. Providence, RI, (1992) 27--40.
4363: \bibitem{FocRos97}V. V. Fock, A. A. Rosly:
4364: Moduli space of flat connections as a Poisson manifold.
4365: Advances in quantum field theory and statistical mechanics:
4366: 2nd Italian-Russian collaboration (Como, 1996).
4367: {\em Internat. J. Modern Phys.} B {\bf 11} (1997) no.~26-27, 3195--3206.
4368: \bibitem{FroNic94} C. Frohman, A. Nicas:
4369: An intersection homology invariant for knots in a rational
4370: homology 3-sphere, {\em Topology} {\bf 33} (1994) no.~1, 123-158.
4371:
4372: \bibitem{Gil01} P. Gilmer: Integrality for TQFTs. Preprint 2001
4373: {\tt math.QA/0105059}.
4374:
4375: %\bibitem{GilKer} P. Gilmer, T. Kerler: in preparation.
4376:
4377: \bibitem{GooWal98}
4378: R. Goodman, N.R. Wallach:
4379: Representations and invariants of the classical groups.
4380: {\em Encyclopedia of Mathematics and its Applications} {\bf 68} Cambridge University
4381: Press, 1998.
4382: \bibitem{GriHar78} P. Griffith, J. Harris: Principles of algebraic geometry.
4383: {\em Wiley Classics Library}. John Wiley \& Sons, Inc.,
4384: New York (1978/1994) xiv+813 pp. ISBN: 0-471-05059-8.
4385: \bibitem{Hen96}
4386: M. Hennings:
4387: Invariants of links and $3$-manifolds obtained from Hopf algebras,
4388: {\em J. London Math. Soc. (2)} \textbf{54} (1996), no.~3, 594--624.
4389: \bibitem{KauRad95}
4390: L.H. Kauffman, D.E. Radford: Invariants of $3$-manifolds derived
4391: from finite-dimensional Hopf algebras,
4392: {\em J. Knot Theory Ramifications} \textbf{4} (1995) no.~1, 131--162.
4393: \bibitem{Kau94} L.H. Kauffman:
4394: Gauss codes, quantum groups and ribbon Hopf algebras,
4395: {\em Rev. Math. Phys.} {\bf 5} (1993) no.~4, 735--773.
4396: \bibitem{Ker89}
4397: T. Kerler: Darstellungen der Quantengruppen und Anwendungen.
4398: Diploma thesis, ETH-Z\"urich, 1989, unpublished.
4399: \bibitem{Ker94}
4400: T. Kerler: Mapping class group actions on quantum doubles,
4401: {\em Commun. Math. Phys.} {\bf 168} (1994) 353-388.
4402: \bibitem{Ker96}
4403: T. Kerler : Genealogy of nonperturbative quantum-invariants of
4404: 3-manifolds: The surgical family. In `Geometry and Physics', {\em Lecture
4405: Notes in Pure and Applied Physics} {\bf 184}, Marcel Dekker (1997)
4406: 503-547.
4407: \bibitem{Ker98}
4408: T. Kerler: On the connectivity of cobordisms and half-projective
4409: TQFT's,
4410: \emph{Commun. Math. Phys.} {\bf 198} No. 3 (1998) 535-590.
4411: \bibitem{Ker99}
4412: T. Kerler: Bridged links and tangle presentations
4413: of cobordism categories, {\em Adv. Math.} {\bf 141} (1999) 207-281.
4414: \bibitem{KerRes} T. Kerler: Resolutions of $p$-Modular TQFT's and Representations of Symmetric Groups.
4415: {\tt math.GT/0110006}
4416:
4417: \bibitem{Kerr=5}
4418: T. Kerler: The Structure of the Fibonacci TQFT. In preparation.
4419:
4420:
4421: \bibitem{KerKyoto} T. Kerler: $p$-Modular TQFT's,
4422: Milnor-Torsion, and the Casson-Lescop Invariant.
4423: {\tt math.GT/0203256}. {\em Geom. Topolo. Monogr.} (in press).
4424:
4425: \bibitem{KerLub00} T. Kerler, V.V. Lyubashenko: Non-semisimple topological quantum field theories
4426: for 3-manifolds with corners.
4427: {\em Lecture Notes in Mathematics} 2001 Springer-Verlag (to appear)
4428: 380 pages, 200 illustrations.
4429: \bibitem{Kir78}
4430: R. Kirby: A calculus for framed links in $S^3$,
4431: {\em Invent. Math.} {\bf 45} (1978), 35-56.
4432:
4433:
4434: \bibitem{Kup96} G. Kuperberg:
4435: Noninvolutory Hopf algebras and $3$-manifold invariants.
4436: Duke Math. J. {\bf 84} (1996), no. 1, 83--129.
4437:
4438:
4439: Involutory Hopf algebras and $3$-manifold invariants.
4440: Internat. J. Math. {\bf 2} (1991), no. 1, 41--66.
4441:
4442: \bibitem{LarSwe69}
4443: R.G. Larson, M.E. Sweedler: An associative orthogonal
4444: bilinear form for Hopf algebras {\em Amer. J. Math.} {\bf 91} (1969) 75--94.
4445:
4446: \bibitem{Les} {Lescop, C.}:
4447: Global surgery formula for the Casson-Walker invariant.
4448: {\em Annals of Mathematics Studies}, {\bf 140}.
4449: Princeton University Press, Princeton, NJ, 1996.
4450:
4451:
4452: \bibitem{Lik97} W.B.R. Lickorish: An Introduction to Knot Theory.
4453: {\em Graduate Texts in Mathematics} {\bf 127}. Springer Verlag, 1997.
4454:
4455: \bibitem{Lub95}
4456: V.V. Lyubashenko : Invariants of 3-manifolds and projective
4457: representations of mapping class groups via quantum groups at roots
4458: of unity, {\em Commun. Math. Phys. }\textbf{172} (1995) 467--516.
4459:
4460: \bibitem{LubMaj94}
4461: V.V. Lyubashenko, S. Majid: {Braided groups and quantum
4462: Fourier transform}, J. Algebra {\bf 166} (1994), n.~3, 506--528.
4463:
4464: \bibitem{Maj93}
4465: S. Majid: {Braided groups},
4466: J. Pure Appl. Algebra {\bf 86} (1993) n.~2, 187--221.
4467:
4468: \bibitem{massey} W. S. Massey: A Basic Course in Algebraic Topology.
4469: {\em Graduate Texts in Mathematics} {\bf 127}. Springer Verlag, 1991.
4470:
4471: \bibitem{MasRob97} G. Masbaum, J.D. Roberts:
4472: A simple proof of integrality of quantum invariants at prime roots of unity.
4473: Math. Proc. Cambridge Philos. Soc. {\bf 121} (1997), no. 3, 443--454.
4474:
4475:
4476:
4477: \bibitem{MatPol94}
4478: S. Matveev, M. Polyak: A geometrical presentation of the surface
4479: mapping class group and surgery,
4480: {\em Commun. Math. Phys.} {\bf 160} (1994) 537--550.
4481:
4482: \bibitem{MengTaub} G. Meng, C. Taubes:
4483: $\underline{\rm SW}=$ Milnor torsion.
4484: {\em Math. Res. Lett.} {\bf 3} (1996), no. 5, 661--674.
4485:
4486: \bibitem{Mil61}
4487: J. Milnor: A duality theorem for Reidemeister torsion.
4488: {\em Ann. of Math.} (2) {\bf 76} (1962) 137--147.
4489:
4490: \bibitem{Mur94} H. Murakami: Quantum $\,SU(2)$-invariants dominate
4491: Casson's $\,SU(2)$-invariant. Math. Proc. Camb. Phil. Soc. {\bf 115} (1994),
4492: 83--103.
4493:
4494: \bibitem{Oh96}
4495: T. Ohtsuki: A polynomial invariant of rational homology $3$-spheres. Invent.
4496: Math. 123 (1996), no.~2, 241--257.
4497:
4498:
4499: \bibitem{Rad76} D. Radford: The Order of the Antipode of a Finite
4500: Dimensional Hopf Algebra is Finite.
4501: {\em American J. Math.} (1976) {\bf 98}, no.~2, 333-355.
4502:
4503:
4504: \bibitem{Rad94}
4505: D. Radford: The trace function and Hopf algebras. {\em J. Algebra} {\bf 163} (1994) 583--622.
4506:
4507:
4508:
4509:
4510: \bibitem{ResTur91}
4511: N.Yu. Reshetikhin, V.G. Turaev:
4512: {Invariants of 3-manifolds via link polynomials and quantum groups},
4513: Invent. Math. {\bf 103} (1991), 547--597.
4514: \bibitem{RosSal93}
4515: L. Rozansky, H. Saleur:
4516: $S$- and $T$-matrices for the super ${\rm U}(1,1) {\rm WZW}$ model. Application to surgery
4517: and $3$-manifolds invariants based on the Alexander-Conway polynomial.
4518: {\em Nuclear Phys. B} {\bf 389} (1993) no. 2, 365--423.
4519:
4520:
4521: \bibitem{Tur76} Turaev, V. G.
4522: Reidemeister torsion and the Alexander polynomial. (Russian)
4523: Mat. Sb. (N.S.) {\bf 18(66)} (1976), no. 2, 252--270.
4524:
4525: \bibitem{Tur94}
4526: V. Turaev: {Quantum invariants of knots and 3-manifolds}, de Gruyter
4527: Stud. Math. {\bf 18} Walter de Gruyter \& Co., Berlin (1994) 588 pp.
4528:
4529: \bibitem{Tur98} V. Turaev:
4530: A combinatorial formulation for the Seiberg-Witten invariants of $3$-manifolds.
4531: Math. Res. Lett. {\bf 5} (1998), no. 5, 583--598.
4532:
4533: \bibitem{TurVi92} V. Turaev, O. Viro:
4534: State sum invariants of $3$-manifolds and quantum $6j$-symbols.
4535: Topology {\bf 31} (1992), no. 4, 865--902.
4536:
4537: \bibitem{Waj83}B. Wajnryb: A simple presentation for the mapping class group of an orientable surface, {\em Israel J. Math.}{\bf 45 }(1983) no. 2-3, 157--174.
4538: B. Wajnryb: An elementary approach to the mapping class group of a surface,
4539: {\em Geom. Topol.}{\bf 3} (1999) 405--466 (electronic).
4540:
4541: \bibitem{Wi89} E. Witten:
4542: Quantum field theory and the Jones polynomial.
4543: {\em Comm. Math. Phys.} {\bf 121} (1989), no.~3, 351--399.
4544:
4545: \bibitem{Yet97}
4546: D. Yetter:
4547: Portrait of the handle as a Hopf algebra.
4548: Geometry and physics. (Aarhus, 1995),
4549: {\em Lecture Notes in Pure and Appl. Math.} {\bf 184},
4550: Dekker, New York, 1997, pp. 481--502.
4551:
4552: \end{thebibliography}
4553: }
4554:
4555: {\sc\footnotesize The Ohio State University,
4556:
4557: Department of Mathematics,
4558:
4559:
4560: 231 West 18th Avenue,
4561:
4562: Columbus, OH 43210, U.S.A. }
4563:
4564: {\em E-mail: }{ \tt kerler@math.ohio-state.edu}
4565: \end{document}
4566: \newpage
4567:
4568: \begin{comment}
4569: The combinatorial description of ${\cal V}_{\cal N}$ can be
4570: greatly simplified with a few modifications of the Hennings calculus.
4571: For example, in the substitution in (\ref{fig-Dec-cross}) we want to
4572: reinterpret the indifferent crossing as the ordinary, original crossing with
4573: ${\cal Z}={\cal Z}^{\dagger}$ still inserted. Hence, the sum $\sum_j$ in (\ref{fig-Dec-cross})
4574: contains only the two terms $1\otimes 1$ and $\th\otimes K\tb$. One easily verifies
4575: that the new crossings with ${\cal Z}$ still fulfill the strict Reidemeister
4576: moves of curves in the plane without the extra element $G=K$ for the first move.
4577: Moreover, we can move $\th$ and $\tb$ through a
4578: crossing at the expense of adding a $K$ to the opposing strand. Other useful rules
4579: are that we only need to consider summands for which every closed
4580: component has exactly one $\th$ and one $\tb$, and a component that contains a
4581: decoration $\rho$ can be removed after multiplication with $i$. Further simplifications
4582: include rules for the insertion of the tensor $\tb\otimes\th-\th\otimes\tb$ on
4583: different strands. Summarily, we have an effective calculus at our hands that allows us to compute
4584: the relevant homological data for ${\cal V}^{FN}$ from a surgery diagram.
4585:
4586: %$\,B_1(x\otimes y)\;=\;\;\;{\sf T}(x)\otimes y\;=\;
4587: % \;\;\;x\otimes y +\mu(x)1\otimes y\;$
4588: %\medskip
4589: \end{comment}
4590:
4591:
4592:
4593: +++++++++++++++++++++++
4594:
4595: In the constructions for the $U(1)$-case are generalized to
4596: $PU(n)$-representations, yielding TQFT's,
4597: ${\cal V}^{FN}_{n,k}$, and knot invariants, . The vector spaces
4598: for connected surfaces with one boundary component are given as
4599: $$
4600: {\cal V}^{FN}_{n,k}(\Sigma_{g,1})\;=\;
4601: IH^m_*\Biggl(p_*\Bigl({\rm Hom}_k(\pi_1(\Sigma_{g,1}),SU(n))/SU(n)\Bigr)\Biggr)\;.
4602: $$
4603: Here, ${\rm Hom}_k$ stands for all those representation
4604: $\rho: \pi_1(\Sigma_{g,1})\to SU(n)$, which map the class $z\in \pi_1$
4605: of the loop around the hole to the central element
4606: $e^{2\pi i\frac kn}\in SU(n)$. $\ldots/SU(n)$ denotes conjugacy classes with
4607: respect to the adjoint action on $SU(n)$, and $p_*$ is the map induced
4608: by $p:SU(n)\to PU(n)$. Finally, $IH^m_*$ denotes the Goresky-MacPherson
4609: intersection homology in middle perversity $m$.
4610:
4611:
4612: +++++++++++++++++++++++++
4613:
4614: Here corresponding coupons are labeled with the same letter.
4615: Given orientations on the strands we thus have the same situation as in the classification
4616: of 2-dimensional surfaces, see \cite{massey}. The coupons can be slid into the standard
4617: form with torus blocs as in the picture below.
4618:
4619:
4620:
4621: \begin{center}
4622: \leavevmode
4623: \epsfbox{skeineval.eps}
4624: \end{center}
4625: %\caption{Skein Evaluation}\lbl{fig-skeineval}
4626: %\end{figure}
4627:
4628: