1: \documentclass[12pt]{article}
2: \usepackage{amsmath,amsfonts,theorem}
3: \input epsf
4: \pagestyle{myheadings}
5: \newcommand{\ntag}[1]{} %\tag{#1}}
6: \numberwithin{equation}{section}
7: \newtheorem{proof}{Proof}
8: \newtheorem{prop}{Proposition}
9: \newtheorem{pro}{Proposition}
10: \renewcommand{\thepro}{}
11: \newtheorem{cor}{Corollary}
12: \newtheorem{co}{Corollary}
13: \renewcommand{\theco}{}
14: \newtheorem{thm}{Theorem}
15: \newtheorem{lem}{Lemma}
16:
17: \theorembodyfont{\rmfamily}
18: \newtheorem{dfn}{Definition}
19: \newtheorem{mcdn}{Main Condition}
20: \newtheorem{warning}{Warning}
21: \renewcommand{\thewarning}{}
22: \newtheorem{rmk}{Remark}
23: \renewcommand{\thermk}{}
24:
25: \newenvironment{pf}{\paragraph{Proof}}{\par\bigskip}
26: \newcommand{\qed}{\ifhmode\unskip\nobreak\fi\quad\ensuremath\square}
27:
28: \newenvironment{digr}[1]{\subsubsection*{#1}}{\quad$\diamondsuit$\par\bigskip}
29:
30: \newcommand{\Span}[1]{\left< #1 \right>}
31: \newcommand{\half}{\textstyle\frac12}
32: \newcommand{\dd}{\mathrm{d}}
33: \newcommand{\id}{\operatorname{id}}
34: \newcommand{\pr}{\operatorname{pr}}
35: \newcommand{\ch}{\operatorname{ch}}
36: \newcommand{\td}{\operatorname{td}}
37: \newcommand{\hF}{\operatorname{hF}}
38: \newcommand{\KH}{\mathrm{KH}}
39: \newcommand{\LC}{\mathrm{LC}}
40: \newcommand{\CLRep}{\mathrm{CLRep}}
41: \newcommand{\II}{\mathrm{II}}
42: \newcommand{\gen}{\mathrm{gen}}
43: % \newcommand{\can}{\mathrm{can}}
44: % \newcommand{\irr}{\mathrm{irr}}
45: \newcommand{\res}{\mathrm{res}}
46: % \newcommand{\red}{_{\mathrm{red}}} % reduced subscheme
47: % \newcommand{\mi}{_{\mathrm{min}}} % minimal surface
48: \newcommand{\rest}[1]{_{{\textstyle{|}}#1}} % restriction of map to
49:
50:
51: \newcommand{\op}{\overline{\partial}}
52: \newcommand{\ubar}{\overline{u}}
53: \newcommand{\Ibar}{\overline{I}}
54: \newcommand{\p}{\partial}
55: \newcommand{\w}{\widetilde}
56: \newcommand{\tensor}{\otimes}
57: \newcommand{\ov}{\overline}
58:
59: % Script letters
60: \newcommand{\sA}{\mathcal A} % sheaf of algebras A
61: \newcommand{\sB}{\mathcal B} % conductor ideal
62: \newcommand{\sG}{\mathcal G} % sheaf G
63: \newcommand{\sD}{\mathcal D}
64: \newcommand{\sJ}{\mathcal J} % sheaf J
65: \newcommand{\sI}{\mathcal I}
66: \newcommand{\sC}{\mathcal C}
67: \newcommand{\sU}{\mathcal U}
68: \newcommand{\sH}{\mathcal H}
69: \newcommand{\sK}{\mathcal K}
70: \newcommand{\sL}{\mathcal L}
71: \newcommand{\sM}{\mathcal M}
72: \newcommand{\Oh}{\mathcal O}
73: \newcommand{\sR}{\mathcal R}
74: \newcommand{\sS}{\mathcal S}
75: \newcommand{\sF}{\mathcal F}
76: \newcommand{\sE}{\mathcal E}
77: \newcommand{\sP}{\mathcal P}
78:
79: % short Greeks
80: \newcommand{\al}{\alpha}
81: \newcommand{\be}{\beta}
82: \newcommand{\de}{\delta}
83: \newcommand{\ep}{\varepsilon}
84: \newcommand{\fie}{\varphi}
85: \newcommand{\ga}{\gamma}
86: \newcommand{\ka}{\kappa}
87: \newcommand{\om}{\omega}
88: \newcommand{\si}{\sigma}
89: \newcommand{\De}{\Delta}
90: \newcommand{\Ga}{\Gamma}
91: \newcommand{\La}{\Lambda}
92: \newcommand{\Om}{\Omega}
93: \newcommand{\la}{\lambda}
94: \newcommand{\Si}{\Sigma}
95: \newcommand{\na}{\nabla}
96:
97: \newcommand{\PP}{\mathbb P}
98: \newcommand{\C}{\mathbb C}
99: \newcommand{\HH}{\mathbb H}
100: \newcommand{\Q}{\mathbb Q}
101: \newcommand{\R}{\mathbb R}
102: \newcommand{\Z}{\mathbb Z}
103:
104: % \mathops
105: \newcommand{\ad}{\operatorname{Ad}} % adjoint bundle
106: %\newcommand{\PP}{\operatorname{\mathcal P}}
107: % \newcommand{\alg}{\operatorname{alg}}
108: % \newcommand{\res}{\operatorname{res}}
109: \newcommand{\rk}{\operatorname{rank}}
110: \newcommand{\hcf}{\operatorname{hcf}}
111: \newcommand{\mir}{\operatorname{mir}}
112: \newcommand{\codim}{\operatorname{codim}} % codimension
113: % \newcommand{\diff}{\operatorname{diff}}
114: \newcommand{\topo}{\operatorname{top}}
115: % \newcommand{\var}{\operatorname{var}}
116: % \newcommand{\vcodim}{\operatorname{v.codim}}
117: \newcommand{\vdim}{\operatorname{v.dim}}
118: \newcommand{\Diff}{\operatorname{Diff}}
119: % \newcommand{\Ext}{\operatorname{Ext}}
120: % \newcommand{\Gr}{\operatorname{Gr}}
121: % \newcommand{\Grass}{\operatorname{Grass}} % Grassmann variety
122: % \newcommand{\Mod}{\operatorname{Mod}}
123: % \newcommand{\Mon}{\operatorname{Mon}}
124: \newcommand{\Pic}{\operatorname{Pic}} % Picard scheme
125: \newcommand{\Tr}{\operatorname{Tr}}
126: \newcommand{\End}{\operatorname{End}}
127: \newcommand{\GFT}{\operatorname{GFT}}
128: \newcommand{\sGFT}{\operatorname{sGFT}}
129: \newcommand{\LaGr}{\operatorname{\La_{\uparrow}\!}}
130: % \newcommand{\orGr}{\operatorname{Gr_{\uparrow}\!}}
131: \newcommand{\Hom}{\operatorname{Hom}}
132: % \newcommand{CLRep}{\operatorname{CLRep}}
133: \renewcommand{\Im}{\operatorname{Im}}
134: \renewcommand{\Re}{\operatorname{Re}}
135: \newcommand{\Sing}{\operatorname{Sing}}
136: % \newcommand{\Spec}{\operatorname{Spec}}
137: \newcommand{\Vol}{\operatorname{Vol}}
138: % \newcommand{\Stab}{\operatorname{Stab}} % Stabiliser group
139: \newcommand{\Conj}{\operatorname{Conj}}
140: \newcommand{\CL}{\operatorname{CL}}
141: % Lie groups
142: \newcommand{\SL}{\operatorname{SL}}
143: \newcommand{\SO}{\operatorname{SO}}
144: %\newcommand{U(1)}{\operatorname U(1)} % unitary group
145: \newcommand{\PU}{\operatorname{PU}}
146: \newcommand{\SU}{\operatorname{SU}}
147: \newcommand{\fsl}{\operatorname{\mathcal{\frak{sl}}}}
148: \newcommand{\fsu}{\operatorname{{\frak{su}}}}
149: % \newcommand{\Spin}{\operatorname{Spin}}
150: % \newcommand{\SP}{\operatorname{Sp}}
151: % \newcommand{\odd}{\mathrm{odd}}
152: \newcommand{\pt}{\mathrm{pt}}
153:
154: \renewcommand{\labelenumi}{(\arabic{enumi})}
155:
156: \begin{document}
157:
158: \title{Quantization, Classical and Quantum Field Theory and Theta-Functions.}
159:
160: \author{\fbox{Andrei Tyurin}}
161: \date{24.02.1940 -- 27.10.2002}
162:
163:
164: \maketitle
165:
166: \begin{center}
167: {\em To Igor Rostislavovich Shafarevich on his 80th birthday}
168: \end{center}
169: \bigskip
170:
171:
172: \section*{Introduction}
173:
174: Arnaud Beauville's survey "Vector bundles on Curves and Generalized
175: Theta functions: Recent Results and Open Problems"
176: \cite{Be} appeared 10 years ago. This elegant survey is short (16 pages) but
177: provides a complete introduction to a specific part of algebraic geometry.
178: To repeat his succes now we need more pages, even though we assume that
179: the reader is already acquainted with the material presented
180: there. Moreover, in Beauville's survey the relation
181: between generalized theta functions and conformal field theories (classical and
182: quantum) was presented already.
183:
184: Following Beauville's strategy we do not provide any proof or motivation.
185: But we would like to propose {\it all constructions} of
186: this large domain of mathematics in
187: such a way that the proofs can be guessed
188: from the geometric picture. Thus this text is not a mathematical monograph
189: yet, but rather a digest of a field of mathematical investigations.
190:
191: In the abelian case ( the
192: subject of several beautiful classical books (\cite{B},
193: \cite{C}, \cite{Wi}, \cite{F1} and many others ) fixing some
194: combinatorial structure
195: (a so called
196: theta structure of level $k$) one obtains special basis in the space
197: of sections of powers of the canonical polarization powers on jacobians.
198: These sections can be presented as holomorphic functions on the "abelian
199: Schottky" space $(\C^*)^g$.
200: This fact provides various applications of these concrete analytic formulas
201: to integrable
202: systems, classical mechanics and PDE's (see the references in \cite{DKN} ).
203:
204: Our practical goal
205: is to do the same in the non-abelian case, that is, to give the answer to
206: the final question of the Beauville's survey (Question 9 in
207: \cite{Be}).
208:
209: It has been observed many time
210: that the {\it construction} of theta functions with characteristics is
211: intricately
212: related to the paradigm of the quantization procedure (which is
213: a quantum field theory in dimension 1).
214: New features came from Conformal Field Theory (which is a field theory
215: in dimension 2 = 1+1).
216: This new stream brings the
217: standard physical paradigm of "symmetries, fields, equations etc ... and
218: gluing properties corresponding to local
219: Lagrangians". New CFT methods provided powerful computational tools while
220: "algebraic
221: geometers would have never dreamed of being able to perform such
222: computations" (A. Beauville).
223:
224: In future we hope to extend this digest to a mathematical monograph
225: with the title "VBAC".
226:
227: {\it Acknowledgments.} These notes were written for a series of lectures given at the Centre de Recherches Mathematiques at the Universite de Montreal in September 2002. I am grateful to J. McKay, J. Hurtubise, J. Harnad, D. Korotkin
228: and A. Kokotov who made my stay in Montreal very agreeable and productive. I would like to express my gratitude to
229: my collaborators C.~Florentino, J.~Mourao, J.P.~Nunes and
230: participants of Quantum Gravity seminar in IST (Lisboa, 2000). I would like to thank
231: Jean Le Tourneux and Andre Montpetit for their assistance in the preparation of these notes. Special thanks go to my daughter, Yulia Tiourina,
232: for catching numerous mistakes and misprints in this huge text.
233:
234: \tableofcontents
235:
236:
237: \section{Quantization procedure}
238:
239: \subsection{The Framework}
240:
241: The first question: what we want to quantize? The
242: framework of geometric quantization is usually described as
243: follows ( see, for example \cite{GS1}). Let $(M, \om)$ be a
244: symplectic manifold which represents a classical mechanical
245: system with finite number of degrees of freedom or the bialgebra
246: $C^\infty(M)$ with the usual commutative multiplication $f_1
247: \cdot f_2$ and Poisson structure given by well known construction:
248: every $f \in C^\infty(M)$ defines a Hamiltonian vector field
249: \begin{equation}
250: H_f = \om^{-1}(df)
251: \end{equation}
252: and a pair $(f_1, f_2)$ defines the function
253: \begin{equation}
254: \{f_1, f_2\} = \om(H_{f_1}, H_{f_2})
255: \end{equation}
256: -Poisson brackets, such that
257: \begin{equation}
258: d \om = 0 \implies \text{ Jacobi equality}.
259: \end{equation}
260:
261: If $M$ is compact, then even abelian multiplication is enough to
262: reconstruct $M$ and then the Lie algebra structure reconstructs
263: $\om$.
264:
265: So in this case the pair $(M, \om)$ and the (observables) algebra
266: $C^\infty(M)$ are equivalent objects.
267:
268: (Pre BRST)-quantization rules is a map $Q$ sending classical
269: objects to corresponding quantum objects:
270: \begin{enumerate}
271: \item $Q(M) = \sH$ is a Hilbert space (of wave functions);
272: \item $Q(f) \in Op(\sH)$ where $Op(\sH)$ is the space of
273: self adjoint operators
274: \item which
275: should satisfy the {\it correspondence principle}
276: \begin{equation}
277: [Q(f_1), Q(f_2)] = i \cdot h \cdot Q(\{ f_1, f_2 \})
278: \end{equation}
279: (Dirac correspondence) and the representation of our Poisson
280: Lie-algebra be irreducuble.
281: \end{enumerate}
282:
283: Unfortunately, such construction couldn't exist at all due to van
284: Hove theorem. Nevertheless one usually uses two basic examples:
285: Souriau - Kostant quantization doesn't satisfy the irreducibility
286: condition while Berezin quantization doesn't satisfy the
287: correspondence principle.
288:
289: But in both approaches we extend classical mechanical
290: date by
291: a {\it prequantization date} $(L, \nabla)$, where $L$ is a
292: complex line bundle (with a Hermitian structure $h$) and $\nabla$
293: is a unitary connection on $L$ with the curvature form
294: \begin{equation}
295: F_\nabla = 2 \pi \cdot i \cdot \om.
296: \end{equation}
297:
298: This equality implies very strong constraint to the
299: symplectic structure:
300: \begin{equation}
301: [\om] \in H^2(M, \Z)
302: \end{equation}
303: that is the cohomology class of the symplectic form has to be
304: integer.
305: Such quadruple
306: \begin{equation}
307: (M, \om, L, \nabla)
308: \end{equation}
309: is ready to be quantized.
310: First of all the space of wave functions is the space of sections
311: \begin{equation}
312: \sH = \Ga^\infty (L)
313: \end{equation}
314: with the inner product given by the formula
315: \begin{equation}
316: <s, s'> = 1/n! \int_M (s, s')_h \om^{1/2 dim M},
317: \end{equation}
318: (usually one extends this space to $L^2$ - sections). For every
319: function (a classical observable) we use the operator
320: \begin{equation}
321: Q(f) = \nabla_{H_f} + 2 \pi i \cdot f,
322: \end{equation}
323: that is
324: $$
325: Q(f)(s) = \nabla_{H_f} s + 2 \pi i \cdot f \cdot s.
326: $$
327: Two simple exercises:
328: \begin{enumerate}
329: \item Dirac equality holds (2.4);
330: \item the representation of
331: our Poisson Lie-algebra is very far from being irreducible (even
332: for the simplest case $M = \R^2 = T^{*} \R)$
333: \end{enumerate}
334:
335: %\begin{figure}[tbn]
336: %\centerline{\epsfxsize=2in\epsfbox{18.eps}}
337: %\caption{\sl complete integrable}
338: %\label{F1}
339: %\end{figure}
340:
341: Well known remedy is a choice of a {\it polarization}. A
342: polarization is a distribution that is a sub bundle of the
343: complexified tangent space
344: \begin{equation}
345: F \subset T^\C M , \quad rk F = 1/2 dim M;
346: \end{equation}
347:
348:
349: Fixing such polarization the wave function space can be obtained
350: as
351: \begin{equation}
352: Q(M)= \sH^F = \{ s \in \Ga^\infty (L) \vert \nabla_F s = 0 \}.
353: \end{equation}
354: There are two natural choices:
355: \begin{enumerate}
356: \item $\overline{F} = F$ is a {\it real polarization};
357: \item $\overline{F} = F^\perp $ is a
358: {\it complex polarization}, that is, a choice of an almost complex
359: structure $I$ such that $(F = T^{0,1})$.
360: \end{enumerate}
361:
362: Of course for the general choice of a polarization we don't know
363: so much but there is a couple of good choices:
364:
365: in the real case: choose $F$ integrable and moreover completely
366: integrable that is defined by a Lagrangian fibration:
367: \begin{equation}
368: \pi \colon M \to B , \quad \pi^{-1}(b_{gen}) = T^n
369: \end{equation}
370: that is, the generic fiber here is a Lagrangian torus of
371: dimension $n = 1/2 dim M$ (see Fig.1);
372:
373: and in the complex case: choose an
374: almost complex structure which is integrable and compatible with
375: $\om $ such that the pair $(\om, I)$ gives a Kahler metric with
376: condition (1.6) that is a Hodge metric.
377:
378:
379: \begin{figure}[tbn]
380: \centerline{\epsfxsize=3in\epsfbox{22c.eps}}
381: \caption{\sl Completely integrable system}
382: \label{F2}
383: \end{figure}
384:
385:
386: There is a couple of equivalent conditions for both of these
387: cases: for a real polarization, for any two vector fields
388: \begin{equation}
389: \p_1, \p_2 \in F \subset T^\C M
390: \end{equation}
391: the commutator
392: \begin{equation}
393: [\p_1, \p_2] \in F \subset T^\C M
394: \end{equation}
395: i.e. vector fields from our real distribution form a Lie
396: subalgebra.
397: (Otherwise their commutators form a quadratic form on $ F$ with
398: coefficients in $T^\C M / F$ just like the second quadratic
399: form in the standard differential geometry of a surface
400: in $\R^3$ ).
401:
402: For the complex case we have to add the equality of the
403: same nature
404: \begin{equation}
405: [I \p_1, I \p_2] = [\p_1, \p_2] + I [I \p_1, \p_2]
406: + I [\p_1, I \p_2]
407: \end{equation}
408: where we consider $I$ as a section of $End TM$ with
409: \begin{equation}
410: I^2 = -1
411: \end{equation}
412: $$
413: \om(\p_1, I \p_2) = - \om (I \p_1, \p_2)
414: $$
415: $$
416: \om(\p_1, I \p_2) > 0.
417: $$
418:
419:
420: These cases are lying in different domains of mathematics: the
421: real polarizations are contained in Symplectic Geometry and the
422: complex polarization belongs to Algebraic Geometry. A
423: quantization is {\it perfect} if it admits both of these
424: polarizations. So the perfect quantization belongs to the union of
425: Algebraic and Symplectic Geometries.
426:
427:
428: \subsection{The real polarization.(Symplectic Geometry)}
429:
430:
431:
432:
433: Recall that a subcycle $T \subset M$ is Lagrangian if it is a
434: middle dimensional subcycle such that
435: \begin{equation}
436: \om \vert_T = 0.
437: \end{equation}
438: Thus the restriction of the pair $(L, \nabla)\vert_T$ is the
439: trivial
440: line bundle with a flat connection
441: \begin{equation}
442: F_\nabla \vert_T = 2 \pi i \om \vert_T = 0.
443: \end{equation}
444: Recall that a gauge class of a flat connection is given by
445: a character
446: of the fundamental group of $T$
447: \begin{equation}
448: \chi \colon \pi_1(T) \to U(1).
449: \end{equation}
450:
451:
452: \begin{dfn} A Lagrangian cycle $T$ is Bohr-Sommerfeld
453: if $\chi = 1$, that is, if there exists a covariant constant
454: section of $(L, \nabla)\vert_T$.
455: \end{dfn}
456: To kill a character, we need $rk H_1(T, \Z)$ conditions
457: so we may expect that
458: in a family
459: of Lagrangian cycles the subfamily of Bohr-Sommerfeld cycles has
460: this codimension. Moreover, it was shown
461: that in any small
462: Darboux-Weinstein
463: neighborhood of a smooth
464: Bohr-Sommerfeld cycle $T$ every other Bohr-Sommerfeld cycle $T'$
465: intersects
466: $T$.
467:
468: For example, our real polarization (1.13) defines a family
469: of Lagrangian
470: cycles parameterized by a space $B$ of dimension $n = 1/2 \dim M$.
471: But a general fiber of any real polarization fibration $\pi$
472: (1.13) is a n-dimensional torus, that is
473: $$
474: rk H_1(T^n) = n = dim B.
475: $$
476: Thus the set of Bohr-Sommerfeld fibers is discrete and if $B$ is
477: compact it forms
478: a finite set $BS \subset B$ of Bohr-Sommerfeld fibers of $\pi$.
479:
480: Then the Hilbert space of wave functions is given as
481: the direct sum of lines
482: \begin{equation}
483: Q_F(M) = \sH_\pi = \oplus_{b \in BS \subset B} \C \cdot
484: s_b
485: \end{equation}
486: where $s_b$ is a covariant constant section of $(L,
487: \nabla)\vert_{\pi^{-1}(b)}$).
488:
489: %\begin{figure}[tbn]
490: %\centerline{\epsfxsize=2in\epsfbox{23.eps}}
491: %\caption{\sl twice integrable system}
492: %\label{F4}
493: %\end{figure}
494:
495: Recall that every covariant constant section is defined up
496: to $U(1)$-scaling,
497: so we can define naturally an Hermitian scalar product on this space.
498:
499: We can extend the notion of
500: Bohr-Sommerfeld Lagrangian cycles:
501:
502: \begin{dfn} A Lagrangian cycle $T$ is Bohr-Sommerfeld of level $k$ (or $BS_k$ - cycle)
503: if $\chi^k = 1$, that is, if there
504: exists a covariant constant section of $(L^k, \nabla_k)\vert_T$,
505: where $\nabla_k $ is the connection generated by $\nabla$ on the
506: tensor power $L^k$ of the line bundle $L$.
507: \end{dfn}
508:
509: Reasoning the same way as before, we can prove that the subset
510: $BS_k \subset B$ is finite and that the Hilbert space of wave
511: functions of level $k$ is given by the direct sum of lines
512: \begin{equation}
513: Q_F^k(M) = \sH_\pi^k = \oplus_{b \in BS_k \subset B} \C \cdot
514: s_b
515: \end{equation}
516: where $s_b$ is a covariant constant section of $(L^k,
517: \nabla_k)\vert_{\pi^{-1}(b)}$).
518:
519: But we are quantizing a classical dynamical system $(M, \om)$!
520: So the result does not have to depend on additional data like
521: $(L, \nabla, \pi, I)$ and so on.
522:
523: In our construction the wave space $\sH_\pi$ depends on a gauge
524: class of a connection $\nabla$ only and this class is given by
525: the curvature form if $M$ is simply connected (see \cite{AB}).
526: (Recall that in non-simple connected case
527: we have to fix additionally a point of the
528: "jacobian" $H^1(M, \R) / H^1(M, \Z)$ of $M$ (see below).
529:
530: In any case
531: our wave function spaces depend up to a shift of characters on the
532: curvature tensor which is $\om$ and on a choice of a real
533: polarization $\pi$.
534:
535: \begin{figure}[tbn]
536: \centerline{\epsfxsize=3in\epsfbox{23c.eps}}
537: \caption{\sl twice integrable system}
538: \label{F4}
539: \end{figure}
540:
541:
542: But the most important thing is that the projectivization of the
543: wave functions
544: space doesn't depend on a real polarization $\pi$ too.
545: For any other Lagrangian fibration $\pi' \colon M \to B'$,
546: Kostant defined the canonical (up to constant )
547: pairing between
548: $\sH_\pi$ and $\sH_{\pi'}$ (see for example \cite{JW2}). So we
549: get the identification
550: \begin{equation}
551: \PP \sH_\pi = \PP \sH_{\pi'}
552: \end{equation}
553: of projective spaces.
554:
555: Thus $\sH_\pi$ is the wave space of a quantization of $(M, \om)$
556: and of nothing else.
557:
558: \subsection{Kahler quantization. (Algebraic Geometry) }
559:
560: In this case we have a quadruple $(M, \om, \nabla, I)$ where a
561: complex structure $I$ is compatible with $\om$. Thus the pair
562: $(\om, I)$ defines a Kahler metric with the Kahler form $\om$.
563: Moreover, $\om$ is of Hodge type (1,1). Thus the connection
564: $\nabla $ on $L$ has the curvature form of type (1,1) and defines
565: a holomorphic structure on $L$ (see for, example, \cite{GH}).
566: As
567: Kodaira (1958) proved, a Kahler metric with integer cohomology
568: class $[\om] \in H^2(M, \Z)$ of the Kahler form is a Hodge
569: metric, hence $M_I$ is a projective algebraic variety and $L =
570: \Oh_M (1)$ is the class of hyperplane section of $M_I$.
571:
572: So, the complex polarizations bring us to Algebraic Geometry!
573:
574:
575: Naturally, the wave function space
576: \begin{equation}
577: Q_F(M) = \sH_I = H^0(M_I, L)
578: \end{equation}
579: is the space of holomorphic sections of the polarization $L$ and
580: the projectivization $\PP H^0(M_I, L)$ is the classical
581: complete linear system on $M_I$.
582:
583: The Hermitian structure on $L$ gives an Hermitian structure on
584: $H^0(M_I, L)$ and provides the identification
585: \begin{equation}
586: \sH_I = H^0(M_I, L) = H^0(M_I, L)^*
587: \end{equation}
588: with changing of the complex structure to the complex
589: conjugate.
590:
591: The standard construction of the Algebraic Geometry gives the
592: following map
593: \begin{equation}
594: \phi \colon M_I \to \PP H^0(M_I, L)^*,
595: \end{equation}
596: that is, under the previous identification
597: \begin{equation}
598: \ov{\phi} \colon M_I \to \PP H^0(M_I, L).
599: \end{equation}
600:
601: Images of points of $M_I$ are called {\it coherent states}. A
602: priori such image is defined up to a transformation
603: from $PGL( \PP H^0(M_I, L))$. But its every
604: {\it stable} orbit contains unique
605: unitary orbit preserving our form $\om$. If our form $\om$ is the
606: restriction of
607: the Fubini-Study form
608: from $ \PP H^0(M_I, L)$, then we can switch on to Geometric
609: Invariant Theory and reduce the theory of coherent states to it.
610: So in algebraic geometrical set up we can assume that our form is
611: a restriction of the Fubini-Study form. In the general case we
612: have to be more careful and to add so called J.Rawnsley
613: $\ep$-function of
614: (see \cite{R}).
615:
616: Of course we lose the Dirac correspondence and have to use the
617: Berezin-Toeplitz quantization rules (see for example \cite{R}).
618: Thus we introduce the level parameter $k \in \Z^+$. The space of
619: wave function of level $k$ is the space of holomorphic sections
620: \begin{equation}
621: \sH_I^k = H^0(M_I, L^k).
622: \end{equation}
623: The Dirac correspondence can be restored in the
624: {\it quasi classical limit} $k \to \infty$ or $1/k \to 0$.
625: So, our complex polarization becomes a polarization in
626: algebraic geometrical sence. Recall that if $k$ is large enough,
627: then $\ov{\phi}$ (1.27) is an embedding and the dimension of the
628: corresponding projective space is given by the Riemann-Roch
629: formula as a pure topological invariant.
630:
631: So, every underlying symplectic manifold $(M, \om)$ of
632: a polarized algebraic variety $M_I, \Oh(1)$ with a
633: natural complex polarization $\Oh(1)$ gives the collection
634: of wave functions
635: spaces
636: $H^0(M_I, L^k) = H^0(M_{I}, \Oh(k))$
637:
638: But the quantization set up disturbs
639: the Algebraic Geometry Peace.
640: The main question is:
641:
642: do these spaces appear as a result of a successful quantization
643: procedure of a classical system given by the underlying
644: symplectic manifold $(M, \om)$?
645:
646: That is,
647:
648: is the projectivization of this space independent on the complex
649: structure $I$ ?
650:
651: (In physical symbols, $\frac{\p \PP H^0(M_I, L)}{\p I} = 0$ .)
652:
653: Mathematically correct question is the following:
654: let $ \sM $ be the moduli space of such complex structures, that
655: is, the moduli space of polarized (in the algebraic geometryical
656: sense) varieties and
657: \begin{equation}
658: p \colon V_k \to \sM, \quad p^{-1}(M_I) = H^0(M_I, L^k)
659: \end{equation}
660: - a vector bundle of holomorphic sections of polarization line
661: bundles. Note that by the Kodaira vanishing theorem it is a vector
662: bundle indeed! Then
663: there exists a holomorphic projective flat connection on $V$.
664:
665: Existence of such connection constraints the topology type of the
666: vector bundle: if a vector bundle $V$ admits a projective flat
667: connection, then
668: \begin{equation}
669: c_t(V) = (1 - c \cdot t)^{rk V}.
670: \end{equation}
671: In particular,
672: \begin{equation}
673: c_1(V) = rk V \cdot c, \quad c \in H^2(\sM, \Z).
674: \end{equation}
675: This is very strong topological constraint!
676:
677: Thus problems of new type are
678: given inside of Algebraic Geometry:
679:
680: When a pair $M_I, \Oh(1)$ of an algebraic variety with
681: a polarization is a result
682: of a quantization procedure
683: for the underlying symplectic manifold ?
684:
685: Before considering the simplest example when $\dim M = 2$ we give
686: some general remarks:
687: \begin{enumerate}
688: \item of course our vector bundles $V_k$ (1.29) are
689: holomorphic but the required projective flat connection could be
690: non holomorphic;
691: \item so, it is quite reasonable to construct may be not
692: holomorphic projective Hermitian connection;
693: \item in all classical examples
694: holomorphic projective flat connections are Hermitian by the
695: construction;
696: \item but in the last non-abelian case (which is of the main
697: interest here) this fact isn't known
698: up to now;
699: \item if such new connection exists automatically we get a Higgs field
700: on every vector bundle (1.29) on moduli spaces of complex
701: structures.
702: \item In the physical set up it may be that we do not obtain vector
703: bundles (1.29) but have the collection of projective bundles
704: which {\it aren't projectivizations of vector bundles};
705: \item even in the infinite dimensional case (when Hilbert vector
706: bundles are always trivial) the projective bundles can be non trivial
707: and are given by a cohomology class from $H^3(\sM, \Z)$. This
708: Brauer group gives us non trivial analog of Atiyah's K-functor.
709: \end{enumerate}
710:
711: In the conformal field theory context this connection always {\it
712: has to be holomorphic} but we do not have space and time to
713: discuss it here.
714:
715:
716: Because of the time-space problem this very interesting subject
717: is out of our considerations here.
718:
719:
720: \subsection{Extended Kodaira-Spencer theory}
721:
722: The theory of deformations of complex structures was developed by
723: Kodaira and Spencer many years ago. Trying "theoretically" to
724: solve the
725: problem of successful
726: Kahler quantization we have to lift this theory to the problem
727: of
728: holomorphic deformations of a triple $(I, L_I, (s)_0)$,
729: where $I$ is an integrable complex structure, $L_I$ is
730: our holomorphic line bundle and $s \in H^0(L_I)$ is
731: zero divisor of a holomorphic section of this line bundle.
732: The best reference for such lifting is Welters paper
733: \cite{W}. Recall that the holomorphic structure of $L_I$
734: codes the connection $\nabla$ from the quadruple $(1.7)$ by the decomposition
735: \begin{equation}
736: \nabla = \nabla^{1, 0} + \nabla^{0, 1}
737: \end{equation}
738: where
739: \begin{equation}
740: \nabla^{1, 0} = 1/2 \cdot(1 - i I) \nabla, \quad \nabla^{0, 1} =
741: 1/2 \cdot (1 + i I) \nabla.
742: \end{equation}
743: Our history starts with an infinitesimal deformation of the
744: complex structure $\dot I$. Since it is an anti involution $I^2 =
745: -1$ we have the equality
746: \begin{equation}
747: \dot I I + I \dot I = 0
748: \end{equation}
749: and one can consider $\dot I$ as a projection of $-i$ eigenspace
750: of $I$ to the $+i$-eigenspace. That is
751: \begin{equation}
752: \dot I \in \Om^{0, 1} (M_I,T^{1, 0})
753: \end{equation}
754: (just as usual Beltrami differential).
755:
756: Linearization of the integrability condition (1.16) gives us
757: \begin{equation}
758: (\op \dot I)^{2, 0} = 0,
759: \end{equation}
760: that is, the (0, 2)-component of this tensor is trivial.
761:
762: Suppose that we have the following list of properties.
763: \begin{enumerate} \item
764: Our polarization $L$ is a rational part of the canonical class $K$
765: of $M_I$. That is
766: \begin{equation}
767: - [K] = n [c_1(L)].
768: \end{equation}
769: \item Further on we suppose that the Levi-Civita connection of
770: the Kahler
771: metric $\om(\quad , I \quad )$ (1.17) on $K $ is proportional to
772: the unitary connection on $L$.
773: \item We consider a sufficiently large level $k$ such that
774: \begin{equation}
775: i >o \implies H^i(M_I, \Oh(k \cdot c_1(L) + 1/2 K)) = 0.
776: \end{equation}
777: (For example, $M_I$ is a Fano-variety of index $n$ and
778: $L = \Oh(k)$).
779: \item Then we have the symplectic Kahler form
780: \begin{equation}
781: ( k + n/2) \om.
782: \end{equation}
783: \end{enumerate}
784:
785: From this it is easy to see (for example, by local
786: representation of our tensor) that
787: there exists a section
788: \begin{equation}
789: W_0 \in \Ga^\infty (S^2 T^{1, 0}M_I) \in (S^2)^{2, 0}
790: \end{equation}
791: where $S^2$ is the second symmetrical power of the
792: holomorphic tangent bundle such that the convolution of
793: this tensor with our symplectic form gives us $\dot I$:
794: \begin{equation}
795: W_0 * (k + n/2) \om = \dot I.
796: \end{equation}
797: (Recall that $\om$ is of a type $(1.1)$ with respect to the
798: complex structure $I$.)
799:
800: To see this very important tensor we have to represent tensors
801: as homomorphisms of tensors vector bundles!
802: By this notation we
803: emphasize
804: that this construction is due to Welters \cite{W}.
805: So our symplectic form
806: \begin{equation}
807: (k+n/2)\om \colon T^{1, 0} \oplus T^{0, 1} \to \Om^{1, 0} \oplus \Om^{0, 1}
808: \end{equation}
809: is of the Hodge type $(1,1)$ and it means that
810: \begin{equation}
811: \om(T^{1, 0}) \in \Om^{0,1} \text{ and } \om(T^{0, 1}) \in \Om^{1,0}.
812: \end{equation}
813: On the other hand,
814: \begin{equation}
815: \dot I \colon T^{0,1} \to T^{1,0}.
816: \end{equation}
817: Thus the composition of these homomorphisms
818: \begin{equation}
819: \dot I \circ \frac{1}{k+n/2} \om^{-1} = W \colon \Om^{1,0} \to T^{1,0}
820: \end{equation}
821: is the very one symmetric tensor.
822:
823:
824:
825:
826: Now consider the cup product map
827: \begin{equation}
828: H^0(M_I, T^{1, 0}) \otimes H^1(M_I, \Om^{0, 1}) \to H^1(M_I, \Oh)
829: \end{equation}
830: and the restriction of it to
831: $H^0(M_I, T^{1, 0}) \otimes [(k+n/2)\om]$:
832: \begin{equation}
833: [(k+n/2)\om] \colon H^0(M_I, T^{1, 0}) \to H^1(M_I, \Oh).
834: \end{equation}
835: Suppose that this in an {\it isomorphism}. For example
836: \begin{equation}
837: H^0(M_I, T^{1, 0}) = H^1(M_I, \Oh) = 0
838: \end{equation}
839: or $M_I$ is a polarized abelian variety.
840:
841: In this case our holomorphic line bundle $L_I$ can't be deformed
842: continuously
843: under deformation of the complex structure.
844: So we have to investigat deformations of
845: holomorphic sections only.
846:
847:
848: Having such symmetric form and following Nigel Hitchin let us
849: infinitesimally deform a section $s \in H^0(M_I, L_I)$ to a $\dot
850: I$-holomorphic section. Differentiating along a "time" $t$ the
851: condition $(1 + i I_t) \nabla s_t = 0$ we get the equality
852: \begin{equation}
853: i \dot I \nabla^{1,0} s = \nabla^{0,1} a
854: \end{equation}
855: where $-a$ is a section of $L$. Such section depends linearly on
856: $s$ and $\dot I$ and defines a connection on the bundle $V$
857: (1.29).
858: The main observation of
859: the extended Kodaira-Spencer theory is the fact that
860: this equation admits
861: a cohomological interpretation. Namely LHS and RHS of (1.49) are
862: results of applying first order
863: differential
864: operators on $L$. So now we have to consider the vector
865: bundle $D^1(L_I)$ of
866: holomorphic differential operators on $L_I$ applied to
867: a holomorphic section
868: in LHS and to the potential of a connection.
869:
870: Let the exact sequence
871: \begin{equation}
872: 0 \to \Om^{1,0} \otimes L \to J^1(L) \to L \to 0
873: \end{equation}
874: be the extension defining 1-jet bundle of our line bundle $L$. By
875: the Atiyah (see \cite{A2}) theorem, the cocycle of this extension
876: is
877: \begin{equation}
878: [c_1(L)] = [\om] \in H^1(M_I, \Om^{1,0}).
879: \end{equation}
880:
881: Now
882: \begin{equation}
883: D^1(L) = Hom (J^1(L), L )
884: \end{equation}
885: is the bundle of the sheaf of germs of differential operators
886: on $L$ of order
887: $\leq 1$. So this sheaf is the extension
888: \begin{equation}
889: 0 \to \Oh \to D^1(L) \to T^{1, 0} \to 0
890: \end{equation}
891: (given by the class $- [\om]$).
892:
893: The long exact cohomology sequence of this exact triple
894: decays in two parts
895: \begin{equation}
896: 0 \to H^0(M_I, \Oh) \to H^0(M_I,D^1(L)) \to 0
897: \end{equation}
898: and
899: \begin{equation}
900: 0 \to H^1(M_I, D^1(L)) \to H^1(M_I, T^{1, 0}) \to H^2(M_I, \Oh)
901: \dots
902: \end{equation}
903: because they are divided by the homomorphism (1.44).
904: The first exact sequence states
905: that an every globally defined first order holomorphic operator
906: on $L$ is
907: a {\it multiplication by a constant} and the second states
908: that the {\it
909: symbol map} $H^1(M_I, D^1(L)) \to H^1(M_I, T^{1,0})$ is a
910: { \it monomorphism}.
911:
912: Now for a holomorphic section $s \in H^0(M_I, L_I)$ the
913: valuation morphism
914: \begin{equation}
915: d_s \colon D^1(L) \to L , \quad d_s (D) = D s,
916: \end{equation}
917: gives the complex
918: \begin{equation}
919: 0 \to D^1(L) \to L \to 0
920: \end{equation}
921: which hypercohomology $H_h^1(d_s)$ defines linear infinitesimal
922: deformations of a triple $(\dot I, L_I, s)$.
923:
924: Each class of such hypercohomology can be done as a pair
925: \begin{equation}
926: (\{ \dot s_i\}, \{\al_{i,j} \}) \in C^0(\sU, L) \oplus C^1(\sU, D^1(L))
927: \end{equation}
928: where $\sU$ is a Cech covering. Indeed it is a 1-cocycle of the
929: total complex, associated with the double complex
930: \begin{equation}
931: \pm d_s \colon C^i(\sU, D^1(L)) \to C^i(\sU, L).
932: \end{equation}
933: The spectral sequence of hypercohomology (of a double complex)
934: yields the exact sequence
935: \begin{equation}
936: H^0(M_I, D^1(L)) = \C \to H^0(L_I) \to H_h^1(d_s) \to H^1(D^1(L)) \to H^1(L_I)
937: \end{equation}
938: corresponding to the pairs (1.56).
939:
940: On the other hand, $d_s$-complex (1.57) defines the exact
941: quadruple
942: \begin{equation}
943: 0 \to ker_s \to D^1(L) \to L \to coker_s \to 0
944: \end{equation}
945: and the second spectral sequence gives
946: \begin{equation}
947: 0 \to H^1(ker) \to H_h^1(d_s) \to H^0(coker) \to H^2(ker).
948: \end{equation}
949: It is easy to see (this observation due to Welters \cite{W})
950: that the support
951: of the sheaf $coker_s$ can be very small:
952: \begin{equation}
953: Supp F = Sing (s)_0
954: \end{equation}
955: -the singularity locus of the zero-divisor of a section $s$. In
956: particular, if a zero divisor
957: is smooth (this is a general case)
958: then $F=0$ and the exact quadruple (1.60) is in fact a triple.
959:
960: Now let us return to our main equation (1.49).
961: The RHS admits a solution if the
962: \begin{equation}
963: (\op LHS )^{0,2} = 0
964: \end{equation}
965: (by the Dolbeault lemma). So we have to compute this RHS Hodge
966: type. But by ( 1.41 ) it is easy to see that
967: \begin{equation}
968: (\op (i \dot I \nabla^{1,0}))^{0,2} = 0.
969: \end{equation}
970: Here we are using
971: \begin{equation}
972: i \dot I \nabla^{1,0} \in \Om^{0,1} (M_I, D^1(L))
973: \end{equation}
974: and the infinitesimal integrability condition (1.34).
975: Now we have the interpretations of $i \dot I \nabla^{1,0}$ and
976: $-a$ from (1.49)
977: as (0, 1) - forms with coefficients in $D^1(L)$ and we have
978: the complex
979: \begin{equation}
980: C^i = \Om^{0,i} (M_I, D^1(L)) \oplus \Om^{0, i-1}(M_I, L_I)
981: \end{equation}
982: with the differential
983: \begin{equation}
984: d(s) (D, a) = (\op D, \op a + (-1)^{i-1} D s)
985: \end{equation}
986: Now
987: \begin{equation}
988: \op s = 0 \implies d(s)^2 = 0
989: \end{equation}
990: and our pair
991: \begin{equation}
992: (i \dot I \nabla^{1,0}, -a) \in C^1.
993: \end{equation}
994: Moreover, equations (1.65) and (1.49) give
995: \begin{equation}
996: d(s) (i \dot I \nabla^{1,0}, -a) = 0.
997: \end{equation}
998: Thus, every solution to (1.49) defines a hypercohomology class
999: in $H_h^1(d_s)$!
1000:
1001: Now suppose we can lift every pair
1002: \begin{equation}
1003: (\dot I, s) \in \Om^{0,1}(M_I, T^{1,0}) \oplus H^0(M_I, L_I)
1004: \end{equation}
1005: to a cocycle from $H_h^1(d_s)$ such that the natural "symbol" map
1006: \begin{equation}
1007: \si \colon H_h^1(d_s) \to H^1(M_I, T^{1,0})
1008: \end{equation}
1009: sends our pair to the cohomology class $[\dot I] \in H^1(M_I, T^{1,0})$.
1010: That is there exists a map
1011: \begin{equation}
1012: A \colon \Om^{0,1}(M_I, T^{1,0}) \oplus H^0(M_I, L_I) \to H_h^1(d_s)
1013: \end{equation}
1014: such that the composition $\si A$ is the map to the cohomology class.
1015: Then such map defines a projective connection on the vector bundle
1016: of holomorphic sections (1.29).
1017:
1018: Indeed, we can presents a hypercohomology class as a pair (1.58)
1019: \begin{equation}
1020: (D, s') \in \Om^{0,1} (M_I, D^1(L)) \oplus \Om^{0, 0}(M_I, L_I).
1021: \end{equation}
1022: But now the symbols of operators $-i \si (D)$ and $i \dot I
1023: \nabla^{1,0}$ are cohomological and, therefore the corresponding
1024: operators are cohomological in $\Om^{0,1}(M_I, D^1(L))$ because
1025: map to symbols is a monomorphism (1.55). Thus there exists a
1026: global operator
1027: $\sD \Om^0(M_I, D^1(L))$ such that
1028: \begin{equation}
1029: D - i \dot I \nabla^{1,0} = \op \sD.
1030: \end{equation}
1031: Since $(D, s')$ is a cocycle ($d(s)(D, s') = 0 $, (see (1.71))
1032: one gets
1033: \begin{equation}
1034: Ds + \op s' = i \dot I \nabla^{1,0} s + \op (\sD s + s').
1035: \end{equation}
1036: Thus
1037: \begin{equation}
1038: -a = \sD s + s'
1039: \end{equation}
1040: is a solution for (1.49).
1041:
1042: Now for two such solutions hypercocycles
1043: $(i \dot I \nabla^{1,0}, a_1)$
1044: and $(i \dot I \nabla^{1,0}, a_2)$ are cohomological and
1045: the hypercocycle
1046: \begin{equation}
1047: (0, a_1 - a_2) = d(s) \sD'.
1048: \end{equation}
1049: But from the definition of the complex (1.59) we have
1050: \begin{equation}
1051: \op \sD' = 0 \implies a_1 - a_2 = \la s
1052: \end{equation}
1053: since every first order holomorphic operator is the multiplication
1054: by a constant. Geometrically this means that we get a holomorphic
1055: connection on
1056: the projectivization of our vector bundle.
1057:
1058: Thus our last task is to construct the lifting $A$ (1.74).
1059: But there is a canonical way to associate such lifting
1060: with a holomorphic symmetrical tensor
1061: $W$ (1.40).
1062:
1063: Consider the sheaf $D^2(L)$ of second order holomorphic
1064: operators given by the
1065: standart extension
1066: \begin{equation}
1067: 0 \to D^1(L) \to D^2(L) \to S^2(T^{1,0}) \to 0.
1068: \end{equation}
1069: Evaluating differential operators on a given section $s \in
1070: H^0(M_I, L_I)$ we get two homomorphisms
1071: \begin{equation}
1072: d_s \colon D^1(L) \to L
1073: \end{equation}
1074: $$
1075: d'_s \colon D^2(L) \to L
1076: $$
1077: giving two complexes: our complex (1.57) and its analog
1078: \begin{equation}
1079: 0 \to D^2(L) \to L \to 0.
1080: \end{equation}
1081: Considering the "trivial" third complex
1082: \begin{equation}
1083: 0 \to S^2(T^{1,0}) \to 0 \to 0
1084: \end{equation}
1085: we get the homomorphism of the vertical triples to the triple
1086: $$
1087: 0 \to L \to L \to 0 \to 0
1088: $$
1089: and the exact triple of these complexes. The corresponding exact
1090: sequence of hypercohomology has the form
1091: \begin{equation}
1092: 0 \to H_h^0(d_s) \to H_h^0(d'_s) \to
1093: \end{equation}
1094: $$
1095: \to H^0(M_I, S^2(T^{1,0})) \to H_h^1(d_s) \to H_h^1(d'_s) \to \dots
1096: $$
1097: and every holomorphic quadratic form on the tangent bundle defines
1098: a 1-hypercohomology class (which may be zero). Of course our
1099: required map depends on a choice of a section $s$, so we denote
1100: this map by the symbol
1101: \begin{equation}
1102: \phi_s \colon H^0(M_I, S^2(T^{1,0})) \to H_h^1(d_s).
1103: \end{equation}
1104: The composition of this map with the forgetful homomorphism $\si$
1105: (1.73) gives the homomorphism
1106: \begin{equation}
1107: \mu_L \colon H^0(M_I, S^2(T^{1,0})) \to H^1(M_I, T^{1,0})
1108: \end{equation}
1109: which doesn't depend on the choice of $s$. More precisely
1110: \begin{equation}
1111: \mu_L (W_0) = -W * [c_1(L)] + \mu_{\Oh}
1112: \end{equation}
1113: It is quite natural to call it the "cohomological heat equation".
1114: For example
1115: for polarized abelian varieties the map $\mu_\Oh = 0$. Moreover,
1116: all extensions $D^n(\Oh)$ are trivial hence all
1117: coboundary homomorphisms are trivial.
1118:
1119:
1120: In particular, from this equality and (1.41) we can see that
1121: \begin{equation}
1122: \mu_\Oh (W_0) = - W_0 * 1/2 K
1123: \end{equation}
1124: because we can say more about sheaves of differential operators
1125: on the trivial line bundle. Namely the extension class for all
1126: jet-bundles is equal to
1127: \begin{equation}
1128: 1/2[K_{M_I}] \in \Om^{1,1}.
1129: \end{equation}
1130: From this it is easy to see that
1131: \begin{equation}
1132: \mu_L (W_0) = -W_0 * ([c_1(L)] + 1/2 [K_{M_I}])
1133: \end{equation}
1134: where $K_{M_I}$ is the canonical class of $M_I$ and the
1135: cohomology class
1136: $[K_{M_I}]$ depends on $\om$ only (and not on the choice of an
1137: admissible complex
1138: structure).
1139:
1140: Of course we can get more precise formula using representatives
1141: of classes by
1142: forms and using the Levi-Civita connections for tensor bundles
1143: (see \cite{H1})
1144:
1145:
1146:
1147: So the extended Kodaira-Spencer theory gives us (canonically) a
1148: holomorphic projective connection on the vector bundles (1.29) but
1149: of course to compute a curvature of one we have to present it as
1150: a Dolbeault or Cech class.
1151:
1152: An existence of Welters tensor $W_0$ (1.40) gives strong geometrical
1153: constraints. In particular $M_I$ can't be a variety of general
1154: type. More precisely the canonical class $K$ of $M_I$ can't be
1155: positive.
1156:
1157: \subsection{Faithful functors}
1158:
1159: Suppose our $M_I$ which we consider as a point $I$ of
1160: the moduli space $\sM$
1161: of all its deformation defines a new polarized algebraic variety
1162: $M'_{I'}$ which also we consider as a point $I'$ of the
1163: moduli space $\sM'$
1164: such a way, that
1165: we can reconstruct $M_I$ (from the geometry of $M'_{I'}$).
1166: Geometrically, mapping $M_I$ to $M'_{I'}$ defines a holomorphic
1167: embedding of moduli spaces
1168: \begin{equation}
1169: J \colon \sM \to \sM',
1170: \end{equation}
1171: and $J(I) = I'$. If a morphism of polarized varities from $\sM $
1172: induces a morphism of corresponding varieties from $\sM'$ then
1173: the map $M_I \to M'_{I'}$ is called a {\it faithful
1174: functor}. From geometrical point of view these geometrical
1175: images are undistinguishable in spite of the fact
1176: that these varieties can have different dimensions and so on.
1177: An algebraic geometer is skillful enough if he can recognize
1178: such geometrically undistinguishable varieties of different
1179: dimensions and different shapes.
1180:
1181: For example, a smooth intersection
1182: \begin{equation}
1183: X = Q_1 \cap Q_2 \subset \PP^5
1184: \end{equation}
1185: of two quadrics in five dimensional projective space is
1186: geometrically equivalent
1187: to an algebraic curve $\Si_I$ of genus 2.
1188: To reconstruct $\Si_I$ from $X$ it is enough to remark that
1189: \begin{enumerate} \item By the Lefshetz theorem $X$ is
1190: simply connected;
1191: \item $\Oh_X(-2) = K_X$ - the canonical class of $X$.
1192: Hence the embedding of $X$ into $\PP^5$ is canonical.
1193: \item The space of all quadrics in $\PP^5$ through $X$ is
1194: a pencil
1195: \begin{equation}
1196: \vert 2 H - X \vert = \PP^1
1197: \end{equation}
1198: that is parameterized by the projective line.
1199: \item The subset of singular quadrics in this pencil is
1200: six different points
1201: \begin{equation}
1202: W =p_1 \cup ... \cup p_6 \subset \PP^1.
1203: \end{equation}
1204: \item The double cover
1205: \begin{equation}
1206: \phi \colon \Si_I \to \PP^1
1207: \end{equation}
1208: with ramification in these six points $W$ is the required curve
1209: $\Si_I$ of genus 2.
1210: \end{enumerate}
1211:
1212: Note that the previous double cover is given by the canonical
1213: linear system of $\Si_I$.
1214: Now a simple exercise in linear algebra shows that there exists
1215: just one pencil of quadrics in $\PP^5 $ up to a linear
1216: transformation with the subset of singular
1217: quadrics which is equal to the ramification locus of the
1218: double cover.
1219:
1220: If $M_I \to M'_{I'}$ is a faithful functor and geometries of both
1221: varietes are undistinguishable it is quite reasonable to consider
1222: a quantization of $\sM'$ as a quantization of $\sM$, that is,
1223: to say
1224: \begin{equation}
1225: Q(M_I) = Q(M'_{I'})
1226: \end{equation}
1227: and so on ... We will say that such quantization is successful if
1228: the
1229: quantization of $\sM$ is successful, that is, corresponding
1230: vector bundles (1.29)
1231: on $\sM'$ admit projective flat holomorphic connections.
1232:
1233:
1234: \subsection{Perfect quantization}
1235:
1236: Suppose our algebraic variety $(M_I, \Oh(1))$ with a Hodge form
1237: $\om$ admits as a symplectic manifold a real polarization that is
1238: a Lagrangian fibration:
1239: \begin{equation}
1240: \pi \colon M \to B , \quad \pi^{-1}(b_{gen}) = T^n
1241: \end{equation}
1242: that is the generic fiber is a Lagrangian torus of dimension
1243: $n = 1/2 dim M$;
1244:
1245: Again we may suppose that the subset $BS_k \subset B$ is
1246: finite and the Hilbert space of wave functions of level $k$
1247: is given by the direct sum of lines
1248: \begin{equation}
1249: Q_F^k(M) = \sH_\pi^k = \oplus_{b \in BS_k \subset B} \C \cdot
1250: s_b
1251: \end{equation}
1252: where $s_b$ is a covariant constant section of $(L^k,
1253: \nabla_k)\vert_{\pi^{-1}(b)}$).
1254:
1255: Suppose we can construct a natural isomorphism
1256: \begin{equation}
1257: \PP Q_F^k(M) = \PP H^0(M_I, \Oh(k)).
1258: \end{equation}
1259: Then we have a perfect quantization providing a lot of
1260: beautiful and very important
1261: properties.
1262: \begin{enumerate}
1263: \item First of all, both kinds of quantization are successful
1264: : indeed, the LHS of the equality (1.100) doesn't depend on a complex
1265: polarization at all. On the other hand, RHS doesn't depend on
1266: a real polarization.
1267: \item LHS of the equality is decomposed into a sum of lines
1268: so the RHS as a space of sections admits a special basis
1269: (may be after fixing some additional structure like so called
1270: theta - structure for abelian varieties \cite{Mum}).
1271:
1272: \item A basis of such type is called the Bohr-Sommerfeld basis.
1273: \item Such situation can be called {\it perfect quantization.}
1274: \end{enumerate}
1275:
1276: Moreover, the first example of a perfect quantization is given by
1277: the classical theory of theta-functions.
1278:
1279:
1280: Obviously it is not the simplest example.
1281: The simplest example is the following: consider the standard two
1282: sphere $S^2$ realized as the complex Riemann sphere with the
1283: standard $U(1) = S^1$-action (rotations around North and South
1284: Poles). Now $U(1)$-invariant symplectic form $\om$ of volume 1
1285: gives the phase space of the classical mechanical system $S^2,
1286: \om$. To quantize this one we have to fix a complex structure, but
1287: it is unique and the real polarization
1288: \begin{equation}
1289: \pi \colon S^2 \to [-1, 1]
1290: \end{equation}
1291: is given by the projection to the rotation axis.
1292:
1293: As a prequantization date we get the line bundle $\Oh(1)$ of
1294: degree 1 with the standard $U(1)$-invariant unitary connection
1295: $\nabla$. Then
1296: the wave function spaces
1297: \begin{equation}
1298: \sH^k_I = H^0(\Oh(k))
1299: \end{equation}
1300: and $U(1)$- action on this space defines the special eigenbasis of
1301: Fourier polynomials. Fourier monomials
1302: are in 1-1 correspondence with
1303: Bohr-Sommerfeld fibers of the projection (1.101) in angle
1304: coordinates.
1305: Thus in this case we have the perfect quantization.
1306:
1307: It is easy to see that in this case all "good" theoretical
1308: conditions
1309: as (1.48) are
1310: disturbed but the result of quantization is successful.
1311:
1312: If we put out the word "natural" from the statement (1.100)
1313: we get the statement
1314: about ranks of both wave functions spaces. In this case we say that
1315: the quantization
1316: is {\it numerically} perfect. The geometry behind this
1317: numerical coincidence
1318: will be described in (2.25) - (2.30) and \cite{T3}.
1319: Such rank's equalities
1320: were proved for K3 surfaces cases and many others. But historically first case
1321: was the Gelfand-Cetlin system (see \cite{GS2}).
1322:
1323:
1324:
1325: \section{Algebraic curves = Riemann surfaces}
1326:
1327: \subsection{Direct approach}
1328:
1329:
1330:
1331: Let $\Si$ be a compact smooth oriented Riemann surface and
1332: $\Si_I$ be a complex structure on it, thus, one has an algebraic
1333: curve of genus $g > 1$. Such curves admit the canonical
1334: polarization by the cotangent bundle $T^* \Si_I = \Oh(K)$
1335: where $K$ is the canonical class of $\Si_I$
1336: (see for example \cite{DSS}). Moreover, if we fix a metric
1337: on $\Si$ with the conformal
1338: class given by the complex structure $I$ then
1339: the Levi-Civita connection on the cotangent
1340: bundle gives the prequantization data $(T^*\Si, \nabla_{LC})$.
1341: This pair is equivalent to
1342: the holomorphic pair $(\Si_I, \Oh(K))$.
1343:
1344: The union of all spaces $H^0(\Oh(k K))$ of holomorphic sections
1345: gives the collection of vector bundles
1346: \begin{equation}
1347: p_k \colon V_k \to \sM_g, \quad p^{-1}(M_I) = H^0(M_I, \Oh(k K))
1348: \end{equation}
1349: on the moduli space $\sM_g$ of curves of genus $g$.
1350: (Recall that $\dim \sM_g= 3g-3$).
1351:
1352: %\begin{figure}[tbn]
1353: %\centerline{\epsfxsize=2in\epsfbox{24.eps}}
1354: %\caption{\sl Riemann surface = curve}
1355: %\label{F5}
1356: %\end{figure}
1357:
1358: In particular, for $k=2$ the vector bundle $V_{2} = T^* \sM_g$ is
1359: the cotangent bundle of the moduli space, that is, $V_{2}$
1360: is the bundle with spaces of quadratic differentials as fibers.
1361: In particular the canonical class $K_{\sM_{g}}$ is the first Chern
1362: class of the vector bundle $V_{2}$:
1363: \begin{equation}
1364: c_1 (V_2) = K_{\sM_g}.
1365: \end{equation}
1366: Now according to Mumford this canonical class has
1367: the decomposition
1368: \begin{equation}
1369: K_{\sM_g} = 13 \Theta + \text{boundary divisors $\dots$}
1370: \end{equation}
1371: where $\Theta$ is the diviser of zero theta constants.
1372:
1373: If the vector bundle $V_{2}$ admits a holomorphic projective flat
1374: connection then by
1375: (1.31) the number 13 has to be divided by $3g-3$.
1376: This isn't the case thus $V_{2}$
1377: doesn't admit
1378: any flat connection and an algebraic curve as a polarized
1379: algebraic variety isn't a
1380: result of
1381: a successful quantization procedure of the classical dynamical
1382: system $(\Si, \om)$ where $\om$ is the restriction of the
1383: Fubini-Study form.
1384:
1385: But the history isn't finished! We can find a "faithful functor".
1386:
1387: \subsection{Jacobians}
1388:
1389: May be the first faithful functor for algebraic curve is mapping
1390: of
1391: the curve to the
1392: polarized jacobian
1393: $(J(\Si_I), \Theta )$. There exists a lot of constructions but
1394: for us it will be very convinient
1395: to consider $J(\Si_I)$ as the {\it moduli space of holomophic
1396: topological trivial line bundles}. Fixing a point $p_0 \in \Si_I$
1397: we consider for every other point $p\in \Si_I$ the topological
1398: trivial line bundle $\Oh(p - p_0)$, where we consider points as
1399: divisors of degree 1. So we get the map
1400: \begin{equation}
1401: a_{p_0} \colon \Si_I \to J(\Si_I).
1402: \end{equation}
1403: By the Riemann theorem this map is an embedding (if genus $g>0$)
1404: and induces the isomorphism
1405: \begin{equation}
1406: a \colon H_1(\Si, \Z) \to H_1(J(\Si_I), \Z).
1407: \end{equation}
1408: The moduli space of line bundles is a group (with respect to
1409: the tensor product of line bundles). So, we can consider the Pontrjagin map
1410: \begin{equation}
1411: a_{p_0}^{g-1} \colon \Si_I^{[g-1]} \to J(\Si_I)
1412: \end{equation}
1413: sending a finite set of points $\{p_1, ... , p_{g-1}\}$ to
1414: the line bundle
1415: \begin{equation}
1416: \Oh(p_1 + ... + p_{g-1} - (g-1)p_0) \in J(\Si_I).
1417: \end{equation}
1418: The image of this map is a divisor $\Theta$ (so called
1419: {\it theta diviser}) and the
1420: corresponding line bundle admits only one (up to $\C^*$)
1421: holomorphic section (with
1422: the zero set $\Theta$).
1423:
1424: To reconstruct the curve $\Si_I$ from the pair
1425: $(\Theta \subset J(\Si_I))$ we have to consider
1426: the Gauss map
1427: \begin{equation}
1428: G \colon \Theta \to \PP T^* J(\Si_I)
1429: \end{equation}
1430: to the projectivization of the cotangent bundle sending
1431: non singular point of $\Theta$ to the tangent hyperplane
1432: to $\Theta$ at this point.
1433: But the cotangent bundle (of a group) is
1434: trivial $\PP T^* J(\Si_I) = \PP^{g-1} \times J(\Si_I)$, and
1435: the composition of the projection to $\PP^{g-1}$ and
1436: the Gauss map gives the rational cover
1437: \begin{equation}
1438: G \circ p \colon \Theta \to \PP^{g-1}.
1439: \end{equation}
1440: Now the ramification diviser of this cover is the dual divisor
1441: to
1442: the canonical embedding of $\Si_I$ into $(\PP^{g-1})^*$ (for
1443: hyperelliptic curves we use other arguments). By the classical
1444: projective duality theory we can reconstruct a curve by its dual
1445: divisor. Thus the sending of a curve $\Si_I \to (J(\Si_I),
1446: \Theta)$ is a faithful functor.
1447:
1448: Now using our rule (1.97) we may quantize the polarized variety
1449: $(J(\Si_I), \Theta)$.
1450:
1451: The role of wave function spaces now are played by the spaces of
1452: holomorphic sections:
1453: \begin{equation}
1454: \sH_{I}^k = H^0(J(\Si_I), \Oh (k \Theta)).
1455: \end{equation}
1456: These spaces are called spaces of theta function of
1457: the Riemann surface $\Si_I$ of level $k$.
1458:
1459: This quantization is successful! Corresponding vector bundles
1460: on
1461: the moduli space admit projective flat connections (see
1462: \cite{W}). But the situation is much better than just a successful
1463: quantization.
1464:
1465:
1466:
1467: \subsection{Algebro geometrical theory of theta functions}.
1468:
1469: Consider a principal polarized abelian variety $(A, \Theta)$ of
1470: dimension $g$. Then the line bundle $\Oh(k \Theta)$ admits
1471: the finite group $K$ of translations preserving this line bundle.
1472: It is easy to see that
1473: \begin{equation}
1474: K = A_k \subset A
1475: \end{equation}
1476: is the subgroup of points of order $k$ in the abelian group $A$.
1477: Recall that we have a symplectic form on $K$ given by the wedge
1478: product of 1-cycles coupled by the polarization $\Theta$.
1479: The decomposition
1480: \begin{equation}
1481: K = K_+ \times K_-
1482: \end{equation}
1483: by two isotropic subgroups of this group is called
1484: theta-structure of level $k$.
1485:
1486: The beautiful classical result reproduced by Mumford is the following
1487: \begin{enumerate}
1488: \item In the complete linear system $\vert k \Theta \vert$ there
1489: exists only one divisor $D_+$ which is invariant with respect to
1490: the subgroup $K_+$ - action.
1491: \item Every translation $a \in K_-$ defines a divisor $D_+ + a$
1492: and the rational function $\theta_a$ with pole at the divisor
1493: $D_+$ and zeros at the divisor $D_+ + a$. This function is called
1494: theta function with characteristic $a \in K_-$.
1495: \item The space of holomorphic sections
1496: \begin{equation}
1497: H^0(A, \Oh (k \Theta)) = H^0(A, \Oh(D_+)) = \oplus_{a \in K_-} \C \theta_a.
1498: \end{equation}
1499: \item This is the theta functions with characteristics decomposition.
1500: \end{enumerate}
1501:
1502: Note, that the cardinalities of our finite abelian groups are
1503: given by formulas
1504: \begin{equation}
1505: \vert K \vert = k^{2g} , \quad \vert K_{\pm} \vert = k^{g}.
1506: \end{equation}
1507:
1508: Now it is quite useful to apply the extended Kodaira-Spenser
1509: theory to this situation. First of all for every abelian variety
1510: \begin{equation}
1511: H^0(A, T) \to H^2(A, \Oh)
1512: \end{equation}
1513: is an isomorphism thus the condition (1.48) holds. Moreover, from
1514: cohomology sequence of (1.50) we have
1515: \begin{enumerate}
1516: \item an isomorphism
1517: \begin{equation}
1518: H^0(A, S^2(T)) = H^1(A,D^1(k \Theta)));
1519: \end{equation}
1520: \item
1521: \begin{equation}
1522: H^0(A, \Oh) = H^0(A, D^1(\Oh(k \Theta)));
1523: \end{equation}
1524: \item and from (1.60)
1525: we have
1526: \begin{equation}
1527: 0 \to H^0(A, \Oh(k \Theta) )/ \C \cdot s \to H_h^1(d_s) \to
1528: H^1(A, D^1(\Oh(k \Theta))) \to 0;
1529: \end{equation}
1530: \item so under a deformation of the pair $(A, \Oh(k \Theta))$
1531: for every section $s$ there exits unique deformation of this
1532: section which is {\it defined by the heat equation} (see the
1533: standard presentation (2.49) below).
1534: \end{enumerate}
1535:
1536: \subsection{Symplectic-combinatorial theory of theta functions}
1537:
1538: A principal polarization $\Theta$ on an abelian variety $A$ can
1539: be done by
1540: some unimodular skew symmetrical integer form
1541: \begin{equation}
1542: \om \in H^2(A, \Z)
1543: \end{equation}
1544: because $A $ is a $2g$-dimensional real torus:
1545: \begin{equation}
1546: \om \in \wedge^2 H^1(A, \Z) = H^2(A,\Z).
1547: \end{equation}
1548: For a jacobian $J(\Si)$ this form is induced by the intersection
1549: form on 1-cycles on the Riemann surface $\Si$.
1550:
1551: The theory of unimodular integer skew symmetrical forms predicts
1552: that there are two isotropic $\Z$-sublattices in $H^1(A, \Z)$
1553: such that our form $\om$ has the standard symplectic shape.
1554:
1555: Recall that the underlying smooth manifold of our abelian variety
1556: is nothing else but a real torus $T^{2g}$ and the decomposition
1557: of $H^1(A, \Z)$ by two isotropic $\Z$-submodules induces
1558: the corresponding decomposition of this torus into two families of Lagrangian
1559: subtori
1560: \begin{equation}
1561: T^{2g} = T^g_+ \times T^g_-.
1562: \end{equation}
1563: Considering the projection to the second $g$-torus
1564: \begin{equation}
1565: \pi \colon T^{2g} \to T^g_-
1566: \end{equation}
1567: we obtain an integrable system that is a real polarization
1568: of the torus $T^{2g} = A$.
1569:
1570: %\begin{figure}[tbn]
1571: %\centerline{\epsfxsize=2in\epsfbox{35.eps}}
1572: %\caption{\sl tori fibration}
1573: %\label{F6}
1574: %\end{figure}
1575:
1576: \begin{figure}[tbn]
1577: \centerline{\epsfxsize=3in\epsfbox{35c.eps}}
1578: \caption{\sl Bohr-Sommerfeld fibers}
1579: \label{Fig 2}
1580: \end{figure}
1581:
1582: The direct interpretation of the prequantization line bundle
1583: shows that the subset of $k$-Bohr-Sommerfeld fibers
1584: \begin{equation}
1585: BS_k = (T^g_-)_k \subset T^g_-
1586: \end{equation}
1587: is the subgroup of points of order $k$ on our second torus
1588: $T^g_-$ (see Fig.5). So, the wave function space of the
1589: Bohr-Sommerfeld
1590: quantization is the sum of lines
1591: \begin{equation}
1592: Q_\pi^k (T^{2g}, \om) = \oplus_{a \in (T^g_-)_k} \C \cdot s_a
1593: \end{equation}
1594: where $s_a$ are covariant constant sections of restrictions of
1595: the prequantization line bundle.
1596:
1597:
1598: We can see that the results of the Kahler quantization and the
1599: Bohr-Sommerfeld quantization give wave function spaces of the
1600: same rank so the quantization is numericaly perfect (see the end
1601: of subsection 1.5).
1602:
1603: This is the first amazing fact in this story.
1604: We would like to say couple of words about the geometry behind
1605: such coincidence.
1606:
1607: Returning to the fibration (2.22) we can construct so called
1608: {\it dual fibration}
1609: \begin{equation}
1610: \pi' \colon (T')^{2g} \to T^g_-
1611: \end{equation}
1612: changing every torus-fiber $T^g_+$ by the dual torus $(T^g_+)^*$.
1613: This dual torus is nothing else but the space of gauge classes of
1614: all $U(1)$-flat connections on the trivial line bundle on
1615: $T^g_+$. We have got a priori a new $2g$-torus with new
1616: symplectic form $\om'$ of the same type as before.
1617:
1618: Now consider any holomorphic line bundle $L$ on our old
1619: complex torus $A$ with $U(1)$-connection which curvature
1620: form has to be proportional to $\om$. (Just like our
1621: prequantization bundle $\Oh(k)$). Then for every point
1622: $a \in T^g_+$ the restriction of $L$ to the Lagrangian fiber
1623: $\pi^{-1}(a)$ gives a flat line bundle, that is the point of the
1624: dual fibers
1625: \begin{equation}
1626: s_L \in (T^g_+)^* = (\pi^{-1}(a))^*.
1627: \end{equation}
1628: Thus every such line bundle defines a section
1629: \begin{equation}
1630: s_L \colon T^g_+ \to (T')^{2g},
1631: \end{equation}
1632: that is a middle dimensional submanifold $s_L$ (more precisely
1633: $s_L(T^g_+)$) in the manifold $(T')^{2g}$.
1634:
1635: Of course trivial connection defines zero section of the
1636: new fibration (2.25) that is we have a middle dimensional
1637: submanifold
1638: \begin{equation}
1639: s_{\Oh(0)} \subset (T')^{2g}.
1640: \end{equation}
1641:
1642: Obviously, the intersection set
1643: \begin{equation}
1644: s_{\Oh(0)} \cap s_{\Oh(k)} = BS_k
1645: \end{equation}
1646: is the set of Borh-Sommerfeld fibers of level $k$. Moreover, it
1647: is easy to see that both our cycles are Lagrangian, oriented
1648: and admit positive index intersections only. Thus
1649: the number of Bohr-Sommerfeld fibers is given by the
1650: intersection index of cohomology classes
1651: \begin{equation}
1652: [s_{\Oh(0)}] \cdot [ s_{\Oh(k)}] = \# BS_k.
1653: \end{equation}
1654: Now using the first Chern class of the polarization we can easily
1655: compute these numbers.
1656:
1657: The moral is the following: an abelian variety $A = T^{2g}$ has
1658: a symplectic partner $(T')^{2g}$
1659: such that coherent sheaves geometry on $A$ can be encoded by the
1660: geometry of Lagrangian cycles (or super cycles) on the partner
1661: $(T')^{2g}$. Usually this partner is called
1662: {\it mirror partner}.
1663: Mirror symmetry now is far developed domain so we
1664: can't dive to this subject (see \cite{CK}).
1665: In spite of the coincidence of our previous construction with
1666: the Strominger-Yau-Zaslow mirror symmetry construction we would
1667: like to stop here discussion of this subject but formally it can be extended
1668: along this line (see \cite{T3}).
1669:
1670: Of course the dimensional coicidence can be also extended. Moreover there are
1671: theta bases in both of wave function spaces.
1672: To compare them we have to construct
1673: a holomorphic object, namely, a section
1674: of the theta bundle of level $k$ using only
1675: Bohr-Sommerfeld fiber of the projection $\pi$ (2.22) (see Figure 1).
1676:
1677: We will do this using so called {\it coherent state
1678: transform} or the slightly generalized
1679: Segal-Bargmann isomorphism introduced in the
1680: context of the quantum theory as a transform from the Hilbert
1681: space of square integrable functions on the configuration
1682: space to the space of holomorphic functions on the phase space.
1683:
1684: Considering zero fiber $T^g_+$ of the polarization map $\pi$
1685: (2.22) we can interpret our complex torus $A$ as a
1686: complexification of this real torus and input our geometrical
1687: situation to the situation of the classical coherent state
1688: transform.
1689:
1690: In the finite dimensional context a configuration space is
1691: just $\R^g$ as the real part of the phase space $\C^g$.
1692: The coherent state transform
1693: \begin{equation}
1694: CST_t \colon L^2(\R^g, d^g x) \to L^2(\C^g, d\mu_t) \cap \sH(\C^g)
1695: \end{equation}
1696: is a unitary isomorphism.(Here $d\mu_t$ is the Gaussian measure
1697: and $\sH$ is the space of holomorphic functions.)
1698:
1699: For our abelian case we have to replace $\R^g$ by $T^g = U(1)^g$ and $\C^n$
1700: by our complex
1701: torus $A$. Remark
1702: that $T^g$ in our case is a special Lagrangian
1703: subtorus (see, for example \cite{T4}).
1704:
1705:
1706: B. C. Hall proposed a generalization of the CST where $\R^n$ is replaced
1707: by $\SU(2)$ and $\C^n$ by $SL(2, \C)$ (see
1708: \cite{Ha}) and we will use it later for non-abelian case. But here we use this
1709: construction for abelian group $U(1)^g \subset (\C^*)^g$.
1710:
1711: First of all, the decomposition (2.21) induces the decomposition
1712: \begin{equation}
1713: H^1(A, \Z) = H^1(T^g_+, \Z) \oplus H^1(T^g_-, \Z).
1714: \end{equation}
1715: Let us fix a basis $(a_1, \dots, a_g)$ in the lattice $H^1(T^g_+, \Z)$ and
1716: a basis $(b_1, \dots, b_g)$ in the lattice
1717: $ H^1(T^g_-, \Z)$ such that the full system $(a_1, \dots, a_g ,b_1, \dots, b_g)$
1718: is a standard basis of 1-homology of $A$. For a complex Riemann
1719: surface $\Si$
1720: consider the period matrix
1721: \begin{equation}
1722: \Om = \Vert \Om_{ij}\Vert = Re \Om + i \cdot Im \Om
1723: \end{equation}
1724: as a point of the Siegel space $H$.
1725: Let $\vec x =(x_1, \dots, x_g) \in [0,1]^g$
1726: be periodic coordinates on $U(1)^{g} = T^g_+$. Then we have
1727: $U(1)$-invariant
1728: Laplacian
1729: \begin{equation}
1730: \De^{Im \Om} = \sum_{i,j = 1}^g \frac{Im \Om_{ij}}{2 \pi} \frac{\p^2}{\p x_i
1731: \p x_j}.
1732: \end{equation}
1733: The complexification of $T^g_+$ is $(\C^*)^g$ with coordinates $(e^{2\pi i z_1},
1734: \dots,
1735: e^{2\pi i z_g})$ for $\vec z = \vec x + i \vec y$. Then the Haar mesure on
1736: $(\C^*)^g$ is $d \vec x d \vec y$ and we have $U(1)^g$-invariant
1737: complex
1738: Laplacian
1739: \begin{equation}
1740: \Delta^{Im \Om}_\C = \sum_{i,j = 1}^g \frac{Im \Om_{ij}}{2 \pi} (\frac{\p^2}{\p x_i
1741: \p x_j} + \frac{\p^2}{\p y_i
1742: \p y_j}).
1743: \end{equation}
1744:
1745: As the first step consider the fundamental solution $\mu_t$ at the
1746: identity of the
1747: heat equation on $(\C^*)^g$
1748: \begin{equation}
1749: 4 \frac{\p u}{\p t} = \Delta^{Im \Om}_\C u
1750: \end{equation}
1751: and average the measure $\mu_t d \vec x d \vec y$ with respect to the action of
1752: $U(1)^g$ :
1753: \begin{equation}
1754: \nu_t(\vec z) d \vec x d \vec y =
1755: (\int_{U(1)^g} \mu_t(\vec z + \vec x) d \vec x) \quad d \vec x d \vec y.
1756: \end{equation}
1757: The explicit formula is
1758: \begin{equation}
1759: \nu_t(\vec z) = (\frac{2}{t})^{g/2} (det (Im \Om))^{-1/2} e^{\frac{\pi}{2t}
1760: \sum_{i,j}
1761: (z_i - \ov{z_i}) (Im \Om)^{-1}_{ij}(z_j - \ov{z_j})}.
1762: \end{equation}
1763: Now consider non-self adjoint Laplace operator
1764: \begin{equation}
1765: \Delta^{-i \Om} = \sum_{i,j = 1}^g \frac{i}{2 \pi} \Om_{ij}\frac{\p^2}{\p x_i
1766: \p x_j}.
1767: \end{equation}
1768: C. Florentino, J. Mourao and J. Nunes proved in \cite{FMN} the
1769: following
1770: \begin{prop} Let $\sC$ be the analytic continuation from $U(1)^g$ to
1771: $(\C^*)^g$. Then the transform
1772: \begin{equation}
1773: C_t^{(-i \Om)} = \sC \circ e^{t/2 \Delta^{(-i \Om)}} \colon L^2(U(1)^g, d \vec x)
1774: \to
1775: \end{equation}
1776: $$
1777: \to L^2((\C^*)^g, d\nu_t) \cap \sH((\C^*)^g)
1778: $$
1779: is unitary.
1780: \end{prop}
1781: So for any function $f \in L^2(U(1)^g, d \vec x)$ given by its Fourier
1782: decomposition
1783: \begin{equation}
1784: f(\vec x) = \sum_{\vec n \in \Z^g} a_n e^{2\pi i \vec x \cdot \vec n}
1785: \end{equation}
1786: we have
1787: \begin{equation}
1788: (C_t^{(-i \Om)} f)(\vec z) =
1789: \sum_{\vec n \in \Z^g} a_n e^{ti\pi \vec n \Om \vec n}\cdot
1790: e^{2\pi i \vec x \cdot \vec n}.
1791: \end{equation}
1792: This function is the analytic continuation to $(\C^*)^g$ of the solution of
1793: the complex heat equation
1794: \begin{equation}
1795: 2 \frac{\p u}{\p t} = \Delta^{(-i \Om)} u
1796: \end{equation}
1797: on $(\C^*)^g$ with the initial condition given by f.
1798:
1799: But the coherent states transform (2.40) can be extended from
1800: $L^2(U(1)^g, d \vec x)$ to the space of distributions $(C^\infty(U(1)^g))' $
1801: given by Fourier series of the form (2.41) such that there exists an integer
1802: $N > 0$ such that
1803: \begin{equation}
1804: lim_{\vec n \cdot \vec n \to \infty} \frac{| a_n |}{(1 + \vec n \cdot \vec n)^N}
1805: = 0.
1806: \end{equation}
1807: The Laplace operator and its powers act as continuous linear operators on this space of distributions
1808: (by duality from the corresponding action on $C^\infty(U(1)^g)$)
1809: and define for ($t > 0$) the action of the operator $e^{t/2} \Delta^{(-i \Om)}$ on
1810: distributions $f$ of the form (2.41) as
1811: \begin{equation}
1812: e^{t/2} \Delta^{(-i \Om)} (\sum_{\vec n \in \Z^g} a_n e^{2\pi i \vec x \vec n}) =
1813: \end{equation}
1814: $$
1815: = \sum_{\vec n \in \Z^g} a_n e^{t i \pi \vec n \Om \vec n}
1816: e^{2\pi i \vec z \vec n}.
1817: $$
1818:
1819:
1820: In the same paper \cite{FMN} the authors proved the following
1821: \begin{prop} If a series $f \in (C^\infty(U(1)^g))' $ then the RHS of (2.45)
1822: defines a holomorphic function on
1823: $(\C^*)^g$.
1824: \end{prop}
1825: Now we are ready to map every Bohr-Sommerfeld fiber of the
1826: projection (2.22) to an analytic function on $(\C^*)^g$. It is
1827: enough to do it for zero torus $T^g_+ = U(1)^g$ of this
1828: projection.
1829:
1830: For this let us consider the distribution
1831: \begin{equation}
1832: \theta^\R_0 (x) = \sum_{\vec n \in \Z^g} e^{2\pi i k \vec x \vec n}.
1833: \end{equation}
1834: It is nothing else but the delta-function at the identity of level
1835: $k$. Let us apply CST (2.40) to this distribution:
1836: \begin{equation}
1837: C_{1/k}^{(-i \Om)} (\theta_k^\R) =
1838: \end{equation}
1839: $$
1840: = \sum_{\vec n \in \Z^g} e^{\pi i k \vec n \cdot \Om / k \vec n k } \cdot e^{
1841: 2 \pi i k\vec n \vec z}.
1842: $$
1843: For every $l \in \Z^g / k \Z^g$ we can start with the function
1844: \begin{equation}
1845: \theta^\R_l (x) = \sum_{\vec n \in \Z^g} e^{2\pi i \vec x (l+ k\vec n)}
1846: \end{equation}
1847: to get the analytic function
1848: \begin{equation}
1849: C_{1/k}^{(-i \Om)} (\theta_l^\R) = \theta_l(z, \Om) =
1850: \end{equation}
1851: $$
1852: = \sum_{\vec n \in \Z^g} e^{\pi i (l+ k \vec n )\cdot \Om / k (l
1853: + k\vec n )} \cdot e^{ 2 \pi i (l + k\vec n ) \vec z}.
1854: $$
1855:
1856: Remark that we are substitute our continious
1857: positive parameter $t$ coming from
1858: heat kernels by its discrete conterpart $1/k$.
1859:
1860:
1861: We would like to emphasize that in spite of 150 years of
1862: development of the classical theory of theta functions this
1863: "comparing quantization" approach was realized first quite
1864: recently by C. Florentino, J. Mourao and J. Nunes in \cite{FMN}.
1865:
1866: What we have to do now is just to compare our holomorphic functions on
1867: $(\C^*)^g$ with holomorphic sections of the $\Theta$-line bundle on
1868: $A$.
1869:
1870: \subsection{Abelian holomorphic flat connections}
1871:
1872: The usual way to define a line bundle $L$ is
1873: to consider some finite cover $\{U_i\}$ on the base and
1874: a collection of {\it transition functions} $f_{ij}$ which are
1875: regular
1876: and regular inversed on $U_i \cap U_j$.
1877: On the intersection $U_i \cap U_j \cap U_k$ we have the cocycle
1878: equality
1879: \begin{equation}
1880: f_{ij} \cdot f_{jk} \cdot f_{ki} = 1.
1881: \end{equation}
1882: A collection $f_{ij}$ is equivalent to a collection $f'_{ij}$
1883: iff there exists a collection $\{\beta_i\}$ of functions, where each
1884: $\beta_i$ is regular and regular inversed on $U_i $ such that
1885: \begin{equation}
1886: f'_{ij} = \beta_i f_{ij} \beta_j^{-1}.
1887: \end{equation}
1888:
1889:
1890:
1891: The idea is to make these functions (2.51) as simple as possible using
1892: this equivalence: for example take all $f'_{ij}$
1893: constant.
1894:
1895: Starting with any collection of functions (2.50) let us
1896: consider the collection of differentials
1897: forms
1898: \begin{equation}
1899: \{ f_{ij}^{-1} d f_{ij} \} \in Z^{1}(\{U_{i} \} )
1900: \end{equation}
1901: Obviously, this cochain is a cocycle from $H^1(\Om) = H^{1,1}$
1902: whose cohomology class
1903: \begin{equation}
1904: c(L) \in H^{1,1}
1905: \end{equation}
1906: is the first Chern class of $L$.
1907:
1908: If this class is zero then the cocycle is trivial and there are
1909: matrices of differential forms $h_i$ such that
1910: \begin{equation}
1911: f_{ij}^{-1} d f_{ij} = h_i - h_j.
1912: \end{equation}
1913: \begin{dfn} A collection of differential forms $h_i$
1914: on $U_i$ is called flat holomorphic connection of the
1915: vector bundle $L$ given by the collection of transition
1916: functions
1917: $f_{ij}$.
1918: \end{dfn}
1919:
1920: Having such collection we can solve the system of
1921: linear equations on $U_i$
1922: \begin{equation}
1923: \beta_i^{-1} \cdot d \beta_i = h_i
1924: \end{equation}
1925: and obtain a new collection of transition functions (2.51).
1926: Now using all equations we can see that functions
1927: $f'_{ij} = \beta_i \al_{ij} \beta_j^{-1}$ are constant, that is
1928: \begin{equation}
1929: d f'_ij = 0.
1930: \end{equation}
1931: If transition functions are constant then this vector bundle is
1932: called
1933: {\it local system of coefficients} and we can trivialize it over
1934: any simply connected open set. Thus we have a character of
1935: the fundamental group of our base:
1936: \begin{equation}
1937: \rho \colon \pi_1 \to \C^*.
1938: \end{equation}
1939:
1940: For a curve we've proved the classical result of Poincare:
1941:
1942: \begin{prop}
1943: Every $L \in J(\Si)$ admits holomorphic flat
1944: connection
1945: given by a character $\chi \colon \pi_1(\Si) \to \C^*$.
1946: \end{prop}
1947:
1948: How many characters give the same vector bundle?
1949: Just as many as flat connections admitted by this bundle there are.
1950: The difference of any two holomorphic connections can be
1951: identified over
1952: $U_i \cap U_j$, thus
1953: \begin{equation}
1954: \{h_i\} - \{ h'_i\} \in H^0(\Oh_\Si (K_\Si)).
1955: \end{equation}
1956:
1957: By Serre duality
1958: \begin{equation}
1959: H^1(\Oh)^* = H^0( \Oh( K_\Si))
1960: \end{equation}
1961: is a fiber of
1962: the cotangent bundle $T^* J(\Si)$.
1963:
1964: The full space $A$ of all
1965: holomorphic flat connections is the space of all characters
1966: \begin{equation}
1967: A = (\C^*)^{2g}, \quad dim_\C A = 2g
1968: \end{equation}
1969:
1970:
1971: Every character defines a holomorphic vector bundle
1972: by the standard construction
1973: \begin{equation}
1974: L = U \times \C / (\pi_1,\rho)
1975: \end{equation}
1976: where $U$ is the universal cover of our base with the natural
1977: action of the fundamental group of a base on $U$ and the $\rho$-action
1978: on $\C$ .
1979:
1980: Thus we have the forgetful map
1981: \begin{equation}
1982: f \colon A \to J(\Si).
1983: \end{equation}
1984: Every fiber of this map is the set of holomorphic flat connections
1985: on a fixed line bundle. Thus the map $f$ sending each character
1986: to the corresponding line bundle provides on $A$ the structure of an
1987: {\it affine bundle over the cotangent bundle}.
1988:
1989: So fibers of $f$ are affine spaces over $H^0(J(\Si),\Om)$ (but in
1990: this case the cotangent bundle is the trivial $g$-dimensional
1991: vector bundle). Over every small open set $U_i$ our affine bundle
1992: admits a section that is can be identified with the restriction
1993: of the cotangent bundle. Any descrepance exists only on the
1994: intersections $U_i \cap U_j$. The collection of differences of
1995: sections gives a 1-cocycle
1996: \begin{equation}
1997: \ep \in H^1(J(\Si), \Om).
1998: \end{equation}
1999: If this cocycle is trivial, our affine bundle coincides with its
2000: vector bundle while non trivial affine bundle is defined by this
2001: cocycle uniquely. By the Dolbeault theorem in our case the cocycle
2002: $\ep_A \in H^{1,1} = H^2(J(\Si), \C)$. Poincare proved that
2003: \begin{equation}
2004: \ep_A = [\Theta] = [\om].
2005: \end{equation}
2006: To prove this equality we have to use the functional equation
2007: for theta functions.
2008:
2009:
2010: The affine bundle (2.62) admits {\it non holomorphic sections}:
2011: let $U(1) \subset \C^*$ be the subgroup of unit norm numbers.
2012: There is a map
2013: \begin{equation}
2014: mod \colon \C^* \to U(1), \quad z \to \frac{z}{|z|}.
2015: \end{equation}
2016: Then we have the subspace
2017: \begin{equation}
2018: U(1) (\C^*)^{2g}
2019: \end{equation}
2020: of unitary characters of $\pi_1(\Si)$ which is a {\it section} of
2021: the projection $f$ (2.62) of the affine vector
2022: bundle $(\C^*)^{2g}$. It follows immediately from the maximum
2023: principle for a compact complex manifold.
2024:
2025:
2026:
2027: So we have the smooth identification
2028: \begin{equation}
2029: J(\Si) = U(1)^{2g}.
2030: \end{equation}
2031:
2032:
2033: Let $H_1(\Si, \Z) = <a_1, ... ,a_g, b_1,... ,b_g> $ be as before
2034: (see (2.32)).Such
2035: presentation defines
2036: very important subspace of the character space, so
2037: called {\it abelian Schottky subspace}:
2038: \begin{equation}
2039: S^a_g = (\C^*)^g = \{ \chi \in (\C^*)^{2g} \vert \chi(a_i) = 1 \}.
2040: \end{equation}
2041: and
2042: \begin{equation}
2043: uS^a_g = (U(1))^g = \{ \chi \in U(1)^{2g} \vert \chi(a_i) = 1 \}
2044: \end{equation}
2045: is
2046: {\it abelian unitary Schottky space}.
2047:
2048: The restrictions of the forgetful map $f$ (2.62) onto this space has the
2049: following properties
2050: \begin{enumerate}
2051: \item $f \colon S^a_g = (\C^*)^g \to J(\Si)$ is infinite ($\Z^g$) cover,
2052: \item $ f^* (\Oh (\Theta))$ is trivial, so holomorphic sections =
2053: holomorphic functions = theta functions;
2054: \item this line bundle is given by the {\it automorphy factors}
2055: \begin{equation}
2056: e_{a_i}(z) = 1
2057: \end{equation}
2058: $$
2059: e_{b_i}(z) = e^{-2\pi z_i - \pi i \Om_{ii}}
2060: $$
2061: where $\vec z$ are complex coordinates of the universal cover $U$ of $A$
2062: described after formula (2.34), $\Om$ is
2063: the period matrix of $\Si$ and so on (see the set up around (2.32) - (2.35));
2064: let $\{ u_i\}$ be the basis in $U$ dual to $\{ a_i\}$ and
2065: \begin{equation}
2066: u_{g + i} = \sum_1^g \Om_{ij} u_i;
2067: \end{equation}
2068: \item thus using these automorphy factors we see that the space $H^0(J(\Si),
2069: \Oh(k \Theta))$ is naturaly identifyed with the space of holomorphic functions
2070: $\theta$
2071: on $(\C^*)^g$ such that $\theta(z + u_i) = \theta$ and
2072: \begin{equation}
2073: \theta(z + u_{g+i}) = e^{-2\pi k z_i - \pi i k
2074: \Om_{ii} };
2075: \end{equation}
2076: \item thus any $\theta(z)$ admits the following decomposition
2077: \begin{equation}
2078: \theta (z) = \sum_{l \in \Z^g / k \Z^g} \theta_l(z, \Om)
2079: \end{equation}
2080: where theta functions with characteristics are functions (2.49).
2081: \end{enumerate}
2082:
2083: \subsection{Perfect quantization}
2084:
2085: Summarizing all results and constructions, we have
2086:
2087: \begin{thm} Principal polarized abelian varieties
2088: (and hence algebraic curves) admit the perfect
2089: quantization by theta functions.
2090: \end{thm}
2091:
2092: In particular, we proved very important fact (which we want to
2093: generalize for the non-abelian case):
2094:
2095: \begin{prop}
2096: The Bohr-Sommefeld wave space (2.24) doesn't depend on the
2097: decomposition (2.21) determining projection $\pi$ (2.22). This
2098: decomposition defines a basis in this space only and the result
2099: of Bohr-Sommerfeld quantization doesn't depend on the choice of a
2100: real polarization of an "arithmetical" type (2.22).
2101: \end{prop}
2102: Of course this observation is obvious in the set-up of the classical theory
2103: of theta functions, but in the set-up of non-abelian theta functions it is giving
2104: an important "independence condition" in CQFT and low dimensional topology.
2105:
2106:
2107:
2108:
2109: Let us list the sequence of tasks necessary to get a perfect quantization:
2110: \begin{enumerate}
2111: \item a $g$-torus $U(1)^g$ with the marked point $0 \in U(1)^g$ we identify with
2112: the zero fiber of the fibration $\pi : A \to T^g_-$ (2.22) containing zero point;
2113: \item for zero point $0 \in U(1)^g$ we construct $\de$-function $\theta_0^\R$
2114: as the Fourier
2115: serie (2.46);
2116: \item using the period matrix $\Om$ of $A$ we construct CST
2117: $C^{(-i \Om)}_{1/k}$ (2.40) , (2.41);
2118: \item applying this transform to the distribution $\theta_0^\R$
2119: we get the collection of holomorphic theta functions with characteristics $\{
2120: \theta_l(z, \Om)\}$ (2.49) on the complexification $(\C^*)^g$ of $U(1)^g$;
2121: \item using the periods matrix $\Om$ we construct the cover $(\C^*)^g \to A$
2122: and get sections of the line bundle $\Oh(k \Theta)$ as functions subjecting to
2123: automorphity factors conditions (2.70);
2124: \item to check that functions $\{
2125: \theta_l(z, \Om)\}$ (2.49) correspond to sections of $\Oh(k \Theta)$.
2126: \end{enumerate}
2127:
2128: This is the end of story for the moduli spaces of topological trivial
2129: vector bundles
2130: of rank one on algebraic curves=Riemann surfaces. But any curve $\Si$ also
2131: has the other
2132: "moduli space", namely the moduli space of topological trivial
2133: vector bundles of rank 2. This is a faithful functor too. To implement
2134: this theory we have to plug Conformal Field Theory in dimension
2135: $D = 2$. We will use the previous list of tasks as a pattern of
2136: much more sophisticated efforts in this "non-abelian" case.
2137:
2138:
2139: Note that $H^0(J(\Si), \Oh ( k \Theta))$ is the space of the irreducible
2140: representation of the Heisenberg
2141: group of level $k$.
2142:
2143:
2144: This statement is in the very heart of the classical theory of theta
2145: functions. Actually the group $K$ (2.12) is the quotient of the
2146: Heisenberg group $\Ga_k$ with the kernal $\Z_k$ (with respect to the
2147: multiplicative action). More precisely, the Heisenberg group acts
2148: on the line bundle $\Oh(k \Theta)$ and on the space of its
2149: holomorphic sections. This is an
2150: irreducible representation of $\Ga_k$ and such representation is
2151: unique up to projectiviazation. The existence of the special basis is provided
2152: by this
2153: action. ( In particular, $rk H^0(J(\Si), \Oh ( k \Theta)) = k^g$).
2154: This property (to be the space of unique irreducible
2155: representation of some group or algebra) is true for non-abelian
2156: case too. But in this case the space of theta functions is the
2157: space
2158: of irreducible representation of the gauge algebra
2159: of WZW CQFT.
2160:
2161:
2162:
2163: \section{Non-abelian theta functions}
2164:
2165:
2166: The direct generalization of jacobians of algebraic curves
2167: as a faithful functor is the sending of an algebraic curve $\Si_I$
2168: to the {\it moduli space} $M^{ss}$ of semi-stable
2169: topologicaly trivial rk 2 vector bundles on this curve.
2170: Of course, we can consider much more complicated situation,
2171: but this case is the first {\it non-commutative} case which
2172: is expressive enough to see all new features of the
2173: geometrical situation.
2174:
2175: Following our pattern we can quantize a Riemann surface $\Si$
2176: using the faithful functor
2177: \begin{equation}
2178: \Si \to M^{ss}(\Si)
2179: \end{equation}
2180: sending a Riemann surface to the moduli space of semi-stable
2181: vector bundles on it.
2182:
2183: For this we have to repeat all steps of the abelian (classical) theory of
2184: theta functions.
2185:
2186:
2187: \subsection{Algebraic geometry of moduli spaces of vector bundles}
2188:
2189:
2190: The algebro-geometric part of the
2191: non-abelian theory of theta functions is well developed by
2192: Narashimhan, Beauville,
2193: Laszlo, Pauly, Oxbury, Ramanan, Sorger and many others.
2194: In non-abelian context important new notion of semi-stability
2195: appears. Recall that for our case semi-stability of $E$ means
2196: that $E$ doesn't contain linear subbundles of positive degree
2197: (that is, with positive $c_1$). To work with rk 2 vector bundles
2198: we need some information about the structure of coherent sheaves
2199: on algebraic curves and more from the homological algebra. We
2200: have to know that every coherent sheaf on a smooth algebraic
2201: curve is a direct sum of a local free sheaf (= vector bundle) and
2202: a torsion sheaf with support in a finite set of points. From the
2203: homological algebra we need Riemann-Roch theorem and first
2204: cohomology of sheaves. All these facts can be found in any
2205: survey on this subject, for example in \cite{DSS}
2206:
2207: Let $M^{ss} (\Si)$ be the moduli space of topologically trivial
2208: semi-stable holomorphic bundles of rank 2 on $\Si$ where every
2209: non-stable (but semi stable) bundle is presented by the direct sum
2210: of two opposite
2211: topological trivial line bundles.
2212: Then $ dim M^{ss}(\Si) = 3g-3$.
2213:
2214: The tangent space to the moduli space at a point $E$
2215: has well known shape
2216: \begin{equation}
2217: TM^{ss}_E = H^1(\ad E)
2218: \end{equation}
2219: where $\ad E$ is the traceless part of the vector bundle of
2220: endomorphisms of $E$. For any stable bundle $E$ one has
2221: $H^0(\Si_I, \ad E) = 0$ and thus by the Riemann-Roch theorem
2222: \begin{equation}
2223: dim M^{ss} = rk TM^{ss}_E = rk H^1(\ad E) = 3g-3.
2224: \end{equation}
2225:
2226: Definitions of theta divisors are absolutely parallel to (2.6):
2227: let $\si \in Pic_{g-1}(\Si)$ be a line bundle on $\Si$ of degree
2228: $g-1$; then we have
2229:
2230: \begin{prop}
2231: \begin{enumerate}
2232: \item $\Theta = \{ E \in M^{ss}(\Si) \vert H^0(\Si, E(\si))
2233: \neq 0\}$;
2234: \item as a variety $\Theta$ is birationally equivalent
2235: to the projective space $\PP^{3g-4}$;
2236: \end{enumerate}
2237: \end{prop}
2238:
2239: Indeed, a general vector bundle from the divisor $\Theta$ twisted by $\si \in
2240: Pic_{g-1}(\Si)$
2241: admits a section. (Recall that in VBAC-slang "twisted" means
2242: tensor product with the line bundle corresponding to $\si$.)
2243: We may expect that this section has no zeros, thus our general
2244: vector bundle $E$ can be represented as an extension
2245: \begin{equation}
2246: 0 \to \Oh(-\si) \to E \to \Oh(\si) \to 0.
2247: \end{equation}
2248: Of course there exists trivial extension $E = \Oh(-\si) \oplus
2249: \Oh(\si)$ but this bundle isn't stable at all.
2250:
2251: Any other extension is given by non zero cocycle of the space
2252: \begin{equation}
2253: Ext^1(\Oh(\si), \Oh(-\si)) = H^1(\Si_I, \Oh(-2\si)) = \C^{3g-3}
2254: \end{equation}
2255: by the Riemann-Roch theorem. Automorphisms of a subbundle act
2256: on these cocycles by constant multiplication.
2257: Thus if $E(\si)$ has no other section (what happens in general case),
2258: our vector bundle $E$ is defined uniquely by a point of
2259: the projective space $\PP H^1(\Si_I, \Oh(-2\si)) = \PP^{3g-4}$.
2260: It is a simple exercise to estimate the dimension of spaces of
2261: bundles admitting a section with non trivial zero set.
2262:
2263: The moduli space $M^{ss}$ is close to be rational itself. Let us
2264: consider the space $Pic_g$ of classes of divisors of
2265: degree $g$ (instead of $g-1$ as in Proposition 5) and twist
2266: all bundles from $M^{ss}$ by some divisor $\si \in Pic_g(\Si)$.
2267: Then for every vector bundle $E \in M^{ss}$
2268: we have
2269: \begin{equation}
2270: rk H^0(\Si_I, E (\si)) \geq 2.
2271: \end{equation}
2272: Now we have the canonical (evaluating) homomorphism
2273: \begin{equation}
2274: can \colon H^0(\Si_I, E(\si)) \otimes \Oh \to E(\si)
2275: \end{equation}
2276: Again by dimensional computations it is easy to see that this
2277: moduli space contains a Zariski open set $M_0$ of bundles
2278: for which this homomorphism
2279: can be extended to the exact sequence
2280: \begin{equation}
2281: 0 \to \C^2 \otimes \Oh \to E(\si) \to \oplus_{i=1}^{2g}
2282: \Oh_{p_i} \to 0
2283: \end{equation}
2284: where $2g$ points $\{ p_i \}$ form the sum which is an effective
2285: divisor equivalent to $2\si$:
2286: \begin{equation}
2287: p_1 + p_2 + .... + p_{2g} \in \vert 2\si \vert = \PP H^0(\Si_I,
2288: \Oh(2\si)) = \PP^g.
2289: \end{equation}
2290: Why does this sum of points give an effective divisor from $\vert
2291: 2\si\vert$? Well, the homomorphism
2292: \begin{equation}
2293: \wedge^2 can = det \colon \Oh \to \Oh(2 \si) = c_1(E(\si))
2294: \end{equation}
2295: and the effective divisor (3.9) is just zero set of this
2296: homomorphism.
2297:
2298:
2299:
2300: Sending a vector bundle $E \in M_0$ to such effective divisor
2301: we
2302: obtain the map (of course algebraic)
2303: \begin{equation}
2304: det \colon M_0 \to \PP^g.
2305: \end{equation}
2306: The fiber can be obtained by the canonical homomorphism data:
2307: \begin{equation}
2308: ker can_{p_i} \colon \C^2 \to E_{p_i}.
2309: \end{equation}
2310: So for every point $p_i$ we have a point on the projective line
2311: \begin{equation}
2312: \ga_i = \PP ker can_{p_i} \in \PP \C^2 = \PP^1.
2313: \end{equation}
2314: Thus the fiber of the map (3.11) is given by a collection of
2315: points
2316: \begin{equation}
2317: (\ga_1, .... , \ga_{2g}) \subset \PP^1
2318: \end{equation}
2319: up to a projective linear transformation.
2320:
2321: We can see that any fiber of the map $det$ is rational too
2322: and $M^{ss}$ is an algebraic fibration with rational
2323: $g$-dimensional base and rational $2g-3$-dimensional fiber.
2324:
2325: The same trick we can use for vector bundles of higher ranks. But
2326: we would like to recall the old problem of the birational
2327: geometry:
2328:
2329: {\it Is the moduli space $M^{ss}$ rational?}
2330:
2331: In spite of many attempts to solve this rationality problem,
2332: the answer isn't known up to now.
2333:
2334: \begin{dfn}
2335: The space $H^0(M^{ss}(\Si), \Oh ( k \Theta))$ is called the space
2336: of non-abelian theta functions of level $k$.
2337: \end{dfn}
2338: Continuing the parallel description of the geometry of moduli
2339: spaces we get
2340: \begin{prop}
2341: \begin{enumerate}
2342: \item Theta divisors are ample. Moreover
2343: \item $H^2(M^{ss}, \Z) = \Z$.
2344: \item (Faltings 92) $H^0(M^{ss}(\Si), \Oh ( k \Theta))$ is the space
2345: of irreducible representation of the gauge algebra
2346: of WZW CQFT.
2347: \end{enumerate}
2348: \end{prop}
2349: Such spaces have played a noted role in Conformal Field Theory.
2350: The Proposition will be proved below.
2351:
2352:
2353: \subsection{Holomorphic flat connections}
2354:
2355: Up to now we have considered any vector bundle as a local free
2356: sheaf.
2357: As far as for line bundles any vector bundle is defined by
2358: some finite cover $\{U_i\}$ of the base and by
2359: a collection of {\it transition matrices} $\al_{ij}$ regular
2360: and regular inversed on $U_i \cap U_j$.
2361: On the intersection $U_i \cap U_j \cap U_k$ we have again the
2362: cocycle equality
2363: \begin{equation}
2364: \al_{ij} \cdot \al_{jk} \cdot \al_{ki} = 1.
2365: \end{equation}
2366: A collection $\al_{ij}$ is equivalent to a collection $\al'_{ij}$
2367: iff there exists a collection $\beta_i$ matrices, each $\{\beta_i\}$ is
2368: regular and regular inversed on $U_i $ such that
2369: \begin{equation}
2370: \al'_{ij} = \beta_i \al_{ij} \beta_j^{-1}.
2371: \end{equation}
2372:
2373:
2374: For the matrix case "as simple as possible" representatives can be
2375: \begin{enumerate}
2376: \item triangle matrices (the theory of extensions, see (3.4));
2377: \item constant matrices (the theory of flat holomorphic
2378: connections).
2379: \end{enumerate}
2380: Again let us
2381: consider the collection of matrices of differential
2382: forms
2383: \begin{equation}
2384: \{ \al_{ij}^{-1} d \al_{ij} \} \in Z^{1}(\{U_{i} \} )
2385: \end{equation}
2386: Obviously this cochain is a cocycle from $H^1(\Om (End E))$ which
2387: cohomology class
2388: \begin{equation}
2389: c(E) \in H^{1}(\Om(End E))
2390: \end{equation}
2391: is the full Chern class of $E$.
2392:
2393: The automorphisms group $Aut E$ acts on the space $H^{1}(\Om(End E
2394: ))$ preserving this class.
2395:
2396: For example, on a curve by the Serre duality
2397: \begin{equation}
2398: H^{1}(\Om(End E))^{*} = H^{0}(End E)
2399: \end{equation}
2400: and it is easy to see that
2401:
2402: \begin{prop}
2403: \begin{equation}
2404: <c(E), \si> \neq 0 \implies \quad \si \quad \text{is an idempotent
2405: in } \ H^0(End E)
2406: \end{equation}
2407: \end{prop}
2408: This simple observation belonging to Atiyah gives plenty results
2409: about the structure of multidimensional bundles: the vector bundle
2410: $End E$ splits as
2411: \begin{equation}
2412: End E = \Oh \oplus ad E
2413: \end{equation}
2414: where the first component $\Oh$ is the trivial line bundle
2415: corresponding to homotheties and $ad E$ corresponds to traceless
2416: endomorphisms. Thus we have the decomposition
2417: \begin{equation}
2418: \Om(End E) = \Om \oplus ad E \otimes \Om
2419: \end{equation}
2420: where the first component sends endomorphisms to their traces.
2421: Thus if $H^0(End E)$ contains one indempotent then it contains a
2422: second too and such vector bundle is the direct sum of line
2423: bundles.
2424:
2425: Vector bundle $E$ is called {\it indecomposable} if $E \neq E_{1}
2426: \oplus E_{2}$.
2427:
2428: For such bundles
2429: \begin{equation}
2430: c(E) \in H^{1}(\Om)
2431: \end{equation}
2432: since $H^0(End E)$ doesn't contain any idempotents.
2433:
2434: If $\{ \al_{ij}^{-1} \cdot d \al_{ij} \}$ is the cochain (3.15)
2435: defining $c(E)$ then
2436: \begin{equation}
2437: tr (\al_{ij}^{-1} d \al_{ij}) \in H^{1} (\Om).
2438: \end{equation}
2439: By the Dolbeault isomorphism
2440: \begin{equation}
2441: H^1(\Om) = H^{1,1} \in H^2(\C).
2442: \end{equation}
2443: The formal matrix equality
2444: \begin{equation}
2445: tr (\al_{ij}^{-1} \cdot d \al_{ij}) = (det \Vert\al_{ij}
2446: \Vert)^{-1} d (det \Vert \al_{ij} \Vert)
2447: \end{equation}
2448: shows that
2449: \begin{equation}
2450: c(E) = c(det E)
2451: \end{equation}
2452: for any indecomposable vector bundle on curve.
2453:
2454:
2455: It is easy to see (just by the direct interpretation) that this
2456: cohomology class is integer and moreover
2457: \begin{equation}
2458: \{\al_{ij}^{-1} d \al_{ij}\} = c_1(E)
2459: \end{equation}
2460: where $E$ is the indecomposable vector bundles given by
2461: $\{\al_{ij}\}$. If this class is zero as in our case of topologically
2462: trivial bundles then the cocycle is trivial and there are
2463: matrices of differential forms $h_i$ such that
2464: \begin{equation}
2465: \al_{ij}^{-1} d \al_{ij} = h_i - h_j.
2466: \end{equation}
2467: \begin{dfn} Collection of matrix differential forms $h_i$
2468: on $U_i$ is called flat holomorphic connection on the
2469: vector bundle $E$ given by the collection of transition
2470: functions
2471: $\al_{ij}$.
2472: \end{dfn}
2473:
2474: Having such collection we can solve the system of linear
2475: equations on $U_i$
2476: \begin{equation}
2477: \beta_i^{-1} \cdot d \beta_i = h_i
2478: \end{equation}
2479: and get a new collection of transition matrices. Now using all
2480: equations we can see that matrices $\al'_{ij} = \beta_i \al_{ij}
2481: \beta_j^{-1}$ are constant, that is
2482: \begin{equation}
2483: d \al'_ij = 0.
2484: \end{equation}
2485: If matrices are constant, this vector bundle is called
2486: {\it local system of coefficients} and we can trivialize it over
2487: any simply connected open set and obtain the representation of
2488: the fundamental group of the base:
2489: \begin{equation}
2490: \rho \colon \pi_1 \to SL(2, \C).
2491: \end{equation}
2492: Actually, because of the equivalence relations we obtain a class of
2493: representations only. That is the representation $\rho$ up to
2494: the adjoint action of $SL(2, \C)$.
2495:
2496:
2497: For curve we have the classical result:
2498:
2499: \begin{prop} (A. Weil) Every $E \in M^{ss}(\Si)$ admits a
2500: holomorphic flat connection
2501: given by a class of representations $\rho \colon \pi_1(\Si) \to SL(2, \C)$.
2502: \end{prop}
2503: The author emphasizes again that he've proved this statement
2504: already.
2505:
2506: \begin{rmk}
2507: In the influential paper \cite{We} A. Weil proposed
2508: \begin{enumerate}
2509: \item the proof of the previous Proposition;
2510: \item the idea that high dimensional vector bundles should play
2511: the role of non-abelian analogue of Jacobians;
2512: \item the notion of {\it class of matrix divisor} which is
2513: equivalent to the notion of vector bundle.
2514: \end{enumerate}
2515: \end{rmk}
2516:
2517:
2518:
2519:
2520: Many classes of representations give the same vector
2521: bundle:
2522: the difference of two holomorphic connections can be
2523: identified over
2524: $U_i \cap U_j$, thus
2525: \begin{equation}
2526: \{h_i\} - \{ h'_i\} \in H^0(End E \otimes \Om).
2527: \end{equation}
2528: An element of the last space is a homomorphism
2529: \begin{equation}
2530: \phi \colon E \to E \otimes \Om
2531: \end{equation}
2532: and for case of curves (when $K_\Si = \Om_\Si$) it is called
2533: {Higgs field}.
2534:
2535: Note that by Serre duality
2536: \begin{equation}
2537: H^1(\ad E)^* = H^0( \ad E \otimes K)
2538: \end{equation}
2539: that is the vector bundle over the moduli space $M^{ss}$ with
2540: fibers equal to spaces of Higgs fields is nothing else but the
2541: cotangent bundle $\Om$ of $M^{ss}$.
2542:
2543:
2544: For rank 2 vector bundles with fixed determinant we have to
2545: consider traceless Higgs fields only:
2546: \begin{equation}
2547: \phi \in H^0(\Si, \ad E \otimes K).
2548: \end{equation}
2549:
2550:
2551: The full space $A^{na}$ of all
2552: holomorphic flat connections
2553: \begin{equation}
2554: A^{na} = CLRep(\pi_1(\Si), SL(2, \C))
2555: \end{equation}
2556: is the space of all classes of representations and
2557: every representation defines a holomorphic vector bundle
2558: just by the standard construction
2559: \begin{equation}
2560: E = U \times \C^2 / (\pi_1,\rho)
2561: \end{equation}
2562: where $U$ is the universal cover of $\Si$ with the natural
2563: action of the fundamental group of athe base.
2564:
2565: Of course in the set of such bundles there are non-stable but
2566: indecomposable
2567: vector bundles and even semi-stable but indecomposable. Let us remove
2568: such representations and get the space $A_{na}^{ss}$ of classes
2569: of representations which gives stable and decomposable semi-stable bundles.
2570:
2571: Consider the forgetful map
2572: \begin{equation}
2573: f \colon A^{na}_{ss} = CLRep(\pi_1(\Si), \SU(2)), SL(2, \C)) \to M^{ss}.
2574: \end{equation}
2575: This is the affine bundle over the cotangent bundle (see (2.62
2576: )). Again such bundle is given by a cocycle
2577: \begin{equation}
2578: \ep_{na} \in H^1(M^{ss}, \Om) = H^{1,1}(M^{ss}, \C)
2579: \end{equation}
2580:
2581: \begin{prop}
2582: \begin{equation}
2583: \ep_{na} = [\Theta_{na}] = [\om_{na}]
2584: \end{equation}
2585: \end{prop}
2586: That is, the answer is precisely the same as in the abelian case.
2587: We will prove this fact a little bit latter using special
2588: subspaces.
2589:
2590: Both of affine bundles (abelian and non-abelian) (2.62) and (3.39)
2591: admit {\it non holomorphic sections}. Again constructions of these
2592: sections are quite parallel (see (2.67) for the abelian case) :
2593: the including $\SU(2) \subset SL(2, \C)$ of complexification
2594: defines
2595: the subspace
2596: \begin{equation}
2597: CLRep (\pi_1(\Si), \SU(2)) \subset CLRep (\pi_1(\Si), SL(2, \C)
2598: \end{equation}
2599: which is the section of the projection $f$ (3.39). (Again it
2600: follows from the maximum principle for the compact curve.) But the
2601: new feature of non abelian
2602: case is the following:
2603:
2604: we don't know if the restriction of forgetfull map to
2605: $CLRep (\pi_1(\Si), \SU(2))$ is
2606: onto. Does every semi-stable bundle admits Hermitian flat connection?
2607:
2608: We will prove this statement later, and note that the restriction map defines
2609: the Narasimhan-Sesadri isomorphism
2610: \begin{equation}
2611: NS \colon CLRep(\pi_1(\Si), \SU(2)) \to M^{ss}.
2612: \end{equation}
2613:
2614:
2615: So we have differential identification
2616: $$
2617: CLRep(\pi_1(\Si), \SU(2)) = M^{ss}.
2618: $$
2619:
2620:
2621: Let $\pi_1 = <a_1, ... ,a_g, b_1,... ,b_g \vert \prod_{i=1}^g [a_i, b_i] = 1>$
2622: be a standard presentation
2623: of the fundamental group. Such presentation defines a couple of
2624: very important subspaces of the spaces of representations, so
2625: called Schottky subspaces:
2626: \begin{enumerate}
2627: \item $S_g = \{ \rho \in A^{na} \vert \rho (a_i) = 1 \}$ is
2628: non-abelian complex Schottky space.
2629: \item $S_g = SL(2, \C)^g / Ad_{diag}SL(2, \C)$
2630: \item $dim S_g = dim M^{ss}$
2631: \item $uS_g = S_g \cap CLRep(\pi_1(\Si), \SU(2))$ is unitary Schottki space;
2632: \item $dim \quad uS_g = 1/2 dim M^{ss}$.
2633: \end{enumerate}
2634:
2635: We saw in the abelian case that the restriction of the forgetful
2636: map to the Schottki space is a cover giving theta sections as
2637: functions.
2638:
2639:
2640: For the non-abelian case the situation is much more complicated:
2641: \begin{enumerate}
2642: \item the forgetful map
2643: \begin{equation}
2644: f \colon S_g \to M^{ss}(\Si)
2645: \end{equation}
2646: is meromorphic only
2647: (because of existence of non stable flat bundles).
2648: For example, the Schottky uniformization of curve gives the
2649: non stable bundles.
2650: \item Properties of this map are unknown yet.
2651: C. Florentino proved only that the differential of this map at the
2652: unitary Schottky space is non degenerated and in section 9 we show that for
2653: general curve $\Si$ the image of $f$ (3.44) is Zariski dense in $M^{ss}(\Si)$.
2654: \end{enumerate}
2655: To avoid this gap, we have to use Symplectic and Lagrangian
2656: geometry and sophisticated analysis.
2657:
2658: \subsection{Moduli of stable pairs
2659: and the desingularization of moduli spaces}
2660:
2661: The construction of the moduli spaces was achived in the 1960's,
2662: mainly by D.Mumford and the mathematicians of Tata institute.
2663: But after appearing of the gauge theory approach to vector bundles
2664: these moduli spaces were input in the collection of moduli spaces
2665: of {\it stable pairs} related by a chain of {\it flips} (see
2666: \cite{Re}). This chain of flips was used by Thaddeus, Bertram and
2667: Bredlow-Daskalopoulos for the computations of the cohomology ring
2668: of moduli spaces and by Seshadri \cite{Se} for the construction
2669: of a desingularisation of moduli spaces.
2670:
2671: New object to study will be a pair $(E, s)$ consisting of a
2672: vector bundle $E$ on a curve $\Si$ and a nonzero section
2673: $s \in H^{0}(E)$. S. Bredlow defined new stability condition for
2674: such pairs and proved theorem relating stable pairs to vortices
2675: on the Riemann surface. The vortex equation depends on positive
2676: real parameter $t$ and so the stability condition also depends on
2677: $t$.
2678: \begin{rmk}
2679: The vortex equation requests a Kahler metric, not a conformal
2680: class only. We consider the algebro-geometrical interpretation
2681: here. We fix the normalization of such metric such that the
2682: volume $vol \Si = 4 \pi$.
2683: \end{rmk}
2684:
2685:
2686:
2687: Our vector bundles admit nonzero sections so we have to consider
2688: some positive line bundle $D$ of degree $d$ such that $det E =
2689: D$. We obtain this case by twisting vector bundles from $M^{ss}$
2690: by an appropriate line bundle. The stability condition for pair
2691: $(E, s)$ with rk 2
2692: vector bundle is the
2693: following: a pair is semi-stable if for any line bundle $ L
2694: \subset E$
2695: \begin{equation}
2696: deg L \leq 1/2 d - t \quad \text{ if } s \in H^{0}(L)
2697: \end{equation}
2698: $$
2699: deg L \leq 1/2 d +t \quad \text{ otherwise. }
2700: $$
2701: The fondation of this theory can be found in \cite{BD},
2702: \cite{Ber}, \cite{Th2}.
2703:
2704: Let $M(D, t)$ be the moduli space of stable pairs. It's easy to
2705: see that
2706: \begin{enumerate}
2707: \item $M(D, t)$ is nonempty if and only if $t \leq d/2$;
2708: \item if $t$ is irrational then $M(D, t)$ is compact;
2709: \item if for $(E_{i}, s_{i}) \in M(D, t), \quad i=1,2$ and if there
2710: exists a homomorphism $\phi \colon E_{1} \to E_{2}$ such that
2711: $\phi(s_{1}) = s_{2}$ then $\phi$ is an isomorphism;
2712: \item for $(E, s) \in M(D, t)$ there are no endomorphisms of $E$
2713: annihilating the section $s$ exept 0 and no endomorphisms preserving $s$
2714: exept identity;
2715: \item if $(E, s) \in M(D, t)$ then $(E(\xi), s(\xi)) \in M(D(\xi), t)$
2716: for any effective divisor $\xi$.
2717: \end{enumerate}
2718:
2719: The deformation theory of stable pairs is very simple: let $t$ be
2720: irrational, then for $(E, s) \in M(D, t)$
2721: \begin{enumerate}
2722: \item there is a natural exact sequence
2723: \begin{equation}
2724: o \to H^{0}(ad E) \to H^{0}(E) \to TM(D,t)_{(E,s)} \to H^{1}(ad
2725: E) \to H^{1}(E);
2726: \end{equation}
2727: where homomorphisms are defined by the section $s$ .
2728: \item $rk
2729: TM(D,t)_{(E,s)} = d - g -2.$
2730: \end{enumerate}
2731: So for irrational $t$ the space $M(D, t)$ is smooth and we
2732: concentrate our attention on these smooth moduli spaces. More
2733: precisely we have the case if
2734: \begin{equation}
2735: t_{i} \in (max(0, d/2 - i - 1, d/2 - i)).
2736: \end{equation}
2737: For all $t_{i}$ from these intervals we have the same compact
2738: smooth moduli spaces $M(D, t_{i})$ related by the chain of flips
2739: (see \cite{Re}). We will use them later.
2740:
2741:
2742: In general case the stability condition for pair $(E, s)$ is the
2743: following: a pair is semi-stable if for any proper subbundle $
2744: E_{1} \subset E$
2745: \begin{equation}
2746: \mu (E_{1}) \leq \mu (E) - t \quad \text{ if } s \in H^{0}(E_{1})
2747: \end{equation}
2748: $$
2749: \mu E_{1} \leq \mu(E) +t \quad \text{ otherwise. }
2750: $$
2751: where $\mu$ as usual is slope of the bundle. The fondations of
2752: this theory can be find in \cite{BD}.
2753:
2754:
2755: For moduli spaces of stable pairs we have the same collections
2756: of statements as before.
2757:
2758: To apply this method to desingularization of $M^{ss}$ we have to
2759: send a vector bundle $E \in M^{ss}$ to the pair $(End E, s)$
2760: where $s \in H^{0}(End E)$. For a stable $E$ the choice of
2761: section is unique: it has to be the line subbundle of
2762: homotheties. For
2763: semi-stable bundles there are two cases: $H^{0}(End E) = 2, 4$.
2764: Considering the parameter $t$ which is very close to $0$ we get
2765: the moduli space of stable pairs $\widetilde{M^{ss}}$ with the
2766: birational projection to $M^{ss}$ which is a smooth variety.
2767: Always below {\it we will consider this canonical desingularization
2768: will not mention that specially}.
2769:
2770: \begin{rmk}
2771: This desingularization was constructed by Seshadri \cite{Se}. He
2772: considered a {\it parabolic structure} on vector bundle over
2773: fixed point of a curve and choose special parameters of such
2774: structure to get the same non singular variety with a birational
2775: map to the $M^{ss}$.
2776: \end{rmk}
2777:
2778:
2779:
2780:
2781:
2782:
2783:
2784: \subsection{Holomorpic symplectic geometry of Higgs fields.}
2785:
2786:
2787: Returning to Higgs fields (3.36),
2788: we can send every pair $(E, \phi)$ where $\phi$ is a Higgs
2789: field
2790: to the space of holomorphic quadratic differentials:
2791: \begin{equation}
2792: \wedge^2 \phi \colon \Oh \to \wedge^2 (E \otimes K_{\Si}).
2793: \end{equation}
2794: if $E$ is stable our pair is stable too and $T^* M^{ss}$ is a
2795: moduli space of such pairs. Then
2796: we get
2797: the holomorphic map
2798: \begin{equation}
2799: \pi \colon T^* M^{ss} \to H^0(\Si, K_{\Si_I}^2).
2800: \end{equation}
2801:
2802:
2803: As usual the cotangent bundle admits the holomorphic "action
2804: 1-form" $\al$ such that the holomorphic differential of it $d
2805: \al = \Om$ is the holomorphic symplectic form defining the
2806: holomorphic skew symmetrical isomorphism
2807: \begin{equation}
2808: \Om \colon T T^* \to T^* T^*.
2809: \end{equation}
2810: One can see that we just reproduce notions and constructions of
2811: the classical symplectic geometry in the holomorphic set up. In
2812: particular, we can expect existence of "completely integrable
2813: systems" that is a holomorphic map of holomorphic symplectic
2814: variety $\sM$
2815: \begin{equation}
2816: \pi \colon \sM \to B
2817: \end{equation}
2818: such that every fiber is Lagrangian with respect to $\Om$ and the
2819: generic fiber is a middle dimensional complex torus or an abelian
2820: variety.We call such map as holomorphic moment map.
2821:
2822: The brilliant observation of Hitchin is that the map (3.50) gives
2823: such integrable system. Moreover, the variety $T^* M^{ss}$ can be
2824: slightly extended as a holomphic symplectic variety such a way
2825: that fiber of $\pi$ become compact.
2826:
2827: \begin{dfn} Higgs pair $(E, \phi)$ is called stable if it doesn't
2828: exist a line
2829: subbundle $L \in E$ of positive degree invariant with respect to
2830: $\phi$. That is, the restriction of $\phi$ to $L$ hasn't the form
2831: \begin{equation}
2832: \phi \vert_L \colon L \to L \otimes K_\Si \subset E \otimes K_\Si.
2833: \end{equation}
2834: \end{dfn}
2835:
2836: It can be shown (see \cite{H2}) that the moduli space $MH_\Si$ of
2837: stable Higgs pairs
2838: \begin{enumerate}
2839: \item contains the cotangent bundle $T^* M^{ss}$ as a Zariski
2840: dense subset;
2841: \item $T^* M^{ss}$ is different from $MH_\Si$ in codimension 2
2842: and we can use Hartog's principle;
2843: \item the holomorphic symplectic form can be extended to such
2844: form (with the same notation) $\Om$;
2845: \item the map $\pi$ (3.50) can be extended on $MH_\Si$ to
2846: a holomorphic
2847: moment map
2848: \begin{equation}
2849: \pi \colon MH_\Si \to H^0(\Si, \Oh(2K_\Si))
2850: \end{equation}
2851: with compact fibers.
2852: \end{enumerate}
2853:
2854: All these statements can be proved directly and we just
2855: describe general
2856: fiber of the map $\pi$. General Higgs field we can consider as an
2857: endomorphism of the
2858: projectivization
2859: of our vector bundle $E$
2860: $$
2861: \phi \colon \PP E \to \PP E = \PP E \otimes K_\Si.
2862: $$
2863: If such map is birational then over every point $p \in \Si$
2864: \begin{enumerate}
2865: \item either $\phi_p$ is a linear automorphism of $\PP E_p$ and
2866: has two fixed points $p_1, p_2 \in \PP E_p$,
2867: \item or the endomorphism $\phi_p$ is degenerated and the
2868: image gives us one point $p \in \PP E_p$.
2869: \end{enumerate}.
2870: Thus we have double cover
2871: \begin{equation}
2872: \phi \colon \Si_\phi \to \Si
2873: \end{equation}
2874: with ramification divisor
2875: \begin{equation}
2876: \xi_\phi = (\wedge^2 \phi)_0 \in \vert 2K_\Si \vert
2877: \end{equation}
2878: (recall that all our endomorphisms are traceless).
2879:
2880: The curve $\Si_\phi$ is called spectral curve. It lies on the
2881: ruled surface $\PP E$ and we have the standard adjoint sequence of
2882: sheaves on this surface
2883: \begin{equation}
2884: 0 \to \Oh(- \Si_{\phi}) \to \Oh \to \Oh_{\Si_{\phi}} \to 0.
2885: \end{equation}
2886:
2887: Multiplying this sequence by the Grothendieck line bundle $H$ on
2888: $\PP E$ (recall
2889: that for the projection to base $R^0pr(H) = E, R^1pr (H) = 0 $) we get the
2890: exact triple
2891: \begin{equation}
2892: 0 \to \Oh(- \Si_\phi) \otimes H \to H \to \Oh_{\Si_\phi}(H) \to
2893: 0.
2894: \end{equation}
2895: The exact sequence of this triple for direct images gives us the
2896: isomorphism
2897: \begin{equation}
2898: 0 \to E \to R^0(\Oh_{\Si_{\phi}} (H)) \to 0
2899: \end{equation}
2900: because of the restriction
2901: $$
2902: \Oh(-\Si_phi) (H )\vert_{\PP E_p} = \Oh(-2) \otimes \Oh(1) =
2903: \Oh(-1)
2904: $$
2905: on the projective line.
2906:
2907:
2908: So on the spectral curve $\Si_\phi$ our vector bundle $E$ defines
2909: the line bundle $\Oh_{\Si_{phi}} (H) \in Pic (\Si_\phi)$.
2910:
2911: For double cover $\phi$ (3.55) the norm map
2912: \begin{equation}
2913: Nm \colon J(\Si_\phi) \to J(\Si)
2914: \end{equation}
2915: has connected preimage of zero
2916: \begin{equation}
2917: Nm^{-1}(0) = Pr_\phi
2918: \end{equation}
2919: which is an abelian variety which is called {\it Prym variety}.
2920:
2921: The isomorphism (3.59) shows that
2922: \begin{equation}
2923: Nm (\Oh_{\Si_\phi} (H)) = \wedge^2 E = \Oh.
2924: \end{equation}
2925: Thus
2926: \begin{equation}
2927: \Oh_{\Si_\phi} (H) \in Pr_\phi.
2928: \end{equation}
2929:
2930:
2931: So, general quadratic differential
2932: $w \in H^0(\Si, \Oh(2 K_\Si)$ has
2933: zero-set
2934: \begin{equation}
2935: (w)_0 = p_1, \dots , p_{4g-4}
2936: \end{equation}
2937: and defines the double cover
2938: \begin{equation}
2939: \phi \colon \Si_w \to \Si
2940: \end{equation}
2941: with the ramification locus (3.64).
2942:
2943: It is easy to see that every $L \in Pr_\phi$ defines rk2 vector
2944: bundle on $\Si$:
2945: \begin{equation}
2946: E_L = R^0 L
2947: \end{equation}
2948: with $\wedge^2 E_L = \Oh.$ This bundle is stable as a Higgs
2949: bundle if the cover is irreducible.
2950:
2951: Thus the fiber of the map $\pi$ (3.54) is
2952: \begin{equation}
2953: \pi^{-1} (w) = Pr_\phi.
2954: \end{equation}
2955:
2956:
2957:
2958:
2959:
2960: \subsection{Gauge theory on Riemann surface. Higgs fields.}
2961:
2962:
2963: The new features of the gauge theory is a necessity to
2964: consider objects of different dimensions.
2965: For example,
2966: Riemann surface bounds threefold (handlebody) and we come to
2967: 3-dimensional theory to get a result in 2-dimensional case. In
2968: the same vein sometimes it is reasonable to consider graphs
2969: (1-dimensional complexes)
2970: instead of Riemann
2971: surfaces and so on.
2972:
2973: A classical field theory on a manifold $M$ has three
2974: ingredients:
2975: \begin{enumerate}
2976: \item a collection $\sA$ of {\it fields} on $M$, which are
2977: geometric
2978: objects, such as sections of vector bundles, connections on
2979: vector bundles,
2980: maps from $M$ to some auxiliary manifold (the target space) and
2981: so on;
2982: \item an {\it action functional}
2983: $ S \colon \sA\to\C$ which is an integral of a function $L$ (the
2984: Lagrangian) of fields;
2985: \item a collection of observable functionals on the space of
2986: fields,
2987: $ w \colon \sA\to\C.$
2988: \end{enumerate}
2989: For example: {\it Chern--Simons functional}. Here
2990: \begin{enumerate}
2991: \item
2992: $M$ is a
2993: 3-manifold, \item the collection of fields
2994: \[
2995: \sA=\Om^1(M) \tensor \fsu (2)
2996: \]
2997: is the space of 1-form with coefficients in the Lie algebra,
2998: \item the action functional
2999: \begin{equation}
3000: S(a)=\frac{1}{8} \pi^2\int_M tr(a \cdot d a + \frac{2}{3} a^3),
3001: \end{equation}
3002: \item as an observable we can consider a {\it Wilson loop},
3003: given by some knot $K\subset M$:
3004: \[
3005: W_K(a)= tr(Hol_K(a))
3006: \]
3007: - the trace of holonomy of connection $a$ around knot $K$.
3008: \end{enumerate}
3009: \begin{rmk} These functions on the space of fields $\sA$ form
3010: an algebra because of the Mandelstam constraint:
3011: $$
3012: tr(A ) \cdot tr(B) = tr(A \cdot B) + tr( A^{-1} \cdot B).
3013: $$
3014: \end{rmk}
3015: Our space $CLRep(\pi_1(\Si), \SU(2))$ consists of classes of
3016: $\SU(2)$-representations of the fundamental group of a Riemann
3017: surface of genus $g$. This space contains the subspace of
3018: reducible representations
3019: \begin{equation}
3020: CLRep(\pi_1(\Si), \SU(2))^{red} \subset CLRep(\pi_1(\Si),\SU(2)).
3021: \end{equation}
3022: To get this space as the phase space of some mechanical system,
3023: consider a
3024: compact smooth Riemann surface $\Si$ of genus $g > 1$ and trivial
3025: Hermitian vector bundle $E_h$ of rank 2 on it.
3026: As usual, let $\sA_h$ be the
3027: affine space (over the vector space $\Om^1(\End E_h)$)
3028: of Hermitian
3029: connections and $\sG_h$ is the corresponding Hermitian gauge
3030: group. This space has the subspace:
3031: \[
3032: \sA_h^{red} \subset \sA_h
3033: \]
3034: where the left-hand side is the subset of reducible connections.
3035: As usual,
3036: let
3037: \[
3038: \sA_h^*=\sA_h-\sA_h^{red}.
3039: \]
3040:
3041: The sending of connection to its curvature tensor defines a
3042: $\sG_h$-equivariant map
3043: \begin{equation}
3044: F \colon \sA(E_h)\to\Om^2(End E_h)=Lie(\sG_h)^*
3045: \end{equation}
3046: to the coalgebra Lie of the gauge group.
3047:
3048: We can consider this map as the moment map with respect to the
3049: action of $\sG_h$. The subset
3050: \begin{equation}
3051: F^{-1}(0)=\sA_F
3052: \end{equation}
3053: is the subset of flat connections.
3054:
3055: For a connection $a\in \sA_F$ a tangent vector to $\sA_h$ at $a$ is
3056: \begin{equation}
3057: \om\in\Om^1(\End E_h)=T \sA_h,
3058: \end{equation}
3059: and for it the following condition holds
3060: \begin{equation}
3061: \na_a (\om)=0.
3062: \end{equation}
3063: The trivial vector bundle $E_h$ admits zero connection
3064: $\theta $,
3065: which is interesting and important from many points of view. In
3066: particular it provides possibility to identify $\sA_h$ with
3067: $\Om^1(End E_h)$ sending connection $a$ to the
3068: form $a - \theta$. We
3069: will identify forms and connections this way.
3070:
3071: The space $\sA_h=\Om^1(End E_h)$ is the {\it collection of fields} of
3072: YM-QFT with the {\it Yang--Mills functional}
3073: \[
3074: S(a)=\int_{\Si} |F_a|^2.
3075: \]
3076: Thus $\sA_F$ is a {\it classical phase space}, that is, the space
3077: of solutions of the {\it Euler--Lagrange equation} $\de S(a)=0$.
3078:
3079: There exists a symplectic structure on the affine space $\sA_h$,
3080: induced by the canonical 2-form given on the tangent space
3081: $\Om^1(End E_h)$ at a connection $a$ by the formula
3082: \begin{equation}
3083: \om (\ga_1,\ga_2)=\int_{\Si} tr (\ga_1 \wedge\ga_2).
3084: \end{equation}
3085: This form is $\sG_h$-invariant, and its restriction to
3086: $\sA_F^{irr}$ is degenerate along $\sG_h$-orbits: at a connection
3087: $a$, for a tangent vector $\ga \in\Om^1(End E_h)$, we have
3088: \[
3089: \om\in T\sG_h \text{if and only if} \om=\na_a \phi \quad\text{for $
3090: \phi\in\Om^0(\End E_h)=Lie(\sG_h)^*$,}
3091: \]
3092: and
3093: \[
3094: \int_{\Si} tr (\na_a \fie \wedge\om)=\int_{\Si}tr(\fie\wedge\na_a\om)=0.
3095: \]
3096: Hence
3097: \begin{equation}
3098: \om\in T \sA_F \text{if and only if} \Om_0 (\na_a \fie,\om)=0.
3099: \end{equation}
3100:
3101: Interpreting (3.70) as a moment map and using symplectic reduction
3102: arguments, we get a non degenerate closed symplectic form $\om$ on
3103: the space
3104: \begin{equation}
3105: \sA_F /\sG_h= CLRep(\pi_1(\Si), \SU(2))
3106: \end{equation}
3107: of classes of $\SU(2)$-representations of the fundamental group
3108: of the Riemann surface, and a stratification of this space. The
3109: form $\om$ defines symplectic structure on $CLRep(\pi_1(\Si),
3110: \SU(2))^{irr})$ and symplectic orbifold structure on
3111: $CLRep(\pi_1(\Si), \SU(2))$.
3112:
3113: \begin{rmk}
3114: The symplectic structure $\om$ is defined in pure topological
3115: way (see \cite{G1}).
3116: \end{rmk}
3117:
3118:
3119: On the other hand, the form $\om_0$ on $\sA_h$ is the differential
3120: of the
3121: 1-form $D$ given by the formula
3122: \begin{equation}
3123: D(\om)=\int_{\Si} tr((a) \wedge\om).
3124: \end{equation}
3125: We consider this form as a unitary connection $A_0$ on the trivial
3126: principal \hbox{$U(1)$-bundle} $L_0$ on $\sA_h$.
3127:
3128: To descend this Hermitian bundle and its connection to the
3129: orbit space,
3130: one defines $\Theta$-cocycle (or $\Theta$-torsor) on the
3131: trivial line
3132: bundle (see \cite{RSW}). This cocycle is $U(1)$-valued function
3133: $\Theta$ on $\sA_h\times\sG_h$ defined as follows: for any triple
3134: $(\Si,a,g)$ where $(a,g)\in\sA_h\times\sG_h$, we can find a
3135: triple $(Y,A,G)$ where $Y$ is a smooth compact 3-manifold, $A$ a
3136: $\SU(2)$-connection on the trivial vector bundle $\sE$ on $Y$ and
3137: $G$ is a gauge transformation of it, such that
3138: \begin{equation}
3139: \p Y=\Si, \quad a=A \vert_{\Si} \quad \text{and} \quad g=G\vert_{\Si}.
3140: \end{equation}
3141: Then
3142: \begin{equation}
3143: \Theta (a, g)=e^{i(CS(A^G)-CS(A))}.
3144: \end{equation}
3145:
3146: Recall that the Chern--Simons functional on the space $\sA
3147: (\sE_h)$ of unitary connections on the trivial vector bundle is
3148: given by the formula
3149: (3.68).
3150: It can be checked that the function (3.79) does not depend
3151: on the
3152: choice of the triple $(Y,A,G)$ (see \cite{RSW}, \S2).
3153:
3154: The differential of $\Theta$ at $(a, g)$ is given by the formula
3155: \begin{multline}
3156: d\Theta(\om,\phi)= \\
3157: \frac{\pi i}{4}\,\Theta\int_{\Si}\Bigl(tr(g^{-1} d g\wedge g^{-1}\om
3158: g)-tr(a\wedge\na_{a^g}\phi)+2tr(F_{a^g} \wedge \phi) \Bigr),
3159: \end{multline}
3160: where $\om\in\Om^1(End E_h)$ and $\phi\in\Om^0(\End
3161: E_h)=Lie(\sG_h)$.
3162:
3163: But the restriction of this differential to the subspace of flat
3164: connections is much simpler:
3165: \begin{equation}
3166: d \Theta (\om, \phi)=\frac{\pi i}{4}\,\Theta\int_{\Si} tr(g^{-1}d g
3167: \wedge g^{-1}\om g),
3168: \end{equation}
3169: and is independent of the second coordinate.
3170:
3171: That this function is a cocycle ndeed results from the functional
3172: equation
3173: \begin{equation}
3174: \Theta (a, g_1 g_2)=\Theta (a, g_1) \Theta (a^{g_1}, g_2).
3175: \end{equation}
3176: Using this function as a torsor $\sA_h \times_{\Theta}U(1)$, we
3177: get a principal $U(1)$-bundle $S^1(L)$ on the orbit space
3178: $\sA_h/\sG_h$:
3179: \begin{equation}
3180: S^1(L)=(\sA_h\times S^1)/\sG_h,
3181: \end{equation}
3182: where the gauge group $\sG_h$ acts by
3183: \begin{equation}
3184: g(a, z)=(a^g, \Theta(a,g) z),
3185: \end{equation}
3186: on the line bundle $L$ with a Hermitian structure.
3187:
3188: Following \cite{RSW}, let us restrict this bundle to the subspace
3189: of flat connections $\sA_F$. Then one can check that the
3190: restriction of the form $D$ (3.77) to $\sA_F$ defines a
3191: $U(1)$-connection $A_{CS}$ on the line bundle $L$.
3192:
3193: By definition, the curvature form of this connection is
3194: \begin{equation}
3195: F_{A_{CS}}=i\cdot\Om.
3196: \end{equation}
3197:
3198: Thus the quadruple
3199: \begin{equation}
3200: (CLRep(\pi_1(\Si), \SU(2)),\om, L, A_{CS})
3201: \end{equation}
3202: is a {\it prequantum system} and we are ready to perform the
3203: Geometric Quantization Procedure described in section 1.
3204:
3205:
3206:
3207:
3208:
3209: \subsection{Complex polarization of $CLRep(\pi_1(\Si), \SU(2))$}
3210:
3211: The standard way of getting a complex polarization is to give to
3212: Riemann surface $\Si$ of genus $g$ a conformal structure $I$.
3213: We get a complex
3214: structure on the space of classes of representations
3215: $CLRep(\pi_1(\Si), \SU(2))$ as follows: let $E$ be our complex vector
3216: bundle and $\sA$ the space of all
3217: connections on
3218: it. Every connection $a\in\sA$ defines the correspoding covariant
3219: derivative $\na_a\colon\Ga(E)\to\Ga(E\tensor T^*X)$, a first
3220: order differential operator with the ordinary derivative $d$ as
3221: the principal symbol. A complex structure gives the decomposition
3222: $\p+\op$, so any covariant derivative can be decomposed as
3223: $\na_a= \nabla^{1,0} + \na^{0,1}$,
3224: where
3225: $\na^{1,0}\colon\Ga(E)\to\Ga(E\tensor\Om^{1,0})$ and $\na^{0,1}
3226: \colon\Ga(E)\to\Ga(E\tensor\Om^{0,1})$. Thus the space of
3227: connections admits a decomposition
3228: \begin{equation}
3229: \sA=\sA'\times \sA'',
3230: \end{equation}
3231: where $\sA'$ is an affine space over $\Om^{1,0}(ad E)$ and $\sA''$
3232: is an affine space over $\Om^{0,1}(ad E)$.
3233:
3234: The group $\sG$ of all automorphisms of $E$ acts as the group of
3235: gauge transformations, and the projection $pr \colon \sA\to\sA''$
3236: to
3237: the space
3238: $\sA''$ of $\op$-operators on $E$ is equivariant with respect to
3239: the $\sG$-action.
3240:
3241: Giving $E$ a Hermitian structure $h$, we get the subspace
3242: $\sA_h\subset \sA$ of Hermitian connections, and the restriction
3243: of the
3244: projection $pr$
3245: to $\sA_h$ is one-to-one. Under this Hermitian metric $h$, every
3246: element $g\in\sG$ gives an element $\overline{g} = (g^*)^{-1}$
3247: such that
3248: \[
3249: \overline{g} = g \text{if and only if} g\in\sG_h.
3250: \]
3251: Now for $g\in\sG$, the action of $\sG$ on the component $\sA''$ is
3252: standard:
3253: \[
3254: \na^{0,1}_{g(a)}=g\cdot \na^{0,1}_a\cdot g^{-1}=\na^{0,1}_a-(\na^{o,1}_a g)
3255: \cdot g^{-1};
3256: \]
3257: and the action on the first component $\sA'$ of $\p$-operators is
3258: \[
3259: \na^{1,0}_{g(a)}=\overline{g}\cdot\na^{1,0}_a\cdot\overline{g}^{-1}=
3260: \na^{1,0}_a-((\na^{0,1}_a g)\cdot g^{-1})^*.
3261: \]
3262: It is easy to see directly that the action just described
3263: preserves unitary connections:
3264: \begin{equation}
3265: \sG(\sA_h)=\sA_h,
3266: \end{equation}
3267: and that the identification $\sA_h=\sA''$ is equivariant with
3268: respect to this action.
3269:
3270: It is easy to see that $(\na^{0,1}_a)^2 = 0\in\Om^{0,2}(ad E)$.
3271: Thus the orbit
3272: space
3273: \begin{equation}
3274: \sA'' /\sG=\bigcup \sM_i
3275: \end{equation}
3276: is the union of all components of the moduli space of
3277: topologically trivial $I$-holomorphic bundles on $\Si_I$. (Although this
3278: union doesn't admit any good structure, as it contains all
3279: unstable vector bundles, however we can use the notion {\it stack}
3280: for such situation .) Finally, the image of $\sA_F\in \sA_h$ is
3281: the component $M^{ss}$ of maximal dimension ($3g-3$) of s-classes
3282: of semi-stable vector bundles. Thus by classical technique of GIT
3283: of Kempf--Ness type we get:
3284:
3285: \begin{prop} {\it (Narasimhan--Seshadri)}
3286: \[
3287: CLRep(\pi_1(\Si), \SU(2))=M^{ss}.
3288: \]
3289: \end{prop}
3290:
3291: \begin{prop} The form $F_{A_{CS}}$ (3.85) is a $(1,1)$-form and
3292: the line bundle $L$ admits unique holomorphic structure
3293: $\Oh(\Theta)$ compatible with the Hermitian connection $A_{CS}$.
3294: \end{prop}
3295:
3296: On the other hand, complex structure $I$ on $\Si$ defines a
3297: K\"ahler metric on $M^{ss}$ (the so-called Weil--Petersson metric)
3298: with K\"ahler form
3299: \begin{equation}\om_{\mathrm{WP}}=i F_{A_{CS}}=i\cdot\om.
3300: \end{equation}
3301: This metric defines the Levi-Civita connection on the complex
3302: tangent bundle $T M^{ss}$, and hence a Hermitian connection
3303: $A_{LC}$ on the line bundle
3304: \begin{equation}
3305: \det T M^{ss}= \Oh(4 \Theta),
3306: \end{equation}
3307: and a Hermitian connection $\frac{1}{4}A_{LC}$ on $L$ compatible
3308: with the holomorphic structure on $L = \Oh(\Theta)$. Thus we have
3309:
3310: \begin{prop}
3311: \[
3312: \frac{1}{4} A_{LC}=A_{CS}.
3313: \]
3314: \end{prop}
3315:
3316: Finally, considering $M^{ss}$ as a family of $\op$-operators, we
3317: get the Quillen determinant line bundle $L$ with a Hermitian
3318: connection $A_Q$ with curvature form
3319: \begin{equation}
3320: F_{A_Q}=i\cdot\om.
3321: \end{equation}
3322: Hence we can extend the equality of Proposition 4:
3323: \begin{prop}
3324: \[
3325: \frac{1}{4} A_{LC}=A_{CS}=A_Q.
3326: \]
3327: \end{prop}
3328:
3329: Now we would like to describe Higgs fields in gauge terms.
3330:
3331: A holomorphic Higgs field $\phi$ is given by a section from
3332: $H^0(\Si_I, ad E \otimes K)$ and corresponding quadratic
3333: differential
3334: \begin{equation}
3335: \pi (E, \phi) = tr(\phi^2).
3336: \end{equation}
3337: Every quadratic differential $w$ is a holomorphic function on the
3338: dual space $\Om^{0,1}(\Si_I, K^{-1})$ of Beltrami differentials.
3339: This function is given
3340: by the integral
3341: \begin{equation}
3342: H_w = \int_\Si tr(\phi^2) \cdot \dot I.
3343: \end{equation}
3344: Obviously this function from the cotangent bundle
3345: $T^* M^{ss}$ can be extended to the full moduli space of Higgs pair
3346: $MH_\Si$.
3347:
3348: \begin{prop} For every two holomorphic quadratic differentials
3349: $w, w'$
3350: functions $H_w$ and $H_{w'}$ are Poisson commuting with respect
3351: to the holomorphic symplectic structure.
3352: \end{prop}
3353: Indeed this question can be lifted to the affine space of
3354: connections or Cauchy- Riemann operators on $E$ (3.87). The new
3355: affine space is
3356: \begin{equation}
3357: T^* \sA'' = \sA'' \times \Om^{0,1}(ad E).
3358: \end{equation}
3359: But every function $H_w$ as a function on this flat space
3360: doesn't depend on $\sA''$ - variables. But $\sA''$-variables and
3361: $\Om^{0,1}(ad E)$ - variables are conjugate with respect to the
3362: canonical symplectic form on the flat space. The zero-level of
3363: the moment map for the complex gauge group action (see (3.88)) is
3364: precisely the space of holomorphic Higgs fields. Thus $H_w$ and $H_{w'}$
3365: Poisson commute.
3366:
3367: \subsection{Computations of ranks}
3368:
3369: Computations of ranks in non-abelian setup is in a sense
3370: absolutely new procedure using a new idea - consideration
3371: of full collection of moduli spaces
3372: \begin{equation}
3373: M^{ss}_{2}, M^{ss}_{3}, \dots , M^{ss}_{g}, \dots
3374: \end{equation}
3375: for all genus. In the complex polarization case ideas are coming
3376: from the Donaldson theory for 4-folds (see, for example, \cite{T2}
3377: and \cite{T3}) where the second Chern class $c_{2}(E)$ is playing
3378: the role of the parameter $g$. In the real polarization case
3379: constructions are coming from Conformal Field Theory.
3380: \begin{rmk}
3381: It will be very useful to relate the Donaldson theory and CFT directly
3382: as it was done by Nakajima for Hilbert schemes $Hilb^{g} S$ of
3383: an algebraic surface $S$.
3384: \end{rmk}
3385:
3386: So let $\Si$ be a smooth algebraic curve of genus $g$ and
3387: \begin{equation}
3388: \sS (\Si)= Sym (H_{even}(\Si,\R)) \otimes \Lambda(H_{odd}(\Si, \R))
3389: \end{equation}
3390: be the graded algebra where $\deg(\si)=4-i$ if $\si \in
3391: H_i(\Si)$. Let $M^{ss}_g$ be the moduli space of rank 2
3392: semi-stable vector bundles on $\Si$ with trivial determinant. On
3393: the direct product $ \Si \times M^{ss}_g$ there isn't the
3394: universal bundle $\sE$ but there exists the bundle $\ad \sE$ with
3395: its Chern class
3396: \begin{equation}
3397: - p_1(\sE)= c_1(\sE)^2 - 4 c_2(\sE)= c_2(\ad \sE) \in H^4(\Si \times
3398: M^{ss}_g).
3399: \end{equation}
3400: This class is defined by the slant product, that is, the
3401: homomorphism $\mu\colon H_*(\Si)\to H^*(M^{ss}_g)$ defined by
3402: \begin{equation}
3403: \mu (\si)=-\frac{1}{4} p_1(\sE)/ \si.
3404: \end{equation}
3405: So, $ \deg(\al)=i \implies \mu(\al) \in H^i(M^{ss}_g)$.
3406:
3407: More precisely,
3408: \begin{equation}
3409: -p_1(\sE)=-2[\Si] \otimes [\Theta]+\pt \otimes x - 4
3410: \sum_{i=1}^g(a_i \otimes \mu(a_i)+b_i \otimes \mu(b_i)),
3411: \end{equation}
3412: where the cohomology class of the theta diviser
3413: \begin{equation}
3414: [\Theta] \in H^2(M^{ss}_g, \Z);\quad x \in H^4(M^{ss}_g, \Z);
3415: \end{equation}
3416: $[\Si]$ is the fundamental class of $C$; $[\pt]$ is the class of a
3417: point of $\Si$ and $ a_i, b_i, i=1,\dots,g$ is a symplectic basis
3418: of $H^1(\Si, \Z)$.
3419:
3420: \begin{dfn}
3421: The Newstead invariant is a linear function
3422: $$
3423: N_g \colon \sS (\Si) \to \Q
3424: $$
3425: sending a typical monomial $\al= \si_1 \cdot \cdot \cdot \si_r$
3426: where $\si_i \in H_{n_i}(\Si)$ of total degree
3427: $d(\al)=\sum_{i=1}^{r} (4-n_i)$
3428: to 0 if $d \neq 6g-6= dim M^{ss}_g$ and to the index of
3429: intersection
3430: \begin{equation}
3431: \mu(\si_1) \cdot \cdot \cdot \mu(\si_r)
3432: \end{equation}
3433: if $d=6g-6= dim M^{ss}_g$.
3434: Such homogeneous polynomial of degree $3(g-1)$ we will call {\it
3435: Newstead polynomial}.
3436: \end{dfn}
3437: \begin{rmk}
3438: Of course we can apply this constructions to any
3439: family of vector bundles, that is, to any vector bundle $\sE$ over
3440: $ \Si \times B $ where $B$ is a base of this family. Of course in this
3441: case we have one
3442: polynomial which we will denote by $ N(\sE)$.\end{rmk}
3443:
3444: Every Newstead polynomial is $Diff \Si$-invariant. From this we
3445: can see its shape:
3446: \begin{enumerate}
3447: \item $$
3448: \quad N_g( \mu([\Si])^m \cdot \mu([\pt])^n \cdot \mu(a_{i_{\min}
3449: })\cdot \mu(b_{j_{\min} }) \cdot \cdot \cdot) \neq 0 \implies i_{\min}
3450: =j_{\min}.
3451: $$
3452: Indeed, we use the commutation relations,
3453: the diffeomorphism invariance of
3454: the Newstead polynomials and the existence of an orientation
3455: preserving
3456: diffeomorphism $\phi \colon \Si \to \Si$ satisfying
3457: $$
3458: \phi^* ( a_i)= -a_i;
3459: \phi^*(b_i)=-b_i; \phi^*(a_j)=a_j, \phi^*(b_j)=b_j, i\neq j
3460: $$
3461: (the
3462: Dehn twist) and get the statement.
3463: \item The value
3464: $ N_g( \mu([\Si])^m \cdot \mu([\pt])^n \cdot \mu(a_{i_1
3465: })\cdot \mu(b_{i_1}) \cdots )$ is independent on the choice of
3466: handles $i_j$ because of
3467: the existence of an orientation preserving diffeomorphism
3468: interchanging any pair of handles;
3469: \item the expression $ \ga=2
3470: \sum_{i=1}^{g} \mu(a_i) \cdot \mu(b_i)$ is independent on the
3471: choice of symplectic basis $\{a_i, b_i \}$.
3472: \end{enumerate}
3473:
3474: From this it is easy to see that if
3475: \begin{equation}
3476: \ga =2 \sum_{i=1}^g \mu(a_i) \cdot \mu(b_i) \in H^6(M^{ss}_g )
3477: \end{equation}
3478: then
3479: \begin{equation}
3480: N_g \in \C[[\pt],[\Si], \ga].
3481: \end{equation}
3482: Thus this polynomial is nonzero only on the monomials of the type
3483: $ [C]^m \pt^n \ga^p;\text{ for } m+2n+3p=3g-3 $.
3484:
3485: Since the nondegenerate pairing (3.102) only involves the classes
3486: $[\Si], [\pt], \ga $ and does not depend on the conformal
3487: structure on $\Si$, we can reduce our of the graded algebra (3.97)
3488: to the "universal" algebra
3489: \begin{equation}
3490: \Bbb S_3= Sym ( \al \cdot \Z \oplus \be \cdot \Z \oplus \ga \cdot
3491: \Z),
3492: \end{equation}
3493: where $\deg \al=1; \deg \be=2, \deg \ga=3 $. Then the collection
3494: of the polynomials $N_g$ becomes the linear map
3495: \begin{equation}
3496: N \colon \Bbb S_3 \to \Q
3497: \end{equation}
3498: sending a monomial of degree $d$ not divisible by 3 to 0 and a
3499: monomial of degree $3n$ to $N_{n+1}$ (3.102) of this monomial. So
3500: we have non homogeneous polynomial $N$ with homogeneous components
3501: of degree $3g-3$.
3502:
3503: Moreover, there is the following recurrence relation between homogeneous
3504: components of $N$ (3.106):
3505: \begin{equation}
3506: N_g (\ga \cdot \al)=g \cdot N_{g-1}(\al)
3507: \end{equation}
3508: (see for example \cite{Th2}, Proposition 26). This is very
3509: important point: we get the inductive statement relating the
3510: Newstead polynomials for different genus.
3511: \begin{rmk}
3512: We will see later the fusion rule of the same type in the real
3513: polarization case.
3514: \end{rmk}
3515:
3516: It is convenient to use the following natural normalization
3517: \begin{equation}
3518: N^0_{3n}=\frac{1}{(n+1)!}
3519: N_{n+1}.
3520: \end{equation}
3521: We call such function the normalized Newstead polynomial. For
3522: this polynomial
3523: \begin{equation}
3524: N^0\colon \Bbb S_3 \to \Q
3525: \end{equation}
3526: we have the equality
3527: \begin{equation}
3528: N^0(\ga \cdot z) = N^0(z)
3529: \end{equation}
3530: for the generator $\ga$ and arbitrary $z \in \Bbb S$.
3531: So we can exclude the generator $\ga$ and consider $N^0$ as a
3532: linear map
3533: \begin{equation}
3534: N^0 : \Bbb S_2= Sym_*(\al \cdot \Z \oplus \be \cdot \Z) \to \Q.
3535: \end{equation}
3536: The so-called Newstead Conjecture ( see [Th2]) predicts that
3537: \begin{equation}
3538: N^0(\al^m \cdot \be^k) \neq 0 \implies k \le \frac{1}{3}(m+2k).
3539: \end{equation}
3540: So it is
3541: convenient to let $\al\cdot\be=\om$ and such that the Newstead
3542: polynomial is
3543: \begin{equation}
3544: N^0 : Sym_*(\al \cdot \Z \oplus \om \cdot \Z) \to \Q
3545: \end{equation}
3546: satisfying the relation
3547: \begin{equation}
3548: N^0(\al^{3k} \cdot \om^{n})=N_{3(n+k)+1}(\al^{n+3k } \cdot \be^n).
3549: \end{equation}
3550: From this we have the main equality:
3551: \begin{equation}
3552: N^0(\al^{3k} \cdot \om^{n})=\frac{(-4)^{n+k}\cdot
3553: (n+3k)!}{(n+k+1)!} \cdot (2k)! \cdot(4^k-2)\cdot B_{2k},
3554: \end{equation}
3555: where $B_i$ is the $i$th Bernoulli number (see \cite{Th2}, formula
3556: (29)).
3557:
3558: For any linear form $L \in
3559: Sym_*(\al \cdot \Z \oplus \be \cdot \Z) \oplus \ga \cdot \Z$, the
3560: interior product by any homogeneous element $x$ gives the differential
3561: operator
3562: \begin{equation}
3563: \frac{\partial}{\partial x} L(z)=(\deg z+\deg x) L(x \cdot z).
3564: \end{equation}
3565: Thus we can consider the relations of the type (3.110)
3566: as a differential equation. From this we get the Newstead
3567: polynomials as solutions to such equations. Moreover we
3568: recognize the
3569: generating functions of the pairing (3.102) as the standard elementary
3570: functions (see for example \cite{Th1}).
3571:
3572: \begin{rmk} \begin{enumerate} \item
3573: For the simplest graded algebra $ Sym_*(\Z \oplus \cdots
3574: \oplus \Z)$ the product $ z_1 \cdot z_2 \in Sym_{d_1+d_2}$of
3575: two homogeneous elements $ z_i \in Sym_{d_i}$ is defined as
3576: $$
3577: z_1 \cdot z_2(\si_1, \dots, \si_{d_1+d_2})=
3578: $$
3579: \begin{equation}
3580: = \frac{1}{(d_1+d_2)!}\sum_{g \in \Sigma_{d_1+d_2}}
3581: z_1(\si_{g(1)}, \dots, \si_{g(d_1)}) \cdot z_2(\si_{g(d_1+1)},
3582: \dots, \si_{g(d_1+d_2)}),
3583: \end{equation}
3584: where $\Sigma_n$ is the symmetric group on $n$ letters
3585: \item Witten's approach deals with the volumes of moduli spaces
3586: \begin{equation}
3587: vol (M_{g}^{ss}) = N^{0}(\al^{3g-3})
3588: \end{equation}
3589: and the other values of the Newstead polynomials can be read off
3590: from these formulae.
3591: \item Historically the
3592: procedure
3593: was the following: Thaddeus reconstructed values of the Newstead
3594: polynomial (3.115)
3595: from the beautiful form of the generating functions for
3596: ranks
3597: of conformal block spaces given by the Verlinde formula
3598: (see below).
3599: \item To use these computations Thaddeus proposed the following
3600: conjecture: {\it The quantization of $M^{ss}_{g}$ is numerically
3601: perfect} (see subsection 1.5).
3602: \item To avoid Thaddeus's conjecture we use the
3603: Donaldson's construction from \cite{Do}.
3604: \item Other regular way to get the pairing formuli and the computation of
3605: the collection of Newstead polynomials is the Geometric
3606: Approximation Procedure: there is a chain of birational
3607: transformations or flips (see \cite{Re})
3608: \begin{equation}
3609: \text{MP}_{\max}\leftrightarrow\dots\leftrightarrow\text{MP}_1
3610: \leftrightarrow\text{MP}_0,
3611: \end{equation}
3612: where every $\text{MP}_i$ is the moduli space of $t$-stable pairs
3613: (see \cite{Th2}, \cite{BD} or \cite{Ber}), and can be realised as
3614: a family of vector bundles over $\Si$. Thus using the same
3615: construction we can define polynomials $NP_{g,i}$ analogeous of
3616: the Newstead polynomials, and the linear forms $NP_i$, the
3617: analogs of (3.114) and (3.115). Recall that the moduli space on
3618: the extreme left admits the $\PP^{1}$-bundle structure that is
3619: regular map
3620: \begin{equation}
3621: \text{MP}_{\max}\to M_g^{ss}
3622: \end{equation}
3623: to the moduli space of stable bundle with trivial determinant.
3624: From this it is easy to see that the form
3625: \begin{equation}
3626: NP^{0}_{\max}\cdot H=N^0
3627: \end{equation}
3628: where $H$ is the Grothendieck class of the $\PP^{1}$-bundle,
3629: $N^{0}$ is the Newstead form.
3630: \item
3631: Now the moduli space $\text{MP}_0$ on the extreme right-hand end
3632: of (3.119) is the projective space of dimension $3g-2$.
3633: \begin{equation}
3634: \text{MP}_0=\PP^{3g-2}.
3635: \end{equation}
3636: \item Using the description of every flip in the chain (3.119) we can
3637: reconstruct all polynomials step by step begining with
3638: $\PP^{3g-2}$.
3639: \end{enumerate}
3640: \end{rmk}
3641:
3642: Using the Hirzebruch-Riemann-Roch theorem we have
3643: \begin{equation}
3644: \sum_{i} rk H^{i}(M_{g}^{ss}, \Oh(k \Theta)) = (ch(\Oh(k\Theta))
3645: \cdot td (M_{g}^{ss}))_{3g-3}.
3646: \end{equation}
3647: Since the divisor $\Theta$ is positive the LHS of previous
3648: equality is equal to
3649: $$
3650: rk \sH^{r}_{I} = rk H^{0}(M_{g}^{ss}, \Oh(k \Theta))
3651: $$
3652: by Kodaira vanishing theorem. On the other hand, Newstead showed
3653: (see \cite{Ne}) that
3654: \begin{equation}
3655: c_{1}(M^{ss}) = 4 \cdot [\Theta] \quad \text{ and } p(M_{g}^{ss})
3656: = (1 + \beta)^{2g-2}.
3657: \end{equation}
3658:
3659: Hence the Todd class
3660: \begin{equation}
3661: td(M^{ss}) = exp(2 [\Theta]) (\frac{1/2 \sqrt{\beta}}{sinh(1/2
3662: \sqrt{\beta})})^{2g-2}
3663: \end{equation}
3664: Using values of the Newstead polynomials we get the RHS of (3.123)
3665: :
3666: \begin{equation}
3667: rk H^{0}(M_{g}^{ss}, \Oh(k \Theta)) = (\frac{k+2}{2})^{g-1}
3668: \sum_{j}\frac{1}{sin (\frac{j \pi}{k+2})^{2g-2}}.
3669: \end{equation}
3670:
3671: \begin{rmk}
3672: All of the computations are due to Zagier and Macdonald (see
3673: \cite{Th1} for the corresponding references).
3674: \end{rmk}
3675:
3676: \subsection{Hitchin's connections}
3677:
3678: Here we will see that
3679: \begin{enumerate}
3680: \item The sending $ \Si_I$ to $ M^{ss}(\Si_I)$ (3.1) is a
3681: faithful functor and
3682: \item the sending $\Si_I \to H^0(M^{ss}(\Si_I), \Oh(k \Theta))$ is a
3683: successful quantization.
3684: \end{enumerate}
3685: That is, for $g > 2$ we can reconstruct $\Si_I$ from $M^{ss}$ and
3686: every vector bundle from the collection of vector bundles on the
3687: moduli space $\sM_{g}$ of curves of genus $g$
3688: \begin{equation}
3689: p_k \colon V_k \to \sM_g ; \quad p_k^{-1}(\Si_I) = H^0(M^{ss}(\Si_I),
3690: \Oh(k \Theta))
3691: \end{equation}
3692: admits a projective flat connection. So, locally the
3693: projective space $\PP H^0(M^{ss}(\Si_I), \Oh(k \Theta))$ doesn't
3694: depend on $I$.
3695:
3696: The first statement is almost obvious since
3697: \begin{equation}
3698: Sing M^{ss} = K(\Si_I)
3699: \end{equation}
3700: is the Kummer variety of our curve $\Si_I$. Blowing up the
3701: finite set $Sing K(\Si_I)$, considering the double cover
3702: with ramification in the exceptional divisor and then blowing
3703: down exceptional
3704: divisor we get the Jacobian $J(\Si)$. It is easy to see that
3705: the restriction of non-abelian theta divisor $\Theta$ to
3706: the Kummer variety gives the divisor $2 \Theta$ on the
3707: Jacobian.
3708: Now the classical Torelli theorem gives us our curve $\Si_I$.
3709:
3710: Recall that in the abelian case under deformations of
3711: a complex structure on $\Si$ we get vector bundles
3712:
3713: $\pi \colon V_k \to \sM_g, \quad \pi^{-1}(\Si) = H^0(J(\Si), \Oh
3714: (k \Theta))) $.
3715:
3716: It is the classical result ( we saw this already in
3717: terms of the extended Kodaira-Spencer theory \cite{W}) that $V_k$
3718: (1.29) admits a projective flat connection.
3719:
3720: In parallel to this fact Hitchin proved that even for
3721: non-abelian
3722: theta function case $V_k$ admits
3723: a projective flat connection.
3724:
3725: To construct the holomorphic projective connection we can use
3726: the extended Kodaira-
3727: Spencer theory again. First of all our goodness conditions (1.48)
3728: holds:
3729: \begin{equation}
3730: H^0(T M^{ss}) = H^1(M^{ss}, \Oh) = 0
3731: \end{equation}
3732: by Narasimhan-Ramanan result \cite{NR}. Now we have to construct
3733: a holomorphic section $W_0 \in H^0(S^2 (T M^{ss}))$. To do this,
3734: consider a semi-stable vector bundle $E \in M^{ss}$. Then the fiber
3735: \begin{equation}
3736: T M^{ss}_E = H^1(\Si, \ad E)
3737: \end{equation}
3738: where $ad E \subset End E$ is the traceless part of the sheaf of
3739: endomorphisms. The
3740: dual space (using the Serre duality)
3741: \begin{equation}
3742: T^* M^{ss}_E = H^0(\Si, \ad E \otimes K_\Si)
3743: \end{equation}
3744: where $K_\Si$ is the canonical class of $\Si$.
3745:
3746: Now a tangent vector to the moduli space $\sM_g$ of
3747: Riemann surfaces at a point
3748: $\Si_I$ is given by a Beltrami differential
3749: \begin{equation}
3750: \dot I \in H^1(\Si_I, K_\Si^*).
3751: \end{equation}
3752: We have the cup-product map
3753: \begin{equation}
3754: H^1(\Si_I, K_\Si^*) \otimes H^0(\Si, \ad E \otimes K_\Si) \to
3755: H^1(\Si, \ad E)
3756: \end{equation}
3757: which is nothing else but a quadratic form on $T^* M^{ss}_E$ and
3758: being extended on full moduli space $M^{ss}$ it gives the required
3759: section of the symmetric power of the tangent bundle. Using the
3760: standard machinery of the end of section 5 we get
3761: a holomorphic projective connection.
3762:
3763: The next question is about the curvature. Here it is quite
3764: reasonable to return to the gauge theory terminology.
3765:
3766: Recall that the affine space of all unitary connections $\sA_h$
3767: is canonically a symplectic manifold. Since the tangent space to
3768: $\sA$ at a connection $a \in \sA$ is given by the space of forms
3769: $\Om^1(\fsu(2))$, the symplectic form is given by the formula
3770: (3.74). This form doesn't depend on $a$, so it is closed. The
3771: gauge group $\sG$ acts on the space $\sA_h$ in the usual way and
3772: preserves the form $\om$.
3773:
3774: For the moment map (3.70) the symplectic quotient is
3775: \begin{equation}
3776: m^{-1} (0) / \sG = CLRep(\pi_1(\Si), \SU(2)).
3777: \end{equation}
3778: (see (3.70)). The symplectic reduction of the symplectic form
3779: defines a symplectic structure on this space. Thus
3780: \begin{equation}
3781: (CLRep(\pi_1(\Si), \SU(2)), \om)
3782: \end{equation}
3783: is a phase space of the classical mechanical system. Moreover we
3784: have the prequantization line bundle (3.83) with the connection
3785: (3.85).
3786:
3787: For every connection $a$ its covariant derivative $\nabla_a$
3788: defines the complex
3789: \begin{equation}
3790: \Om^0(\fsu(2)) \to \Om^1(\fsu(2)) \to \Om^2(\fsu(2))
3791: \end{equation}
3792: and the tangent space to $CLRep(\pi_1(\Si),\SU(2))$ at a point $a$
3793: is given as the first cohomology group of this complex
3794: \begin{equation}
3795: T CLRep(\pi_1(\Si), \SU(2))_a = H^1_a.
3796: \end{equation}
3797:
3798: Kahler polarization is just the choice of a conformal structure
3799: which can be done using
3800: the Hodge star operator $*$ acting on the space of forms
3801: \begin{equation}
3802: * \colon \Om^1 \to \Om^1.
3803: \end{equation}
3804: Our complex structure $I$ acts as a endomorphism
3805: \begin{equation}
3806: I = - *
3807: \end{equation}
3808: and
3809: \begin{equation}
3810: \om( \ga, I \ga) = - \om( I \ga, \ga) = - \int_\Si tr(\ga, *\ga).
3811: \end{equation}
3812:
3813: Such complex structure defines the decomposition
3814: \begin{equation}
3815: \nabla_a = \nabla^{1,0} + \nabla^{0,1}
3816: \end{equation}
3817: for every $a \in \sA$ and identifies $\sA$ with the complex
3818: affine space
3819: of Cauchy-Riemann operators on $E$ taking (0,1)-part of $\nabla_a$. Thus the
3820: the tangent space
3821: \begin{equation}
3822: T^{1,0}_a CLRep(\pi_1(\Si), \SU(2))_I = \Om^{0,1} (\fsu(2)).
3823: \end{equation}
3824: This complex structure is invariant with respect to the gauge
3825: group $\sG$ action thus we get an integrable complex structure as
3826: the result of the symplectic reduction (3.89).
3827:
3828: This complex structure $*$ on $\Si$ induces the complex structure
3829: $I$ on the space of classes of representations $CLRep(\pi_1(\Si),
3830: \SU(2))$ and we have the main identification
3831: \begin{equation}
3832: CLRep(\pi_1(\Si))_I = M^{ss} (\Si_{-*}).
3833: \end{equation}
3834:
3835: The last remark which we make before the curvature computations is
3836: following:
3837: \begin{prop} The map
3838: \begin{equation}
3839: f \colon H^1(\Si_I, T^{1,0}) \to H^0(M^{ss}, S^2(T^{1,0} M^{ss}))
3840: \end{equation}
3841: given by the formula
3842: \begin{equation}
3843: f(\dot I)(\phi) = \int_\Si tr(\phi^2) \wedge \dot I
3844: \end{equation}
3845: is an isomorphism.
3846: \end{prop}
3847:
3848: Indeed by the Hartog's principle we can extend every holomorphic
3849: quadratic form from $T^* M^{ss}$ to $MH_\Si$. But every
3850: holomorphic form is constant on the fibers of the map $\pi$
3851: (3.54) (which are compact abelian varieties). Hence the space of
3852: holomorphic quadratic forms is the dual space to the space of
3853: holomorphic quadratic differentials on $\Si_I$ and we are done.
3854:
3855: Now let $0 \in B \subset \C^m$ be an open set and
3856: \begin{equation}
3857: p \colon \tilde{M^{ss}} \to B , \quad p^{-1} (0) = M^{ss}_I
3858: \end{equation}
3859: be a body of holomorphic deformations of complex structure.
3860: Consider a section
3861: \begin{equation}
3862: t \in H^0(B, TB)
3863: \end{equation}
3864: that is a holomorphic vector field on $B$. By the classical
3865: Kodaira-Spencer
3866: theory it define a section
3867: \begin{equation}
3868: \xi \in H^0(B, R^1 T\tilde{M^{ss}}_{/p})
3869: \end{equation}
3870: that is the family of Kodaira-Spencer classes for deformations
3871: along $t$. Moreover considering our vector field $t$ as a
3872: differential operator on elements of a small open cover of fibers
3873: we get the section
3874: \begin{equation}
3875: D_t \in H^0(B, R^1(D^1(L))_{}/p)
3876: \end{equation}
3877: that is the family of cohomology classes $H^1(M^{ss}_b, D^1(L))$
3878: over points of $B$. This section describes deformations of a
3879: polarization line bundles over $t$.
3880:
3881: Having the holomorphic quadratic form (3.133)
3882: \begin{equation}
3883: W \in H^0(B, R^0(S^2(T_{/p})
3884: \end{equation}
3885: we can lift it to the family of Laplace-Beltrami operators
3886: $\Delta$ (using the Levi-Civita
3887: connection for the metric (3.140)) over elements of open covers
3888: of fibers of $p$,
3889: and their differences define
3890: \begin{equation}
3891: D \in H^0(B, R^1(D^2(L)_{/p}))
3892: \end{equation}
3893: because the operators of second order have the same principal
3894: symbols. Described in (1.86) general construction gives us the
3895: projective connection
3896: on the vector bundles $V_k \vert_B$ (3.146) with the principal symbol
3897: \begin{equation}
3898: - \frac{1}{2k - 4} \si \cdot \de (W) \in H^0(B, R^1
3899: T\tilde{M^{ss}}_{/p}).
3900: \end{equation}
3901: Moreover, the combination of these operators
3902: \begin{equation}
3903: \p_t + 1/(2k -4) i \Delta +i D
3904: \end{equation}
3905: gives globally defined holomorphic {\it heat operator} on $L_I$
3906: \begin{equation}
3907: \p_t + P_t
3908: \end{equation}
3909: and the principal symbol of $P_t$ is a quadratic form lifted from
3910: the target of the moment map $\pi$ (3.93). Thus by Proposition 14
3911: these symbols commute as functions on the cotangent
3912: bundle.
3913:
3914:
3915: What we have to do now is just to compute the commutator of these
3916: operators for two holomorphic vector fields $t$ and $t'$ from
3917: (3.147):
3918: \begin{equation}
3919: \p_t P_{t'} - \p_{t'} P_t - [P_t, P_{t'}].
3920: \end{equation}
3921: The last term commutator is of order 3 a priori. But its
3922: principal symbol is the Poisson brackets of symbols of operators
3923: (3.154)
3924: (see \cite{GS1}
3925: ). They are Poisson commuting. Thus this commutator
3926: \begin{equation}
3927: \p_t P_{t'} - \p_{t'} P_t - [P_t, P_{t'}] \in H^0(M^{ss}, D^2(L))
3928: \end{equation}
3929: for any complex structure from $B$.
3930:
3931: \begin{prop} This operator has to be constant.
3932: \end{prop}
3933:
3934: Indeed, in the cohomology sequence of the triple (1.81) there is
3935: the part
3936: \begin{equation}
3937: \to H^0(M^{ss}, S^2 T) \to H^1(M^{ss}, D^1(L))
3938: \end{equation}
3939: which is a monomorphism because the combination of this
3940: homomorphism with the symbol projection $H^1(M^{ss}, D^1(L)) \to
3941: H^1(T)$ is an isomorphism by Proposition 15 (3.144). Thus the
3942: beginning part of this exact sequence
3943: \begin{equation}
3944: H^0(M^{ss}, D^1(L)) \to H^0(M^{ss}, D^2(L)) \to 0.
3945: \end{equation}
3946: In particular,
3947: \begin{equation}
3948: H^0(M^{ss}, D^1(L)) = H^1(M^{ss}, D^2(L))
3949: \end{equation}
3950: and in the same vein
3951: \begin{equation}
3952: H^0(M^{ss}, D^1(L)) = \C
3953: \end{equation}
3954: Thus the commutator in (1.154) is a constant and our projective
3955: connection is projectively flat.
3956:
3957: Problem: is this connection unitary? Or is there another unitary
3958: projective flat
3959: connection?
3960:
3961: Summarizing that we have
3962:
3963: \begin{thm} The Kahler quantization of moduli spaces of topological trivial
3964: rk2 vector bundles is succesful.
3965: \end{thm}
3966:
3967:
3968:
3969:
3970: \section{Symplectic geometry of moduli spaces of vector bundles
3971: }
3972:
3973:
3974: \subsection{Goldman $U(1)$-action}
3975:
3976: The symplectic-geometrical part of the
3977: non-abelian theory of theta functions is well developed by W.
3978: Goldman, J. Weitsman,
3979: L. Jeffrey and J. Hurtubise.
3980:
3981: The starting point is the Narasimhan-Seshadri identification
3982: $ M^{ss}(\Si) = CLRep(\pi_1, \SU(2))$ as a result of the
3983: Symplectic Reduction described in the previous subsection:
3984: $\sA_h $ is the space of $\SU(2)$-connections, $\sG_h$ is
3985: the $\SU(2)$-gauge group, and
3986: $ \om (\ga_1, \ga_2) = \int_\Si tr (\ga_1 \wedge \ga_2)$ is
3987: the $\sG_h$-invariant symplectic form on $\sA_h$ (see (3.74)).
3988:
3989: The map $F \colon \sA \to \Om^2(ad)$ (3.66) sending
3990: connection to its curvature form is
3991: the moment map for $\sG_h$ - action. Hence
3992: $F^{-1}(0) = \sA_F$ is the space of flat connections and
3993: $( CLRep(\pi_1, \SU(2)), \om )$ is the result of
3994: the symplectic reduction with respect to this action.
3995:
3996: Any simple closed loop $C$ on $\Si$ defines the Goldman
3997: function on the space of classes of representations
3998: $CLRep(\pi_1, \SU(2))$:
3999: \begin{equation}
4000: c_C(\rho) = 1/\pi \cdot cos^{-1}(1/2 tr(\rho([C]))) \in [0, 1].
4001: \end{equation}
4002: In \cite{G2} Goldman proved that $c_C$ is a Hamiltonian of
4003: $U(1)$-action on $CLRep(\pi_1, \SU(2))$. This action can be
4004: described as follows:
4005: \begin{enumerate}
4006: \item for every $g \in \SU(2) , \quad g \neq \pm id$ there are exist
4007: unique element
4008: \begin{equation}
4009: log\quad g \in \fsu(2), \quad |log\quad g|^2 = 2,
4010: \end{equation}
4011: and a real number $t \in \R$ such that
4012: \begin{equation}
4013: g = e^{t \cdot log\quad g};
4014: \end{equation}
4015: \item obviously $[g, e^{t \cdot log\quad g}] = 0$ for every
4016: $t \in \R$.
4017: \item Suppose that the class $[C] = a_1 $ in a standard presentation of
4018: the fundamental group of $\Si$:
4019: \begin{equation}
4020: \pi_1 = <a_1, \dots ,a_g, b_1, \dots , b_g \vert
4021: \prod_{i=1}^g [a_i, b_i] = id>,
4022: \end{equation}
4023: \item for any $\rho \in CLRep(\pi_1, \SU(2))$
4024: such that $\rho (a_1) \neq \pm id$ the formula
4025: \begin{equation}
4026: \rho (b_1) \to \rho (b_1) \cdot e^{t \cdot log \rho(a_1)}
4027: \end{equation}
4028: defines Hamiltonian $U(1)$-action on $CLRep(\pi_1, \SU(2))$.
4029: \end{enumerate}
4030: It is easy to see that
4031: \begin{equation}
4032: C_1 \cap C_2 = \emptyset \implies \{c_{C_1}, c_{C_2} \}_\om = 0
4033: \end{equation}
4034: where the Poisson brackets are taken with respect to the canonical
4035: symplectic structure $\om$ (3.74). Thus we have to require a
4036: maximal set of disjoint homotopy inequivalent loops which we call
4037: circles . It easy to see (you will see this in a minute) that the
4038: maximal number of circles in such set is equal to 3g-3. Moreover,
4039: let a maximal set of disjoint inequivalent circles be
4040: \begin{equation}
4041: \{ C_1, ... , C_{3g-3} \}.
4042: \end{equation}
4043:
4044: \begin{figure}[tbn]
4045: \centerline{\epsfxsize=3in\epsfbox{27c.eps}} \caption{\sl Maximal
4046: collection of simple loops} \label{Fig 2}
4047: \end{figure}
4048:
4049:
4050: Then removing these circles we get
4051: \begin{equation}
4052: \Si - \{ C_1, ... , C_{3g-3} \} = \cup_{i=1}^{2g-2} \tilde{v_i}
4053: \end{equation}
4054: is a finite set of trinions (or "pairs of pants").
4055:
4056: \begin{figure}[tbn]
4057: \centerline{\epsfxsize=3in\epsfbox{28c.eps}} \caption{\sl Pairs
4058: of pants} \label{Fig 2}
4059: \end{figure}
4060:
4061: A trinion (or "pair of pants") plays the role of points
4062: in algebraic geometry under investigations of geometry of
4063: connections: just like we restrict functions to a point we can
4064: restrict connections to a trinion. Roughly speaking, a trinion is
4065: the minimal geometrical object admitting non trivial non-abelian
4066: gauge theory.
4067:
4068: Riemann surface with a trinion decomposition (4.8) is called
4069: {\it marked} Riemann surface.
4070: Any marked surface defines the {\it dual
4071: trivalent graph} $\Ga$
4072: \begin{enumerate}
4073: \item the set of vertices of which
4074: \begin{equation}
4075: V(\Ga) = \{ v_i \} = \{ \tilde{v_i} \}
4076: \end{equation}
4077: is the set of trinions;
4078: \item and with the set of edges
4079: \begin{equation}
4080: E(\Ga) = \{ e_i \} = \{ C_i \};
4081: \end{equation}
4082: \item and two vertices $v_i$ and $v_j$ are joined by the
4083: edge $e_l$ iff the corresponding circle
4084: \begin{equation}
4085: C_l = \p \tilde{v_i} \cap \p \tilde{v_j}.
4086: \end{equation}
4087: \end{enumerate}
4088:
4089:
4090: \subsection{Real polarization}
4091:
4092: Every marked Riemanian surface $\Si_\Ga$ defines a real
4093: polarization of the prequantized phase space
4094: \begin{equation}
4095: (CLRep(\pi_1, \SU(2)),\om, L, A_{CS}).
4096: \end{equation}
4097:
4098: It is given in a very geometric way in the set-up of perturbation
4099: theory of 3-dimensional Chern--Simons theory. As we saw, a mark
4100: of Riemann surfaces is given by the choice of maximal
4101: collection of disjoint, noncontractable, pairwise nonisotopic
4102: smooth circles on $\Si$. Any such system contains $3g-3$ simple
4103: closed circles (4.7).
4104:
4105:
4106: The isotopy class of a trinion decomposition $\{C_i\}$ gives the
4107: map
4108: \begin{equation}
4109: \pi_\Ga \colon CLRep(\pi_1, \SU(2)) \to \R^{3g-3}
4110: \end{equation}
4111: with fixed coordinates $(c_1, \dots, c_{3g-3})$ such that
4112: \begin{equation}
4113: c_i (\pi_{\{C_i\}} (\rho))=\frac{1}{\pi} cos^{-1}(\frac{1}{2} tr
4114: \rho([C_i]))\in [0, 1].
4115: \end{equation}
4116: which are just Goldman functions (4.1).
4117:
4118: Thus
4119: \begin{enumerate}
4120: \item
4121: this map is the moment map for the Hamiltonian
4122: $U(1)^{3g-3}$-action on $CLRep(\pi_1, \SU(2))$, thus gives a real
4123: polarization of the system (4.12).
4124: \item The coordinates $c_i$ are {\it action coordinates} for this
4125: Hamiltonian system ( \cite{GS1}).
4126: \end{enumerate}
4127:
4128:
4129: These functions $c_i$ are continuous on all $CLRep(\pi_1,
4130: \SU(2))$ and smooth inside of the unit cube $[0,1]^{3g-3}$.
4131:
4132: Thus we have
4133: \begin{enumerate}
4134: \item The image of $CLRep(\pi_1, \SU(2))$ under $\pi_\Ga$ is
4135: a convex polyhedron
4136: \begin{equation}
4137: \Delta_\Ga\subset [0, 1]^{3g-3}.
4138: \end{equation}
4139: \item The symplectic volume of $CLRep(\pi_1, \SU(2))$ is equal to
4140: the Euclidean volume of $\Delta_\Ga $:
4141: \begin{equation}\int_{CLRep(\pi_1, \SU(2))}\om^{3g-3}=Vol
4142: \Delta_\Ga.
4143: \end{equation}
4144: \item The set of Bohr--Sommerfeld orbits of the real
4145: polarization $\pi_\Ga$ of level $k$
4146: \begin{equation}
4147: BS_{k}(\pi_{\Ga},CLRep(\pi_1, \SU(2)),\om,L,A_{CS})
4148: \end{equation}
4149: is the set of $1/2k$- integer points in
4150: the polyhedron $\Delta_\Ga$,
4151: and
4152: \begin{equation}
4153: lim_{k\to\infty} k^{3-3g}\cdot N_{k-BS}=
4154: \int_{CLRep(\pi_1, \SU(2))}\om^{3g-3}
4155: =Vol \Delta_\Ga
4156: \end{equation}
4157: just as predicted by general quantization rules.
4158: \end{enumerate}
4159:
4160: \begin{figure}[tbn]
4161: \centerline{\epsfxsize=3in\epsfbox{31c.eps}} \caption{\sl Moment
4162: map} \label{Fig 2}
4163: \end{figure}
4164:
4165:
4166: To see the statements about Bohr-Sommerfeld fibers remark that
4167: every vector $w\in \R^{3g-3}$ can be interpreted as a {\it
4168: differential $1$-form} on $\R^{3g-3}$, and by the usual
4169: construction using the symplectic form $\Om$, this defines a
4170: vector field $\xi_w$ tangent to the fibers of $\pi$. Integrating
4171: such vector fields, one defines the collection of transformations
4172: \begin{equation}
4173: \{t_w\}= \{e^{\xi_w} \} \subset Diff^+ (CLRep(\pi_1, \SU(2)).
4174: \end{equation}
4175:
4176: These transformations preserve the curvature form of the
4177: connection $A_{CS}$. Thus (because $CLRep(\pi_1, \SU(2))$ is
4178: simple connected), there exists a collection of gauge
4179: transformations $\al_w\in\sG_L$ of $L$ such that
4180: \begin{equation}
4181: (t_w)^*(A_{CS})=A_{CS}^{\al_w}
4182: \end{equation}
4183: (see \cite{AB} for the description of the gauge-orbit with fixed
4184: curvature form)
4185:
4186: We can view such gauge transformations as $U(1)$-{\it torsors},
4187: just as in the description of the formulas for classical theta functions
4188: for Abelian varieties (2.49). From this
4189: the subset $BS_{k} \subset \Delta_{\Ga}$
4190: of Bohr-Sommerfeld fibers of level $k$ is the subset of
4191: $1/2k$-integer points of $\Delta$.
4192:
4193: To describe the polyhedron $\Delta_{\Ga}$ we have to describe the
4194: space of flat $\SU(2)$-connections on a trinion $pp$ (pair of
4195: pants): let
4196: \begin{equation}
4197: \p pp = C_1 \cup C_2 \cup C_3.
4198: \end{equation}
4199: Then we have the map
4200: \begin{equation}
4201: \pi \colon CLRep(\pi_1(pp), \SU(2)) \to \R^3
4202: \end{equation}
4203: with fixed coordinates $(c_1, c_2, c_3)$ just like in (3.139). It
4204: is easy to check (see \cite{JW1}) the following fact
4205:
4206: \begin{figure}[tbn]
4207: \centerline{\epsfxsize=3in\epsfbox{32c.eps}} \caption{\sl Moment
4208: tetrahedron} \label{Fig 2}
4209: \end{figure}
4210:
4211: \begin{prop} \begin{enumerate}
4212: \item the image
4213: \begin{equation}
4214: \pi (CLRep(\pi_1(pp), \SU(2)) = \De_\Theta
4215: \end{equation}
4216: is the tetrahedron given by the following inequalities
4217: \begin{equation}
4218: 0 \leq c_i \leq 1, \quad |c_1-c_2| \leq c_3 \leq min(c_1 + c_2, 2
4219: - c_1 -c_2);
4220: \end{equation}
4221: \item $\pi$ is 1-1 map, smooth inside of the tetrahedron and continious
4222: everywhere.
4223: \end{enumerate}
4224: \end{prop}
4225:
4226: Using the dual 3-valent graph $\Ga$ of the trinion decomposition
4227: (4.8) we can describe the polyhedron $\De_\Ga$ (4.15):
4228:
4229: \begin{prop} The polyhedron $\De_\Ga$ is given by inequalities (4.24) for every
4230: triple $(c_i, c_j, c_k)$ of the coordinates in $\R^{3g-3}$ corresponding to
4231: a boundary of any trinion in the decomposition (4.8).
4232: \end{prop}
4233:
4234: \subsection{ Bohr-Sommerfeld fibers}
4235:
4236:
4237: In these terms we can describe Bohr-Sommerfeld fibers (using
4238: action coordinates $c_i$ ) as functions
4239: \begin{equation}
4240: w\colon E(\Ga) \to\frac{1}{2k}\{0,1,2,\dots,k\}
4241: \end{equation}
4242: on the collection of edges of the 3-valent graph $\Ga $ to the
4243: collection of $\frac{1}{2k}$ - integers, such that, for any three
4244: edges $e_{1}, e_{2}, e_{3}$ meeting at a vertex $v$, the following
4245: 3 conditions hold:
4246: \begin{enumerate}
4247: \item $w(e_{1}) + w(e_{2}) + w(e_{3})\in \frac{1}{k}\cdot \Z$;
4248:
4249: \item $w(e_{1}) + w(e_{2}) + w(e_{3}) \leq 1$;
4250:
4251: \item and
4252: \begin{equation}
4253: |w(e_{1})-w(e_{2})|\leq w (e_{3}) \leq min( w(e_{1}) + w(e_{2}),
4254: 2 - w(e_{1} - e_{2})).
4255: \end{equation}
4256: \end{enumerate}
4257: Such a function $w$ is called an {\it admissible integer weight of
4258: level} $k$ on the graph $\Ga$.
4259:
4260: \begin{figure}[tbn]
4261: \centerline{\epsfxsize=3in\epsfbox{34c.eps}} \caption{\sl
4262: Bohr-Sommerfeld fibers} \label{Fig 2}
4263: \end{figure}
4264:
4265: Let $W^{k}(\Ga)$ be the set of such weights.
4266:
4267: \begin{rmk}
4268: The geometrical meaning of the last inequality (4.26) is the
4269: following: it is the triangle inequality for triangles on
4270: 2-sphere. Thus if we do our graph flat that is fix a cyclic
4271: orientation around every vertex and write down around every
4272: vertex a triangle with sides orthogonal to edges and of lengths
4273: $w(e_{1}), w(e_{2}), w(e_{3}))$, identifying equal sides we get
4274: triangulated Riemann surface
4275: $\Si_{w}$ which is
4276: a finite combinatorial cover of $S^{2}$
4277: \begin{equation}
4278: \phi_{w} \colon \Si_{w} \to S^{2}.
4279: \end{equation}
4280: For small genus, for example for the graph $\Theta$, this cover
4281: has degree one and gives just a triangulation of the basic sphere
4282: $S^{2}$. But from metric point of view this is a $1/2k$-integer
4283: triangulation.
4284: \end{rmk}
4285: Here is the list of simple but very important properties of
4286: weights:
4287: \begin{prop}
4288: \begin{enumerate}
4289: \item the number $|W^k(\Ga)|$ of admissible weights of level $k$
4290: is independent of the graph $\Ga$ but depends on genus $g$ only;
4291: \item
4292: \begin{equation}
4293: |W^{k}(\Ga)| =|W_g^k| =\frac{(k+2)^{g-1}}{2^{g-1}}
4294: \sum_{n=1}^{k+1}
4295: \frac{1}{(\sin(\frac{n\pi}{k+2}))^{2g-2}}
4296: \end{equation}
4297: is the Verlinde number (3.126).
4298: \end{enumerate}
4299: \end{prop}
4300:
4301:
4302: To describe the geometry of Bohr-Sommerfeld fibers given by an
4303: admissible integer weight $w$ (4.25) we have to consider the
4304: levels of this function:
4305: for $\al\in \{0, \frac{1}{2k},\dots,1\}$ let
4306: \begin{equation}
4307: w^{-1}(\al)=\Ga_1(\al)\cup\cdots\cup\Ga_n(\al)\subset\Ga
4308: \label{eq5.1}
4309: \end{equation}
4310: be the decomposition into connected components. Then every
4311: component $\Ga_i(\al)$ is a 3-valent graph with $n_i$ univalent
4312: vertices $a_1,\dots, a_{n_i}$ (see section 6).
4313:
4314: Every gauge class of connections contains a connection $a_0$
4315: adapted to a trinion decomposition (see \cite{JW1},
4316: Definition~2.2). Fix the filtration
4317: \begin{equation}
4318: Z(\SU(2))=\Z_2\subset U(1)\subset \SU(2),
4319: \end{equation}
4320: and view it as the triple $ \{\Z_2, U(1), \SU(2)\}$
4321:
4322: For a connection $[a]\in \pi^{-1}(w)$, we have the function
4323: \begin{equation}
4324: e_w\colon E(\Ga)\to \{\Z_2, U(1), \SU(2)\}
4325: \end{equation}
4326: sending every loop $C_j$ to the element of $G$ conjugate to the
4327: stabilizer of the monodromy of $[a]$ around this loop, and the
4328: function
4329: \begin{equation}
4330: v_w\colon V(\Ga)\to \{\Z_2, U(1), \SU(2)\}
4331: \end{equation}
4332: sending trinion $P_i$ to the stabilizer of the flat connection
4333: $a\rest{P_n}$. Of course,
4334: \begin{align}
4335: & C_j\subset \partial P_n \implies v_w(P_n)\subset e_w(C_j); \notag \\
4336: & C_1 \cup C_2 \cup C_3=\partial P_n \quad \text{and} \quad
4337: e_w(C_1)=e_w(C_2)=\SU(2) \implies
4338: \\
4339: & \qquad\qquad e_w(C_3)=\SU(2) \implies v_w(P_n)=\SU(2), \notag
4340: \end{align}
4341: and so on.
4342:
4343: Obviously
4344: \begin{equation}
4345: e_w(C_j)= \text{$U(1)$ or $\SU(2)$.}
4346: \end{equation}
4347:
4348: \begin{prop} The functions $e_w$ and $v_w$ depend on $w$ and not on
4349: the choice of $[a]\in \pi^{-1}(w)$.
4350: \end{prop}
4351:
4352: More precisely, they depend on the combinatorics of the
4353: decomposition (4.29).
4354:
4355: Thus $w$ defines the direct products
4356: \[
4357: \prod_{C\in E(\Ga)} e_w(C) \quad\text{and}\quad
4358: \prod_{P\in V(\Ga)} v_w(P),
4359: \]
4360: and $\prod_{P\in V(\Ga)} v_w(P)$ acts on $\prod_{C\in E(\Ga)}
4361: e_w(C)$ as follows: for
4362: \begin{gather*}
4363: g=(g_1,\dots, g_{2g-2})\in \prod_{P\in V(\Ga)} v_w(P)
4364: \quad\text{with $g_i\in v_w(P_i)$, and}\\
4365: (t_1,\dots, t_{3g-3})\in \prod_{C\in E(\Ga)} e_w(C)
4366: \quad\text{with $t_n\in e_w(C_n)$,}
4367: \end{gather*}
4368: if $C_n\subset \partial P_i\cap\partial
4369: P_j$ then
4370: \begin{equation}
4371: g(t_n)=g_i \circ t_n \circ g_j^{-1}.
4372: \end{equation}
4373:
4374: \begin{rmk}
4375: This action described in detail in subsection 6.3 under the
4376: consideration of the gauge theory on graphs.
4377: \end{rmk}
4378:
4379: \begin{prop}[\cite{JW1}, Theorem~2.5] The fibre $\pi^{-1}(w)$ is given by
4380: \begin{equation}
4381: \pi^{-1}(w)=\prod_{C\in E(\Ga)} e_w(C) \Bigm/ \prod_{P\in V(\Ga)} v_w(P).
4382: \end{equation}
4383: \end{prop}
4384:
4385:
4386: \begin{cor}
4387: The fibre $\pi^{-1}(w)$ is isomorphic to
4388: \begin{equation}
4389: \pi^{-1}(w) = T^t \times [(S^3)^p \times (S^2)^s] / G_w,
4390: \end{equation}
4391: where $t, p$ and $s$ are nonnegative integers and $G_w$ is the
4392: finite Abelian group defined by $w$, or more precisely by the
4393: combinatoric data (4.1); moreover,
4394: \begin{equation}
4395: H_1(\pi^{-1}(w))=\Z^t \oplus \Z_2^p.
4396: \end{equation}
4397: \end{cor}
4398:
4399: Translations along the torus $T^t$ in (4.37) are induced by
4400: Hamiltonians lifted from the target space $\R^{3g-3}$ of $\pi$.
4401:
4402: Jeffrey and Weitsman \cite{JW1} use the normalization of the
4403: action coordinates via branched covers to construct a covariant
4404: constant section $s_w$ of the restrictions of $(L^k,A_{CS})$ to
4405: $\pi^{-1}(w)$.
4406:
4407:
4408:
4409:
4410:
4411:
4412:
4413:
4414: Following the real polarization quantization rules from section 1
4415: (formula (1.21)) the wave function space is given by the direct
4416: sum
4417: \begin{equation}
4418: \sH^{k}_{\Ga} = \oplus_{b \in BS_{k}} \C \cdot s_{b}
4419: \end{equation}
4420: where $s_{b}$ is a covariant constant section of the restriction
4421: of our line bundle and connection to $\pi^{-1}(b)$.
4422:
4423: Geometrically this combinatorial description of Bohr-Sommerfeld
4424: fibers can be done
4425: as follows: consider the space $\R^{3g-3}$ with action coordinates
4426: $c_i$ (4.14). This space contains the integer sublattice
4427: $\Z^{3g-3}\subset \R^{3g-3}$, and we can consider the {\it action
4428: torus}:
4429: \begin{equation}
4430: T^A=\R^{3g-3} / \Z^{3g-3}.
4431: \end{equation}
4432:
4433: In particular, we get a map
4434: \begin{equation}
4435: \pi_A \colon CLRep(\pi_1, \SU(2)) \to T^A
4436: \end{equation}
4437: which glues at most points of the boundary of $\Delta_{\Ga}$.
4438:
4439:
4440: Now every integer weight $w$ (4.25) satisfying (1) and (2), but
4441: {\it a priori} without the Clebsch--Gordan conditions (3) (4.26),
4442: defines a point of order $2k$ on the action torus
4443: \[
4444: w\in T^A_{2k}.
4445: \]
4446: In particular, the collection $W^k(\Ga)$ of admissible integer
4447: weights (subject to (3) (4.26)) can be considered as a subset of
4448: points of order $2k$ on the action torus:
4449: \begin{equation}
4450: W^k(\Ga)\subset T^A_{2k}.
4451: \end{equation}
4452: Thus for different $\Ga$ of genus $g$ we have {\it a priori}
4453: different subsets $\{W^{k}(\Ga)\}$ of the set of points of order
4454: $2k$ on the action torus $T^{A}$ (4.40).
4455:
4456: The linear decomposition of the wave function space (4.39)
4457: defines the space uniquely up to the coordinatewise
4458: $|W^k(\Ga)|$-torus action because every covariant constant
4459: section is defined up to $U(1)$-action. But using admissible
4460: integer weights we can define this space as a space
4461: with a fixed basis
4462: \begin{equation}
4463: \sH^{k}_{\Ga} = \oplus_{w \in W^{k}_\Ga} \C \cdot w
4464: \end{equation}
4465: and the integer sublattice
4466: \begin{equation}
4467: \Z^{|W^k_\Ga|} \subset \sH^{k}_{\Ga}.
4468: \end{equation}
4469: All of these spaces are embedded in one large space
4470: \begin{equation}
4471: \sH^{k}_{\Ga} \subset \sH^k = \oplus_{w \in T^A_{2k}} \C \cdot w
4472: \end{equation}
4473: (see (4.42)) with the integer lattice
4474: \begin{equation}
4475: \Z^{|T^A_{2k}|} \subset \sH^{k}.
4476: \end{equation}
4477:
4478:
4479: \subsection{Delzant model}
4480:
4481:
4482: Returning to the moment map (3.148) we come into the set-up of the
4483: toric symplectic geometry: let $(M, \om)$ be a compact
4484: symplectic smooth manifold of dimesion $2n$ with a {\it smooth}
4485: Hamiltonian action of $n$-dimensional torus $T^n$. Then
4486: action-angle coordinates define the moment map
4487: \begin{equation}
4488: \pi \colon M \to \De \subset \R^n
4489: \end{equation}
4490: whose image is a convex polyhedron $\De$ in Euclidean $n$-space.
4491: This polyhedron satisfies so called Delzant conditions and
4492: contains complete information on the symplectic geometry of $(M,
4493: \om)$. That is, $\De$ determines the manifold, the symplectic form
4494: and the $T^n$-action (see \cite{D}). Moreover, if a polyhedron
4495: $\De \subset \R^n$ satisfies the Delzant conditions then we can
4496: construct a smooth symplectic manifold with Hamiltomian
4497: $T^n$-action such that the image of the moment map is precisely
4498: $\De$.
4499:
4500: Moreover, if a phase space $(M, \om)$ is prequantized
4501: and $M$ has a Hodge structure whose Kahler form is $\om$, then
4502: this Hodge structure can also be reconstructed from $\De$ (see
4503: \cite{G1}).
4504:
4505: We saw that our space $CLRep(\pi_1(\Si), \SU(2))$ admits the
4506: Hamiltonian $T^{3g-3}$-torus action, but
4507: \begin{enumerate}
4508: \item our phase space $CLRep(\pi_1(\Si), \SU(2))$ is singular if $g>2$;
4509: \item the action isn't smooth over the boundary of $\De$.
4510: \end{enumerate}
4511:
4512: Nevertheless the polyhedron satisfies Delzant conditions hence
4513: determines the smooth compact Hodge manifold $DM_\Ga$ with the
4514: smooth $T^{3g-3}$-action which is called {\it Delzant model of
4515: moduli spaces} (see \cite{T8}). For a complex Riemann surface $\Si_I$ the moduli
4516: space $M^{ss}(\Si_I)$ is an algebraic variety. Every such variety
4517: admits a birational morphism
4518: \begin{equation}
4519: f \colon M^{ss}(\Si_I) \to DM_\Ga.
4520: \end{equation}
4521:
4522: The list of properties is the following
4523: \begin{enumerate}
4524: \item The Delzant model $DM_\Ga$ is canonically polarized by the line bundle
4525: $H$ corresponding to the theta polarization of every $M^{ss}$.
4526: \item The Delzant model is rigid as a complex manifold and hence
4527: the Kahler quantization of $(DM_\Ga, H)$ is succesful. The wave
4528: function spaces are given as spaces of holomorphic sections
4529: \begin{equation}
4530: H^0(DM_\Ga), H^k)
4531: \end{equation}
4532: and dimensions of these spaces can be computed by the
4533: Duistermaat-Heckman formula.
4534: \item Bohr-Sommerfeld fibers for $CLRep(\pi_1(\Si), \SU(2))$ and $DM_\Ga$
4535: as sets of points of $\De_\Ga$ coincide.
4536: \item The rank of wave spaces (4.49) is equal to the number
4537: of Bohr-Sommerfeld fibers
4538: thus the quantization of the dynamical system $(DM_\Ga, \om)$ is
4539: numericaly perfect (see \cite{T8}).
4540: \end{enumerate}
4541:
4542:
4543: \section{Two versions of CQFT}
4544:
4545: Result of any quantization procedure is creation of a quantum
4546: mechanics which is QFT in dimension d = 1 when there is only one
4547: variable - a time.
4548: But our main geometrical object - Riemann surfaces - are also the basic object
4549: of conformal field theories in dimension 2 (or, more precisely, (1,1)).
4550:
4551: Some years ago the constraint of conformal invariance was proposed as
4552: the fundamental principle of quantum field theory. All two-dimensional
4553: conformal field theories are duing to the theory of
4554: critical phenomena and to the string theory. So, the new powerful
4555: mathematical tool appeared since large data on the geometry of
4556: Riemann surfaces is encoded in this rich mathematical structure.
4557: Moreover, the results of a succesful Kahler quantization procedure
4558: = the wave function spaces $H^0(M^{ss}(\Si), \Oh ( k \Theta))$
4559: appeared first as spaces of conformal blocks of CQFT.
4560:
4561: On the
4562: other hand,
4563: the spaces
4564: $\sH_\pi^k = \oplus_{b \in BS_k \subset B} \C \cdot
4565: s_b$ as results of the Bohr-Sommerfeld quantization procedure
4566: appeared as fibers of projective flat vector bundles over
4567: the moduli space $\sM_g$ of curves of genus $g$ in the
4568: combinatorial version of CQFT in
4569: \cite{FS} and \cite{MS}. Moreover, a 2-dimensional conformal field theory
4570: was formulated as algebraic geometry of the moduli spaces of
4571: Riemann surfaces.
4572:
4573: A very important case of that a pinched Riemann sphere was
4574: investigated by Belavin-Knijnik-Zamolodchikov in \cite{KZ} which
4575: produce the general ideology as well as thechnical fondations.
4576:
4577: We can start with the simple "classical" version of CFT.
4578:
4579: \subsection{WZW-version}
4580:
4581:
4582: Recall that the space of classical theta functions has very
4583: important property to be a space of unique irreducible
4584: representation of some group or algebra (namely the Heisenberg
4585: group). The space of holomorphic sections $H^0(M^{ss}(\Si), \Oh (
4586: k \Theta))$ is also the space
4587: of the irreducible representation of the gauge algebra
4588: of WZW RCFT. Such spaces have played a noted role in conformal
4589: field theory.
4590:
4591: It turn out that in rational conformal field theories, one encounters the
4592: "conformal blocks" of Belavin, Polyakov and Zamolodchikov coincidencing with
4593: our spaces $H^0(M^{ss}(\Si), \Oh ( k \Theta))$ .
4594:
4595: First of all recall Wess-Zumino-Witten CQFT
4596: (WZW for short) set up: let $t$ be
4597: a local parameter in a point $p \in \Si$. Then
4598: the {\it loop algebra}
4599: \begin{equation}
4600: sl_2^t = sl(2, \C)\otimes \C[t, t^{-1}]
4601: \end{equation}
4602: is the {\it horisontal algebra} of WZW.
4603:
4604: The central extention by the cocycle $k \cdot c$
4605: \begin{equation}
4606: \ov{sl_2^t} = sl(2, \C) \otimes \C[t, t^{-1}] \oplus \C \cdot e
4607: \end{equation}
4608: is the k-{\it gauge algebra of} WZNW and the bracket of two
4609: elements of the loop algebra is given by the formula
4610: \begin{equation}
4611: [X \otimes f, Y \otimes g] = [X, Y] \otimes fg + k \cdot tr(XY)
4612: Res_0(gdf)
4613: \end{equation}
4614:
4615: The theory of Kac-Moody algebras tells us that this algebra admits
4616: unique irreducible representation $\sH^k$ such that there
4617: exists a non-zero vector annihilated by
4618: $sl(2, \C)\otimes \C[t]$. Remark that the central element
4619: of $\ov{sl_2^t}$ acts by multiplication by $k$ (which is the
4620: level number).
4621:
4622: On the other hand, the field $\C (\Si)$ of meromorphic functions
4623: on $\Si$ has the natural embedding:
4624: \begin{equation}
4625: i \colon sl(2, \C) \otimes \C (\Si) \to \ov{sl_2^{t}}.
4626: \end{equation}
4627: This field $\C (\Si)$ contains the subring $\C (\Si)_p$ of
4628: regular functions in $\Si - p$. The structure of algebra on
4629: \begin{equation}
4630: sl(2, \C) \otimes \C (\Si)_p
4631: \end{equation}
4632: is given by the formula
4633: \begin{equation}
4634: [X \otimes f, Y \otimes g] = [X, Y] \otimes fg
4635: \end{equation}
4636: The restriction of the embedding (4.4) to this subring realizes it as the
4637: subalgebra of $\ov{sl_2^t} $.
4638:
4639: Thus the algebra $sl(2, \C) \otimes \C (\Si)_p$ acts on the
4640: representation $\sH^k$
4641: described before.
4642:
4643: \begin{dfn} The space of conformal blocks $V_k$ of level $k$
4644: is the subspace of the dual
4645: $(\sH^k)^*$ annihilated by the Lie algebra
4646: $sl(2, \C) \otimes \C (\Si)_p$.
4647: \end{dfn}
4648:
4649: Why such space is related to vector bundles on $\Si$? First of
4650: all, we note, that every vector bundle $E$ on $\Si - p$ is
4651: trivial and of course it is trivial on a
4652: small disc $D_p$ containing $p$. Fixing
4653: these trivializations we send a vector bundle $E$ to the Laurent
4654: series
4655: \begin{equation}
4656: \ga_E \in SL(2, \C)\otimes \C[t, t^{-1}].
4657: \end{equation}
4658: To get rid of the trivializations, we have to mod out by the
4659: automorphisms groups of trivial bundles on $D_p$ and $\Si - p$.
4660: So the set of vector bundles is
4661: \begin{equation}
4662: \{ E \} = SL(2, \C) \otimes \C (\Si)_p \backslash \fsl(2,
4663: \C)\otimes \C[t, t^{-1}] / SL(2, \C)\otimes \C[t].
4664: \end{equation}
4665: Remark that we don't discuss the stability condition and our
4666: double coset space is so called {\it algebraic stack},
4667: not a moduli space.
4668:
4669: In any case our, stack $\{ E \}$ is the quotient of the universal
4670: homogeneous space
4671: \begin{equation}
4672: Q = SL(2, \C) \otimes \C[t, t^{-1}] / SL(2, \C)\otimes \C[t]
4673: \end{equation}
4674: and the line bundle $\Oh( \Theta)$ can be lifted to the universal
4675: line bundle $L_\chi$ on this homogeneous space. The theorem of
4676: Kumar and Mathueu provides the isomorphism
4677: \begin{equation}
4678: H^0(Q, L_\chi^k) = \sH^k
4679: \end{equation}
4680: and $H^0(M^{ss}, Oh(k \Theta))$ coincides with the subspace of
4681: $(\sH^k)^*$ invariant under Lie algebra
4682: $sl(2, \C) \otimes \C (\Si)_p$ action.
4683:
4684: To see that this space is finite dimensional over $\C$, we need
4685: only the standard experience with matrix divisors (for example,
4686: like in the survey \cite{T2}). But to compute the dimension we
4687: have to use more sophisticated versions of all objects.
4688:
4689: First of all, on our Riemann surface $\Si$ we choose a finite set
4690: of different points
4691: \begin{equation}
4692: p_1, p_2, ....., p_n \subset \Si
4693: \end{equation}
4694: different from our fixed point $p \in \Si$ and labeled by the set of
4695: irreducible representations of $sl(2, \C)$, that is, by positive
4696: integers $n_1, n_2, ...., n_n$ (twice spin-numbers).
4697:
4698: Now our algebra $sl(2, \C) \otimes \C (\Si)_p$ acts on the space
4699: $V_{n_i}$ of the irreducible representation $n_i$ by the formula
4700: \begin{equation}
4701: (X \otimes f) v = f(p_i) X v
4702: \end{equation}
4703: where $p_i$ is the point corresponding to the representation
4704: $n_i$ and $v \in V_{n_i}$.
4705:
4706: New vector space is given now by the formula
4707: \begin{equation}
4708: V_k (\{ (p_i, n_i) \})_\Si = Hom_{sl(2, \C) \otimes \C (\Si)_p}(
4709: \sH^k, \otimes_{i=1}^n V_{n_i}).
4710: \end{equation}
4711:
4712: These spaces have a number of good properties.
4713: For example, they don't depend on moduli of curves, the choice
4714: of points and so on. But the most important property is the
4715: behavior with respect to a simplest degeneration of
4716: curve $\Si$ to curve $\Si_0$ with one double point.
4717: Geometrically this means that we have a smooth curve $\Si_0$
4718: of genus $g-1$ with a fixed pair of points $(q_+, q_-)$.
4719: Then
4720: \begin{equation}
4721: V_k (\{ (p_i, n_i) \})_\Si = \oplus_{n \in \Z^+} V_k (\{ (p_i,
4722: n_i), (q_+, n), (q_-, n) \})_{\Si_0}.
4723: \end{equation}
4724: These properties provide the induction by genus (just like formula (3.107)
4725: ).
4726:
4727: Now we are interested only in ranks of previous spaces.
4728: We may forget about curves, points and others geometrical
4729: features (how we did after (3.109) in subsection 3.7) and consider the monoid $M$ of formal sums
4730: \begin{equation}
4731: N = n_1 + n_2 + ... + n_n \in M
4732: \end{equation}
4733: providing the present labeling of points. Every rank of space $V_k (\{ (p_i,
4734: n_i) \})_\Si$ can be done as a symbol
4735: \begin{equation}
4736: rk_g(N) , \quad \text{where}\quad N = n_1 + n_2 + ... + n_n \in
4737: M.
4738: \end{equation}
4739: The list of properties of these numbers is the following:
4740: \begin{enumerate}
4741: \item $rk_0(0) = 1;$
4742: \item $rk_g(N) = \sum_{n \in \Z^+}rk_{g-1}(N +n +n)$;
4743: \item let $g = g' + g''$, then
4744: \begin{equation}
4745: rk_g(N' + N'') = \sum_{n \in \Z^+} rk_{g'} (N' + n)\cdot
4746: rk_{g''}(N'' + n).
4747: \end{equation}
4748: \end{enumerate}
4749:
4750: Using these equalities, we can decent genus $g$ to 0 and get so
4751: called {\it fusion rules}:
4752: \begin{enumerate}
4753: \item $rk_0(0) = 1$;
4754: \item $rk_0(N' + N'') = \sum_{ n \in \Z^+} rk_0(N' + n) rk_0(N'' + n)$.
4755: \end{enumerate}
4756:
4757: These fusion rules give the structure of commutative ring on our
4758: monoid $M$:
4759: \begin{equation}
4760: N \cdot N' = \sum_{ n \in \Z^+} rk_0(N + N' + n) \cdot n.
4761: \end{equation}
4762:
4763: This commutative ring is called the {\it fusion ring}.
4764:
4765: \begin{rmk} We don't prove here the factorisation rules (5.17) in the algebraic
4766: geometrical set-up, that is, in the set-up of vector bundles. We
4767: will get them as bypass results at the end of our long story. But
4768: remark that there are exist approaches to prove these rules in
4769: purely geometrical way.
4770: \end{rmk}
4771:
4772: We saw that our fusion ring is very close to be the ring of
4773: irreducible representations $R(sl(2, \C))$ where the multiplication
4774: is the tensor product of representations. But there is
4775: a small difference coming from level $k$ condition.
4776:
4777: According to the Clebsh-Gordan rules, a tensor product of
4778: irreducible representation has the decomposition:
4779: \begin{equation}
4780: n_1 \otimes n_2 = (n_1 + n_2) \oplus (n_1 + n_2 - 2) \oplus ....
4781: \oplus (n_1 - n_2)
4782: \end{equation}
4783: (here $n_1 \geq n_2$). But the level condition is $n \leq k$ and
4784: the multiplication is given by the formulas:
4785: \begin{enumerate}
4786: \item $ n_1 + n_2 \leq k \implies n_1 \circ n_2 = n_1 \otimes n_2$;
4787: \item $n_1 + n_2 \geq k \implies n_1 \circ n_2 = (2k - n_1 - n_2) \oplus ... \oplus (n_1-n_2)$.
4788: \end{enumerate}
4789: From this it is easy to see that the fusion ring $M_k$ of level
4790: $k$ is the quotient of
4791: $R(sl(2, \C))$ by the ideal generated by $(k+1)$.
4792:
4793: The fusion ring $M_k$ has the element
4794: \begin{equation}
4795: c = \sum_{n=0}^k n \circ n.
4796: \end{equation}
4797:
4798: Let $Spec M_k$ be the set of characters (homomorphisms to $\C$)
4799: of $M_k$. Then after elementary computations we get the equality
4800: \begin{equation}
4801: rk_g(\emptyset) = \sum_{\chi \in Spec M_k} \chi(c)^{g-1}
4802: \end{equation}
4803: and
4804: \begin{equation}
4805: \chi(c) = \sum_{n=0}^k \vert \chi(n) \vert^2.
4806: \end{equation}
4807:
4808: Recall that every character of the representations ring $R(sl(2,
4809: \C))$ is given by $z \in \C$ and
4810: \begin{equation}
4811: \chi_z(n) = \frac{sin((n+1)z)}{sin z}.
4812: \end{equation}
4813: This character is from $Spec M_k$, that is, it vanishes on $(k+1)$
4814: iff
4815: \begin{equation}
4816: z = \frac{n \pi}{k+2}, \quad (1 \leq n \leq k+1).
4817: \end{equation}
4818: Now elementary computations and formula (8.35) give
4819: \begin{equation}
4820: rk_g(\emptyset) = rk V_k =\frac{(k+2)^{g-1}}{2^{g-1}}
4821: \sum_{n=1}^{k+1}
4822: \frac{1}{(\sin(\frac{n\pi}{k+2}))^{2g-2}}
4823: \end{equation}
4824: which is the Verlinde number again.
4825:
4826: In the conformal blocks set up very important observation is
4827: the absence of Hermitian structure on all vector spaces of this
4828: theory. Indeed we are working always with "sl" objects, not
4829: with "su". This feature is distiguishing from the quantization
4830: set up where Hermitian structure appears manifestly.
4831:
4832: \subsection{WZW-connection}
4833:
4834: On the moduli space $\sM_{g,1}$ of curves of genus $g$ with a fixed
4835: (pinched) point we
4836: have vector bundles (1.29) interpreted as vector bundles of
4837: conformal blocks. This identification induces another interpretation of the
4838: Hitchin's projective flat connection coming from CQFT (see
4839: \cite{TK} or beautiful survey \cite{S}).
4840:
4841: The construction and investigation of the connection in these
4842: terms are almost parallel to the constructions described in
4843: subsection 3.8.
4844: The difference is following: instead of the moduli space of
4845: Higgs bundles (which is nothing else as the space of $SL(2,
4846: \C)$-representations of the fundamental group of Riemann
4847: surface) we consider the infinite dimensional space $Q$ (5.9) and
4848: lift
4849: all of our problems to it. Actually before that we have to develop a
4850: technique of working with
4851: infinite dimensional objects of such type rigorously. This was done
4852: in \cite{BL}. Saveing the time-space here we describe this
4853: approach only briefly.
4854:
4855: Recall that the ind-homogeneous ("ind" is a term from \cite{BL})
4856: space $Q$ (5.9) describes
4857: $\SU(2)$-vector bundles trivialized on $\Si - p$ where $(\Si, p)
4858: \in \sM_{g, 1}$. Of course we consider the dense open subset
4859: \begin{equation}
4860: Q_0 \subset Q
4861: \end{equation}
4862: corresponding to semi-stable bundles. For this space the
4863: following fact was proved in \cite{DS}: the natural map
4864: \begin{equation}
4865: f \colon Q_0 \to M^{ss}(\Si)
4866: \end{equation}
4867: is a locally trivial torsor in the etale topology. (The proof of
4868: this fact is contained also in unpublished preprint of Beilinson
4869: and Kazhdan.)
4870:
4871: In this set up Yves Laszlo constructed a family of second order
4872: differential operators $\theta (\dot I)$ over the family of
4873: deformations of complex structures. This construction commutes
4874: with the construction given by the composite morphisms $\phi_s$
4875: (1.86) with the cohomology sequence of the exact triple (1.81)
4876: which gives
4877: \begin{equation}
4878: H^1(\Si, T \Si) \to H^0(S^2(TM^{ss}(\Si))) \to H_h^1(d_s).
4879: \end{equation}
4880: This lifting is, in a sense, a lifting of the Sugawara tensor
4881: $T(\dot I)$ (modulo the ind-scheme technique). Namely for every
4882: function $f$ on $\sM_{g,1}$ and a local section $\tilde{e}$ of
4883: the family of dual to (5.10) spaces, the WZW-connection is given
4884: by the formula
4885: \begin{equation}
4886: \nabla ( f \otimes \tilde{s}) = \dot I(f) \cdot \tilde{s} +
4887: \frac{1}{2k-4} \cdot T(\dot I) \tilde{s} \otimes f \quad mod
4888: (\tilde{s} \otimes f).
4889: \end{equation}
4890: From this point the equality of both connections is obvious
4891: because the only non trivial term in the previous formula is the
4892: Sugawara term just as in the formula (3.152). Thus the Hitchin's
4893: connection is the WZW-connection indeed.
4894:
4895:
4896: \subsection{Monodromy representations}
4897:
4898:
4899: All of our projective and vector connections are on vector bundles
4900: over the Teichmuller space $\tau_g$. The modular
4901: group
4902: \begin{equation}
4903: Mod_g = Diff^+ \Si / Diff_0 \Si
4904: \end{equation}
4905: acts on the Teihmuller space $\tau_g$ such that
4906: \begin{equation}
4907: \sM_g = \tau_g / Mod_g.
4908: \end{equation}
4909: Thus all of them define projective or vector representation of
4910: the modular group $Mod_g$
4911: \begin{equation}
4912: \rho \colon Mod_g \to PGL(N_k, \C)
4913: \end{equation}
4914: where $N_k$ is rank of the vector bundle $V_k$.
4915:
4916: For example, for the classical theory of theta functions we have
4917: two vector representations
4918: \begin{equation}
4919: Mon_w^k \colon Mod_g \to PGL(k^g, \C)
4920: \end{equation}
4921: given by the Welters projective flat connection, that is, by
4922: applying the full procedure of extended Kodaira-Spencer theory
4923: (see \cite{W}). On the other hand, the symplectic theory
4924: yields another combinatorial connection (see \cite{K1}
4925: and another monodromy representation
4926: \begin{equation}
4927: Mon_{BS}^k \colon Mod_g \to PGL(k^g, \C).
4928: \end{equation}
4929:
4930: Fortunately, the perfect quantization identifies both of these
4931: connections and representations to give unique abelian
4932: representation
4933: \begin{equation}
4934: Mon_w^k = Mon_{BS}^k = Mon^k_a.
4935: \end{equation}
4936: Moreover, the monodromy group of this representation is finite
4937: \begin{equation}
4938: Mon_k^a = Sp(2g, \Z_{2k}).
4939: \end{equation}
4940:
4941: For non-abelian case we have the representation
4942: \begin{equation}
4943: Mon_H^k \colon Mod_g \to PGL(rk_g(\emptyset),\C)
4944: \end{equation}
4945: given by the Hitchin connection.
4946:
4947: On the other hand, using identifications from \cite{TUY} and \cite{TK}
4948: Kohno constructed this
4949: representation in the pure combinatorial way
4950: \begin{equation}
4951: Mon_c^k \colon Mod_g \to PGL(rk_g(\emptyset),\C).
4952: \end{equation}
4953: (The subscript
4954: $"c"$ here stands for "combinatorial".) The lines decompositions (4.39)
4955: of fibers
4956: is parallel to the decomposition of the highest
4957: weight representation of the affine Lie algebra of $\fsl(2,\C)$
4958: by eigenspaces of the operator $L_0$ from Sugawara construction
4959: of the representation of the Virasoro Lie algebra (see
4960: \cite{K2}). Manifestly this representation is projective
4961: Hermitian. It was investigated by Macsbaum and Funar. The main
4962: result is the following: the monodromy group of this
4963: representation
4964: is infinite if $k \neq 2,3,4,6$ and $k \neq 10$ for $g=2$.
4965:
4966: Moreover, in paper \cite{An} J. Andersen proved {\it asymptotic
4967: faithfulness} of the collection of representations that is the intersection
4968: of kernels
4969: $$
4970: \bigcap_{k=1}^\infty ker Mon^k_H = 1 \text{ or } \Z_2 \text{ for } g = 2.
4971: $$
4972: (Of course this result is proved for any rank of vector bundles.)
4973: More precisely he proved that that Berezin-Toeplitz endomorphisms associated to smooth functions on $M^{ss}$ are asymmptotically flat.
4974:
4975: The appearance of the Sugawara construction in the procedure of
4976: definition of this connection suggested the equivalence of
4977: these representations. It was proved using the
4978: Borthwick-Paul-Uribe method in
4979: \cite{T3}. This method is quite different from CQFT and it will be
4980: productive to use generalized Sugawara construction \cite{Sc}
4981: to get this coincidence directly
4982: as the generalization of Kniznik-Zamolodchikov construction.
4983:
4984:
4985:
4986:
4987: \section{3-valent graphs}
4988:
4989: \subsection{Spin networks}
4990:
4991: Friedan and Shenker \cite{FS}, Moor and Seiberg
4992: \cite{MS} proposed a new approach to CFT - the {\it axiomatic
4993: description of a rational
4994: conformal field theory}. We will call it the MS-axiomatic. The main ingredient
4995: of this theory is 3-valent graphs which appeared in constructions of
4996: section 4 already. We saw that every 3-valent graph is equivalent
4997: to a trinion-decomposed Riemann surface (see (4.9)-(4.11)).
4998:
4999:
5000: In a similar way, we can extend these constructions to an
5001: orientable surface $\Si$ of genus $g$ with $n$ holes.
5002: Geometrically such surface with a trinion decomposition is
5003: equivalent to its dual 3-1-valent graph $\Ga$ or a history
5004: (3-valent graph of
5005: genus $g$ with $n$ parabolic edges which we describe in the special
5006: subsection).
5007: In this case admissible integer weights are defined in the same
5008: way as in (4.25)
5009: and (4.26) but the set of weights
5010: \begin{equation}
5011: W^k(\Ga, v_1, \dots, v_n)
5012: \end{equation}
5013: is the set of functions $w$ (4.25)-(4.26) satisfying the
5014: conditions
5015: \begin{equation}
5016: w(a_i) = v_i
5017: \end{equation}
5018: for every parabolic edge $a_i, \quad 1 \leq i \leq n$, that is, the
5019: values of these functions on parabolic edges are fixed constants.
5020: So in the same vein we have spaces
5021: \begin{equation}
5022: \sH^{k}_{\Ga, v_1, \dots, v_n} = \oplus_{w \in W^{k}(\Ga, v_1,
5023: \dots, v_n)} \C \cdot w
5024: \end{equation}
5025: The {\it
5026: Clebsch--Gordan conditions} (1), (2) and (3) from (4.26) provide a very
5027: surprising
5028: point of all these constructions: namely, let us multiply all our integer
5029: weights $w \in W^{k}(\Ga)$ by level $k$. We get the maps
5030: \begin{equation}
5031: k \cdot w = j \colon E(\Ga) \to 1/2 \Z_{\geq 0}.
5032: \end{equation}
5033: But we can consider the target set as the set of irreducible
5034: representations of $\SU(2)$:
5035: \begin{equation}
5036: \widehat{\SU(2)} = 1/2 \Z_{\geq 0}.
5037: \end{equation}
5038: So every integer weight can be considered as a coloring of edges by
5039: irreducible representations of $\SU(2)$. The conditions (1), (2)
5040: and (3) (4.26) give that for every triple $e_{1}, e_{2}, e_{3}$
5041: via one vertex $v$ the tensor product of representations
5042: \begin{equation}
5043: 0 \in j(e_{1}) \otimes j(e_{2}) \otimes j(e_{3})
5044: \end{equation}
5045: that is this tensor product contains the trivial representation.
5046: Or the tensor product $j(e_{1}) \otimes j(e_{2})$ admits an
5047: intertwiner
5048: \begin{equation}
5049: i_{v} \colon j(e_{1}) \otimes j(e_{2}) \to j(e_{3})
5050: \end{equation}
5051: as $\SU(2)$-equivariant homomorphism.
5052:
5053: Such coloring of graph edges is called $\SU(2)$-{\it spin
5054: network} by Penrose \cite{P}( see also \cite{T1}). Thus the set
5055: $W^{k}(\Ga)$ is the set of spin networks of level $k$ over a graph
5056: $\Ga$. We will see in the following section that for every
5057: integer weight $w \in
5058: W^{k}(\Ga)$ the function
5059: \begin{equation}
5060: 2k \cdot w = t \colon E(\Ga) \to \Z_{\geq 0}.
5061: \end{equation}
5062: is the type of a loop on our graph $\Ga$. It means, that a
5063: colouring can be defined purely in terms of the geometry of graph
5064: as a {\it topological space}.
5065:
5066: The main fact here is the identification of the set $SNW^k(\Ga)$
5067: of $\SU(2)$-spin networks of level $k$
5068: $$
5069: SNW^k_\Ga = W^k(\Ga)
5070: $$
5071: and the set of integer weights $W^k(\Ga)$.
5072:
5073: Our first task is
5074: to compare spaces $\sH^{k}_{\Ga}$
5075: for different graphs of genus $g$.
5076: As a result of this comparing we will construct, following Kohno
5077: \cite{K1}, the holomorphic vector bundle with a holomorphic
5078: projective flat connection over the moduli space $\sM_{g}$ of
5079: Riemann surfaces of genus $g$. This construction is purely
5080: topological. So first of all we have to construct combinatorial
5081: models of the Teichmuller space and the trivalent graphs complex.
5082: These ideas and constructions came from low dimensional topology
5083: and classical and quantum Conformal Field Theories.
5084:
5085: \begin{figure}[tbn]
5086: \centerline{\epsfxsize=3in\epsfbox{30c.eps}} \caption{\sl Pumping
5087: up trick} \label{Fig 2}
5088: \end{figure}
5089:
5090:
5091: \subsection{3-dimensional topology}
5092:
5093:
5094: Recall again that any 3-valent graph is equivalent to a trinion decomposition
5095: of a Riemann surface. It is easy to see that such surface is a boundary
5096: of uniquely determined handlebody.
5097: We explain this fact more carefully.
5098:
5099: Returning to a maximal set of disjoint inequivalent curves $ \{
5100: C_1, ... , C_{3g-3} \} $ (4.7) and removing these curves we get
5101: a finite set of trinions $ \Si - \{ C_1, ... , C_{3g-3} \} =
5102: \cup_{i=1}^{2g-2} \tilde{v_i} $ (4.8) (see Fig.2).
5103:
5104: Every oriented trinion defines the handlebody inside it and
5105: gluing (sewing) these handlebodies together we get the handlebody
5106: $H_\Ga$ such that
5107: \begin{equation}
5108: \p H_\Ga = \Si_\Ga
5109: \end{equation}
5110: Any diffeomorphism $h \in Mod_g$ (where $Mod_g$ is the modular group (5.30))
5111: gives us a compact 3-manifold
5112: \begin{equation}
5113: X_{\Ga} = H_\Ga \cup_h H_\Ga
5114: \end{equation}
5115: where we are gluing two handlebodies with the boundaries by the
5116: diffeomorphism
5117: \begin{equation}
5118: h \colon \p H_\Ga = \Si_\Ga \to \Si_\Ga = \p H_\Ga.
5119: \end{equation}
5120: For the 3-manifold such presentation is called a {\it Heegaard splitting}.
5121: The subgroup
5122: \begin{equation}
5123: M_\Ga \subset M_g
5124: \end{equation}
5125: of diffeomorphisms which can be extended to diffeomorphisms of
5126: the handlebody $H_\Ga$, acts on $Mod_g$ on the left and right and we
5127: have the set of 3-manifolds
5128: \begin{equation}
5129: \{ X_\Ga \} = M_\Ga \backslash Mod_g \slash M_\Ga
5130: \end{equation}
5131: as the set of double cosets.
5132:
5133: In Heegard splitting (6.10) let us add a new handle to the surface and extend
5134: the diffeomorphism $g$ by the identity on this new handle. This procedure
5135: doesn't
5136: change our 3-manifold and is called {\it stabilization}.
5137: Singer proved in \cite{Si} the following statement
5138:
5139: \begin{prop} The equivalence relation between elements $h \in Mod_g$ when $g \to
5140: \infty$ creating homeoomorphic 3-manifolds, is generated by isotopies, left
5141: and right
5142: actions of $M_\Ga$ and a stabilization.
5143: \end{prop}
5144:
5145: To construct an invariant of 3-manifolds we have to send a marked
5146: Riemann surface $\Si_\Ga$ to the wave function vector space $\sH^k_\Ga$
5147: (4.43) with fixed vector $w_{H_\Ga}$ given by the function (4.25) such that
5148: \begin{equation}
5149: w_{H_\Ga}(e) = 0 \quad \text{for every} \quad e \in E(\Ga).
5150: \end{equation}
5151: Here
5152: \begin{equation}
5153: w_{H_\Ga} \in \sH^k_\Ga
5154: \end{equation}
5155: is called the {\it vacuum vector} for the handlebody $H_\Ga$.
5156:
5157: The space $\sH^k_\Ga$ admits the standard Hermitian structure which identifies
5158: this space with the dual.
5159:
5160: Now if we have a linear representation
5161: \begin{equation}
5162: \rho \colon Mod_g \to GL (\sH^k_\Ga)
5163: \end{equation}
5164: then
5165: \begin{equation}
5166: I(X_{\Ga, g}) = \frac{< w_{H_\Ga}, \rho(g)(w_{H_\Ga})>}{\Vert
5167: w_{H_\Ga} \Vert^2}
5168: \end{equation}
5169: has to be an invariant of $X_{\Ga}$ as 3-manifold since it
5170: easy to see that
5171: \begin{enumerate}
5172: \item the vector $w_{H_\Ga}$ is invariant with respect to $M_\Ga$-action and
5173: \item under normalization (6.17) this vector is invariant under stabilization.
5174: \end{enumerate}
5175:
5176: But the situation is a slightly more complicated then just
5177: described. Namely, Kohno constructed nonlinear representation
5178: $\rho$ but made it slightly projective,
5179: that is, determined up to the usual phase amplitude
5180: \begin{equation}
5181: e^{\frac{\pi i \cdot k}{4 (k+2)} }.
5182: \end{equation}
5183: Thus the invariant (6.17) is determined up to the same phase
5184: amplitude.
5185:
5186: To construct such slightly projective representation (6.16) one has to
5187: treat combinatorial constructions in the next special section.
5188:
5189:
5190:
5191: \subsection{Geometry of graphs}
5192:
5193:
5194:
5195:
5196:
5197: A graph is one
5198: dimensional complex containing edges -
5199: cells of one dimensional skeleton and vertices -cells of
5200: zero dimensional
5201: skeleton.
5202:
5203: The combinatorial invariant of a pinched Riemann surfaces with
5204: a trinion decomposition
5205: is a (3-1)-valent graphs or a {\it history}. Such graphs have
5206: 3-valent and 1-valent vertices. The last vertices are half
5207: intervals with one vertex. Vertices of such edges are called
5208: {\it parabolic}.
5209:
5210: Graphs and histories are Feynman diagrams - objects of the standard
5211: formalism
5212: in fields (for 3-valent graph case, $\phi^3$ - theories).
5213:
5214: \begin{figure}[tbn]
5215: \centerline{\epsfxsize=3in\epsfbox{2c.eps}} \caption{\sl Standard
5216: Feynman diagrams} \label{Fig 2}
5217: \end{figure}
5218:
5219:
5220: In classical and quantum field theory a 3-valent graph without
5221: parabolic vertices is called a vacuum-polarization graph. So we will
5222: consider such
5223: graph as a geometrical object. Note that
5224: an closed 1-valent graph is a disjoint finite union of compact
5225: intervals and
5226: a closed 2-valent graph is a disjoint finite union of compact
5227: circles.
5228:
5229: Every graph $\Ga$ has the following ingredients:
5230: \begin{enumerate}
5231: \item $E(\Ga)$ is the set of edges;
5232: \item $V( \Ga)$ is the set of vertices,
5233: \item $S(v)$ is the star of a vertex $v$ that is the set of local edges incident to $v$,
5234: \item $F(\Ga)=\{v\in e \}$ is the set of flags (edge + vertex).
5235: It is the same as
5236: the set of oriented edges $\vec E(\Ga)$,
5237: \item $L( \Ga)$ is the set of circles, that is, of edges with only one vertex.
5238: \end{enumerate}
5239: We will consider homogeneous graphs that is the graphs of the same valence $d$ at every vertex.
5240:
5241: For example for 3-valent graphs two projections $p_e \colon F(\Ga)\to E(\Ga)$ and
5242: $v \colon F(\Ga)\to V(\Ga)$ are finite covers
5243: of degrees
5244: 2 and 3 with the same ramification locus
5245: $W_e = W_v \subset F(\Ga)$, containing pairs
5246: $v\in e$, where $e$ is a loop:
5247: \begin{equation}
5248: R_e = L( \Ga) \subset E( \Ga) ; \quad R_v = L( \Ga) \subset V(
5249: \Ga)
5250: \end{equation}
5251: Let $\vert\ \vert$ be the cardinality of a finite set, then
5252: \begin{equation}
5253: 2 \cdot \vert E(\Ga) \vert-\vert L(\Ga) \vert=\vert F(\Ga) \vert=
5254: 3 \cdot \vert V(\Ga) \vert-\vert L(\Ga)\vert.
5255: \end{equation}
5256:
5257: Thus $ \vert V(\Ga) \vert=2g-2$ and $ \vert E(\Ga)
5258: \vert=3g-3$, where $g>1$ is an integer number called {\it genus}
5259: of
5260: $\Ga$.
5261:
5262: There exists only finite set $TG_{g}$ of 3-valent graphs of genus
5263: $g$. But up to now nobody knows how many graphs it contains
5264: \begin{equation}
5265: \vert TG_{g} \vert = ?
5266: \end{equation}
5267: (besides of small $g = 2, 3,\dots, 11 $).
5268:
5269:
5270: Consider any graph $\Ga$ with vertices of any valence $ > 2$. Let $e \in E(\Ga)$ be
5271: any edge with distinct two vertices $ \p e = v_1 \bigcup v_2$ and $v_1 \neq v_2$. Remove the edge $e$ and identify vertices $v_1$ and $v_2$ to get a single new vertex
5272: $v_{new}$. We get a new graph $\Ga'$ which is called {\it contraction} of $\Ga$
5273: along the edge. Such contraction decreas $\vert E(\Ga) \vert$ and $\vert V(\Ga) \vert$ by one. The minimal under this procedure graph $\Ga_{min}$ has only one vertex. Obviously every graph can be obtained from a 3-valent graph by applying a chain of contractions. The inverse operation is {\it expansion} which is the blow up of a vertex $v$ with $\vert S(v) \vert > 3$ to a new edge $e_{new}$ with
5274: two new vertices $v_{new}, v'_{new} = \p e_{new}$ and partitions
5275: $$
5276: (S(v_{new} - e_{new}) \bigcup ( S(v'_{new}) - e_{new}) = S(v).
5277: $$
5278: The maximal under this procedure graph is a trivalent graph $\Ga_{max}$.
5279:
5280: Every 3-valent graph $\Ga$ has {\it closed relatives}: let $v, v'
5281: $ be vertices of one edge $e$ and $e_{1}, e_{2}, e$ are edges via
5282: $v$ and $e'_{1}, e'_{2}, e$ are edges via $v'$. Let us contract
5283: the edge $e$ to obtain 4-valent vertex $v_{new}$. For this vertex the star $S(v_{new}) = e_{1} \bigcup e_{2} \bigcup e'_{1}
5284: \bigcup e'_{2}$. Now divide this quadruple in two pairs, for
5285: example, $(e_{1}, e'_{1})$ and $(e_{2}, e'_{2})$. Then extend
5286: the vertex $v_{new}$ to new edge $e_{new}$ such a way that we get
5287: \begin{enumerate}
5288: \item
5289: two new vertices $v_{1}$ and
5290: $v_{2}$ such that $\p e_{new} = v_{1}, v_{2}$,
5291: \item
5292: with stars $S(v_{1}) = e_{1}, e'_{1}, e_{new}$ and
5293: $S(v_{2}) = e_{2}, e'_{2}, e_{new}$.
5294: \end{enumerate}
5295: We get a new
5296: graph $\Ga'_{e}$. This is the composition of one contraction and one expansion
5297: preserving valence of the graph.
5298:
5299: \begin{dfn} The transformation of graphs
5300: $$
5301: \Ga \to \Ga'_{e}
5302: $$
5303: is called {\it elementary transformation}.
5304: \end{dfn}
5305: (We will see later that this transformation lies under {\it elementary move for trinion decompositions.})
5306:
5307: Sending the edge $e$ of $\Ga$ to the new edge $e_{new}$ of $\Ga'_e$ we get the identification
5308: \begin{equation}
5309: \na \colon E( \Ga) \to E(\Ga'_{e}).
5310: \end{equation}
5311:
5312:
5313:
5314:
5315: \begin{figure}[tbn]
5316: \centerline{\epsfxsize=3in\epsfbox{1c.eps}} \caption{\sl
5317: Elementary transformations} \label{Fig 11}
5318: \end{figure}
5319:
5320:
5321: So, around any 3-valent graph there exists the collection of
5322: 3-valent graphs which differ from $\Ga$ by elementary transformation. (This triple corresponds to the triple of vertices of the triangle of Fig. 13.)
5323:
5324: By the induction on $g$ it is easy to prove that any 3-valent
5325: graph $\Ga$ of genus $g$ can be transformed to any other
5326: 3-valent graph $\Ga'$ by a finite chain of elementary transformations.
5327: Thus the elementary transformations act on $TG_{g}$ transitively.
5328:
5329: In the same vein we can consider the composition of two contractions and two
5330: expansions preserving valence of verticies and so on.
5331:
5332: Now for any graph $\Ga$ let
5333: $$
5334: \sE(\Ga) = \{ \Ga_{max} \}
5335: $$
5336: be the set of all expansions of $\Ga$ to a trivalent graph. This set of 3-valent graphs we call {\it nest} of 3-valent graphs.
5337:
5338: Of course we can consider compositions of expansions for every vertex $v \in V(\Ga)$ with $\vert S(v) \vert > 3$ independentely. So for every such vertex
5339: the set of expansions $\sE(v)$ is in 1-1 correspondence with the set of 3-valent trees $T$ with parabolic edges identifyed with the set $S(v)$:
5340: $$
5341: \p T = S(v).
5342: $$
5343: Thus
5344: $$
5345: \vert \sE(v) \vert = \sum_{T \vert \p T = S(v)} \frac{1}{\vert Aut T \vert}.
5346: $$
5347:
5348: So for a graph $\Ga$ with one
5349: 4-valent vertex
5350: $\vert \sE(\Ga) \vert = 3 $
5351: and so on.
5352:
5353: The combinatoric behind of these constructions can be described perfectly by the standard pairing scheme in the {\it Feynman diagram technique}.
5354: This technique was applyed as well to so called
5355: {\it ribbon graphs}. Recall that a ribbon graph structure on a graph is cyclic
5356: orders choosen on all stars of edges. Geometricaly it is equaivalent to an embedding of a graph into an oriented surface. Such cyclic orders are required by applications to matrix models (see \cite{MP} and \cite{Mu}).
5357:
5358: For such graphs contractions and expansions preserve the ribbon structures. For a homogeneous d-valent graph $\Ga$ a ribbon graph structure is equivalent to the choice of a {\it connection} on this graph (see \cite{BGH} for the consideration of graphs as manifolds).
5359:
5360: Recall that a path of lenth 1 on $\Ga$ is just an oriented edge
5361: $\vec e$. Let the set
5362: $P_1(\Ga) = \vec E(\Ga)$ be the set of 1-paths on $\Ga$.
5363: Every such path $\vec e $ has vertices of two types
5364: \begin{equation}
5365: v_s(\vec e), \quad v_t(\vec e) \in V(\Ga)
5366: \end{equation}
5367: {\it source} and {\it target} which are equal for loops.
5368:
5369: A path of length $d$ in $\Ga$ is an ordered sequence $(\vec e_1,
5370: ..., \vec e_d)$ of oriented edges such that for every $i$
5371: \begin{equation}
5372: v_t(\vec e_i) = v_s(\vec e_{i+1}).
5373: \end{equation}
5374: Let $e_{d+1} = e_1$, that is, our order is cyclic. A path $(\vec
5375: e_1, ..., \vec e_d)$ is called irreducible if $e_i \neq
5376: e_{i+1} $
5377: for every
5378: $i$ (including $i = d+1 $).
5379:
5380: A connection $\na$ on a homogeneous graph $\Ga$ is a collection of
5381: identifications of the stars
5382: \begin{equation}
5383: \na_{\vec e} \colon S(v_{s}(\vec e)) \to S(v_{t}(\vec e))
5384: \end{equation}
5385: such that
5386: $$
5387: \na_{\vec e} (e) = e
5388: $$
5389: where $v_{s}, v_{t} = \p \vec e$.
5390:
5391:
5392: A path $(\vec
5393: e_1, ..., \vec e_n)$ is called {\it geodesic} if
5394: \begin{equation}
5395: e_{i+2} = \na_{\vec e_{i+1}} (e_i)
5396: \end{equation}
5397: for every
5398: $i$ (including $i = n+1 $) when we have closed geodesic.
5399:
5400: Obviously for every pair of edges from the star of a vertex there exists
5401: unique closed geodesic passing through the edges. Hence the set of closed geodesics defines our connection
5402: uniquely.
5403:
5404: In the same vein we can say that a subgraph $\Ga' \subset \Ga$ is totally geodesic if every geodesic starting in $\Ga'$ stay within $\Ga$.
5405:
5406: Now for every closed geodesic $l = (\vec
5407: e_1, ..., \vec e_d = \vec e_1)$ we have the set of vertices $V_l$ of the edges in this consequence. For every $v \in V_l$ we have a monodromy permutation
5408: $$
5409: m_{v} (l) \colon S(v) \to S(v)
5410: $$
5411: So every closed geodesic $l$ defines a conjugation class
5412: $$
5413: m(l) \in Conj(S_d)
5414: $$
5415: where $S_d$ is the group of permutations with $d$ elements.
5416:
5417: A closed geodesic is {\it flat} if $m(l) = id$.
5418:
5419: Let $G_\Ga$ be the set of simple closed irreducible geodesics of $\Ga$.
5420: Then the map
5421: $$
5422: c_\na \colon G_\Ga \to Conj(S_d)
5423: $$
5424: is an invariant of connection $\na$ and the preimage of the unit class is the subset $G_\Ga^F$ of flat geodesics.
5425: Moreover this map is an invariant for gauge classes of connections:
5426: for every $v \in V(\Ga)$ consider the group
5427: $S_v = Aut S(v) = S_d $
5428: where the last equality is non canonical, of course, and the product
5429: $$
5430: \sG = \prod_{v \in V(\Ga)} S_v.
5431: $$
5432: This group acts on the set $\sA_\Ga$ of all connections on $\Ga$ by the formula:
5433: for every $\vec e$ with $\p \vec e = v_{s}, v_{t}$
5434: \begin{equation}
5435: g (\na)_{\vec e} = g_{s} \na_{\vec e} g_{t}^{-1}
5436: \end{equation}
5437: where $g_v$ are components of the decomposition of our discrete gauge group.
5438:
5439: Then the set
5440: $$
5441: \sB = \sA / \sG
5442: $$
5443: is the set of orbits with respect to the gauge group and the function $c_\na$
5444: is an invariant of the orbits.
5445:
5446: Now gluing 2-cell to every simple closed flat geodesic we get a 2-complex
5447: $C_\na(\Ga)$ which is a closed oriented surface corresponding to this ribbon graph.
5448:
5449: A ribbon graph is called {\it plane} if the surface $C(\Ga) = S^2$ is 2-sphere.
5450:
5451:
5452:
5453: For ribbon graphs the set of expansions is "reducible". In particular we can
5454: choose from any pair of graphs correlated by an elemetary transformation one
5455: of them. Recall that in the procedure of expansion we identify two expanded graphs if there is a ribbon graph isomorphism from one to the other preserving all local edges. In particular if $v \in V(\Ga)$ is d-valent vertex then
5456: $\vert \sS(v) \vert = \frac{1}{d-1} C^{2d-4}_{d-2}$ which is the Catalan number.
5457:
5458: Now forget the ribbon structure:
5459: let $P_d(\Ga)$ be the set of irreducible paths of length $d$ in
5460: $\Ga$. Every path
5461: $(\vec e_1, ..., \vec e_d) \in P_d(\Ga)$ defines two vertices
5462: \begin{equation}
5463: v_s((\vec e_1, ..., \vec e_d)), \quad v_t((\vec e_1, ..., \vec
5464: e_d)) \in V(\Ga)
5465: \end{equation}
5466: - source and target and two maps
5467: \begin{equation}
5468: v_s \colon P_d(\Ga) \to V(\Ga)
5469: \end{equation}
5470: $$
5471: v_t \colon P_d(\Ga) \to V(\Ga)
5472: $$
5473: $$
5474: (v_s \times v_t) \colon P_d(\Ga) \to V(\Ga) \times V(\Ga).
5475: $$
5476: The preimage of the intersection
5477: \begin{equation}
5478: L_d(\Ga) = (v_s \times v_t)^{-1} ( V(\Ga)_\Delta )
5479: \end{equation}
5480: is the set of oriented irreducible loops of length $d$. In
5481: particular
5482: $$
5483: F(L_1)(\Ga)= L(\Ga),
5484: $$
5485: where $F$ is the projection of a path to the graph. For every
5486: vertex á
5487: $v \in V(\Ga)$ we have the set of irreducible loops marked by $v$
5488: \begin{equation}
5489: L_d(\Ga)_v = (v_s \times v_t)^{-1} ( v).
5490: \end{equation}
5491: The union
5492: \begin{equation}
5493: L_\infty(\Ga)_v = \cup_{d=1}^\infty L_d(\Ga)_v
5494: \end{equation}
5495: admits a group structure
5496: \begin{equation}
5497: \pi_1^C(\Ga)_v = L_\infty(\Ga)_v.
5498: \end{equation}
5499: if we consider only irreducible fragments of compositions.
5500:
5501: Obviously, this group depends on the marking point $v$.
5502:
5503: Let $\pi_1(\Ga)$ be the standard fundamental group of $\Ga$ (as
5504: 1-complex). Then the natural epimorphism $r \colon \pi_1^C(\Ga)
5505: \to \pi_1(\Ga)$ is an isomorphism. Obviously, if $\Ga$ is a 3-valent graph
5506: of genus $g$ then $\pi_1(\Ga) = F_g $
5507: is free group with $g$ generators.
5508:
5509: For any 3-valent graph $ \Ga $ a choice of two edges $e, e'$ (may be not
5510: different) defines a new graph $\Ga_{e, e'}$ by the
5511: following procedure: let us choose a pair of inner points
5512: $v \in e$ and $v' \in e'$ and joint them by the edge
5513: $e_{new}$. We get a new 3-valent graph with the pair of new
5514: vertices and the triple of new edges
5515: $e_{new}, e{1/2} , e_{ 1/2}$, where the last two edges are just halves of the edges
5516: obtained by dividing previous edgesá $e, e'$.
5517:
5518: The inverse operation is much more simple : we just remove an
5519: edge and a pair of its vertices.
5520:
5521: \begin{figure}[tbn]
5522: \centerline{\epsfxsize=3in\epsfbox{3c.eps}} \caption{\sl Two
5523: paths} \label{Fig 3o}
5524: \end{figure}
5525: %\epsfbox{3o.fig}}
5526: If our edges were in two paths $
5527: e \in p = (\vec e_{1}, ...,
5528: \vec e_{n})$ and $e' \in p' = (\vec e'_{1}, ..., \vec e'_{m})$,
5529: then on the new
5530: graph $\Ga_{e, e'}$ we would have (formally) two paths
5531: $p_{new}, p'_{new}$.
5532:
5533: It is easy to see that
5534: \begin{enumerate}
5535: \item $ p \neq p' \implies p_{new} = p'_{new}$;
5536: \item $ p = p' \implies p_{new} \neq p'_{new}$.
5537: \end{enumerate}
5538:
5539:
5540: By induction one obtains:
5541: \begin{enumerate}
5542: \item Any 3-valent graph can be created by a finite chain of such
5543: elementary operations starting with the disconnected union of circles
5544: $0 \cup 0 ... \cup 0$;
5545: \item every path on $\Ga$ can be created by a finite chain of
5546: elementary
5547: operations starting from a path on a disjoint finite set of circles.
5548: \end{enumerate}
5549: These statements give the induction procedure which allows to prove all simple
5550: facts about trivalent graphs. There are two types of such problems: {\it Eulerian paths} and {\it coloring problems}.
5551:
5552: Recall that a path $p = (\vec e_{1}, ...,
5553: \vec e_{n}) \in P_d(\Ga)$ is called {\it Eulerian} if it contains every oriented edge and exactly one time. A system of paths $P_E$ is called Eulerian if every oriented edge is contained and exactly one time in one and exactly one path from
5554: this system. Such system is called minimal if $\vert P_E \vert = min$. This is
5555: an invariant of the graph $e(\Ga)$.
5556:
5557: We know already that an elementary operation tranform an Eulerian system $P_E$
5558: on $\Ga$ to the Eulerian system $P'_E$ on $\Ga_{e, e'}$. Moreover
5559: $$
5560: \vert P'_E \vert = \vert P_E \vert \pm 1.
5561: $$
5562: From this it easy to see that for every 3-valent graph $\Ga$ of genus $g$
5563: \begin{enumerate}
5564: \item $1 \leq e(\Ga) \leq g+1$;
5565: \item $e(\Ga) = g-1 mod 2$;
5566: \item there exists a chain of elemetary operations with result graph $\Ga$ such that $e(\Ga') = 1$.
5567: \end{enumerate}
5568:
5569: We can consider every plane ribbon graph $\Ga$ as a graph on the plane $\R^2$.
5570: Such embedded graph defines the geographic map with connected componets of the complement as the countries. This is the {\it 4-colors setup}. This problem is stated for embedded
5571: graphs.
5572:
5573: Other "coloring" problem for graphs appears in the Feynman diagram technique
5574: (for the problem of dependence on the form of Lagrangian and of limits
5575: in matrix models). Let $C = (c_1, ... , c_N)$
5576: is a set of colors. Any coloring is just a map
5577: $$
5578: c \colon E(\Ga) \to C
5579: $$
5580: such that for every vertex $v \in V(\Ga)$ the map $c \colon S(v) \to C$ is an
5581: {\it embedding}. That is around every vertex all edges have different colors.
5582: If the number $N$ of colors is minimal then it is called the {\it chromatic number } $N_{min} = \chi (\Ga)$. Easely we have to consider graphs without
5583: 1-loops. Then for 3-valent graphs
5584: $$
5585: 3 \leq \chi(\Ga) \leq 4.
5586: $$
5587: A little more efforts you need to prove the following fact
5588:
5589: $\chi(\Ga) = 3$ if and only if there is a set $(l_1, ... , l_n)$ of disjoint
5590: simple loops on $\Ga$ such that
5591: \begin{enumerate}
5592: \item every loop $l_i$ contains an even number of edges;
5593: \item for every vertex $v \in V(\Ga)$ there exists a loop which contains it:
5594: $$
5595: V(\Ga) \subset \bigcup_{i=1}^n l_i.
5596: $$
5597: \end{enumerate}
5598:
5599: Both of these "coloring" problems are related by the following equivalence:
5600: two statements
5601: \begin{enumerate}
5602: \item The 4-colors problem admits positive solution;
5603: \item For every plane 3-valent graph $\Ga$ without 1-loops $\chi(\Ga) = 3$
5604: \end{enumerate}
5605: are equivalent.
5606: These problems of "classical theory" of graphs. Considering every homogeneous
5607: graph as a manifold (in style of \cite{BGH}) and following Segal's ideas we
5608: have a reason
5609: to consider the "Loop space" of every graph.
5610:
5611:
5612:
5613:
5614:
5615: Consider standard interval $[0, 1]$ and divide it by a finite
5616: set of points to a complex $[0, p_1, .... , p_{d-1}, 1]$. We can
5617: consider this as a 2-valent history with two parabolic edges $[0,
5618: p_1]$ and $[p_{d-1}, 1]$. The linking of the parabolic edges
5619: of such history gives a cycle (or a loop) that is a closed 2-valent graph.
5620:
5621: As we know every continuous combinatorial map
5622: \begin{equation}
5623: p \colon [0, p_1, .... , p_{d-1}, 1]\to \Ga
5624: \end{equation}
5625: is called path of length $d$ in $\Ga$.
5626:
5627: If this path is irreducible, the preimage of every edge
5628: \begin{equation}
5629: p^{-1}(e) = g_1 \cup ... \cup g_{col(e)}
5630: \end{equation}
5631: is a disjoint union of a finite set of compact 1-graphs and the
5632: number of components is equal to $col(e)$ (without any orientation).
5633:
5634: Now we consider the set $\sL$ of irreducible loops that is closed
5635: paths (0=1). The type of loop $l \in \sL $ is the non negative
5636: function
5637: \begin{equation}
5638: col \colon E(\Ga) \to \Z_{\geq 0}.
5639: \end{equation}
5640: (recall that we consider irreducible unmarked loops only).
5641:
5642: For any type of loops for every vertex $v \in V(\Ga)$ with the set
5643: of edges
5644: $(e_{1,v}, e_{2,v}, e_{3,v})$ the following constraints hold:
5645: \begin{enumerate}
5646: \item
5647: \begin{equation}
5648: col (e_{1,v}) + col( e_{2,v}) + col( e_{3,v}) = 0 \mod 2;
5649: \end{equation}
5650: \item
5651: \begin{equation}
5652: |col( e_{1,v})-col(e_{2,v})| \leq col ( e_{3,v}) \leq col(
5653: e_{1,v}) + col(e_{2,v}).
5654: \end{equation}
5655: \end{enumerate}
5656: Indeed, consider a vertex $v$ and the star $S(v) = e_1, e_2, e_3$.
5657: Let $n_{12}$ be the number equal to how many times we come to $v$ along
5658: the edge $e_1$ and turn to the edge $e_2$. In the same vein we have
5659: $n_{ij}$ for every pair of edges from the star. Then
5660: \begin{equation}
5661: col(e_i) = \sum_{j \neq i} n_{ij}.
5662: \end{equation}
5663: From this we get the statement immediately.
5664:
5665:
5666: Now we say that a loop $l \in \sL$ has the type of level $k$ if
5667: for
5668: every edge $e$
5669: $$
5670: col (e) \leq k.
5671: $$
5672: In the same vein we can prove that the type function (6.36) gives
5673: a type of
5674: loop of level $k$ iff
5675: for every vertex $v \in V(\Ga)$ with the set of local edges
5676: $(e_{1,v}, e_{2,v}, e_{3,v})$ almost the same constraint
5677: holds:
5678: \begin{enumerate}
5679: \item
5680: $$
5681: col (e_{1,v}) + col( e_{2,v}) + col( e_{3,v}) = 0 mod 2;
5682: $$
5683: \item and a slightly sophisticated condition 2)
5684: \begin{equation}
5685: |col( e_{1,v})-col(e_{2,v})| \leq col ( e_{3,v}) \leq
5686: \end{equation}
5687: $$
5688: min (col( e_{1,v}) + col(e_{2,v}), 2k -
5689: col( e_{1,v}) - col(e_{2,v}).
5690: $$
5691: \end{enumerate}
5692:
5693: Note that the inequality (6.38) is just the triangle inequality
5694: and the inequality
5695: (6.40) is the {\it spherical} triangle inequality
5696: (with $\sqrt{k}$ proportional to the radius of the sphere).
5697:
5698: Note also that it is quite reasonable to call every
5699: loop as {\it knot}.
5700:
5701: So every knot defines the type function (6.36) coloring edges of
5702: our graph in $\Z_{\geq 0}$ colors. {\it Dividing this function by 2 we
5703: get the map}
5704: $$
5705: 1/2 col \colon E(\Ga) \to 1/2 \Z_{\geq 0} = \widehat{\SU(2)}
5706: $$
5707: {\it where the last set is the set of irreducible representations of} $\SU(2)$. Conditions (6.38) and (6.40) mean that any pair $(\Ga, cal)$ is
5708: a $\SU(2)$- {\it spin network}.
5709:
5710: Recall that to get a marked Riemann surface from a trivalent graph $\Ga$ we
5711: need to pump up our graph,
5712: that is, to replace every vertex by a trinion and to glue
5713: trinions by tubes along the edges.
5714:
5715:
5716: \subsection{ Circle complex and Graph complex}
5717:
5718:
5719: Let $\Si_{g,n}$ be an oriented surface of genus $g$ with $n$
5720: punctures and $C$ be a simple, closed, homotopy non-trivial,
5721: non-bounded pinchs loop on $\Si_{g,n}$. We call such loop a
5722: {\it circle} and consider it up to the isotopy class.
5723:
5724: {\it Circle
5725: complex} or {\it Curve complex} $C(\Si_{g,n})$ has the set of
5726: circles as 0-skeleton $C_{0}(\Si_{g,n})$,
5727: \begin{enumerate}
5728: \item 1-skeleton $C_{1}(\Si_{g,n})$ is the set of
5729: pairs of classes of disjoint circles;
5730: \item triangles correspond to triples of classes of disjoint cirles ;
5731: \item every $i$-simplex corresponds to a set of $i+1$ classes
5732: mutually
5733: disjoint circles,
5734: \item so, simplexes of maximal dimension are collections of $3g-3$
5735: disjoint
5736: circles just as (4.7).
5737: \end{enumerate}
5738: The main fact which we will use is the following: the group of
5739: combinatorial automorphisms
5740: \begin{equation}
5741: Aut C(\Si_{g}) = Mod_{g}
5742: \end{equation}
5743: is the modular group (5.30).
5744: \begin{rmk}
5745: By the Roiden's theorem
5746: \begin{equation}
5747: Isom \quad \tau_{g} = Mod_{g}
5748: \end{equation}
5749: where the Teichmuller space $\tau_{g}$ (5.31) is equipped with
5750: the Teichmuller metric. Moreover as a metric space $C(\Si_{g,n})$
5751: is very closed to $\tau_{g}$. In particular, $C(\Si_{g,n})$ is
5752: $\delta$-hyperbolic (\cite{MM}). The same facts are true for the
5753: Weil-Peterson metrics.
5754: \end{rmk}
5755: Recall that in the case of a pinched curve $\Si_{g, n}$ the circle
5756: can't be isotopic to a boundary component.
5757:
5758:
5759:
5760:
5761: \begin{figure}[tbn]
5762: \centerline{\epsfxsize=3in\epsfbox{7c.eps}} \caption{\sl Triangles} \label{Fig 13}
5763: \end{figure}
5764:
5765:
5766:
5767:
5768:
5769: For two special but very important cases $\Si_{0, 4}$ and
5770: $\Si_{1,1}$ the
5771: definition is slightly different
5772: (since there are no disjoint pairs of circles
5773: ): two vertices are connected by an edge if circles have 2
5774: intersections in $\Si_{0,4}$-case and 1 intersection in
5775: $\Si_{1,1}$-case. In both cases
5776: \begin{equation}
5777: C(\Si_{0,4}) = C(\Si_{1,1})
5778: \end{equation}
5779: is the ideal triangulation of 2-disc (so called Farey Graph,
5780: see Fig. 14).
5781:
5782: \begin{figure}[tbn]
5783: \centerline{\epsfxsize=3in\epsfbox{4c.eps}} \caption{\sl Farey
5784: graph (the ideal triangulation of disk)} \label{Fig 13}
5785: \end{figure}
5786:
5787:
5788: {\it Trinion decomposition
5789: graph } $C_1{TD(\Si)}$ has the set of trinion decompositions
5790: as 0-skeleton
5791: \begin{equation}
5792: V(C_{1}(TD(\Si))) = \{(C_{1}, ... , C_{3g-3})\}
5793: \end{equation}
5794: (see (4.7)) and a pair of trinion decompositions is connected by
5795: an edge if the decompositions differ by an {\it elementary move}. That
5796: is, in the system of circles $C_{1}, \dots, C_{3g-3}$ on a
5797: Riemann surface
5798: $\Si$ we replace only one circle, for example $C_{1}$, with
5799: another circle $C'_{1}$ intersecting twice if $C_{1}$ and
5800: $C'_{1}$ don't lie on the same trinion and intersecting
5801: once otherwise (minimal intersections):
5802: \begin{equation}
5803: \{C_{1}, \dots, C_{3g-3}\} \to \{C'_{1}, \dots, C_{3g-3}\}.
5804: \end{equation}
5805: Recall that every trinion decomposition $C_{1}, ... , C_{3g-3}$
5806: defines the 3-valent graph $\Ga$ (4.9)- (4.11). Thus we have the
5807: {\it finite graph} $C_{1}(TG_{g})$ with the set of vertices
5808: $$
5809: V(C_{1}(TG_{g})) = TG_{g}
5810: $$
5811: (see (6.21)) and a pair of graphs $\Ga, \Ga'$ is connected by an
5812: edge if they differ by an elementary transformation
5813: (see Definition 9) corresponding to an elementary move of trinion
5814: decomposition (6.45). The fibers of the map
5815: $$
5816: \ga \colon V(C_{1}(TD(\Si))) \to TG_{g}
5817: $$
5818: are described by the formula
5819: $$
5820: Mod_{g} / Aut H_{\Ga}
5821: $$
5822: where $H_{\Ga}$ is the handlebody (6.9) and $Aut$ is the group of
5823: isotopy classes of diffeomorphisms of $\Si = \Si_{\Ga}$
5824: extendable to diffeomorphisms of the handlebody.
5825:
5826: The finite TG-graph $C_{1}(TG_g)$ is 1-skeleton of the finite $TG_g$-complex
5827: $C(TG_g)$ which can be constructed by the following procedure: first of all let us glue by 2-cells all triangles coming from elementary transformations.
5828: Over such triangles in TG-graph we have triangles from Fig. 13 of the infinite TD-graph. Up to such triangles we have the canonical connection $\na$ (6.22) on
5829: TG-graph and we can apply the "geodesic technique" (6.25)-(6.28).
5830: In particular, we obtain 2-complex $C_\na(TG_g) = C(TG_g)$. The important
5831: fact (see Remark at the end of this subsection) proved in \cite{HT} is the following
5832: \begin{enumerate}
5833: \item $C(TG_g)$ is simply connected that is
5834: $$
5835: C(TG_g) = S^2,
5836: $$
5837: \item hence our TG-graph is plane;
5838: \item hence we may ask about this graph all classical questions, for example,
5839: on the chromatic number and so on.
5840: \end{enumerate}
5841:
5842: However the standard way to check these properties is the following
5843: \begin{enumerate}
5844: \item to construct the infinite complex $C(TD(\Si))$ "by our hands";
5845: \item to check described properties;
5846: \item to desend all constructions to TG-complex using the map $\ga : C_1(TD) \to C_1(TG_g)$ and to get the complex $C(TG_g)$;
5847: \item to check the required properties for this finite complex.
5848: \end{enumerate}
5849:
5850: \begin{rmk}
5851: It is quite natural to apply this geodesic technique to the infinite graph $C_1(TD(\Si))$ directly but we have to go passing the standard way.
5852: \end{rmk}
5853:
5854:
5855: For our special cases of Riemann surfaces it can be proved that
5856: \begin{equation}
5857: C_{1}(TD(\Si_{0,4})) = C_{1}(TG(\Si_{1,2})) = C_{1}(\Si_{0,4})=
5858: C_{1}(\Si_{1,2})
5859: \end{equation}
5860: is the Farey graph that is 1-skeleton of the Farey complex again.
5861:
5862:
5863: \begin{figure}[tbn]
5864: \centerline{\epsfxsize=3in\epsfbox{8c.eps.eps}} \caption{\sl Squares} \label{Fig 13}
5865: \end{figure}
5866:
5867:
5868:
5869: The infinite complex $C(TD)$ is 2-dimensional. Its 2-cells
5870: represent relations between special moves, that is, loops in
5871: $TD$-graph
5872: and they can be triangles, squares,
5873: pentagones and hexagons. Every trinion decomposition of $\Si_{g,n}$
5874: contains trinion decompositions of surfaces with smaller $g$ and $n$ as
5875: fragments of it. We call such fragments subsurfaces.
5876: The relations, coming from such subsurfaces are the following:
5877:
5878:
5879: \begin{figure}[tbn]
5880: \centerline{\epsfxsize=3in\epsfbox{9c.eps.eps}} \caption{\sl Pentagons}
5881: \label{Fig 13}
5882: \end{figure}
5883:
5884:
5885:
5886:
5887: \begin{enumerate}
5888: \item $\Si_{1,1}$ and $\Si_{0,4}$ for triangles;
5889: \item \begin{equation}
5890: \Si_{1,1} \cup \Si_{0,4} = \Si_{1,3}
5891: \end{equation}
5892: for squares;
5893: \item $\Si_{0,5}$ for pentagons and
5894: \item $\Si_{1,2}$ for hexagons
5895: \end{enumerate}
5896: (see Figures 13, 15-17).
5897:
5898:
5899:
5900: \begin{figure}[tbn]
5901: \centerline{\epsfxsize=3in\epsfbox{10c.eps.eps}} \caption{\sl Hexagons} \label{Fig 13}
5902: \end{figure}
5903:
5904:
5905:
5906:
5907: Amazing fact is that we can define 2-cells of this complex in
5908: terms of $TD$-graph. It means that $C(TD)$-complex carries no
5909: more information than its 1-skeleton:
5910: \begin{enumerate}
5911: \item triangles are contained in subcomplexes
5912: $C(TD(\Si_{1,1}))$ or $C(TD(\Si_{0,4}))$ which are
5913: the Farey complexes.
5914: Every edge of this complex is a side of two triangles
5915: corresponding
5916: to two subsurfaces
5917: $C(TD(\Si_{1,1}))$ or $C(TD(\Si_{0,4}))$;
5918: \item moreover, if we fix one vertex in any Farey subgraph of $TGD$-graph,
5919: we get
5920: a circle which is a vertex of $C(\Si_{g,n})$ (indeed, fixing any
5921: Farey subgraph of $TD$-graph we get the collection $\{ C_1, \dots
5922: , C_{3g-3} \}$ of circles of
5923: trinion decompositions with fixed set of circles
5924: $( C_2, \dots , C_{3g-3} )$,
5925: since our Farey subgraph corresponds to a choice of $C_1$ );
5926: \item if $C$ is a vertex of $C(\Si)$ (as $C_1$ before ) then there
5927: exists a Farey subgraph with fixed vertex but not unique (indeed
5928: we can extend
5929: our circle $C_1$ to a complete set $\{ C_1, \dots , C_{3g-3} \}$
5930: and consider all
5931: trinion decompositions with fixed $( C_2, \dots , C_{3g-3} )$;
5932: we get a Farey subgraph $F_{C}$ (just a possible choice) with
5933: fixed point $p \in F_C$);
5934: \item now we can define a map
5935: \begin{equation}
5936: \phi \colon Aut C_1(TD) \to Aut C_{0}(\Si_{g,n})
5937: \end{equation}
5938: sending a circle $C \in C_{0}(\Si_{g,n})$ to unique vertex in
5939: $C_{0}(\Si_{g,n})$
5940: corresponding to the marked Farey graph $g(p) \in g (F_C)$;
5941: \item square, pentagon or hexagon ($n$-gones for small $n$)
5942: can be union of triangle and
5943: $n-1$-gon gluied along an edge. To distinguish this trivial
5944: situation from "non trivial" or {\it
5945: alternating}
5946: we consider a loop with $n$ consecutive points such
5947: that every consecutive triple $p_{i},p_{i+1},p_{i+2} $ and the
5948: pair of edges
5949: between arn't contained in any Farey subcomplex;
5950: \item at last, a hexagon can be {\it almost} alternating when its three
5951: consecutive
5952: vertices can be extended to the set of vertices of a square loop.
5953: \end{enumerate}
5954:
5955: Using these simple rules Margalit proved in \cite{M} that
5956: \begin{enumerate}
5957: \item the map $\phi$ (6.47) can be extended to a
5958: well defined map
5959: \begin{equation}
5960: Aut C_1(TD) \to Aut C(TD),
5961: \end{equation}
5962: \item and this map can be extended to an isomorphism
5963: \begin{equation}
5964: \phi \colon Aut C(TD) \to Aut C(\Si),
5965: \end{equation}
5966: \item such that
5967: \begin{equation}
5968: Aut C_{1}(TD)= Aut C(TD)= Aut C(\Si)= Mod_g.
5969: \end{equation}
5970: \end{enumerate}
5971: TD-complex $C(TD)$ was introduced by Hatcher and Thurston
5972: in \cite{HT} where they proved that this complex is connected and
5973: simple connected. But the equality (6.50) was proved quite
5974: recently in \cite{M}.
5975:
5976: \begin{rmk} Returning to the Rational Conformal Field Theory
5977: WZW we can consider the vector space spanned by different
5978: conformal blocks (see Definition 8) interpreted as the
5979: holomorphic sections of some power of theta-line bundle on the
5980: moduli space of semi-stable vector bundles on a Riemann
5981: surface $\Si$ (see section 3).
5982: This surface can be
5983: formed by gluing a number of trinions. The conformal
5984: blocks are obtained by summing over the intermediate states
5985: passing through the circles along which we glue our trinions.
5986: Different gluing procedures lead to different bases of the same
5987: vector space. Hence vectors of
5988: this space obtained by one gluing procedure can be expressed as
5989: linear combinations of vectors found by another way of gluing
5990: the same surface. These linear transformations are given by
5991: {\it dual matrices}, which have to satisfy some
5992: {\it consistency conditions}. To describe these conditions it is
5993: quite convinient to use the finite complex $C(TG)$.
5994: First of all, we
5995: have some very simple duality matrices defined by the elementary moves. The complex $C(TG)$ is connected, hence every duality matrix can be represented as a
5996: product of the simple ones (maybe not uniquely). To get
5997: an unambiguous description of all duality matrices, we have to
5998: use consistency conditions on simple duality matrices which
5999: are given by loops of simple matrices. So we have to find all
6000: independent conditions. Every loop of simple matrices
6001: corresponds to the statement that the product of the corresponding
6002: duality matrices is equal to one. Fixing a collection of
6003: {\it fundamental loops}, like triangles, squares , pentagons
6004: and hexagons before, and filling faces of such loops
6005: gives us the graphs complex $C(TG)$. The relations for the
6006: loops are complete iff the complex is simply connected.
6007: Thus we get a system of (polynomial) equations for duality
6008: matrices. Solutions to this system of equations give the {\it
6009: moduli space of rational conformal field theories}. Sending
6010: the level $k$ to infinity we get a fusion algebra which for our
6011: $\SU(2)$-case is the representation algebra $R(sl(2,\C))$
6012: from subsection 5.1.
6013: \end{rmk}
6014:
6015: Returning to wave function spaces $\sH^k_\Ga$ (4.43)
6016: we will see that the system
6017: of these spaces defines a vector bundle
6018: \begin{equation}
6019: \sH^k_{TG} \to C(TG_g) = S^2
6020: \end{equation}
6021: on the $TG$- complex with a special structure-some local
6022: linear decomposition on line subbundles.
6023:
6024: Remark that the bundles is a subbundle of the trivial bundle
6025: (see (4.45))
6026: \begin{equation}
6027: \sH^k_{TG} \subset \C^{(2k)^{3g-3}} \times C_{TG} = (\oplus_{\al \in T^A_{2k}} \C_\al)
6028: \times C_{TG}.
6029: \end{equation}
6030:
6031:
6032: All these constructions suggest existence of
6033: a flat connection $a_K$. Since the complex $C_{TG}$ is simply
6034: connected the bundle $\sH^k_{TG}$ has to be trivial with the
6035: trivialization defined up to scaling. From this we get the
6036: projective bundle
6037: \begin{equation}
6038: \PP \sH^k \to \sM_g
6039: \end{equation}
6040: on the moduli space of curves of genus $g$ with the projective
6041: flat connection. But this bundle is a slight projective bundle
6042: with the phase amplitude (6.18) only.
6043:
6044: Returning to the graph $C^{1}_{TG_g}$ we will see that
6045: \begin{enumerate}
6046: \item the system of duality matrices defines a flat connection
6047: on $C_{1}(TG_g)$;
6048: \item this connection has trivial projective monodromy, thus
6049: \item this connection defines a representation
6050: \begin{equation}
6051: \rho \colon Aut C_{1}(TG_g) \to PGL(rk_{g}(\emptyset)), \C).
6052: \end{equation}
6053: \item this representation is equal to the desired Kohno representation.
6054: \end{enumerate}
6055:
6056:
6057:
6058:
6059:
6060: \subsection{Gauge theory on graphs.}
6061:
6062: A connection $a$ on the trivial $\SU(2)$-bundle on $\Ga$ is a map
6063: \begin{equation}
6064: a\colon P_1(\Ga)\to \SU(2)
6065: \end{equation}
6066: such that for the orientation reversing involution $i_e$ we have
6067: \begin{equation}
6068: a(i_e(\vec e))=a(\vec e)^{-1}.
6069: \end{equation}
6070: Then the ``path integral'' is
6071: \begin{equation}
6072: a( (\vec e_1,\dots, \vec e_d))=a(\vec e_1) \cdot \cdots \cdot
6073: a(\vec e_d) \in \SU(2)
6074: \end{equation}
6075: that is, it is the map
6076: \begin{equation}
6077: a\colon P_d(\Ga)\to \SU(2)
6078: \end{equation}
6079: such that for the orientation reversing involution $i_\De$ we have
6080: \begin{equation}
6081: a(i((\vec e_1,\dots, \vec e_d)))=a((\vec e_1,\dots, \vec
6082: e_d))^{-1}.
6083: \end{equation}
6084: In the same vein we have the monodromy map for loops
6085: \begin{equation}
6086: a\colon L_d(\Ga)_v \to \SU(2).
6087: \end{equation}
6088: Obviously every such connection is {\it flat }.
6089:
6090: Let $\sA(\Ga)$ be the space of connections that is the space of
6091: functions (6.56) subjecting to the constraint (6.57):
6092: \begin{equation}
6093: \sA(\Ga)=\{ a\in \SU(2)^{P_1(\Ga)} \vert a(i_e(\vec e))=a(\vec
6094: e)^{-1} \}.
6095: \end{equation}
6096:
6097:
6098:
6099: Every element $\w{g}$ of the gauge transformations group
6100: $\sG(\Ga)$ is a function
6101: \begin{equation}
6102: \w{g}\colon V(\Ga)\to \SU(2)
6103: \end{equation}
6104: that is
6105: \begin{equation}
6106: \sG(\Ga)=\SU(2)^{V(\Ga)}
6107: \end{equation}
6108: with the componentwise multiplication.
6109:
6110: This group acts on the space of connections $\sA(\Ga)$ by the
6111: following rule
6112: \begin{equation}
6113: \w{g}(a(\vec e))=\w{g}(v_s(\vec e)) \cdot a(\vec e) \cdot
6114: \w{g}(v_t(i(\vec e))).
6115: \end{equation}
6116: Recall that $\w{g}(v_t(i(\vec e)))= \w{g}(v_t(\vec e))^{-1}.$
6117:
6118: The space of gauge orbits
6119: \begin{equation}
6120: \sB(\Ga)=\sA(\Ga) /\sG(\Ga)
6121: \end{equation}
6122: is the space of classes of representations of the fundamental
6123: group of $\Ga$ as it is expected because all our connections are
6124: flat:
6125: \begin{equation}
6126: \sA(\Ga) /\sG(\Ga)=CLRep(\pi_1(\Ga), \SU(2))
6127: \end{equation}
6128: is the quotient of the space of representations by the adjoint action
6129: \begin{equation}
6130: CLRep(\pi_1(\Ga), \SU(2))=\Hom(\pi_1(\Ga) , \SU(2)) /Ad \SU(2)
6131: \end{equation}
6132: - the space of classes of representations of this group.
6133:
6134: Obviously the fundamental group of 3-valent graph $\Ga$ of genus
6135: $g$ is the free group with $g$ generators. Thus the quotient
6136: (6.66)
6137: \begin{equation}
6138: CLRep(\pi_1(\Ga), \SU(2)) = uS_g
6139: \end{equation}
6140: is the {\it unitary Schottki space}. (see the definition (4) in 3.2
6141: after the formula (3.44)).
6142:
6143:
6144: The gauge transformation group $\sG(\Ga)$ contains the diagonal
6145: subgroup
6146: \begin{equation}
6147: \SU(2)_\De \subset \sG(\Ga)
6148: \end{equation}
6149: of constant functions (6.63). The action of this subgroup defines
6150: the space of constant orbits
6151: \begin{equation}
6152: CL \sA(\Ga)=\sA(\Ga) / \SU(2)_\De.
6153: \end{equation}
6154:
6155: For every vertex $v\in V(\Ga)$ we have the subgroup
6156: \begin{equation}
6157: \sG(\Ga)_v=\{\w{g}\in \sG(\Ga) \vert \w{g}(v)=\id \}
6158: \end{equation}
6159: of gauge transformations preserving the framing at $v$. This is a
6160: normal subgroup, and
6161: \begin{equation}
6162: \sG(\Ga) /\sG(\Ga)_v=\SU(2)_\De
6163: \end{equation}
6164: Thus the full group of gauge transformations is a semidirect
6165: product of $\sG(\Ga)_v$ and $\SU(2)_\De$.
6166:
6167: The quotient
6168: \begin{equation}
6169: \sA(\Ga) /\sG(\Ga)_v=Rep(\pi_1(\Ga, v), \SU(2))
6170: \end{equation}
6171: is the orbit space of framed (at $v$) connections. This space
6172: depends on
6173: the choice of a point $v$.
6174:
6175: Thus the quotient map (6.67)
6176: \begin{equation}
6177: P\colon \sA(\Ga)\to CLRep(\pi_1(\Ga), \SU(2))
6178: \end{equation}
6179: can be decomposed as follows
6180: \begin{equation}
6181: \sA(\Ga) \xrightarrow{\,P\,} Rep(\pi_1(\Ga), \SU(2))
6182: \xrightarrow{/\SU(2)\De} CLRep(\pi_1(\Ga), \SU(2))
6183: \end{equation}
6184: or
6185: \begin{equation}
6186: \sA(\Ga)\xrightarrow{/\SU(2)_\De} CL\sA(\Ga) \xrightarrow{P_{cl}}
6187: CLRep(\pi_1(\Ga), \SU(2)).
6188: \end{equation}
6189: The involution $i_\De$ acts on the space $\sA(\Ga)$ of connections
6190: \begin{equation}
6191: i^*_\De\colon \sA(\Ga)\to \sA(\Ga)
6192: \end{equation}
6193: and there exists an element
6194: \begin{equation}
6195: \w{g}_i=\begin{pmatrix}0&1\\-1&0\end{pmatrix} \in\SU(2)_\De
6196: \subset \sG(\Ga)
6197: \end{equation}
6198: such that
6199: \begin{equation}
6200: i^*_\De=\w{g}_i.
6201: \end{equation}
6202: Thus the involution $i^*_\De$ (3.23) acts
6203: trivially on $CL\sA(\Ga)$. Recall that a gauge fixing is a
6204: section of the projection $P$ (6.75).
6205:
6206:
6207: Recall that the set $\Conj (\SU(2))$ of conjugacy classes of
6208: elements of $\SU(2)$ can be described as the interval $[0,1]$
6209: with respect to the map
6210: \begin{equation}
6211: conj\colon \SU(2)\to \Conj (\SU(2))=[0,1]
6212: \end{equation}
6213: sending a matrix $g\in \SU(2)$ to
6214: \begin{equation}
6215: conj g=\frac{1}{\pi} \cdot \cos^{-1}\Bigl(\frac{1}{2}\Tr g \Bigr)\in
6216: [0,1].
6217: \end{equation}
6218:
6219: Using this map coordinatewise one obtains the map
6220: \begin{equation}
6221: conj\colon \CL\sA(\Ga)\to [0,1]^{3g-3}=\prod_{e\in E(\Ga)}
6222: [0,1]_e,
6223: \end{equation}
6224: which is obviously surjective.
6225:
6226: The map $conj$ is the composite
6227: \begin{equation}
6228: \sA(\Ga) \xrightarrow{\SU(2)_\De} \CL\sA(\Ga)\to
6229: [0,1]^{3g-3}=\prod_{e\in E(\Ga)} [0,1]_e
6230: \end{equation}
6231: and the involution $i_\De$ preserves its fibers.
6232:
6233:
6234:
6235:
6236: \subsection{Abelian gauge theory and $U(1)$-spin networks}
6237:
6238:
6239:
6240:
6241:
6242: Abelian gauge theory on $\Ga$ is a good model for non-Abelian gauge
6243: theory. Abelian spin networks can be considered as
6244: theta-characteristics under numeration of theta functions.
6245: A $U(1)$-connection $a$ is a map
6246: \begin{equation}
6247: a\colon P_1(\Ga)\to U(1)
6248: \end{equation}
6249: such that
6250: \begin{equation}
6251: a(i_{e}(\vec e))=a(\vec e)^{-1}.
6252: \end{equation}
6253: The ``path integral'' is given by
6254: \begin{equation}
6255: a( (\vec e_1,\dots, \vec e_d))=\prod_{\vec e\in P_1(\Ga)} a(\vec
6256: e)\in U(1),
6257: \end{equation}
6258: and so on. Again let
6259: \begin{equation}
6260: \sA_{U(1)}(\Ga)=\bigl\{ a\in U(1)^{P_1(\Ga)} \bigm| a(i_e(\vec
6261: e))=a(\vec e)^{-1} \bigr\}.
6262: \end{equation}
6263: be the space of $U(1)$-connections. Then we have the same
6264: involution
6265: \begin{equation}
6266: \sA_{U(1)}(\Ga) \xrightarrow{i^*_e} \sA_{U(1)}(\Ga).
6267: \end{equation}
6268: Every element $\w{u}$ of the gauge transformation group
6269: $\sG_{U(1)}(\Ga)$ is a function
6270: \begin{equation}
6271: \w{u}\colon V(\Ga)\to U(1).
6272: \end{equation}
6273: This group acts on the space of connections $\sA_{U(1)}(\Ga)$ by
6274: the same rule:
6275: \begin{equation}
6276: \w{u}(a(\vec e))=\w{u}(v_s(\vec e)) \cdot a(\vec e) \cdot
6277: \w{u}(v_t(i(\vec e))).
6278: \end{equation}
6279:
6280: However, there is one important difference between Abelian and
6281: non-Abelian theories. Namely, the diagonal group
6282: \begin{equation}
6283: U(1)_\De \subset \sG_{U(1)}(\Ga)
6284: \end{equation}
6285: has trivial action. Thus
6286: \begin{equation}
6287: \sB_{U(1)}(\Ga)=\sA_{U(1)}(\Ga) /\sG_{U(1)}(\Ga)=\Hom(\pi_1(\Ga) ,
6288: U(1))=U(1)^g.
6289: \end{equation}
6290: Remark that this is abelian unitary Schottki space, that is, the
6291: zero fiber of the projection $\pi$ (2.22) on which we are
6292: constructing our delta functions (2.46) and (2.48) for the
6293: application of the coherent state transform to get an analytical
6294: presentation of theta functions on the complex Schottki space
6295: $(\C^*)^g$. We would like to do this for non-abelian case too.
6296:
6297: Let
6298: \begin{equation}
6299: P_a\colon \sA_{U(1)}(\Ga)\to U(1)^g
6300: \end{equation}
6301: be the projection map.
6302:
6303: The abelian and non-abelian theories are related by the following
6304: map:
6305: \begin{equation}
6306: d\colon \sA_{U(1)}(\Ga)\to \sA(\Ga)
6307: \quad\hbox{given by}\quad
6308: d(a(\vec e))=\begin{pmatrix} e^{i \phi}, 0\\ 0, e^{- i \phi}
6309: \end{pmatrix}.
6310: \end{equation}
6311: This map $d$ is equivariant with respect to every involution
6312: $i^*_e$. Obviously the image $d (\sA_{U(1)}(\Ga) ) \subset
6313: \sA(\Ga)$ is a 2-section of the projection $conj$, that is, the
6314: composite
6315: \begin{equation}
6316: conj \circ d\colon \sA_{U(1)}(\Ga)\to \prod_{e\in E(\Ga)}
6317: [0,1]_e
6318: \end{equation}
6319: is the factorization by the involution $i^*_\De$.
6320:
6321: It is easy to check
6322:
6323: \begin{prop} In the chain of maps
6324: \begin{equation}
6325: \sA_{U(1)}(\Ga) \xrightarrow{\,d\,} \sA(\Ga) \xrightarrow{\,P\,}
6326: U(1)^g\in \CLRep (\pi_1(\Ga))
6327: \end{equation}
6328: the composite $d \circ P$ is the projection map $P_a$ (5.75).
6329: \end{prop}
6330:
6331:
6332: The set of irreducible representations of $U(1)$ is
6333: \begin{equation}
6334: \widehat{U(1)} = \Z
6335: \end{equation}
6336: and a spin $U(1)$-network on a graph $\Ga$ is given by a function
6337: \begin{equation}
6338: w \colon E(\Ga) \to \widehat{U(1)} = \Z
6339: \end{equation}
6340: (like (4.46)) such that for every triple of edges meeting at a
6341: vertex $v \in V(\Ga)$ the following condition holds
6342: \begin{equation}
6343: w(e_1) + w(e_2) + w(e_3) = 0.
6344: \end{equation}
6345: Thus the triangle inequality (Clebsh-Gordan conditions)
6346: becomes the equality.
6347:
6348: Again a spin network is of level $k$ when $w(e) \leq 2k$.
6349:
6350: Considering the basis $(a_1 , \dots , a_g, b_1, \dots b_g)$
6351: (2.32) we have the decomposition of the jacobian of the curve
6352: $\Si_\Ga$
6353: \begin{equation}
6354: J_{\Si_\Ga} = Hom(\pi_1(\Si_\Ga), U(1)) = \prod_{i=1}^g
6355: U(1)_{a_i}\times \prod_{i=1}^g U(1)_{b_i}
6356: \end{equation}
6357: which coincides with the decomposition (2.21),(2.32).
6358:
6359: Now we can consider the coordinates (4.14) of the map (4.41) as
6360: elements of
6361: the target group $U(1)$. Hence the map $\pi_\Ga$ (4.41) has the
6362: form
6363: \begin{equation}
6364: \pi_\Ga \colon J_{\Si_\Ga} \to U(1)^{E(\Ga)} = T^A
6365: \end{equation}
6366: where the target is the action torus (4.40). The image of this
6367: map is a g-dimensional torus
6368: \begin{equation}
6369: \De_\Ga = T^g_- \subset T^A.
6370: \end{equation}
6371: Now we can send every spin network that is a function (6.98)
6372: satisfying (6.99) to a point of order $k$ on $T^A$ dividing by
6373: $k$
6374: \begin{equation}
6375: w(e) \to 1/k w(e).
6376: \end{equation}
6377: We get the map of the set $SNW_a^k$
6378: of abelian spin works of level $k$
6379: \begin{equation}
6380: 1/k \colon SNW_a^k \to (T^A)_k.
6381: \end{equation}
6382: Amusing fact is that
6383: \begin{equation}
6384: 1/k (SNW_a^k) \subset \De_\Ga = T^g_-.
6385: \end{equation}
6386: Moreover
6387: \begin{equation}
6388: 1/k(SNW_a^k) = W^k(\Ga) = BS_k(\Ga).
6389: \end{equation}
6390: Thus we get the canonical identification of the set $\Z^g/ k\Z^g$ of
6391: theta characteristics of level $k$ (2.48) with the set of spin
6392: networks
6393: \begin{equation}
6394: (\Z^g / k\Z^g) = SNW_a^k.
6395: \end{equation}
6396: So in the classical case we can use $U(1)$-spine networks for the
6397: enumeration of theta characteristics. In the same vein we will use
6398: $\SU(2)$-spin networks
6399: for the enumeration of non-abelian theta characteristics.
6400:
6401: Also we saw that the harmonic analysis on $T^g_+ = U(1)^g$ is
6402: the standard Fourier analysis on a torus. We use the Fourier
6403: decomposition
6404: to define our delta-functions (2.46) and (2.48) which CST $C_t^{-i \Om}$
6405: (2.42) sends to classical theta functions. Now we have to extend these
6406: constructions to the non-abelian case.
6407:
6408: \subsection{Harmonic analysis of $\SU(2)$-spin networks}
6409:
6410: Recall that a spin network defines a map $ j \colon E(\Ga)\to
6411: \widehat{SU(2)}$ with constraints. For a triple of representations
6412: $j_{e_1}, j_{e_2}, j_{e_3}$ an
6413: intertwiner is trivial component of the tensor product
6414: $j_{1} \tensor j_{2} \tensor j_{3}$. Such a component exists iff
6415: under the traditional identification $\widehat{SU(2)}=\half
6416: \Z^{+}$ the Clebsch--Gordan conditions
6417: \begin{equation}
6418: j_{1}+j_{2}+j_{3}\in \Z \quad\text{and}\quad \vert j_{1}-j_{2} \vert \le
6419: j_{3} \le j_{1}+j_{2}
6420: \end{equation}
6421: (the triangle inequality) hold. Function $ j \colon E(\Ga)\to
6422: \widehat{SU(2)}$ defines a spin network $\Ga_j$ iff these
6423: conditions hold for every triple $j_{v,1}, j_{v,2}, j_{v,3}$ of
6424: representations around every vertex $v\in V(\Ga)$.
6425:
6426: Any intertwiner $ i_v\in j_{v,1} \tensor j_{v,2} \tensor j_{v,3}$
6427: is defined uniquely so one omits any labeling of the vertices
6428: and denotes $\SU(2))$-spin network by the symbol $\Ga_{j}$.
6429:
6430: Spin network $\Ga_{j}$
6431: is of level $k$ if $j_{i} \le k$ for every edge $e_i$ and in the
6432: last inequality we take $min(j_{1}+j_{2}, k - j_{1}-j_{2})$
6433: instead of $j_{1}+j_{2}$. Of course if $k' > k$ then every spin
6434: network of level $k$ is of level $k'$ automatically. We would
6435: like to fix a level and denote a spin network of level $k$ by the
6436: symbol $\Ga_{j^k}$ and the set of all spin net works by the symbol
6437: $SNW^k(\Ga)$. Recall that we have the identification $SNW^k(\Ga) = W^k(\Ga)
6438: $ (see the end of subsection 6.1).
6439:
6440:
6441:
6442: Spin network $\Ga_{j, k}$ is a purely combinatorial object but we
6443: can identify it with
6444: a function on the special space - the unitary Schottky space $uS_g$
6445: given by the formula
6446: \begin{equation}
6447: uS_{g}=\SU(2)^{g} / Ad_{\mathrm{diag}} \SU(2)
6448: \end{equation}
6449: where $Ad_{\mathrm{diag}}\SU(2)$ is the diagonal adjoint action
6450: on the direct product. This space doesn't depend on a three
6451: valent graph modulo the choice of appropriate topological data.
6452: Only one invariant of a graph is used, namely its
6453: genus $g$.However, it is quite natural to use a graph $\Ga$
6454: as a starting point of the construction.
6455:
6456: Consider the product
6457: \begin{equation}
6458: \SU(2)^{E(\Ga)}=\prod_{e\in E(\Ga)} \SU(2)_e
6459: \end{equation}
6460: with $\SU(2)$ components enumerated by edges of $\Ga$ which we
6461: can identify
6462: with the space of flat connections $\sA$ (6.88) (under a choice of any
6463: orientation of $\Ga$).
6464:
6465: Let $dx$ be the Haar measure on $\SU(2)$ normalized by the
6466: condition $\int_{\SU(2)}dx=1$ and $\vec dx$ the product measure
6467: on $\SU(2)^{E(\Ga)}$ normalized by $\int \vec dx=1$. Then by the
6468: Peter--Weyl formula, any function $f\in L^{2}(\SU(2)^{E(\Ga)},
6469: \vec dx)$ has the decomposition
6470: \begin{equation}
6471: f(x)=\sum_{\vec{\rho}\in \widehat{\SU(2)^{E(\Ga)}}}
6472: \Tr[B_{\vec{\rho},f} \vec{\rho}(x)],
6473: \end{equation}
6474: where $\widehat{\SU(2)^{E(\Ga)}}$ is the space of irreducible
6475: representations of $\SU(2)^{E(\Ga)}$, and $B_{\vec{\rho},f}$ are
6476: endomorphisms of the space $V_{\vec{\rho}}$ of the representation
6477: $\vec{\rho}$, given by
6478: \begin{equation}
6479: B_{\vec{\rho},f}=\frac{1}{\dim V_{\vec{\rho}}}\int_{\SU(2)^{E(\Ga)}}
6480: f(x) \vec{\rho}\,^{-1}(x) \vec dx.
6481: \end{equation}
6482: Every irreducible representation of $\SU(2)^{E(\Ga)}$ is given by
6483: tensor product of irreducible representations of $\SU(2)$ $
6484: \vec\rho=\rho_{1}\tensor \cdots \tensor\rho_{3g-3}$. So every spin
6485: network $\Ga_{j}$ of genus $g$ defines a representation of
6486: $\SU(2)^{E(\Ga)}$ by the tensor product of all labels
6487: $\vec j=\bigotimes_{e\in E(\Ga)} j_e$.
6488:
6489: Of course this is an analog of the standard Fourier decomposition
6490: (2.41) or even (2.46). Here the representation $\vec \rho$ is a
6491: label of {\it frequency} and an endomorphism $B_{\vec \rho, f}$
6492: is the {\it Fourier coefficient } that is the analog of a {\it
6493: number}. The formula (6.113) is nothing else than the integral
6494: formula for a Fourier coefficient.
6495:
6496: Every endomorphism of the space $V_{\vec j}$ is a vector in the
6497: tensor product
6498: $$
6499: \Bigl(\bigotimes_{e\in E(\Ga)} j_e\Bigr)
6500: \tensor \Bigl(\bigotimes_{e\in E(\Ga)} j_e\Bigr)^*=
6501: \Bigl(\bigotimes_{e\in E(\Ga)} j_e\Bigr)
6502: \tensor \Bigl(\bigotimes_{e\in E(\Ga)} j_e\Bigr),
6503: $$
6504: since we are dealing with $\SU(2)$-representations.
6505:
6506: But after symmetrizing components of the final product can be
6507: labeled by elements of the set $F(\Ga)$, and we can decompose it
6508: as
6509: \begin{equation}
6510: \Bigl(\bigotimes_{e\in E(\Ga)} j_e\Bigr)
6511: \tensor \Bigl(\bigotimes_{e\in E(\Ga)} j_e\Bigr)^*=
6512: \bigotimes_{v\in V(\Ga)}(j_{v,1} \tensor j_{v,2} \tensor j_{v,3}).
6513: \end{equation}
6514:
6515: For every triple of representations around a vertex $v\in V(\Ga)$ we
6516: have a vector $ i_v\in j_{v,1} \tensor j_{v,2} \tensor j_{v,3}$
6517: and their tensor product gives us the vector
6518: \begin{equation}
6519: B(\Ga_j)=\bigotimes_{v\in V(\Ga)} i_v \in End V_{\vec j}\,.
6520: \end{equation}
6521: So one has the function
6522: \begin{equation}
6523: f_{\Ga_{j}}(x)= Tr[B(\Ga_j) \vec j(x)]\in L^{2}(\SU(2)^{E(\Ga)},
6524: d \vec x).
6525: \end{equation}
6526:
6527: This function is invariant with respect to the action on
6528: $\SU(2)^{E(\Ga)} = \sA $ of the gauge group on $\Ga$ $
6529: \SU(2)^{V(\Ga)}=\prod_{v\in V(\Ga)} \SU(2)_v$ with components
6530: enumerated by the vertices and an orientation of $\Ga$.
6531:
6532: It is easy to see that the endomorphism $B(\Ga_j)$ (5.96) is an
6533: intertwiner with respect to such action. Then by the standard
6534: result of harmonic analysis on groups our function $f_{\Ga_{j}}$
6535: is invariant with respect to
6536: this action. That is, $f_{\Ga_{j}}$ is a function on the homogeneous
6537: space
6538: \begin{equation}
6539: uS_g = \sA / \sG.
6540: \end{equation}
6541:
6542: But the set $\{ \Ga_j\} = SNW^k(\Ga) = W^k(\Ga)$ enumerates the
6543: collection
6544: of functions $f_{\Ga_j}$ (6.115) each of which is an analog of the
6545: function
6546: (2.48) enumerated by the set $(\Z^g / k\Z^g)$ of theta characteristics.
6547:
6548: \begin{prop} Graph $\Ga$ and function $f_{\Ga_j}$ determine
6549: spin network
6550: $\Ga_j$ uniquely.
6551: \end{prop}
6552: Indeed,
6553: \begin{enumerate}
6554: \item knowing graph we can reconstruct the map $P \colon \sA \to uS_g$;
6555: \item lifting function $f_{}\Ga_j$ on $\sA = \SU(2)^{E(\Ga)}$ and using
6556: the harmonic decomposition of this function on $\sA$
6557: (6.61)
6558: we determine the representation $j_1 \otimes \dots \otimes j_{3g-3}$ and
6559: the space $V$ of this representation;
6560: \item by the Peter-Weyl formula we determine the endomorphism $B(\Ga_j)$
6561: (6.114);
6562: \item decomposing this endomorphism on blocks (6.113)
6563: correspoding to
6564: vertices we reconstruct the function $j$ (6.4).
6565: \end{enumerate}
6566: and we are done.
6567:
6568: Now we are ready to construct CST for the pair $uS_g \subset S_g$
6569: (see the end of 3.2) since the complex Schottki space $S_g$ is
6570: the complexification of $uS_g$.
6571:
6572:
6573:
6574:
6575:
6576: \subsection{RCFT and Kohno representation}
6577:
6578:
6579: The usual data for RCFT include an algebra $A$ and
6580: a discrete set $R$ of representations
6581: (a finite set for level $k$) $\{ \rho_i \}$
6582: (just like in subsection 5.1) such that
6583: \begin{equation}
6584: \rho \in R \implies \rho^* \in R,
6585: \end{equation}
6586: and a distinguished element $1$ such that $1^* = 1$. Moreover for every
6587: ordered triple $\rho_i, \rho_j, \rho_k$ we have a finite (rationality of RQFT)
6588: dimensional space $V^{\rho_i}_{\rho_j, \rho_k}$.
6589: \begin{equation}
6590: dim V^1_{\rho_i, \rho_j} = 1 \quad \text{ if } \quad \rho_i =
6591: \rho_j^*
6592: \end{equation}
6593: and 0 otherwise.
6594:
6595: \begin{rmk} So we can extract a {\it modular tensor category}
6596: from RCFT.
6597: \end{rmk}
6598:
6599: Having such $RCFT$ and 3-1-valent graph $\Ga$ and a spin network
6600: $\Ga_j$ we get from every vertex $v \in V(\Ga)$ the spaces with
6601: edges $e_1, e_2, e_3$ through it thus we have the space of
6602: intertwiners
6603: \begin{equation}
6604: V^{j_1}_{j_2, j_3} \in j_1 \otimes j_2 \otimes j_3
6605: \end{equation}
6606: (like in (6.6)). Taking tensor products of them together over all
6607: vertices
6608: we get the space
6609: \begin{equation}
6610: \sH^k_{\Ga_j} = \otimes_{v \in V(\Ga)} V^{j_1}_{j_2, j_3}
6611: \end{equation}
6612: and taking a direct product over all possible networks (labelings) we get
6613: the space
6614: \begin{equation}
6615: \sH^k_{\Ga} = \oplus_{j \in SNW^k(\Ga)} \sH^k_{\Ga_j}.
6616: \end{equation}
6617: For $\SU(2)$-case we have
6618: \begin{enumerate}
6619: \item $ dim V^{j_1}_{j_2, j_3} = 1$;
6620: \item $ dim \sH^k_{\Ga_j} = 1$;
6621: \item the space (6.121) is precisely our space (4.39), (4.43)
6622: of wave functions under
6623: Bohr-Sommerfeld quantization.
6624: \end{enumerate}
6625: It is useful to fix bases in all our spaces. (In $\SU(2)$-case
6626: where all spaces are 1-dimensional we can do it using the
6627: standard Wigner 3j-symbols (for recalling and quantum groups
6628: extention see \cite{KR}). We equipe
6629: the wave function spaces with bases.
6630:
6631: This construction hasn't
6632: to be invariant with respect to changes in trinion decompositions
6633: (or changes of bases). But having the complex $TG_g$ we know
6634: the obstackles.
6635:
6636: To get such invariance we have to produce matrices
6637: of transformations to new bases and check that these matrices satisfy the
6638: same algebraic identities as the changes of decompositions themselves.
6639: We have to check that these matrices satisfy the same algbraic
6640: equations as
6641: our transformations itselves (see Fig.11, Definition 9 and 1)-4) around (6.47)).
6642:
6643: Fortunately there are 5 elementary moves only described in \cite{MS} with
6644: traditional symbols
6645: \begin{equation}
6646: F, \quad S, \quad, T, \quad, \Om, \quad , \Theta
6647: \end{equation}
6648: satisfying the required algebraic relations.
6649:
6650:
6651: Recall that elementary transformations of 3-valent graphs admit
6652: relations 1)-4) around (6.47). So our intertwiners vector spaces must
6653: satisfy the system of identities described below:
6654: \begin{enumerate}
6655: \item 3 transformations are diagonal matrices of phases (6.18)
6656: (in an appropriate basis)
6657: \begin{equation}
6658: \Om^{\pm}, \quad \Theta, \quad T \colon V^{\rho_i}_{\rho_j, \rho_k}
6659: \to V^{\rho_i}_{\rho_j, \rho_k};
6660: \end{equation}
6661: \item "genus 1 equality "
6662: \begin{equation}
6663: S \colon V^{\rho_i}_{\rho_j, \rho_j^*} \to
6664: \sum_l V^{\rho_i}_{\rho_l, \rho_l^*};
6665: \end{equation}
6666: \item the fusing matrix
6667: \begin{equation}
6668: F \colon V^{\rho_i}_{\rho_j, \rho_k} \otimes V^{\rho_k}_{\rho_t, \rho_m} \to
6669: \sum_{l} V^{\rho_j}_{\rho_t, \rho_l} \otimes V^{\rho_l}_{\rho_j, \rho_m};
6670: \end{equation}
6671: \item the braiding matrix
6672: \begin{equation}
6673: B^\pm = F \Om^\pm F^{-1}
6674: \end{equation}
6675: subjecting the ordinary Reidemeister relations
6676: \begin{equation}
6677: B_{12} B_{23} B_{12} = B_{23} B_{12} B_{23}.
6678: \end{equation}
6679: \end{enumerate}
6680: We can see that the most general topological (geometrical) setting
6681: for applying of discussed TQFT is a 3-manifold containing a
6682: 3-valent graph (instead a loop = knot). So the most striking
6683: difference between such objects and spin networks is following:
6684: in spin networks the holonomy of one edge around another is
6685: trivial that is any "braiding" relations are absent. In spin
6686: networks it isn't necessary to distinguish over and under crossing
6687: when a 3-valent graph is projected on a plane. Otherwise all
6688: rules for evaluating spin networks closely parallel the skein
6689: rules for the Jones polynomials. Here we are translating RCFT
6690: setting to Penrose's interpretation of the quantum gravity.
6691:
6692: Recall that all braid properties of all matrices in the game are
6693: arising from the holonomy of the Knizhnik-Zamolodchikov equation
6694: (see \cite{KZ}) and the polynimial equations for these matrices
6695: expected from the Conformal Field Theory were proposed by Moore
6696: and Seiberg in \cite{MS}.
6697:
6698: Consider a 1-3-valent graph $\Ga$ which corresponds to a trinion
6699: decomposition of a pinched Riemann sphere $ \Si = \PP^{1} - p_{1}
6700: - \dots - p_{n} $.
6701:
6702: Let
6703: \begin{equation}
6704: \{I_{\mu}\}, \quad \mu = 1,2,3
6705: \end{equation}
6706: be any orthonormal basis of $sl(2,\C)$ with respect to the
6707: Cartan-Killing form and
6708: \begin{equation}
6709: \Om = \sum_{\mu} I_{\mu} \otimes I_{\mu} \in End (1) =
6710: End(sl(2,\C)).
6711: \end{equation}
6712: Let $\pi_{i}(I_{\mu})$ be the action on i-th component of the
6713: tensor product $j_{1} \otimes \dots \otimes j_{n}$ (of some
6714: labeling of the graph) and
6715: \begin{equation}
6716: \Om_{ij} = \sum_{\mu} \pi_{i}(I_{\mu}) \pi_{j}(I_{\mu}) \in
6717: End(j_{1} \otimes \dots \otimes j_{n}).
6718: \end{equation}
6719: Consider the configuration space
6720: \begin{equation}
6721: \C^{n} - \De = \{(z_{1}, \dots, z_{n}) \vert i \neq j \implies
6722: z_{i} \neq z_{j}\}
6723: \end{equation}
6724: and the trivial vector bundle on it with the fiber $ j_{1} \otimes
6725: \dots \otimes j_{n}$. It is easy to check that the KZ-equation
6726: \begin{equation}
6727: \p \Phi / \p z_{i} = \frac{1}{k+2} \sum_{i \neq j}
6728: \frac{\Om_{ij}}{z_{i}-z_{j}} \Phi, \quad 1 \leq i \leq n
6729: \end{equation}
6730: defines a slightly projective flat connection
6731: \begin{equation}
6732: \om = \frac{1}{k+2} \sum_{i < j} \Om_{ij} d log(z_{i}-z_{j})
6733: \end{equation}
6734: on this trivial bundle. Any solution of the KZ-equation gives a
6735: {\it covariant constant} (horizontal) section of this bundle.
6736:
6737: Using the detailed information on the monodromy of this equation
6738: Tsuchiya and Kanie proposed the operator formalism two
6739: dimensional conformal field theory in \cite{TK}. We can use this
6740: information too but we would like to present the matrices of
6741: MS-axiomatic description in more background independent way.
6742:
6743: First of all consider the sphere with 4 holes $\Si_{0,4}$ and
6744: two graphs $\Ga$ and $\Ga'_{e}$ corresponding to two types of
6745: trinion decompositions of it (see Fig.11).
6746:
6747: Symbolically we can note every labeling of $\Ga$ by the matrix
6748: \begin{equation}
6749: \begin{pmatrix}
6750: j_{2}&&&&j_{3}\\
6751: &&j&&\\
6752: j_{1}&&&&j_{4}
6753: \end{pmatrix}
6754: \end{equation}
6755: Then for the graph $\Ga'_{e}$ the matrix has the form
6756: \begin{equation}
6757: \begin{pmatrix}
6758: j_{3}&&&&j_{4}\\
6759: &&j'&&\\
6760: j_{2}&&&&j_{1}
6761: \end{pmatrix}
6762: \end{equation}
6763: Remark that the sets of parabolic edges are the same
6764: \begin{equation}
6765: P(\Ga) = P(\Ga'_{e}) = e_{1}, e_{2}, e_{3}, e_{4}
6766: \end{equation}
6767: but the subsets $E_{v}$ are different. Thus we have two spaces
6768: \begin{equation}
6769: \sH^{k}_{\Ga, j_{1}, j_{2}, j_{3}, j_{4}} = Hom_{\SU(2)}(j_{1}
6770: \otimes j_{2}, j_{3} \otimes j_{4})
6771: \end{equation}
6772: (6.3) with the basis given by the set of weights $ W^{k} (\Ga,
6773: j_{1}, j_{2}, j_{3}, j_{4}) $ (6.1) and
6774: \begin{equation}
6775: \sH^{k}_{\Ga'_{e}, j_{1}, j_{2}, j_{3}, j_{4}} =
6776: Hom_{\SU(2)}(j_{2} \otimes j_{3}, j_{4} \otimes j_{1})
6777: \end{equation}
6778: with the basis $ W^{k} (\Ga'_{e}, j_{1}, j_{2}, j_{3}, j_{4}) $.
6779:
6780: Obviously these spaces of (conformal) blocks coincide but the
6781: bases are different.
6782:
6783: \begin{dfn}
6784: The matrix
6785: \begin{equation}
6786: F_{e} \colon \sH^{k}_{\Ga, j_{1}, j_{2}, j_{3}, j_{4}} =
6787: Hom_{\SU(2)}(j_{1} \otimes j_{2}, j_{3} \otimes j_{4}) \to
6788: \sH^{k}_{\Ga'_{e}, j_{1}, j_{2}, j_{3}, j_{4}} =
6789: Hom_{\SU(2)}(j_{2} \otimes j_{3}, j_{4} \otimes j_{1})
6790: \end{equation}
6791: sending the first basis to the second is called the fusion matrix.
6792: \end{dfn}
6793:
6794: Let the symbol of the first basis be
6795: \begin{equation}
6796: \begin{pmatrix}
6797: j_{2}&j_{3}\\
6798: j_{1}&j_{4}
6799: \end{pmatrix}
6800: \end{equation}
6801: and of the second be
6802: \begin{equation}
6803: \begin{pmatrix}
6804: j_{3}&j_{4}\\
6805: j_{2}&j_{1}
6806: \end{pmatrix}
6807: \end{equation}
6808: Then the fusion matrix
6809: \begin{equation}
6810: F \begin{pmatrix}
6811: j_{2}&j_{3}\\
6812: j_{1}&j_{4}
6813: \end{pmatrix} = \begin{pmatrix}
6814: j_{3}&j_{4}\\
6815: j_{2}&j_{1}
6816: \end{pmatrix}
6817: \end{equation}
6818: can be considered as a transformation of solutions of KZ-equation
6819: with coefficients
6820: \begin{equation}
6821: F_{ij} \begin{pmatrix}
6822: j_{2}&j_{3}\\
6823: j_{1}&j_{4}
6824: \end{pmatrix}
6825: \end{equation}
6826: In these notations it is easy to check the following equalities
6827: \begin{enumerate}
6828: \item \begin{equation}
6829: \sum_{j} F_{ij}\begin{pmatrix}
6830: j_{2}&j_{3}\\
6831: j_{1}&j_{4}
6832: \end{pmatrix} F_{jk}\begin{pmatrix}
6833: j_{3}&j_{4}\\
6834: j_{2}&j_{1}
6835: \end{pmatrix} = \delta_{ik};
6836: \end{equation}
6837: \item \begin{equation}
6838: F_{ij}\begin{pmatrix}
6839: j_{2}&j_{3}\\
6840: j_{1}&j_{4}
6841: \end{pmatrix} = F_{ij}\begin{pmatrix}
6842: j_{4}&j_{1}\\
6843: j_{3}&j_{2}
6844: \end{pmatrix}
6845: \end{equation}
6846: \item for $\Si_{0,5}$ five fusion matrices are subjecting the
6847: pentagone equation (see 3) below (6.47)).
6848: \end{enumerate}
6849:
6850: Now using fusion matrices and simply connectedness of $C(TG)$
6851: from 5.2 we can identify uniquely all spaces (4.39), (4.43) and
6852: (6.121).
6853:
6854: Using the same symbols we can define the collection of {\it
6855: braiding matrices}. Again we can describe it as monodromies of
6856: covariant constant solutions to the KZ-equation. Symbolically we
6857: have the chain of transformations
6858: \begin{equation}
6859: \begin{pmatrix}
6860: j_{2}&j_{3}\\
6861: j_{1}&j_{4}
6862: \end{pmatrix} \to \begin{pmatrix}
6863: j_{1}&j_{3}\\
6864: j_{2}&j_{4} \end{pmatrix} \xrightarrow{ F } \begin{pmatrix}
6865: j_{3}&j_{4}\\
6866: j_{1}&j_{2}
6867: \end{pmatrix} \to \begin{pmatrix}
6868: j_{3}&j_{2}\\
6869: j_{1}&j_{4}
6870: \end{pmatrix}
6871: \end{equation}
6872: \begin{dfn} The composition of these transformations
6873: is called a braid matrix.
6874: \end{dfn}
6875: Again we denote coefficients of this matrix by the symbol
6876: \begin{equation}
6877: B_{ij}\begin{pmatrix}
6878: j_{2}&j_{3}\\
6879: j_{1}&j_{4}
6880: \end{pmatrix}
6881: \end{equation}
6882: This braiding matrix can be diagonalized by the fusion matrix:
6883: \begin{equation}
6884: B \begin{pmatrix}
6885: j_{2}&j_{3}\\
6886: j_{1}&j_{4}
6887: \end{pmatrix} = F \begin{pmatrix}
6888: j_{2}&j_{3}\\
6889: j_{1}&j_{4}
6890: \end{pmatrix}^{-1} D F \begin{pmatrix}
6891: j_{2}&j_{3}\\
6892: j_{1}&j_{4}
6893: \end{pmatrix}
6894: \end{equation}
6895: where $D = \{d_{i}\}$ is the diagonal matrix with elements
6896: \begin{equation}
6897: d_{i} = (-1)^{j_{2} + j_{3} - i} e^{\pi i (\frac{i(i+1)}{k+2} -
6898: \frac{j_{2}(j_{2}+1)}{k+2} - \frac{j_{3}(J_{3}+1)}{k+2})}.
6899: \end{equation}
6900: Moreover
6901: \begin{equation}
6902: B_{ij} \begin{pmatrix}
6903: j_{2}&j_{3}\\
6904: j_{1}&j_{4}
6905: \end{pmatrix} = (-1)^{j_{1} + j_{4} - i - j} e^{\pi i (-1)
6906: (\frac{i(i+1)}{k+2} + - \frac{j(j+1)}{k+2} -
6907: \frac{j_{1}(j_{1}+1)}{k+2} - \frac{j_{4}(J_{4}+1)}{k+2})} F_{ij}
6908: \begin{pmatrix}
6909: j_{1}&j_{3}\\
6910: j_{2}&j_{4}
6911: \end{pmatrix}
6912: \end{equation}
6913: All of these equalities can be checked using monodromies of
6914: KZ-equation for $n = 4$ (or for $n = 3$ plus $\infty$ as in
6915: \cite{K1}).
6916:
6917: Using braid matrix $B$ we can defined matrices $\Om^{\pm}$ (6.123)
6918: by the formula (6.127).
6919:
6920: For every edge $e \in E(\Ga)$ we have the diagonal transformation
6921: of the space $\sH^{k}_{\Ga}$ (4.41) given in the basis $\{w\}$
6922: (4.39) by the formula
6923: \begin{equation}
6924: T_{e}(w) = e^{2\pi i (\frac{w(e)(w(e) + 1)}{k+2} -
6925: \frac{k}{8(k+2)})} w.
6926: \end{equation}
6927:
6928:
6929: Now consider the first "elliptic" elementary transformation in
6930: Fig.11. This transformation preserves the shape of the graph
6931: $\Ga_{1,1}$ but changes "a" and "b" standard generators of the
6932: fundamental group
6933: of 1-torus. This graph has one loop edges $a$, one 3-valent vertex
6934: $v$ and one parabolic edge $p$ and corresponds to 1-holed torus.
6935: For 2-holed torus the graph $\Ga_{1,2}$ has two edges $a$ and
6936: $a'$, two vertices $v_{1}$ and $v_{2}$ and two corresponding
6937: parabolic edges $p_{1}$ and $p_{2}$.
6938:
6939:
6940: So we have the spaces (6.120)
6941: \begin{equation}
6942: \sH^{k}_{\Ga_{1,1}, j} \quad \text{ and } \sH^{k}_{\Ga_{1,2},
6943: j_{1}, j_{2}}.
6944: \end{equation}
6945: (of course the number $j$ has to be integer in this case).
6946:
6947: Using fusion matrices around edges "a" we construct the
6948: isomorphism
6949: \begin{equation}
6950: f \colon \sH^{k}_{\Ga_{1,2}, j_{1}, j_{2}} \to \oplus_{j}
6951: \sH^{k}_{\Ga_{1,1}, j}
6952: \end{equation}
6953: where $j$ satisfies the Clebsh-Gordan conditions with $j_{1}+
6954: j_{2} + j \leq k$. The source space admits the basis $w_{kl}$
6955: which corresponds to the admissible weights $w(a) = k$, $w(a') =
6956: l$. In this basis using braid matrix we define the linear operator
6957: \begin{equation}
6958: \sB (w_{kl}) = \sum_{n} B_{nk} \begin{pmatrix}
6959: j_{1}&j_{2}\\
6960: l & l
6961: \end{pmatrix} w_{nl}
6962: \end{equation}
6963: on the space $\sH^{k}_{\Ga_{1,2}, j_{1}, j_{2}}$.
6964:
6965: Now we are ready to define {\it switching operators} representing
6966: the action of the modular group $Mod_{1,1}$ on the space
6967: $\sH^{k}_{\Ga_{1,1}, j}$. More precisely, it is the diffeomerphism
6968: of the elliptic elementary transformation in Fig. 11.
6969:
6970:
6971:
6972: \begin{dfn}
6973: An endomorphism
6974: \begin{equation}
6975: S_{k}(j) \in End \sH^{k}_{\Ga_{1,1}, j}
6976: \end{equation}
6977: is called a switching operator if the following 3 conditions are
6978: satisfied
6979: \begin{enumerate}
6980: \item \begin{equation}
6981: S_{k}(j)^{2}= (-1)^{j} e^{\pi i \frac{j(j+1)}{k+2}} \circ id;
6982: \end{equation}
6983: \item
6984: \begin{equation}
6985: (S_{k}(j) T_{e=a})^{3} = S_{k}(j)^{2};
6986: \end{equation}
6987: \item let \begin{equation}
6988: S_{k}(j_{1}, j_{2}) = f^{-1} \circ (\oplus_{j}S_{k}(j)) \circ f
6989: \end{equation}
6990: be the endomorphism of $\sH^{k}_{\Ga_{1,2}, j_{1}, j_{2}}$. Then
6991: \begin{equation}
6992: S_{k}(j_{1, j_{2}}) \circ T_{e=a}^{-1} \circ T_{e=a'}\circ
6993: S_{k}(j_{1}, j_{2})^{-1} = \sB.
6994: \end{equation}
6995: \end{enumerate}
6996: \end{dfn}
6997: (Geometrically the operator $S_{k}(j_{1}, j_{2})$ has to be an
6998: intertwiner of diffeomerphsims corresponding to sliding the hole
6999: $p_{2}$ along "a" and "b" paths on 2-torus).
7000:
7001: The existence of solutions for small $k = 1, 2, 3$ can be varified
7002: by elementary computations and for general $k$ can be obtained
7003: from $\SU(2)$-WZW model over 2-holed torus (see for example
7004: \cite{TUY}).
7005:
7006: \begin{rmk}
7007: We can see that for a closed graph (without parabolic edges) we
7008: are using braiding matrices in the equality (6.160) only. Moreover
7009: in this equality we are using the combination $\sB$ of braiding
7010: matrices. The geometrical meaning of this observation is the
7011: following: the construction of switching operators almost
7012: independent on over and under crossings when 3-valent graph is
7013: projected on a plane. So we can hope that in high genus closed
7014: case $\sM_{g}$ evaluation for spin networks and ones for RCFT
7015: coincide.
7016: \end{rmk}
7017:
7018: Now following Kohno \cite{K1} and using the fusion matrices $F$
7019: (6.143), switching operators $S$ (6.156) and diagonal operators
7020: $T_{e}$ (6.152) we construct the slightly projective
7021: representation (6.54).
7022:
7023:
7024: \begin{figure}[tbn]
7025: \centerline{\epsfxsize=3in\epsfbox{6c.eps}} \caption{\sl Graphs $\Theta_g$ and
7026: $\Ga_{g0}$} \label{Fig 14}
7027: \end{figure}
7028:
7029:
7030:
7031:
7032: Consider multi theta graph $\Theta_{g-1}$ of genus $g$ as in
7033: \cite{T3} (see Figure 18). On the Riemann surface
7034: $\Si_{\Theta_{g-1}}$ there exists the collection of simple circles
7035: \begin{equation}
7036: a_{1}, \dots, a_{g}, b_{1}, \dots, b_{g}, \de_{2}, \dots,
7037: \de_{g}, \ep_{2}, \dots, \ep_{g-1}
7038: \end{equation}
7039: as in Fig.12b in \cite{K1}. The collection of circles
7040: \begin{equation}
7041: a_{1}, \dots, a_{g}, \de_{2}, \dots, \de_{g}, \ep_{2}, \dots,
7042: \ep_{g-1}
7043: \end{equation}
7044: is the collection (4.7) of disjoint inequivalent circles. These
7045: curves correspond to edges of the graph $\Theta_{g-1}$. The
7046: identification of these circles with corresponding edges give us
7047: the collection of diagonal operators
7048: \begin{equation}
7049: T_{a_{1}}, \dots, T_{a_{g}}, T_{\de_{2}}, \dots, T_{\de_{g}},
7050: T_{\ep_{2}}, \dots, T_{\ep_{g-1}}
7051: \end{equation}
7052: (6.152) on the space $\sH^{k}_{\Theta_{g-1}}$ (4.39), (4.43).
7053:
7054: These operators represent the Dehn twists around the circles
7055: (6.162).
7056: But it isn't enough to generate full $Mod_{g}$. We have to add
7057: the Dehn twists around $\{ b_{i} \}$. To represent these twists
7058: we have to change the graph.
7059:
7060: Consider the graph $\Ga_{g0}$ (see Fig. 18) which is a chain of circles $C_{1},
7061: \dots, C_{g}$ joint by chain of consequent edges $ B_{1}, \dots,
7062: B_{g-1}$ such that $B_{i}$ joints $C_{i}$ and $C_{i+1}$. Then
7063: circles $C_{2}, \dots, C_{g-1}$ contain two edges
7064: \begin{equation}
7065: C_{i} = C'_{i} \cup C''_{i}.
7066: \end{equation}
7067: such that the set of edges
7068: \begin{equation}
7069: E(\Ga_{g0}) = C_{1}, C_{g}, C'_{2}, \dots, C'_{g-1}, C''_{2},
7070: \dots, C''_{g-1}, B_{1}, \dots, B_{g-1}.
7071: \end{equation}
7072:
7073: On
7074: the Riemann surface $\Si_{\Ga_{g0}}$ there are circles which
7075: correspond to edges $\{ B_{i} \}$. We cut this surface along
7076: these circles to obtain two 1-holed tori $\Si_{1,1}^{1}$ and
7077: $\Si_{1,1}^{g}$ and $g-2$ 2-holed tori $\{\Si_{1,2}^{i}\}, \quad
7078: i= 2,\dots, g-1$. The 1-holed torus $\Si_{1,1}^{1}$ conains the
7079: curve $b_{1}$ and $\Si_{1,1}^{g}$ contains $b_{g}$ from the
7080: collection (6.44). On the
7081: 2-holed torus $\Si_{1,2}^{i}$ we have the curves $ b_{i}, \quad
7082: i = 2,\dots, g-1$. The operator corresponding to the Dehn twist
7083: about $b_{i}$ acts on the correspoding block of the 1-holed and
7084: 2-holed tori (6.153). So we have the collection of representatives
7085: of
7086: these Dehn twists
7087: \begin{equation}
7088: T_{A_{1}} \circ S_{k}(j(B_{1})) \circ T_{A_{1}}, T_{A''_{2}} \circ
7089: S_{k}(j(B_{1}), j(B_{2})) \circ T_{A''_{2}}, \dots,
7090: \end{equation}
7091: $$
7092: T_{A''_{g-1}} \circ S_{k}(j(B_{g-2}), J(B_{g-1})) \circ
7093: T_{A''_{g-1}}, T_{A_{g}} \circ S_{k}(j(B_{g-1})) \circ T_{A_{g}}.
7094: $$
7095: (We identify the space $\sH^{k}_{\Theta_{g-1}}$ to
7096: $\sH^{k}_{\Ga_{g0}}$ with the basis (4.39) enumerated by spin
7097: networks $\{\Ga_{g0}\}$ of level $k$).
7098:
7099: Using the description of generators of $Mod_{g}$ and
7100: relations in \cite{Wa} we can check that all these operators give
7101: the slightly projective representation (5.38) (see \cite{K1}).
7102:
7103: So now we have two slightly projective representations (5.37) and
7104: (5.38) corresponding to two versions of CFT WZW and MS-axiomatic
7105: vertion. Fibers of the corresponding flat vector bundles are wave
7106: function spaces for two different kind of polarizations: complex
7107: and real. In paper \cite{K1} Kohno considered his version of
7108: CFT as the combinatorial description of the space of conformal
7109: blocks that is the space of holomorphic sections (3.127).
7110: But the direct identification can be constructed at the end of
7111: this text only.
7112:
7113:
7114: \section{Analytical aspect of the theory of
7115: non abelian theta functions}
7116:
7117: \subsection{The unitary Schottky space}
7118:
7119:
7120:
7121: This section is absolutely parallel to the subsection 2.4. The
7122: foundation for the following constructions is preprint
7123: \cite{FMNT}.
7124:
7125: The basis of the space $\sH^{k}_{\Ga}$ (6.3) is enumerated by
7126: the set of spin networks corresponding to the set of functions
7127: $\{ f_{\Ga_{j}}\}$ on the unitary Schottki space $uS_{g}$
7128: (6.110). On the other hand the unitary Schottky space is the fiber
7129: of the real polarization (4.13):
7130: \begin{equation}
7131: \pi_{\Ga}^{-1}(0, \dots , 0) = uS_{g}.
7132: \end{equation}
7133:
7134: The unitary Schottky space $uS_g = \SU(2)^{g} /
7135: Ad_{\mathrm{diag}} \SU(2)$ is a singular manifold. To avoid this
7136: problem we will consider geometrical objects on it as the
7137: corresponding objects on the direct product $\SU(2)^{g}$
7138: invariant with respect to
7139: $ Ad_{\mathrm{diag}} \SU(2)$-action.
7140: Our group $\SU(2)$ is embedded as a real part into its
7141: complexification
7142: $\SL(2, \C)$. The principal fact is that the theory of finite
7143: dimensional (non unitary) representations of $\SL(2, \C)$
7144: completely the same answers to all questions. Obviously
7145: \begin{equation}
7146: \widehat{SU(2)}= \widehat{\SL(2, \C)} = \half \Z^{+}
7147: \end{equation}
7148: and we can note both type of representations by the same symbols
7149: $j \in \half \Z^{+}$. Remark that all spaces $V_j$ of
7150: representations of both types of groups are precisely the same,
7151: $rk V_j = (2j+1)$. Moreover all intertwiners in tensor
7152: products of representations are the same. But of course $\SL(2,
7153: \C)$-action doesn't preseves the inner product.
7154:
7155: So here for convenience of the reader and also to establish
7156: notation we start by briefly recalling from \cite{Ha} the
7157: coherent state transform for Lie groups.
7158:
7159: \subsection{$g$-extended Hall's construction of CST for $\SU(2)$}
7160:
7161: Using the heat kernel $\rho_{t}$ for a Laplacian $\Delta$ on
7162: $\SU(2)$ associated with a given bi-invariant metric B. C. Hall
7163: constructed in \cite{Ha} the map
7164: $$
7165: C_{t} : L^{2}(\SU(2),dx)\mapsto {\cal H}(\SL(2, \C))
7166: $$
7167: \begin{equation}
7168: C_{t}(f)(z) = \int_{\SU(2)} f(x) \rho_t (x^{-1} z)dx, \qquad
7169: f\in L^{2}(\SU(2),dx), z\in \SL(2, \C)
7170: \end{equation}
7171: where $dx$ is the normalized Haar measure on $\SU(2)$ and ${\cal
7172: H}(\SL(2, \C))$ is the space of holomorphic functions on $\SL(2,
7173: \C)$. Recall $\rho_t$ has unique analytic continuation to
7174: $\SL(2, \C)$ also denoted by $\rho_t$. This map is called
7175: $\SU(2)$-averaged coherent state transform and for each $f\in
7176: L^{2}(\SU(2),dx)$, the function $C_{t}f(x)$ is just the solution
7177: of the heat equation,
7178: \begin{equation}
7179: \frac{\partial u}{\partial t} = \frac{1}{2}\Delta u,
7180: \end{equation}
7181: with initial condition given by $u(0,x)=f(x)$. Therefore,
7182: $C_{t}(f)$ is given by
7183: \begin{equation}
7184: C_{t}f(z) = ({\cal C}\circ \rho_t \star f)(z) = \left({\cal
7185: C}\circ e^{t\frac{\Delta}{2}}f \right)(z),
7186: \end{equation}
7187: where $\star$ denotes the convolution in $\SU(2)$ and ${\cal C}$
7188: denotes
7189: analytic continuation from $\SU(2)$ to $\SL(2, \C)$.
7190:
7191: Again for every $f\in L^{2}(\SU(2),dx)$ there exists the
7192: expansion given by the Peter-Weyl's theorem:
7193: \begin{equation}
7194: f(x) = \sum_{j \in \widehat{\SU(2)}} \mbox{tr} (j (x)B_{j ,f} )
7195: \end{equation}
7196: where the sum is taken over the set of equivalence classes of
7197: irreducible representations of $\SU(2)$ and $B_{j ,f}\in End \
7198: V_j$ is given by
7199: \begin{equation}
7200: B_{j ,f} = \left( \dim V_j \right) \int_{\SU(2)} f(x) j
7201: (x^{-1})dx,
7202: \end{equation}
7203: $V_j$ being the representation space for $j$. Then we obtain:
7204: \begin{equation}
7205: C_{t}f(z) = \sum_{j \in \widehat{\SU(2)}}
7206: e^{-t\frac{\lambda_{j}}{2}}\mbox{tr} (j (z)B_{j ,f } )
7207: \end{equation}
7208: where $\lambda_{j}$ is the eigenvalue of $-\Delta$ on functions
7209: of the type
7210: $$\mbox{tr} (B_j(x) ), \quad B\in End(V_j).
7211: $$
7212: Remark that in this formula $j \in \widehat{\SL(2, \C)}$ but $B_j
7213: $ is the same endomorphism of the same space as for $j \in
7214: \widehat{\SU(2)}$.
7215:
7216:
7217: It turns out that there exists a natural extension of $C_{t}$ to
7218: appropriately chosen distributions on $\SU(2)$ and we get
7219: holomorphic functions on $\SL(2, \C)$ corresponding to
7220: holomorphic sections of line bundles on it.
7221:
7222: This program can be extended to $\SU(2)^g \subset \SL(2, \C)^g$
7223: wordwisely. Moreover all constructions can be decented to
7224: quotients
7225: \begin{equation}
7226: uS_g = \SU(2)^{g} / Ad_{\mathrm{diag}} \SU(2), \quad S_g =\SL(2,
7227: \C)^g/ Ad_{\mathrm{diag}} \SL(2, \C).
7228: \end{equation}
7229:
7230: A natural $g$-extension of Hall's construction of CST can be done
7231: immediately but an extention of the previous construction to the
7232: complexification
7233: \begin{equation}
7234: uS_g = \SU(2)^{g} / Ad_{\mathrm{diag}} \SU(2) \subset S_g
7235: =\SL(2, \C)^g/ Ad_{\mathrm{diag}} \SL(2, \C)
7236: \end{equation}
7237: admits new parameters. We are starting by precise description of
7238: them.
7239:
7240: Let $\Omega = (\Omega_{\al \be})_{\al , \be = 1}^g$ be a complex
7241: symmetric matrix with positive imaginary part that is a point of
7242: Siegel upper half space $H$ and
7243: \begin{equation}
7244: \Omega = Re \Omega + i\cdot Im \Omega
7245: \end{equation}
7246: be the decomposition on real and imagine parts.
7247:
7248: Consider on $uS_g$ the invariant Laplacian given by
7249: \begin{equation}
7250: \Delta^{(im \Omega)} = \sum_{\al, \be = 1}^g \frac{im \Omega_{\al
7251: \be}}{2\pi i} \frac{\partial^2 }{\partial x_\al \partial x_\be}
7252: \end{equation}
7253: where $x_\al,x_\be $ are invariant vector fields on $\SU(2)$.
7254:
7255: Let $d\nu_t^{(im \Omega)}$ be the averaged heat kernel mesure on
7256: $uS_g$ corresponding to the heat equation
7257: \begin{equation}
7258: \frac{\partial u}{\partial t} = \frac{1}{ 4}\Delta_\C^{(im
7259: \Omega)} u = \sum_{\al , \be = 1}^n \frac{im \Omega_{\al \be}}{8
7260: \pi }\left( \frac{\partial^2}{\partial x_\al \partial x_\be } +
7261: \frac{\partial^2 }{
7262: \partial y_\al \partial y_\be }\right) u
7263: \end{equation}
7264: that is if $\mu_t^{(im \Omega)}$ denotes the fundamental solution
7265: of this heat equation then $d\nu_t^{(im \Omega)}$ is obtained by
7266: averaging $\mu_t^{(im \Omega)} dw$ with respect to the left
7267: action of $uS_g$ where $dw\equiv dxdy$ is the Haar measure on
7268: $S_g$. We get the transform
7269: $$
7270: C_t^{(im \Omega)} : L^2(uS_g,dx) \rightarrow \sH (S_g)\cap
7271: L^2((S_g, d\nu_t^{(im \Omega)})
7272: $$
7273: \begin{equation}
7274: C_t^{(im \Omega)}(f)(w) = \left( {\cal C} \circ e^{\frac{t}{
7275: 2}\Delta^{(im \Omega)}}f \right) (w) \quad , \ t > 0.
7276: \end{equation}
7277: Now consider non-self-adjoint Laplace operator on $uS_g$ of the
7278: form
7279: \begin{equation}
7280: \Delta^{(-i\Omega )} = - \sum_{\al, \be = 1}^g \frac{i }{ 2 \pi
7281: }\Omega_{{\al \be}} \frac{\partial^2}{ \partial x_\al \partial
7282: x_\be}
7283: \end{equation}
7284: It is easy to see that the imaginary part of this Laplacian does
7285: not affect the unitarity properties of the extended Hall's CST
7286: defined in this way. Therefore we have the following
7287: \begin{prop} For any $\Omega\in H_g$ and $t>0$ the transform
7288: $$
7289: C_t^{(-i\Omega)} ={\cal C} \circ e^{\frac{t }{ 2}
7290: \Delta^{-i(\Omega)}} \colon L^2(uS_g,dx) \rightarrow \sH (S_g)
7291: \cap L^2((S_g,(d\nu_t^{(im \Omega)})
7292: $$
7293: is unitary.
7294: \end{prop}
7295: Indeed we can decompose this transform as
7296: \begin{equation}
7297: C_t^{(-i\Omega)} = \left( {\cal C} \circ e^{\frac{t }{
7298: 2}\Delta^{(im \Omega)}} \right) \circ e^{\frac{t }{ 2}
7299: \Delta^{(-i re \Omega)}}.
7300: \end{equation}
7301: Then the unitarity of ${\cal C} \circ e^{\frac{t }{ 2}\Delta^{(im
7302: \Omega)}}$ is well known and that of the operator $e^{\frac{t }{
7303: 2} \Delta^{(-i re \Omega)}} = e^{-i \frac{t }{ 2}\Delta^{(re
7304: \Omega)}}$ (for any $t\in \R$) follows from the fact that
7305: $\Delta^{(re \Omega)}$ is a self-adjoint operator.
7306:
7307:
7308: Again every $f\in L^{2}(uS_g,d\vec x)$ we can consider as $
7309: Ad_{\mathrm{diag}} \SU(2)$-invariant function on $\SU(2)^{g}$
7310: and use the expansion given by the Peter-Weyl theorem:
7311: \begin{equation}
7312: f(x) = \sum_{\vec j \in \widehat{\SU(2)^g}} \mbox{tr} (\vec j
7313: (\vec x)B_{\vec j ,f} )
7314: \end{equation}
7315: where the sum is taken over the set of equivalence classes of
7316: irreducible representations of $\SU(2)^g$ and $B_{\vec j ,f}\in
7317: End \ V_{\vec j}$ is given by the usual formula. It has to be an
7318: intertwiner with respect to $ Ad_{\mathrm{diag}} \SU(2)$. Recall
7319: that every class $\vec j \in \widehat{\SU(2)^g}$ is a tensor
7320: product $\vec j = j_1 \otimes ... \otimes j_g$ where $j_i \in
7321: \widehat{\SU(2)}$ and the same is true for $\SL(2, \C)^g$. Its
7322: image under $C_t^{(-i\Omega)}$ is given by
7323: \begin{equation}
7324: C_{t}^{(-i\Omega)}f(\vec z) = \sum_{\vec j \in \widehat{\SU(2)^g}}
7325: e^{-t\frac{\lambda_{\vec j}^{(-i\Omega)}}{2}}\mbox{tr} (\vec j
7326: (z)B_{\vec j ,f } )
7327: \end{equation}
7328: where $\lambda_{\vec j}^{(-i\Omega)}$ is the eigenvalue of
7329: $-\Delta^{(-i\Omega)}$ on functions of the type
7330: $$\mbox{tr} (B_{\vec j} (\vec z)), \quad B\in End(V_{\vec j}).
7331: $$
7332: Again in this formula $j \in \widehat{\SL(2, \C)^g}$ and $B_{\vec
7333: j} $ is the same endomorphism of the same space as for $j \in
7334: \widehat{\SU(2)^g}$. These endomorphisms are intertwiners with
7335: respect to $ Ad_{\mathrm{diag}} \SL(2, \C)$-action thus
7336: $C_{t}^{(-i\Omega)}f(\vec z)$ is a function on $S_g$.
7337:
7338: This function is the analytic continuation to $\SL(2, \C)^g$ of
7339: the solution of the complex heat equation on $\SU(2)^g$
7340: \begin{equation}
7341: \frac{\partial u }{ \partial t} = \frac{1 }{ 2}
7342: \Delta^{(-i\Omega)} u
7343: \end{equation}
7344: with initial condition given by $f$.
7345:
7346: Now this coherent states transform can be extended from
7347: $L^{2}(uS_g,d\vec x)$ to the space of distributions
7348: $L^{2}(uS_g,d\vec x)'$ (recall that the unitary Schottky space
7349: $uS_g$ is compact). More precisely we have
7350: \begin{prop} For any $\Omega\in H_g$ and $t>0$ the transform
7351: \begin{equation}
7352: C_t^{(-i\Omega)} \colon (C^\infty(uS_g)' \to \sH ((\SL(2,
7353: \C))^g)
7354: \end{equation}
7355: $$
7356: f = \sum_{\vec j \in \widehat{\SU(2)^g}} \mbox{tr} (\vec j (\vec
7357: x)B_{\vec j ,f} ) \mapsto \sum_{\vec j \in \widehat{\SU(2)^g}}
7358: e^{-t\frac{\lambda_{\vec j}^{(-i\Omega)}}{2}}\mbox{tr} (\vec j
7359: (z)B_{\vec j ,f} )
7360: $$
7361: is well defined linear map.
7362: \end{prop}
7363:
7364: Indeed, the action of the Laplace operator and of its powers on
7365: the space of distributions is defined by duality from the
7366: corresponding action on $C^\infty(uS_g)$. For $f \in
7367: (C^\infty(uS_g)'$ with $f = \sum_{\vec j \in \widehat{\SU(2)^g}}
7368: \mbox{tr} (\vec j (\vec x)B_{\vec j ,f} )$ there exists positive
7369: a constant $C_1$ such that for all $\vec j \in
7370: \widehat{\SU(2)^g}$
7371: \begin{equation}
7372: \Vert B_{\vec j ,f} \Vert < C_1 \cdot e^{\Vert \vec j \Vert}
7373: \end{equation}
7374: where
7375: $$
7376: \Vert \vec (j_1, ... , j_g) \Vert = \sqrt{\sum_{l=1}^g (2j_l +
7377: 1)^2}.
7378: $$
7379: To show that the image of any distribution $f$ is an holomorphic
7380: function it is enouph to observe that there exists a positive
7381: constant $C_2$ such that for all $\vec j \in \widehat{\SU(2)^g}$
7382: \begin{equation}
7383: |e^{-t\frac{\lambda_{\vec j}^{(-i\Omega)}}{2}}|\cdot \Vert B_{\vec j ,f}
7384: \Vert < e^{C_2 \cdot (\Vert \vec j \Vert)^2}.
7385: \end{equation}
7386: for $\Vert \vec j \Vert$ sufficiently large. The existence
7387: follows from positivity of imaginary part of $\Om$.
7388:
7389: Our non-abelian theta functions on $S_g$ are images under the CST
7390: transform defined in previous Proposition of linear combinations
7391: of delta functions type distributions.
7392:
7393:
7394: Again let us consider the distribution
7395: \begin{equation}
7396: \theta^{uS_{g}}_{k} (x) = \sum_{\vec j \in \widehat{\SU(2)^g}}
7397: \mbox{tr} (\vec j (\vec x) )
7398: \end{equation}
7399: that is $B_{\vec j ,f} = \id$
7400:
7401:
7402: It is nothing else as delta-function at identity of level $k$.
7403: Let us apply CST (7.20) to this distribution:
7404: \begin{equation}
7405: C_{1/k}^{(-i \Om)} (\theta_k^{uS_{g}}) = \sum_{\vec j \in
7406: \widehat{\SU(2)^g}} e^{-t\frac{k \lambda_{\vec
7407: j}^{(-i\Omega)}}{2}}\mbox{tr} (\vec j (z) )
7408: \end{equation}
7409:
7410: For every spin network $\Ga_{j} \in SNNW^{k}_{\Ga}$ we can start
7411: with the function
7412: \begin{equation}
7413: \theta^{\Ga_{j}} (x) = f_{\Ga_{j}} \sum_{\vec j \in
7414: \widehat{\SU(2)^g}} \mbox{tr} (\vec j (\vec x) )
7415: \end{equation}
7416: to get the analytic function
7417: \begin{equation}
7418: C_{1/k}^{(-i \Om)} (\theta^{\Ga_{j}})
7419: \end{equation}
7420:
7421:
7422: \begin{rmk} Again we are specializing our continious
7423: positive parameter $t$ coming from
7424: heat kernels by its discret contrepart $1/k$.
7425: \end{rmk}
7426:
7427: This collection of holomorphic functions on the Schottki space
7428: \begin{equation}
7429: \{ C_{1/k}^{(-i \Om)} (\theta^{\Ga_{j}}) \}
7430: \end{equation}
7431: depending on a period matrix of {\it any} abelian variety is the
7432: combinatorial version of non abelian theta functions - theta
7433: functions with characteristics $\Ga_{j} \in SNW^{k}_{\Ga}$.
7434:
7435: \section{BPU-map}
7436:
7437:
7438: \subsection{Geometry of Lagrangian cycles}
7439:
7440:
7441: We can see that the ranks of wave spaces of the Kahler
7442: quantization and the Bohr-Sommerfeld quantizations are precisely
7443: the same that is the quantization is numerically perfect.
7444: We can identify projectivizations of both
7445: wave function spaces. But, is this identification natural?
7446: It will be the case if we can send every
7447: covariant constant section to a holomorphic section of the
7448: line bundle by some canonical way.
7449:
7450: The natural way was used firstly that is the coherent state
7451: transform or the Segal-Bargmann isomorphism introduced in the
7452: context of the quantum theory as a transform from the Hilbert
7453: space of of square integrable functions on the configuration
7454: space to the space of holomorphic functions on the phase space.
7455: In the abelian case (see subsection 2.4) considering the zero
7456: fiber $T^g_+$ of the polarization (2.22) we can interpret our
7457: complex torus $A$ as a complexification of this real torus and
7458: input our geometrical situation to the situation of the classical
7459: coherent state transform.
7460:
7461: In the very classical context a configurations space is
7462: just $\R^g$ as the real part of the phase space $\C^g$ but the
7463: construction was extended to the cases when
7464: \begin{enumerate}
7465: \item $\R^n$ is replaced by $T^n$ and $\C^n$ by any complex
7466: torus $A$ such that $T^n$ is a special Lagrangian subtorus (the
7467: case of subsection 2.3);
7468: \item $\R^n$ is replaced by $SU(2)$ and $\C^n$ by $SL(2, \C)$
7469: (Hall, \cite{Ha}) in the section 7;
7470: \item $\R^n$ is replaced by $SU(2)^{[g]} / diag Ad SU(2)$
7471: (= unitar
7472: Schottky space) and $\C^n$ by $SL(2, \C)^{[g]}/diag Ad SL(2, \C)$
7473: (= complex Schottky space) in the subsection 7.1;
7474: \item $\R^n$ is replaced by any Bohr-Sommerfeld Lagrangian cycle
7475: (more precisely a Legendrian cycle in the boundary of pseudo
7476: convex domain) and $\C^n$ by any compact Kahlerian symplectic
7477: manifold (Borthwick-Paul-Uribe map \cite{BPU}).
7478: \end{enumerate}
7479:
7480: The last case is the point of applications of the beautiful part
7481: of calculus (Fourier integral operators, Legendrian distributions
7482: and so on) and is a very strong tool to compare results of the
7483: Bohr-Sommerfeld and Kaehler quantizations.
7484:
7485: Actually the BPU-map is the extention of the classic WKB-method
7486: "acting in the opposite direction": WKB sends a wave function to
7487: a Lagrangian cycle with semi density in a phase space and BPU
7488: sends any Bohr-Sommerfeld cycle with a half form to a wave
7489: function (state) up to $U(1)$-scaling.
7490:
7491: To start comparing the complex and symplectic geometries we
7492: consider a polarized Kahler manifold $(M_{I}, \om, L, \nabla)$
7493: such that the curvature $F_{\na} = 2 \pi i \om$.
7494: We consider every Lagrangian cycle as a class of a
7495: differential map of a compact smooth $1/2 dim M_{I}$-manifold
7496: \[
7497: \phi \colon S \to M_{I} \quad \text{ such that } \quad \phi^* \om
7498: = 0
7499: \]
7500: non degenrated at some point (that is $ker d \phi_{p} = 0$ ) up to
7501: the natural action of the group $\Diff^+$ of orientation
7502: preserving diffeomorphisms of the manifold. Fondations and
7503: constructions of the theory of such cycles are contained in
7504: \cite{GT}.
7505: The main property of any
7506: such cycle is that {\em it can't be contained by any proper
7507: algebraic subvariety} (in particular, in a divisor). Thus any
7508: holomorphic object is uniquely determined by its restriction
7509: ( that is by
7510: the operation $\phi^*$ ) to a Lagrangian cycle $\phi \colon S \to
7511: X$. So restrictions to $S$ can serve as boundary conditions for
7512: holomorphic sections of line bundles with curvature proportional
7513: to $\om$.
7514:
7515:
7516: Geometrically, if we consider Lagrangian cycles as supports of boundary
7517: conditions for holomorphic objects, they have the minimal possible
7518: dimension. Usual boundary conditions deal with boundaries of
7519: complex domains of real codimension 1. Thus it is only for
7520: Riemann surfaces that Lagrangian boundary conditions coincide
7521: with the usual boundary conditions. In this case, in the modern
7522: theory of integrable systems the restriction of holomorphic
7523: objects to a small circle around a point reduces many analytical
7524: problems to algebraic geometry of curves (see for example the
7525: survey \cite{DKN}). It seems reasonable to expect that
7526: restrictions to Lagrangian submanifolds give a higher dimensional
7527: generalization of the modern version of the theory of integrable
7528: systems.
7529:
7530: There are no invariants of an embedding of a Lagrangian
7531: submanifold $S$ in a symplectic manifold. There are two ways of
7532: getting invariants:
7533: \begin{enumerate}
7534: \item consider families of Lagrangian manifolds admitting
7535: invariants
7536: (in particular limit singular subcycles); or
7537: \item to endow the submanifolds an additional structure (such as a
7538: section of
7539: some bundle or a Hermitian connection on the trivial line bundle).
7540: \end{enumerate}
7541: The restriction of the pair $(L, \nabla)$ to a Lagrangian cycle $S$
7542: gives this type of additional structure. It defines the space of
7543: covariant constant sections: $ H^0_\na ((L, \nabla)\rest{S}) $
7544: which can be nontrivial. Indeed, the restriction to any
7545: Lagrangian submanifold $S$ gives topologically trivial line
7546: bundle on $S$ with flat connection. A connection of this type is
7547: defined by its monodromy character $
7548: \chi\colon \pi_1(S)\to U(1),
7549: $
7550: and it admits a covariant constant section iff this character is
7551: trivial.
7552:
7553: We know that a Lagrangian cycle $S$ is a level $k$
7554: Bohr--Sommerfeld (BS$_k$) cycle if the character is trivial.
7555:
7556: Moreover, such a section defines a trivialization of the
7557: restriction , which identifies $C^\infty$ sections with complex
7558: valued functions on $S$:
7559: \[
7560: \Ga (L^{k}\rest{S})=C^{\infty}_\C(S).
7561: \]
7562: Thus the restriction to $S$ defines an embedding
7563: \begin{equation}
7564: \res\colon H^0(M_{I},L^{k}) \hookrightarrow C^{\infty}_\C(S).
7565: \end{equation}
7566: up to constant.
7567: \begin{dfn} The image
7568: \begin{equation}
7569: \res(H^0(M_{I},L^k))=\sH_S\subset C^\infty(S)
7570: \end{equation}
7571: is called {\em analog of the Hardy space} of level $k$.
7572: \end{dfn}
7573:
7574: Fixing a half-form $\hF$ on $S$; we call a pair $(S,\hF)$ a {\em
7575: half-weighted Lagrangian} cycle or a Lagrangian cycle marked with
7576: a half-form. Now we can identify the space of functions with the
7577: space of half-forms
7578: \begin{equation}
7579: \Ga(L^{k})\rest{S} \cdot\hF =\Ga(\De^{1/2}),
7580: \end{equation}
7581: where $\De$ is the complex volume bundle on $S$. This space is
7582: self adjoint with respect to the natural Hermitian form given by
7583: the integral of product of half-forms.
7584:
7585:
7586: \subsection{Legendrian distributions}
7587:
7588:
7589:
7590: Following Borthwick, Paul and Uribe \cite{BPU}, we can construct a
7591: distribution in some completion of $C^\infty(S)\cdot\hF$. Its
7592: restriction to the image of $\sH_I^k$ gives a covector or a
7593: state. BPU method uses usual codimension 1 boundary conditions
7594: rather than Lagrangian boundary conditions, and the original
7595: Hardy spaces for strictly pseudoconvex domains rather than analog
7596: of Hardy space. We refer the reader to the beautiful paper
7597: \cite{BPU} for the details because in the paper \cite{GT} using
7598: Rawnsley's results we related this construction closely to the
7599: standard constructions of Algebraic geometry. The original
7600: construction is the following:
7601: \begin{enumerate}
7602: \item Our
7603: Hermitian connection on $L^*$ defines a contact structure on the
7604: unit circle bundle $P$ of $L^*$.
7605: \item The disc bundle in $L^*$ is a strictly
7606: pseudoconvex domain, and there is the Szeg\"o orthogonal projector
7607: $\Pi\colon L^2(P)\to \sH$ to the Hardy space of boundary values of
7608: holomorphic functions on the disc bundle.
7609: \item The contact manifold $P$
7610: is a principal $U(1)$-bundle, and the natural $U(1)$-action on $P$
7611: commutes with $\Pi$ and gives a decomposition $\sH=\bigoplus_k
7612: H^0(M_{I},L^{k})$ of the Hardy space.
7613: \item If we fix a metaplectic
7614: structure on $P$, we can lift every $BS_{k}$ submanifold to a
7615: Legendrian submanifold $\La\subset P$ over it, marked with the lifted
7616: half-form $\hF$.
7617: \item $\La$ has an associated space of Legendrian
7618: distributions of order $m$, which is the Szeg\"o projection of
7619: space of
7620: conormal distributions to $\La$ of order $m + \half\dim M_{I}$ (see
7621: \cite{BPU}, 2.1).
7622: \item A half-form on $\La$ is identified with
7623: the symbol of a Legendrian distribution of order $m$ (see
7624: \cite{BPU}, 2.2); thus at the level of symbols, all Legendrian
7625: distributions look like delta
7626: functions or their derivatives.
7627: \item For a Legendrian submanifold $\La$
7628: with a half-form we fix the Legendrian distribution of order
7629: $\frac{1}{2}$ with symbol $\hF$ which is the Szeg\"o projection
7630: of the delta function
7631: $\de_\La$.
7632: \end{enumerate}
7633:
7634: So we have:
7635: \begin{enumerate}
7636: \item
7637: For every lifting $\La\subset P$ of a $BS_{k}$ submanifold $S$
7638: marked with a half-form $\hF$ we have a vector
7639: \begin{equation}
7640: BPU_{k}(\La,\hF)=\Pi^k_{\hF} (\de_\La)\in H^0(M_{I},L^{k}),
7641: \end{equation}
7642: where $\Pi^k_{\hF}$ is the Szeg\"o projection to the
7643: $k$th component of the Hardy space of the distribution with
7644: symbol $\hF$.
7645: \item Every such lifting is defined up to $U(1)$-action on $P$;
7646: thus a pair $(S,\hF)$ defines a point of the projectivization
7647: \begin{equation} BPU_{k} (S,\hF)=\PP (\Pi^k_{\hF} (\de_\La))\in \PP
7648: H^0(M_{I}, L^{k}).
7649: \end{equation}
7650: \end{enumerate}
7651: To apply this construction to our case $(M^{ss}_{\Si_{\Ga}}, \om,
7652: \Oh(k \Theta), A_{CS})$ (4.12) we have to observe that
7653: \begin{enumerate}
7654: \item This construction holds literally in the case
7655: when $\phi \colon S
7656: \to M_{I} $ is an immersion or $S$ has the structure of
7657: a smooth orbifold.
7658: \item By (3.92) and Propositions 12 and 13 there exists
7659: the canonical {\em geodesic} lifting of Bohr-Sommerfeld cycles
7660: to Planckian cycles.
7661: \end{enumerate}
7662:
7663: The following analog of Serre's Theorems~A and~B proved in
7664: \cite{BPU}, Section~3:
7665:
7666: \begin{thm} If\/ $n$ is large enough then for any halfweighted
7667: $BS_k$-cycle $(S,hF)$\/ $BPU_{nk}(\La,\hF)\ne0$. \end{thm}
7668:
7669: There are two or three canonical ways to endow any Lagrangian
7670: submanifold $S$ with a half-form:
7671: \begin{enumerate}
7672: \item If $X$ is a K\"ahler manifold with a metaplectic structure. Then
7673: this metaplectic structure defines a metalinear structure on $S$
7674: (see for example Guillemin \cite{Gu2}), and the K\"ahler metric
7675: $g$ defines a half-form $\hF_g$ on $S$. (This method is of course
7676: the most important for our applications.)
7677: \item The graph of a metasymplectomorphism with symplectic volume as
7678: square of the half-form.
7679: \end{enumerate}
7680:
7681: Every half weighted cycle $(S, hF)$ is {\it weighted } by the
7682: volume form $hF^2$ and defines the number
7683: \begin{equation}
7684: v(S, hF) = \int_S hF^2
7685: \end{equation}
7686: -the {\it volume } of this weighted cycle.
7687:
7688:
7689: We write
7690: \begin{equation} BS_{k}^v
7691: \end{equation}
7692: for the family of half-weight $BS$ cycles
7693: of a volme $v$. Then this space is an infinite dimensional {\it
7694: complex Kahler manifold } (see \cite{T5} and \cite{GT}).
7695:
7696: The BPU construction gives a
7697: ``rational'' map
7698: \begin{equation} \PP \fie_k\colon BS^v_{k} \to \PP
7699: H^0(M^{ss}_{I}, L^{k})^*
7700: \end{equation}
7701: the differential of which
7702: is dual to the restriction monomorphism (1.27)
7703: (see \cite{T7} and \cite{GT}).
7704:
7705: Recall that to send a half-weighted $BS_{k}$-cycle $(S, hF)$ to a
7706: section $BPU_{k}(S, hF)$ we have to lift it to a Planckian cycle.
7707: If the canonical class $K_{M_{I}}$ is proportional to
7708: $c_{1}(L^{k}) = k \cdot[\om]$, we can do this up to finite set of
7709: possibilities which we call a {\it fixing of theta-structure of
7710: level} $k$ (see Definition 16 below). More precisely for every level $k$ we have the tower
7711: of covers:
7712: \begin{equation}
7713: M_{k-theta} \to M_{SNW(k)} \to M_{\Ga_{g}} \to \sM_{g}
7714: \end{equation}
7715: where $\sM_{g}$ is the moduli space of curves of genus $g$ (5.31),
7716: $M_{\Ga_{g}}$ is the moduli space of marked Riemann surfaces (4.8),
7717: $M_{SNW(k)}$ is moduli space of marked curves $\Si_{\Ga}$ with
7718: fixed a spin network (4.42) and $M_{k-theta}$ is a curve with fixed
7719: theta-structure for the fixed half-weighted $k$-Bohr-Sommerfeld
7720: fiber of the projection $\pi_{\Ga}$ (4.13). Let
7721: \begin{equation}
7722: \phi_{k-theta} \colon M_{k-theta} \to \sM_{g}
7723: \end{equation}
7724: be the composition of all these covers. Then the preimage of our
7725: vector bundle $\sH^{k} \to \sM_{g}$ (1.29) on this cover
7726: \begin{equation}
7727: \PP \phi^{*}_{k-theta} \sH^{k} = \PP(\oplus_{w \in SNW_{k}(\Ga)} \Oh
7728: \cdot \theta_{w})
7729: \end{equation}
7730: where
7731: \begin{equation}
7732: \theta_{w} = BPU_{k-theta} (\widetilde{\pi_{\Ga}^{-1}(w)},
7733: hF_{can}).
7734: \end{equation}
7735: Here $\widetilde{\pi_{\Ga}^{-1}(w)}$ is the corresponding to the
7736: theta-structure Planckian cycle (8.35).
7737:
7738: The coincidence of the Hitchin, WZW and Kohno connections implies
7739:
7740: \begin{thm} The lifting of the Hitchin connection to the cover
7741: $\phi_{k-theta}$ (8.10) is
7742: Hermitian.
7743: \end{thm}
7744: Obviously this Hermitian structure isn't invariant with respect to this cover.
7745:
7746: To use the BPU construction described in (8.4), (8.5) of this
7747: section we must repeat some of the details. We stay in
7748: the situation of a complex polarization of a prequantized phase
7749: space $(M_{I}, \om, L, \nabla)$ where $(L, \nabla)$ is
7750: $\Oh(\Theta)$.
7751:
7752: Consider the principal $U(1)$-bundle $P$ of the dual line bundle
7753: $L^*$, the unit circle bundle in $L^*$. Let $D\subset L^*, \quad
7754: \partial D=P $
7755: be the unit disc subfibration with boundary.
7756:
7757: Our complex polarization $I$, that is a compatible complex
7758: structure $I$ on $M$, defines a complex structure $D_I$ on $D$ as
7759: a strictly pseudoconvex domain.(Recall that $L$ is a positive line
7760: bundle).
7761:
7762: The Hermitian
7763: connection on $L^*$ iduced by $\nabla$ is given by 1-form $\al$
7764: on $P$ which defines a
7765: contact structure on $P$ with volume form \[ \frac{1}{2\pi} \al \wedge
7766: d\al^n, \quad \text{where} \quad n=\dim_{\C} M_{I}. \] The null space of
7767: $\al$ at a point $p\in P$ is the maximal complex subspace of the tangent
7768: space for any complex structures.
7769:
7770: For every our $I$ the Hardy subspace
7771: \begin{equation}
7772: \sH_I\subset L^2(P)
7773: \end{equation}
7774: consists of boundary values of holomorphic functions on $D_I$. We
7775: have the Szeg\"o orthogonal projector
7776: \begin{equation}
7777: \Pi_I\colon L^2(P)\to \sH_I.
7778: \end{equation}
7779: The natural action of $U(1)= Aut P$ on $P$ as a principal bundle
7780: commutes with $\Pi_I$ and decomposes the space $\sH_I$ as a
7781: Hilbert direct sum of isotypes:
7782: \begin{equation}
7783: \sH_I=\bigoplus_{k=0}^{\infty} \sH_I^k ; \quad \sH_I^k
7784: =H^0(M_{I}, L^{k})),
7785: \end{equation}
7786: with only positive characters. Thus
7787: \begin{equation}
7788: \Pi_I=\bigoplus_{k=0}^{\infty} \Pi_k.
7789: \end{equation}
7790: Remark that in our situation the vector bundle $\sH^k \to \sM_{g}$
7791: (1.29) is embedded
7792: to the trivial
7793: bundle $L^2(P) \times \sM_{g}$ with the space (8.1) as a fibre
7794: and there
7795: exists the Szego projection
7796: \[
7797: \Pi \colon L^2(P) \times \sM \to \sH = \oplus_{k=0}^{\infty} \sH^k,
7798: \]
7799: which is (8.2) fiberwise.
7800:
7801: Recall that the vector bundles $\sH^k \to \sM_g$ admit
7802: projective flat connections. So locally the space $\PP\sH^k_I$ doesn't
7803: depend on complex structure $I$. Then there exists a
7804: central extension $\widetilde{Mod_{g}}$:
7805: \begin{equation} 1\to U(1)\to\widetilde{Mod_{g}}\to Mod_{g} \to 1,
7806: \end{equation}
7807: where the centre $U(1)= Aut P$ acts on $P$ as the proper group of
7808: a principal bundle. This action induces a natural representation
7809: \begin{equation}
7810: \rho\colon \widetilde{Mod_{g}} \to Op(L^2(P))
7811: \end{equation}
7812: to the operator algebra of the space of functions.
7813:
7814: For every $g \in \widetilde{Mod_{g}} $ the projectors $\Pi_I$
7815: (8.4) define the linear transformation
7816: \begin{equation}
7817: p_{I, g(I)} \circ \Pi_{g(I)} \circ \rho(g) \circ \Pi_I \in
7818: Op(\sH_I)
7819: \end{equation}
7820: which commutes
7821: with the action of $U(1)$. So this transformation is decomposed
7822: as a Hilbert direct sum of isotypes
7823: \begin{equation}
7824: Op(\sH_I) = \oplus_k End H^0(L^k).
7825: \end{equation}
7826: These transformations can be projectivized to the representation
7827: \[
7828: \PP\rho_I^k\colon Mod_{g} \to Aut\PP H^0(L^k).
7829: \]
7830: Recall that our $M_{I}$ and the principal bundle $P$ have given
7831: metaplectic structures.
7832:
7833: Now if $(S, hF)$ is a half-weight $BS$-orbifold, the complex
7834: conjugate of a covariant constant section gives a lift of it to a
7835: half weighted Legendrian cycle $(\La,\hF)$, and the BPU --
7836: construction (8.4) defines a section $ BPU(\La,\hF)\in
7837: H^0(M_{I}, L) $.
7838: In the same vein, we get lifting $\La_{k}$ of $S$ from $BS_{k}$
7839: to the Legendrian cycle on $P$ which is a cyclic $(k)$-cover of
7840: it and
7841: the system of sections
7842: \begin{equation}
7843: BPU_{k}(\La_{k},\hF)\in
7844: H^0(X, L^{k}).
7845: \end{equation}
7846:
7847: The Lagrangian cycle $S$ can be reconstructed from the set of
7848: its BPU
7849: images as the quasi-classic limit as $1/k=\text{Planck's
7850: constant}\to0$: the wave fronts of distributions concentrate on
7851: $S \subset M_{I}$ (see \cite{BPU} and the references given there).
7852:
7853: Now for two Lagrangian cycles $(S_1,\hF_1)$ and $(S_2,\hF_2)$, the
7854: asymptotic behaviour of the scalar product
7855: $\Span{BPU_k(S_1,\hF_1), BPU_k(S_2,\hF_2)}$ (see (1.9)) can be
7856: computed in terms of the intersection $S_1\cap S_2$ (see
7857: \cite{BPU}). In particular,
7858: \begin{equation}
7859: S_1\cap S_2=\emptyset \implies BPU_k(S_1,\hF_1) \perp
7860: BPU_k(S_2,\hF_2)
7861: \end{equation}
7862: asymptotically as $k\to\infty$. (For the orbifold case these
7863: asymptotics are somewhat weaker, but are still quite expressive
7864: for geometric corollaries). This asymptotic technique comes from
7865: the physical interpretation of this set-up as the ``classical''
7866: geometric quantization. More precisely the asymptotic analysis of
7867: quantum states gives
7868: \begin{multline}
7869: \Span{BPU_k(S,\hF), BPU_k(S,\hF)} \ \sim\
7870: \left(\frac{k}{\pi}\right)^{\half dim M_{I}}\int_{S} \vert\hF \vert^2 \\
7871: +O(k^{\half(dim M_{I}-1)}),
7872: \end{multline}
7873: and if $S_1\cap S_2=\emptyset$ then
7874: \begin{multline}
7875: \Span{BPU_k(S_1,\hF_1), BPU_k(S_2,\hF_2)} \ \sim\
7876: \left(\frac{k}{\pi}\right)^{\half(dim M_{I}-1)}+ \\
7877: O(k^{\half(dim M_{I}-2)}).
7878: \end{multline}
7879: From this and the description of the differential of BPU-map we get
7880: \begin{prop} Let $\pi\colon M_{I}\to B$ be any real
7881: polarization with transversal set of
7882: BS$_{k}$ - fibers. Then we can fix such half forms on these
7883: fibers then for $k\gg0$, the BPU vectors in $H^0(M_{I},L^{k})$ are
7884: linear independent:
7885: \[ rk
7886: \Span{BPU_{k}(BS_{k} \cap B)}=\#(BS_k \cap B). \]
7887: \end{prop}
7888:
7889: \begin{rmk} A more sophisticated analysis of the asymptotics of quantum
7890: states extends this observation as follows: let
7891: \[
7892: S_1,\dots, S_{N_{\max}}\subset BS_{k}
7893: \]
7894: be a maximal collection of {\em disjoint\/} $BS_{k}$ cycles. Then
7895: \[
7896: N_{\max} \le rk H^0(L^{k}),
7897: \]
7898: and the right-hand side is given by the Riemann--Roch theorem.
7899: \end{rmk}
7900:
7901:
7902: \subsection{Geodesic lifting}
7903:
7904:
7905:
7906:
7907: The lifting of a BS-Lagrangian cycle in $M_{I}$ to a Legendrian
7908: cycle on $P$ we have described is defined up to the natural
7909: $U(1)$-action on $P$ and the states $BPU_k(S,\hF)$ are defined
7910: up to a phase. But in our situation we can do this almost
7911: canonically (up to a finite ambiguity called a choice of a
7912: theta-structure) and get an actual basis of $H^0(M_{I}, L^{k})$.
7913:
7914: To describe this almost canonical lifting we must consider the
7915: Lagrangian Grassmannization of the tangent bundle of $M_{I}$ as
7916: described in \cite{T3}. Pointwise, the tangent space $(TM_{I})_x$
7917: at a point $x\to M_{I}$ is $\C^n$ with constant symplectic form
7918: $\Span{\ \,,\ }=\om_x$ and constant Euclidean metric $g_x$,
7919: giving the Hermitian triple $(\om_x,I_x,g_x)$. Define the
7920: Lagrangian Grassmannian $(LaGr)_x =LaGr(TM_{I})_x$ to be the
7921: Grassmannian of oriented Lagrangian subspaces in $(TM_{I})_x$.
7922: Taking this space over every point of $M_{I}$ gives the oriented
7923: Lagrangian Grassmannization of $TM_{I}$
7924: \begin{equation}
7925: \pi\colon LaGr(TM_{I})\to M_{I}
7926: \quad\text{with} \quad \pi^{-1} (x) =(LaGr)_x.
7927: \end{equation}
7928: A complex structure $I_x$ on $(TM_{I})_x$ gives the standard
7929: identification
7930: \begin{equation}
7931: (LaGr)_x=U(n) / SO(n).
7932: \end{equation}
7933: This space admits a canonical map
7934: \begin{equation}
7935: det\colon(LaGr)_x\to U(1)=S^1_x
7936: \quad\text{sending $u \in U(n)$ to $det u \in U(1)=S^1$.}
7937: \end{equation}
7938: Taking this map over every point
7939: of $M_{I}$ gives the map
7940: \begin{equation}
7941: det\colon LaGr(TM_{I})\to S^1(L_{-K}),
7942: \end{equation}
7943: where $S^1(L_{-K})$ is the unit circle bundle of the line bundle
7944: $\bigwedge^{n} TM_{I} =det TM_{I}$, with the first Chern class
7945: \[
7946: c_1(det TM_{I})=-K_{M_{I}}
7947: \]
7948: the canonical class of $M_{I}$ (see for example \cite{T2} and
7949: \cite{T3}).
7950:
7951: We have already noted that our Lagrangian cycles does not usually
7952: have an orientation defined a priori. Thus we must consider the
7953: Lagrangian Grassmannian $\La(TM_{I})$ forgetting orientations.
7954: Then we get a map
7955: \[
7956: det\colon \La(TM_{I})\to S^1(L_{-K/2})
7957: \]
7958: in place of (8.28).
7959:
7960: Now for every oriented Lagrangian cycle $S\subset X $, we have the
7961: Gaussian lifting of the embedding $i\colon S\to X$ to a section
7962: \begin{equation}
7963: G(i)\colon S\to La(TM_{I})\vert_{S},
7964: \end{equation}
7965: sending $x\in S$ to the subspace $TS_x\subset(TM_{I})_x$. The composite
7966: of this Gauss map with the projection $det$ gives the map
7967: \begin{equation}
7968: {det}\circ G(i)\colon S\to S^1(L_{-K/2})\vert_{S}.
7969: \end{equation}
7970: Thus every Lagrangian cycle $S$ defines a Legendrian subcycle
7971: \begin{equation}
7972: \La={det}\circ G(i) (S)\subset S^1(L_{-K/2}).
7973: \end{equation}
7974: The Levi-Civita connection of the K\"ahler metric defines a
7975: Hermitian connection $A_{LC}$ on $L_{-K/2}$.
7976:
7977: \begin{dfn} A Lagrangian cycle $S$ is almost geodesic if the
7978: Legendrian cycle $\La={det}\circ G(i)(S)$ is horizontal with
7979: respect to the Levi-Civita connection $A_{LC}$ on $L_{-K/2}$.
7980: \end{dfn}
7981:
7982: We use Proposition 13 now. The line bundle $L_{-K/2}$ is $\Oh(2
7983: \Theta )$ and the Levi-Civita connection is induced by the
7984: connection $\nabla$ on $L = \Oh(\Theta)$. Then we have
7985: \begin{prop} A Lagrangian cycle $S$ is $BS$ if and only if it is
7986: almost geodesic.
7987: \end{prop}
7988:
7989: The Hermitian structures of our line bundles define a map
7990: \begin{equation}
7991: \mu_2\colon S^1(L^*)\to S^1(L_{-K/2})
7992: \end{equation}
7993: of the principal $U(1)$-bundles of these line bundles, which
7994: fibrewise is minus isogeny of degree $2$. Thus {\em every}
7995: Lagrangian cycle $S$ defines an oriented Legendrian subcycle (see
7996: (8.31))
7997: \begin{equation}
7998: \La=\mu_2^{-1} ({det}\circ
7999: G(i) (S))\subset S^1(L^*)=P
8000: \end{equation}
8001: (not almost geodesic {\it a priori}).
8002: Now consider the pair of isogenies
8003: \begin{equation}
8004: \mu_{k}\colon S^1(L^2)\to S^1(L^{2 k})
8005: \end{equation}
8006: \[
8007: \mu_2\colon S^1(L^{k})\to S^1(L^{2 k})
8008: \]
8009: and the lifting $ l\colon S\to S^1(L^{k}) $ given by a covariant
8010: constant section over a $BS_{k}$ cycle $S$.
8011: \begin{dfn} The lift $l$ is almost geodesic if
8012: \[
8013: \mu_{k} \circ {det}\circ G(i) (S)=\mu_2 \circ l (S).
8014: \]
8015: \end{dfn}
8016:
8017: The number of geodesic lifts is obviously $\leq 2(k+d)$.
8018:
8019: In summary, let $\La_k$ be the space of Legendrian subcycles of
8020: $P$ the images of whose projection to $M_{I}$ is the $k$th root of
8021: unity cover of a $BS_{k}$ Lagrangian cycle on $M_{I}$ (that is
8022: Planckian cycles). Then the natural projection
8023: $ p\colon \La_{k}\to BS_{k} $
8024: which sends Legendrian cycle to Lagrangian cycle is a principal
8025: $U(1)$-bundle.
8026:
8027: The geodesic lifting
8028: \begin{equation}
8029: l\colon \widetilde{BS_{k}} \to \La_{k}
8030: \end{equation}
8031: we have described is a multisection of this principal bundle and
8032: $ p\colon \widetilde{BS_{k}}\to BS_{k} $
8033: is a finite cyclic cover.
8034:
8035: Consider a real polarization $\pi\colon M_{I}\to \De$ (1.13). Then
8036: we have a finite set of Bohr--Sommerfeld fibres
8037: \[
8038: \De \cap BS_{k}=\{S_i\}, \quad \text{for $i=1,\dots, \vert SNW_{k}$.}
8039: \]
8040:
8041: \begin{dfn} A choice of geodesic lifts
8042: \[
8043: \{\tilde{S_i}\}\subset \La_{k}
8044: \]
8045: is called a choice of theta structure of the real polarization
8046: $\pi_{\Ga}$.
8047: \end{dfn}
8048:
8049: Marking these Lagrangian cycles with the half-forms we get a
8050: finite set
8051: \[
8052: \{\tilde{S_i},\hF_i\}
8053: \]
8054: of half-weight Legendrian cycles.
8055:
8056: We know that
8057: \[
8058: BPU_k (\{\tilde{S_i},\hF_i\})\subset \PP H^0(M_{I},L^{k})
8059: \]
8060: is a linear independent system of vectors (states) if $k\gg0$ and
8061: an accurate choice of half forms. In particular, if
8062: \begin{equation}
8063: \#(BS_{k} \cap B)=rk H^0(M_{I},L^{k}),
8064: \end{equation}
8065: we
8066: get a Bohr--Sommerfeld basis.
8067:
8068: \begin{rmk} For other descriptions and
8069: applications of the geodesic lift from Lagrangian to Legendrian
8070: cycles see \cite{T1}, \cite{T2} and \cite{T3}.
8071: \end{rmk}
8072:
8073: Now on every $BS$-fiber we can fix the canonical half-form. For
8074: this return to the description of $BS$-fibers (4.36) - (4.38).
8075: Remark that groups (4.30) admit bi-invariant half-forms
8076: $\hF_1$ on $U(1)$ and $\hF_2$ on $\SU(2)$. For every $w$ we can
8077: normalize these form $\hF_1(w)$ and $\hF_2(w)$ so that the
8078: half-form
8079: \begin{equation}
8080: \hF_w=(\hF_1(w))^{t-s} \cdot (\hF_2(w))^{p+s}
8081: \end{equation}
8082: is homogeneous of degree 1 on $\pi^{-1}(w)$ (see (4.37)) with
8083: respect to scaling $\hF_i(w)\to t\cdot\hF_i(w)$. We say that such
8084: half-form is homogeneously normalized.
8085:
8086: It's easy to see (\cite{JW2}, 4.7) that a normalized half-form
8087: for a nonsingular $BS_{k}$ fibre is given by a Hamiltonian vector
8088: field with Hamiltonian in $\R^{3g-3}$ of volume 1.
8089:
8090: Thus every $BS_{k}$ fibre is endowed with the covariant constant
8091: half-form (8.37), and we can proceed to construct the
8092: corresponding Legendrian distributions in $P$. Recall that
8093: $M_{I}$ is almost homogeneous with respect to the Goldman torus
8094: action (4.5). Thus the Schwartz kernel of coherent states does
8095: not depend on points and, outside singular points of fibres, they
8096: behave as in the homogeneous case (see \cite{BPU}, (10--13)). By
8097: lifting to $P$ every $BS_{k}$ fibre $\pi^{-1}(w)$ defines
8098: Legendrian subcycle $\La_w\subset P$ marked with the half-form
8099: \begin{equation}
8100: \bar{hF}_w=(\fie_4\circ {det}\circ G(i))^* hF_w,
8101: \end{equation}
8102: and having monodromy a $k$th root of 1.
8103:
8104: To apply the BPU construction the principal bundle $P=S^1(L^*)$
8105: must be given a metaplectic structure. This can be done at once
8106: using the equality $K_{M_{I}} = - 4 \Theta$.
8107:
8108:
8109: Our long trip is finishing at the final statement:
8110:
8111: {\it the collection of functions}
8112: \begin{equation}
8113: \{ \theta^{\Ga_j} \}, \quad \Ga_j \in SNW^k_\Ga
8114: \end{equation}
8115: (7.25) {\it coincides with the collection of sections}
8116: \begin{equation}
8117: \{\theta_w = BPU_{k-theta} (\widetilde{\pi^{}-1}_\Ga(w), \bar{hF}_{w}) \}
8118: \end{equation}
8119: (see (8.12) and (8.38)) {\it under the meromorphic map} (3.44).
8120:
8121:
8122: Now one can prove this up to phases (6.18). May be it is productive to input
8123: these phase ambiguities into the non-abelian theta-structure notion?
8124:
8125: \section{The main weapon}
8126:
8127: \subsection{Large limit curves}
8128:
8129: To see how all the constructions could be exploited in the main game we have to show behavior on the moduli space "pointwise" way. For this we are using the following strategy:
8130: \begin{enumerate}
8131: \item we choose special complex curves which are convinient for applications
8132: of constructions as algebro-geometrical so analytical constructions and
8133: \item extend obtained structures to general curves using the projective connections.
8134: \end{enumerate}
8135: Recall that the analytic geometry formulation of two-dimensional conformal field theory
8136: (see \cite{FS}) predicts the existence of an extension of all geometrical objects like vector
8137: bundles (1.29) to the compactification $\ov{\sM_g}$ of the moduli space of curves. So our "special" curves may be non smooth. Moreover our curves are
8138: "maximally" non smooth and reducible. They look like
8139: maximally unipotent boundary points of moduli spaces of
8140: Calabi-Yau threefolds.
8141: Now we define such type points for moduli spaces of algebraic
8142: curves and investigate the geometry of their deformations. This
8143: geometry reflects
8144: mathematical relations between (1,1)-conformal field theories
8145: and string theories observed by physisists in a number of
8146: papers. Recall that maximally unipotent boundary points of
8147: CY-moduli spaces correspond to large radius limit point according to
8148: mirror symmetry. We have not time to discuss all parallel to mirror symmetry
8149: for algebraic curves here. However we call our curves {\it large limit
8150: curves} (ll-curves for short).
8151:
8152: Let $\Ga$ be any 3-valent graph of genus $g$.
8153: Topologically it is equivalent to
8154: 3-dimensional handlebody $H_{\Ga}$ with boundary
8155: $\p H_{\Ga} = \Si_{\Ga}$ (6.9)
8156: where $\Si_{\Ga}$ is the Riemann surface given by the pumping up
8157: trick (see Fig. 9) : we pump every edge of $\Ga$ to a tube (a {\it channel} in
8158: terms of CQFT) and every vertex to a trinion (= 2-sphere with 3
8159: holes). By the construction our Riemann surface $\Si_{\Ga}$ has a
8160: trinions (or "pair of pants") decomposition given by removing all
8161: tubes.
8162:
8163:
8164: Of course we get a surface without any complex structure.
8165: Thus it is not an algebraic curve, but we can certainly define uniquely a
8166: reducible algebraic curve $P_{\Ga}$. Namely, send every vertex $v \in
8167: V(\Ga)$ to the complex Riemann sphere $P_{v} = \C\PP^{1}$ with a
8168: triple points $p_{e_1}, p_{e_2}, p_{e_3}$ corresponding to edges of
8169: the star $S(v)$. Now for every edge $e$ with vertices $\p e = v, v'$
8170: we identify points $ p_{e}$ of components $P_{v}$ and $P_{v'}$.
8171: As the result we obtain a reducible algebraic curve with
8172: properties:
8173:
8174: \begin{enumerate}
8175: \item arithmetical genus of the connected reducible curve
8176: $P_{\Ga_{g}}$ is equal to $g$;
8177: \item for any 3-valent graph $\Ga_{g}$ the curve $P_{\Ga_{g}}$ is
8178: Deligne - Mumford stable and called the {\it large limit
8179: curve};
8180: \item the curve $P_{\Ga_{g}}$ defines a point of the Deligne-Mumford
8181: compactification $\ov{\sM_{g}}$ (with the same notation);
8182: \item this
8183: point is contained by the boundary divisor $D = \ov{\sM_{g}} -
8184: \sM_{g}$;
8185: \item moreover,
8186: \begin{equation}
8187: \{ P_{\Ga} \} \in D_{0} \cap ( \cap_{i=1}^{g-1} D_{i})
8188: \end{equation}
8189: where general point of divisor $D_{0}$ corresponds to an
8190: irreducible curve with one node and general point of $D_{i}$ is a
8191: bouquet of smooth curves of genus $i$ and $g-i$ and
8192: \begin{equation}
8193: D = D_{0} \cup ( \cup_{i=1}^{g-1} D_{i})
8194: \end{equation}
8195: \end{enumerate}
8196:
8197: So in the Deligne-Mumford compactification $\ov{\sM_{g}}$ of the
8198: moduli space $\sM_{g}$ of smooth curves of genus $g$ we get the finite
8199: configuration of points
8200: \begin{equation}
8201: \sP \subset \ov{\sM_{g}}
8202: \end{equation}
8203: enumerated by the set of 3-valent
8204: graphs $TG_{g}$ - the set of large limit curves now.
8205:
8206: \begin{rmk}
8207: These curves make a problem of the necessary part of two-dimensional conformal
8208: field theory formulation as analytic geometry (see \cite{FS}).
8209: Recall that the correlation functions of all local quantum
8210: fields can be recovered from the partition function when all
8211: channels (tubes) of the surface are constricted down to nodes.
8212: For the surface $\Si_{\Ga}$ the constriction of all 3g-3 tubes
8213: produces our reducuble curve $P_{\Ga}$ with uniquely defined
8214: complex structure. On the other hand from the point of view of dynamic
8215: triangulation theory these curves are the
8216: polymer
8217: phases of asymptotic triangulations.
8218: \end{rmk}
8219:
8220: \subsection{Canonical classes and canonical maps}
8221:
8222: For every curve $P_{\Ga}$ the canonical class $K_{\Ga}$ is a line
8223: bundle: a restriction of it to every component $P_{v}, v \in
8224: V(\Ga)$ is the sheaf of meromorphic differentials $\om$ with
8225: simple poles at $p_{e_1}, p_{e_2}, p_{e_3}$ where $e_1 \bigcup e_2 \bigcup e_3 =
8226: S(v)$. Thus
8227: \begin{equation}
8228: K_{\Ga} \vert_{P_{v}} = K_{P_{v}}(p_{e_1} + p_{e_2} + p_{e_3} ) =
8229: \Oh_{P_{\Ga}}(1).
8230: \end{equation}
8231: and in the Neron-Severi lattice $NS(P_\Ga) = \prod_{v \in V(\Ga) Pic(P_v)}$
8232: one has
8233: \begin{equation}
8234: c (K_{\Ga}) = (1,1,...,1) \in NS(P_\Ga).
8235: \end{equation}
8236: Every holomorphic section $s$ of $K_{\Ga}$ is a collection of
8237: meromorphic differentials $\{ \om_{v}\}$ on the components $P_{v}$
8238: with poles at $\bigcup_{e \in E()\Ga} p_{e}$ with constraints: for every $e$ such that $\p e = v, v'$
8239: \begin{equation}
8240: res_{p_{e}}\om_{v} + res_{p_{v'}} \om_{v'} = 0.
8241: \end{equation}
8242: On the other hand we have 2g-2 linear relations: for every $v \in
8243: V(\Ga) $ with $(e_1,e_2, e_3 )= S(v)$
8244: \begin{equation}
8245: res_{p_{e_1}} \om_{v} + res_{p_{e_2}} \om_{v} + res_{p_{e_3}}
8246: \om_{v} = 0
8247: \end{equation}
8248: The {\it thickness} $th(\Ga)$ of the graph is the minimal number of edges that may be removed to make the graph disconnected. In \cite{Ar} Artamkin proved the following
8249:
8250: \begin{prop} The canonical linear system $\vert K_{P_\Ga} \vert$ is
8251: \begin{enumerate}
8252: \item base points free iff $th(\Ga) \geq 2$;
8253: \item very ample iff $th(\Ga) \geq 3$.
8254: \end{enumerate}
8255: \end{prop}
8256:
8257: For simplicity below we are working with 3-valent graphs of thickness 3 only.
8258: Then we have the canonical embedding
8259: \begin{equation}
8260: \phi_\Ga
8261: \colon P_\Ga \to \PP^{g-1}
8262: \end{equation}
8263: and it is easy to check
8264:
8265: \begin{prop} In $\PP^{g-1}$ dimension of complete linear system
8266: \begin{equation}
8267: \vert 2h - \phi_{K_{\Ga}}(P_{\Ga}) \vert
8268: \end{equation}
8269: of quadrics through the canonical curve is equal to
8270: \begin{equation}
8271: dim \vert 2h \vert - 3g + 3.
8272: \end{equation}
8273: \end{prop}
8274: The proof is quite classical: we just repeat here the classical argument of Castelnuovo.
8275:
8276: \begin{cor}
8277: For 3-valent graph $\Ga_{g}$ the large limit curve $P_{\Ga}
8278: \in \ov{\sM_{g}}$ is a smooth point of $\ov{\sM_{g}}$ as orbifold
8279: and the fiber of the cotangent bundle at this point is given by the following
8280: equality
8281: \begin{equation}
8282: T^{*}_{P_{\Ga}} \ov{\sM_{g}} = H^{0}(P_{\Ga}, \Oh_{P_\Ga} (2K_{P_\Ga}).
8283: \end{equation}
8284: \end{cor}
8285: Now consider the double canonical map of $P_\Ga$ given by the
8286: complete linear system $\vert 2K_{\Ga}\vert$
8287: \begin{equation}
8288: \phi_{2K_{\Ga}} \colon P_{\Ga} \to \PP^{3g-4}.
8289: \end{equation}
8290: Remark that the target projective space has the interpretation
8291: \begin{equation}
8292: \PP^{3g-4} = \vert 2K_{\Ga}\vert^{*} = \PP
8293: T_{P_{\Ga}} \ov{\sM_{g}}
8294: \end{equation}
8295: is the projectivization of the tangent space to the $\ov{\sM_{g}}$
8296: at $P_{\Ga}$.
8297:
8298: As any double canonical curve our configuration $\phi_{2K_{\Ga}} (P_\Ga)$ of conics and
8299: double points is determined up to a projective automorphism of
8300: $\PP^{3g-4} = \PP T_{P_{\Ga}} \ov{\sM_{g}} $ only. But in the next
8301: subsection we show that {\it the double canonical
8302: model of a large limit curve is determined absolutely canonically
8303: without the projective transformation ambiguity} (unlikely of
8304: canonical model of general algebraic curve).
8305:
8306: Indeed, the images of nodes
8307: \begin{equation}
8308: \{ \phi_{2K_{\Ga}}(\bigcup_{e \in E(\Ga)} p_{e}) \}
8309: \end{equation}
8310: define the configuration of
8311: $$3g-3 = rk H^0(P_\Ga, \Oh(2K_{P_\Ga}))$$
8312: linear independent points ( since components are coming to irreducible conics
8313: and every conic is defined by any triple of points on it). Thus this configuration of points gives the decomposition of the tangent space
8314: \begin{equation}
8315: T_{P_{\Ga}} \ov{\sM_{g}}= \bigoplus_{e \in E(\Ga)} \C_{e}
8316: \end{equation}
8317: where $\PP \C_{e} = \phi_{2K_{\Ga}}(p_{e})$.
8318:
8319: In the next
8320: subsection we show that every tangent direction $\C_{e}$ corresponds
8321: to a rational curve in $\ov{\sM_{g}}$. This identification of images of nodes
8322: under the double canonical embedding and the directions of special deformations cancels the projective transformation ambiguity.
8323:
8324: \subsection{Special 1-parameter deformation of large limit curves}
8325:
8326: Let $ e \subset \Ga$ be a 3-valent graph $\Ga$ with
8327: fixed edge $e \in E(\Ga)$ such that $v, v' = \p e$ are two
8328: different points with stars $S(v)= e, e_{1}, e_{2}$ and $S(v') =
8329: e, e'_{1}, e'_{2}$. That is $e$ is not a loop. We call such pair
8330: a simple {\it flag}. Let us blow down this edge in our graph. We
8331: get new graph with 4-valent vertex $v_{new} = v = v'$ with the
8332: star $S(v_{new}) = e_{1}, e_{2}, e'_{1}, e'_{2}$. There are 3
8333: partitions this edges to pairs: the old one $(e_{1}, e_{2}) \vert
8334: (e'_{1}, e'_{2})$ and couple of new ones:
8335: \begin{equation}
8336: (e_{1}, e'_{2} \vert e_{2}, e'_{1}) \text{ and } (e_{1}, e'_{1}
8337: \vert e_{2}, e'_{2}).
8338: \end{equation}
8339: Now we can blow up the vertex $v_{new}$ to the edge $e_{new}$
8340: with vertices $\p e_{new} = v_{new}, v'_{new}$ and stars
8341: \begin{enumerate}
8342: \item $S(v_{new}) = e_{new}, e_{1}, e_{2}$ and $S(v'_{new}) = e_{new},
8343: e'_{1}, e'_{2}$. This is our starting graph $\Ga$.
8344: \item $S(v_{new}) = e_{new}, e_{1}, e'_{2}$ and $S(v'_{new}) = e_{new},
8345: e'_{1}, e_{2}$. This is the first new graph $\Ga'$.
8346: \item $S(v_{new}) = e_{new}, e_{1}, e'_{1}$ and $S(v'_{new}) = e_{new},
8347: e'_{2}, e'_{2}$. This is the second new graph $\Ga''$.
8348: \end{enumerate}
8349: Remark that as a result of this construction every obtained graph
8350: has distinguished edge that is we have the triple of flags
8351: \begin{equation}
8352: (e \subset \Ga), ( e_{new} \subset \Ga'), ( e_{new} \subset
8353: \Ga'').
8354: \end{equation}
8355: Such triple we call a {\it nest} of flags. Every nest is defined
8356: uniquely by any flag $(e \subset) \Ga$ of the triple. Moreover,
8357: let $S(e)$ be the star of an edge $e$ that is the union
8358: of stars of the boundary $v, v' = \p e$
8359: \begin{equation}
8360: S(e) = S(v)\bigcup S(v').
8361: \end{equation}
8362: Then we have the canonical identification of graphs
8363: \begin{equation}
8364: \Ga - S(e) = \Ga' - S(e_{new}) = \Ga'' - S(e_{new}).
8365: \end{equation}
8366:
8367: \begin{rmk}
8368: It is easy to see that if $e$ is a loop this construction gives
8369: the same graph with the same loop $e$ again.
8370: \end{rmk}
8371: Consider two componets $P_{v}, P_{v'}$ of a curve $P_{\Ga}$ with
8372: common point $p_{e}$ where $v, v' = \p e$. Remove this point and
8373: glue $P_{v}$ and $P_{v'}$ by a tube that is, consider the
8374: connected sum
8375: \begin{equation}
8376: P_{v}\#_{p(e)} P_{v'} = P_{v,v'} = S^{2}.
8377: \end{equation}
8378: This is a 2-sphere with two pairs of points $(p_{e_{1}}, p_{e_{2}})$
8379: and $(p_{e'_{1}}, p_{e'_{2}})$ where $(e, e_{1}, e_{2}) = S(v)$
8380: and $(e, e'_{1}, e'_{2}) = S(v')$. If we fix a complex structure
8381: on $S^{2}$ and consider the double cover
8382: \begin{equation}
8383: \phi \colon E \to \PP^{1}
8384: \end{equation}
8385: with ramification points
8386: \begin{equation}
8387: W = p_{e_{1}} \bigcup p_{e_{2}} \bigcup p_{e'_{1}} \bigcup
8388: p_{e'_{1}}
8389: \end{equation}
8390: we obtain an elliptic curve $E$ with a point of second order
8391: \begin{equation}
8392: \si = p_{e_{1}} + p_{e_{2}} - p_{e'_{1}} - p_{e'_{2}}.
8393: \end{equation}
8394: So the moduli space of complex structures on $S^{2}$ is equal to
8395: $\sM_{1}^{2}$ - the moduli space of smooth elliptic curves with fixed
8396: point of order 2.
8397:
8398: Every such complex structure $\tau \in \sM_{1}^{2}$ on $S^{2}$ and the
8399: standard complex structures on all others components define a
8400: stable algebraic reducuble curve $P_{e \subset\Ga, \tau}$. Thus
8401: we obtain an embedding
8402: \begin{equation}
8403: \psi_{e \subset \Ga} \colon \sM_{1}^{2} \to \ov{\sM_{g}}.
8404: \end{equation}
8405: The moduli space
8406: \begin{equation}
8407: \sM_{1}^{2} = \PP^{1} - (\si, \si', \si'')
8408: \end{equation}
8409: is the projective line without 3 points. These 3 points
8410: correspond to 3 possibility to divide 4 points $p_{e_{1}},
8411: p_{e_{2}}, p_{e'_{1}}, p_{e'_{2}}$ in 2 pairs that is a choice of
8412: a point of order 2 on an elliptic curve.
8413:
8414: It is easy to see that
8415:
8416: \begin{prop}
8417: \begin{enumerate}
8418: \item The map $\psi_{e \subset \Ga}$ (9.24) can be extended to the map
8419: \begin{equation}
8420: \psi_{e \subset \Ga} \colon \PP^{1} \to \ov{\sM_{g}};
8421: \end{equation}
8422: \item
8423: \begin{equation}
8424: \psi_{e \subset \Ga} (\si) = p_{e} \in P_{\Ga}
8425: \end{equation}
8426: where $\si$ is given by (9.23) and $P_{\Ga}$ is a large limit curve with
8427: fixed node corresponding to the edge $e$,
8428: \item
8429: \begin{equation}
8430: \psi_{e \subset \Ga} (\si') = P_{\Ga'};
8431: \end{equation}
8432: and
8433: \item
8434: \begin{equation}
8435: \psi_{e \subset \Ga} (\si'') = P_{\Ga''};
8436: \end{equation}
8437: where the triple
8438: \begin{equation}
8439: (e \subset \Ga), (e_{new} \subset \Ga'), ( e_{new} \subset \Ga'')
8440: \end{equation}
8441: is a nest of flags (9.17).
8442: \item Now we can identify the tangent direction
8443: \begin{equation}
8444: T\psi_{e \subset \Ga} (\PP^{1})_{P_{\Ga}} = \C_{e}
8445: \end{equation}
8446: from the decomposition (9.15) with the corresponding node of the
8447: double canonical curve. We obtain the canonical pluricanonical model.
8448: \end{enumerate}
8449: \end{prop}
8450: \begin{rmk} If edge $e$ is a loop with a vertex $v$ such that
8451: the star $S(v) = e, e'$ then $P_{v}$ is a rational curve with one
8452: double point $p_{e}$ and the smooth point $p_{e'}$. The operation
8453: of connected summing around double point $p_{e}$ gives a smooth
8454: 2-torus $T^{2}$ with fixed point $p_{e'}$ and the isotopy class
8455: $a \in H_{1}(T^{2}, \Z)$ which is the class of the neck of gluing
8456: tube. The class $a$ mod 2 gives a point of order 2 on $T^{2}$.
8457: The space $\sM_{1}^{2}$ of complex structures on $T^{2}$ with
8458: such additional data is $\C^{*} - 1$ that is $\PP^{1}$ without 3
8459: points again. In this case the rational curve $\psi_{e \subset
8460: \Ga} (\sM_{1}^{2})$ admits the compactification by the double
8461: point corresponding to $P_{\Ga}$.
8462: \end{rmk}
8463:
8464: So, the Deligne-Mumford compactification $\ov{\sM_{g}}$ contains
8465: the configuration $\sC$ of rational curves
8466: \begin{equation}
8467: \{ C \} = \sC \text{ every } C = \psi_{e \subset \Ga} (\PP^{1})
8468: \end{equation}
8469: for some flag $e \subset \Ga$. We can consider this configuration of rational
8470: curves as a reducible curve
8471: \begin{equation}
8472: \sC = \bigcup C
8473: \end{equation}
8474: that is the union of all components. It is easy to see that
8475:
8476: \begin{prop}
8477: \begin{enumerate}
8478: \item for every component $C$ of $\sC$
8479: \begin{equation}
8480: C \bigcap \sP = P_{\Ga} + P_{\Ga'} + P_{\Ga''}
8481: \end{equation}
8482: where $(\Ga, \Ga', \Ga'')$ is a nest of graphs (9.17 );
8483: \item for a pair $C, C'$ of componets either the intersection $C \bigcap C'$
8484: is empty or it is transversal and
8485: \begin{equation}
8486: C \bigcap C' \in \sP
8487: \end{equation}
8488: \item the set $S(\Ga)$ of components through every point
8489: $P_{\Ga} \in \sP$ is enumerated by the set $E(\Ga)$;
8490: \end{enumerate}
8491: \end{prop}
8492:
8493: We saw that if an edge $e \in E(\Ga)$ is not a loop then the map
8494: $\psi_{e \subset \Ga}$ (9.26) is an embedding and the component $C$
8495: (9.24) of the reducible curve $\sS \subset \ov{\sM_{g}}$ is a
8496: smooth rational curve with fixed 3 points $ ( P_{\Ga}, P_{\Ga'},
8497: P_{\Ga''})$ such that the corresponding 3 graphs form a nest
8498: (9.17). Over every of such points the tangent spaces admit
8499: decompositions (9.15). Geometry of the other curves from the family
8500: of curves parametrized by $C$ is very near to the geometry of
8501: large limit curves: let $e \in E(\Ga)$, $v, v' = \p e$ and $P_{E,
8502: \si}$ is the point of $C$ corresponding to the elliptic curve
8503: (9.21) with a point of order 2 (9.23). Then $P_{E, \si}$ has 2g-4
8504: old components
8505: \begin{equation}
8506: \bigcup_{v'' \neq v, v'} P_{v''}
8507: \end{equation}
8508: with triples of points and 3g-1 nodes $p_{e'}, e \neq e'$ and one
8509: new component $P_{v, v'}$ with the quadruple of points (see
8510: (9.16) and 1), 2) and 3) after this formula). Then
8511:
8512: \begin{enumerate} \item
8513: the canonical class $K_{P_{E, \si}}$ is a line bundle: a
8514: restriction of it to every component $P_{v''}, v'' \neq v, v' $
8515: is the sheaf of meromorphic differentials $\om$ with simple poles
8516: at $p_{e}, p_{e'}, p_{e''}$ where $\{e,e', e''\} = S(v)$;
8517: \item the
8518: restriction of the canonical class to the component $P_{v, v'}$
8519: is the sheaf of meromorphic differentials $\om$ with simple poles
8520: at $p_{e_{1}}, p_{e_{2}}, p_{e'_{1}}, p_{e''_{2}}$ (see (9.16)).
8521: \item
8522: Thus
8523: \begin{equation}
8524: c (K_{P_{E, \si}}) = (2,1,...,1) \in NS_{P_{E, \si}}
8525: \end{equation}
8526: where the first coordinate corresponds to $P_{v, v'}$.
8527: \item Again
8528: every holomorphic section $s$ of the canonical class is a
8529: collection of meromorphic differentials $\{ \om_{v''}\}$ on the
8530: components $P_{v''}$ with poles at $p_{e}, p_{e'}, p_{e''}$ where
8531: $\{e,e', e''\} = S(v)$ and a meromorphic differential $\om_{v,
8532: v'}$ on the component $P_{v, v'}$ with poles at the quadruple
8533: with the same constraints (1.29) and (1.30).
8534: \item
8535: The canonical map defined by the complete linear system $\vert K_{P_{E,
8536: \si }}\vert$
8537: \begin{equation}
8538: \phi_{K} \colon P_{E, \si} \to \PP^{g-1}
8539: \end{equation}
8540: sends $P_{v''}$ to a configuration of lines and $P_{v, v'}$ to a
8541: conic in $\PP^{g-1}$.
8542: \item Again the dimension of the space of quadratic differentials
8543: on $P_{E, \si}$ is equal to 3g-3 and this curve is an orbifold
8544: smooth point of $\ov{\sM_{g}}$.
8545: \item
8546: The double canonical map of $P_{E, \si}$ given by the complete
8547: linear system $\vert 2K_{\Ga}\vert$ is an embedding
8548: \begin{equation}
8549: \phi_{2K_{P_{E, \si}}} \colon P_{E, \si} \to \PP^{3g-4} = \PP
8550: T\ov{\sM_{g}}_{P_{E, \si}}.
8551: \end{equation}
8552: (see (9.11)).
8553: \item The images of nodes
8554: \begin{equation}
8555: \{ \phi_{2K_{P_{E, \si}}}(p_{e'}) \}, \quad e \neq e' \in
8556: E(\Ga)
8557: \end{equation}
8558: define the decomposition of the restriction of the tangent bundle
8559: \begin{equation}
8560: T\ov{\sM_{g}} \vert_{C} = TC \bigoplus (\bigoplus_{e \neq e' \in
8561: E(\Ga)} L_{e'})
8562: \end{equation}
8563: where the fiber of the line bundle $L_{e'}$ over a point is the
8564: component of the decomposition (9.15).
8565: \item Thus every line bundle $L_{e'}$ from the previous decomposition
8566: is the tautological line bundle
8567: \begin{equation}
8568: L_{e'} = \Oh_{C} (-1),
8569: \end{equation}
8570: \item hence the previous decomposition is
8571: \begin{equation}
8572: T\ov{\sM_{g}} \vert_{C} = \Oh_{C}(2)\bigoplus (3g-4) \Oh_{C}(-1).
8573: \end{equation}
8574: \item The restriction to $C$ of the canonical class of $\ov{\sM_{g}}$
8575: is
8576: \begin{equation}
8577: K_{\ov{\sM_{g}}} \vert_C = \Oh_{C}(3(g-2)).
8578: \end{equation}
8579: \end{enumerate}
8580:
8581: \subsection{ Special 2-parameter deformations of large limit curves}
8582:
8583: Let $v \subset \Ga$ be a pair of a 3-valent graph $\Ga$ and fixed
8584: vertex $v \in V(\Ga)$ and in the star $ S(v) = e \bigcup e_{1}
8585: \bigcup e_{2}$ let us choose a pair $e_{1}$ and $e_{2}$ of edges.
8586: Then we have the pair of vertices $v_{1}$ and $v_{2}$ defined by
8587: equalities $\p e_{1} = v, v_{1}$ and $\p e_{2} = v, v_{2}$.
8588:
8589: Thus such choices are given by 3-flag on a graph $\Ga$
8590: \begin{equation}
8591: v \subset e \subset \Ga
8592: \end{equation}
8593: where
8594: \begin{equation}
8595: e_{1} \bigcup e_{2} = S(v) - e
8596: \end{equation}
8597: Now let us blow down
8598: edges $e_{1}$ and $e_{2}$. We obtain a new graph $\Ga_{new}$
8599: with a new 5-valent vertex $v_{new} \in V(\Ga_{new})$ and fixed
8600: additional choices on the star of $v_{new}$:
8601: \begin{enumerate}
8602: \item fixed edge $e \in S(v_{new}) $ through $v_{new}$ and
8603: \item a partition of other quadruple edges from the star $S(v_{new})$
8604: in two pairs $S(v_{1}) - e_{1}$ and $S(v_{2} - e_{1})$.
8605: \end{enumerate}
8606:
8607:
8608: There are exist 15 such choices divided into 5 triples corresponding
8609: to the first choice of $e' \in S(v_{new}) $. For every such a choice
8610: let us blow up from $v_{new}$ two new edges $e'_{1}$ and $e'_{2}$
8611: such a way that
8612: \begin{enumerate}
8613: \item $v_{new}$ becomes a new 3-valent vertex $v'$
8614: and $S(v') = e', e'_{1}, e'_{2}$,
8615: \item two vertices appear $v'_{1}$ and $v'_{2}$ such that
8616: $\p e'_{1} = v', v'_{1}$ and $\p e'_{2} = v', v'_{2}$.
8617: \end{enumerate}
8618: Thus we get a new 3-flag
8619: \begin{equation}
8620: v' \subset e' \subset \Ga'
8621: \end{equation}
8622: and full collection of choices gives a collection of 3-flags
8623: \begin{equation}
8624: (v \subset e \subset \Ga), (v' \subset e' \subset \Ga'), (v''
8625: \subset e'' \subset \Ga''), ........
8626: \end{equation}
8627: We call this collection of 15 3-flags a {\it nest of 3-flags}.
8628:
8629: A partial case of the choice is the following:
8630: \begin{enumerate}
8631: \item let us blow down just one edge, for example, $e_{1}$,
8632: \item then we get 4-valent vertex $v_{new}$ with the star
8633: $S(v') = e, e_{2}, e_{3}, e_{4}$ where $e_{3} \bigcup e_{4} =
8634: S(v_{1}) - e_{1}$.
8635: \item Now let us change $e$ by any edge from the pair $e_{3},
8636: e_{4}$, for example, by $e_{3}$ and blow up the vertex $v_{new}$
8637: to a new edge $e'$ with vertices $\p e' = v', v''$ such that
8638: \begin{equation}
8639: S(v') = e' \bigcup e \bigcup e_{3} \text{ and } S(v'') = e'
8640: \bigcup e_{2} \bigcup e_{4}.
8641: \end{equation}
8642: Then we have a pair of 3-flags
8643: \begin{equation}
8644: ((v \subset e \subset \Ga)) \text{ and } (v' \subset e' \subset
8645: \Ga')
8646: \end{equation}
8647: from the same nest of 3-flags such that the flags
8648: \begin{equation}
8649: (e_{1} \subset \Ga) \text{ and } (e' \subset \Ga')
8650: \end{equation}
8651: belong to the same nest (9.17).
8652: \item if in this construction we replace $e$ by $e_{4}$ we get the
8653: nest triple
8654: \begin{equation}
8655: (e_{1} \subset \Ga), (e' \subset \Ga'), (e'' \subset \Ga'')
8656: \end{equation}
8657: (see (9.17)).
8658: \end{enumerate}
8659:
8660:
8661: Any flag $(v \subset e \subset \Ga)$ defines
8662: \begin{enumerate}
8663: \item the reducible curve $P_{\Ga}$,
8664: \item its irreducible component $P_{v}$,
8665: \item a pair of its points $p_{e_{1}}, p_{e_{2}} \subset P_{v}$ and
8666: \item
8667: two components $P_{v_{1}}$ and $P_{v_{2}}$.
8668: \end{enumerate}
8669: Remove nodes $p_{e_{i}}, i = 1, 2$ and glue components by tubes
8670: that is consider the connected sum
8671: \begin{equation}
8672: P_{v_{1}} \#_{p_{e_{1}}} P_{v} \#_{p_{e_{2}}} P_{v_{2}} = S^{2}.
8673: \end{equation}
8674: This is 2-sphere with fixed 5 points with fixed the point $p_{e}$
8675: and two pairs of points corresponding to the following two pairs of edges
8676: \begin{equation}
8677: (S(v_{1}) - e_{1}), (S(v_{2}) - e_{2}).
8678: \end{equation}
8679: The moduli space of complex structures on $S^{2}$ with fixed 5
8680: points subjecting such partition conditions is
8681: \begin{equation}
8682: S((v \subset e \subset \Ga)) = \PP^{1} \times S^{2}(S^{2}
8683: \PP^{1}) / PGL(2, \C),
8684: \end{equation}
8685: so it is rational affine surface. Every complex line with 5 points can
8686: be uniquely extended to a reducible curve adding the corresponding
8687: complex lines with 3 points. Thus the surface $S((v \subset e
8688: \subset \Ga))$ is embedded
8689: \begin{equation}
8690: \psi_{(v \subset e \subset \Ga)} \colon S((v \subset e \subset
8691: \Ga)) \to \ov{\sM_{g}}
8692: \end{equation}
8693: and its compactification belongs to
8694: \begin{equation}
8695: \ov{\psi_{(v \subset e \subset \Ga)}S((v \subset e \subset \Ga))}
8696: \subset \ov{\sM_{g}}.
8697: \end{equation}
8698: The compactification divisor is
8699: \begin{equation}
8700: \ov{\psi_{(v \subset e \subset \Ga)}S((v \subset e \subset \Ga))}
8701: - \psi_{(v \subset e \subset \Ga)}S((v \subset e \subset \Ga)) =
8702: \bigcup C
8703: \end{equation}
8704: where 15 components of the curve $\sC$ corresponding to the nests
8705: (9.48) and (9.52).
8706:
8707: Again the Deligne-Mumford compactification $\ov{\sM_{g}}$
8708: contains the configuration $\sS$ of rational surfaces
8709: \begin{equation}
8710: \{ S \} = \sS \text{ every } S = \psi_{v \subset e \subset \Ga}
8711: (S((v \subset e \subset \Ga)))
8712: \end{equation}
8713: for some 3-flag $v \subset e \subset \Ga$. We can consider
8714: this configuration of
8715: rational surfaces as a reducible surface
8716: \begin{equation}
8717: \sS = \bigcup S
8718: \end{equation}
8719: that is the union of all components. It is easy to see that
8720:
8721: \begin{prop}
8722: \begin{enumerate}
8723: \item for every component $S$ of $\sS$
8724: \begin{equation}
8725: S \bigcap \sP = \bigcup C
8726: \end{equation}
8727: where the curves $ C $ belongs the collection (9.32);
8728: \item for a pair $S, S'$ of componets the intersection $C \bigcap C'$
8729: is either empty or it is transversal and
8730: \begin{equation}
8731: C \bigcap C' \subset \sC
8732: \end{equation}
8733: \end{enumerate}
8734: \end{prop}
8735:
8736: We saw that if an edge $e \in E(\Ga)$ from 3-flag $v \subset e
8737: \subset \Ga$ is not a loop then the map $\psi_{v \subset e \subset
8738: \Ga}$ (3.12) is an embedding and the component $S$ (3.13) of the
8739: reducible surface $\sS \subset \ov{\sM_{g}}$ is a rational
8740: surface with fixed 15 components of the reducible curve $\sC$.
8741: Over any of such irreducuble curve the restriction of the
8742: tangent bundle of $\ov{\sM_{g}}$ admits decompositions (9.41) and
8743: (9.43). The structure of a general stable curve $P_{s}, s \subset S$
8744: from the 2-dimensional
8745: family parametrized by $S$
8746: is a slightly different from the geometry of the large limit curves:
8747: $P_{s}$ has 2g-5
8748: old components
8749: \begin{equation}
8750: \bigcup_{v'' \neq v, v_{1}, v_{2}} P_{v''}
8751: \end{equation}
8752: and one
8753: new component $P_{v, v_{1}, v_{2}}$ with 5 points. Then
8754:
8755: \begin{enumerate} \item
8756: the canonical class $K_{s}$ is a line bundle: the restriction of it
8757: to every component $P_{v''}, v'' \neq v, v_{1}, v_{2} $ is the
8758: sheaf of meromorphic differentials $\om$ with simple poles at
8759: $p_{e}, p_{e'}, p_{e''}$ where $e \bigcup e' \bigcup e'' = S(v)$;
8760: \item the
8761: restriction of the canonical class to the component $P_{v, v_{1},
8762: v_{2} }$ is the sheaf of meromorphic differentials $\om$ with
8763: simple poles at $p_{e}, p_{e_{1}}, p_{e_{2}}, p_{e'_{1}},
8764: p_{e'_{2}}$.
8765: \item
8766: Thus in terms of subsection 1.1
8767: \begin{equation}
8768: c (K_{P_{s}}) = (3,1,...,1) \in NS_{P_{s}}.
8769: \end{equation}
8770: where the first coordinate corresponds to $P_{v, v'_{1}, v_{2}}$.
8771: \item
8772: The canonical map by the complete linear system $\vert
8773: K_{P_{s}}\vert$
8774: \begin{equation}
8775: \phi_{K} \colon P_{s} \to \PP^{g-1}
8776: \end{equation}
8777: sends $P_{v''}$ to a configuration of lines and $P_{v, v_{1},
8778: v{2}}$ to a cubic in $\PP^{g-1}$.
8779: \item Again the dimension of the space of quadratic differentials
8780: on $P_{s}$ is equal to 3g-3 and this curve is an orbifold smooth
8781: point of $\ov{\sM_{g}}$.
8782: \item
8783: The double canonical map of $P_{E, \si}$ given by the complete
8784: linear system $\vert 2K_{P_{s}}\vert$ is an embedding
8785: \begin{equation}
8786: \phi_{2K_{P_{s}}} \colon P_{s} \to \PP^{3g-4} = \PP
8787: T\ov{\sM_{g}}_{s}.
8788: \end{equation}
8789: \item The images of nodes
8790: \begin{equation}
8791: \{ \phi_{2K_{P_{s}}}(p_{e'}) \}, \quad e \neq e_{1}, e_{2} \in
8792: E(\Ga)
8793: \end{equation}
8794: define the decomposition of the restriction of the tangent bundle
8795: \begin{equation}
8796: T\ov{\sM_{g}} \vert_{S} = TS \bigoplus (\bigoplus_{e \neq e_{1},
8797: e_{2} \in E(\Ga)} L_{e})
8798: \end{equation}
8799: where the fiber of the line bundle $L_{e}$ over a point is the
8800: component of the decomposition (9.43).
8801: \end{enumerate}
8802:
8803: \subsection{Modular configuration}
8804:
8805: We understands the moduli subspaces (9.33) and (9.60) as partial cases of the general construction.
8806: For any graph $\Ga$ with vertices of any valence $ > 2$ consider the corresponding reducible curve
8807: \begin{equation}
8808: P_{\Ga} = \bigcup_{v \in V(\Ga)} P_v
8809: \end{equation}
8810: where each component equals to $\PP^1$ with $\vert S(v) \vert$
8811: fixed points of transversal intersections with other components along the incidence correspondence of the graph. If our graph
8812: isn't 3-valent such curve has moduli. These moduli are defined by the positions of $\vert S(v) \vert$
8813: points on the component $P_v$ and by the combinatoric of the set of expansions $\sE(\Ga)$ (see subsection 6.3). For example, if
8814: $\Ga$ is the result of one contraction of 3-valent graph $\Ga_{max}$ the corresponding moduli space $\sM^2_1$ (9.24) is a 3-cover of the moduli space of smooth elliptic curves.
8815: For graphs with one 5-valent vertex the corresponding moduli space is the rational surface $S$ from the configuration
8816: (9.60) and so on.
8817:
8818: Every such moduli space $\sM_\Ga$ is embedded into the Deligne-
8819: Mumford compactification and it admits itself the Deligne-Mumford compactification
8820: \begin{equation}
8821: \ov{\sM_\Ga} \subset \ov{\sM_g}
8822: \end{equation}
8823: such that
8824: \begin{equation}
8825: \ov{\sM_\Ga} = \sM_v \bigcup (\bigcup_{\Ga'> \Ga}) \ov{\sM_{\Ga'}}
8826: \end{equation}
8827: where graph $\Ga'$ is an expansion of $\Ga$.
8828:
8829: Obviously the maximal dimension of such moduli space
8830: is $2g-3$.
8831:
8832: So in the compactification $\ov{\sM_g}$ of the moduli space of genus $g$
8833: curves we have the configuration of the moduli spaces
8834: \begin{equation}
8835: \sM = \bigcup_{\Ga} \sM_\Ga \subset \ov{\sM_g}
8836: \end{equation}
8837: such that
8838: \begin{enumerate}
8839: \item
8840: \begin{equation}
8841: \ov{\sM_\Ga'} \bigcap \ov{\sM_{\Ga''}}
8842: = \bigcup_{\Ga > \Ga', \Ga'' } \sM_\Ga;
8843: \end{equation}
8844: \item the restriction of the tangent bundle of $\ov{\sM_g}$
8845: to every component $\sM_\Ga$ admits the decomposition like (9.68).
8846: \end{enumerate}
8847:
8848:
8849:
8850:
8851:
8852: \subsection{ $Pic_0(P_\Ga)$ and moduli spaces $M_{vb}^{ss}(P_\Ga)$ of ll-curves}
8853:
8854: For a singular curve the moduli spaces of vector bundles are not compact. These moduli spaces admit the compactifications by torsion free sheaves. These sheaves are not local free only over
8855: nodes but the theory of compactification is very close to
8856: such theory for algebraic surfaces. All details of this theory
8857: can be found in Artamkin paper \cite{Ar}.
8858:
8859: For every torsion free sheaf $F$ on $P_\Ga$ we have the standard exact surface
8860: \begin{equation}
8861: 0 \to F \to F^{**} \to C(F) \to 0
8862: \end{equation}
8863: where
8864: \begin{enumerate}
8865: \item the quotient sheaf has the form
8866: $$
8867: Supp C(F) \subset \bigcup_{e \in E(\Ga)} p_e;
8868: $$
8869: \item the double dual sheaf $F^{**}$ is a vector bundle on the normalization of the curve.
8870: \end{enumerate}
8871: The cardinality of $Supp F$ is the measure of deviation from the local freedom of
8872: $F$. Let this number be
8873: \begin{equation}
8874: \vert Supp C(F) \vert = c_2(F).
8875: \end{equation}
8876: So $c_2(F) = 0$ implies $F$ is local free thus it is a vector bundle.
8877:
8878: A vector bundle $E$ on $P_\Ga$ is called topologicaly trivial if
8879: the restrictions $E \vert_{P_v}$ are trivial for all $v \in V(\Ga)$.
8880: Below we describe the moduli spaces of topologicaly trivial
8881: \begin{enumerate}
8882: \item line bundles $Pic_0 (P_\Ga)$,
8883: \item rk 2 semi-stable bundles $M_{vb}^{ss}$;
8884: \item the compactification $J(P_\Ga) = \ov{Pic_0(P_\Ga)}$ of
8885: $Pic_0(P_\Ga)$ by torsion free sheaves $F$ (for which obviously
8886: $F^{**} \in Pic_0(P_\Ga)$);
8887: \item the compactification $M^{ss}(P_\Ga) = \ov{M^{ss}_{vb}(P_\Ga)}$ of
8888: $Pic_0(P_\Ga)$ by torsion free sheaves $F$ (for which obviously
8889: $F^{**} \in M^{ss}_{vb}(P_\Ga)$).
8890: \end{enumerate}
8891:
8892:
8893: Since each component $P_{v}$ is projective line the restriction of any topologicaly trivial line bundle equals
8894: \begin{equation}
8895: L \vert_{P_{v}} = \Oh_{P_{v}}.
8896: \end{equation}
8897:
8898:
8899:
8900: To describe a line bundle on $P_\Ga$ consider a collection of any trivializations of $L$ on all components $P_v$ of $P_\Ga$ and denote the line bundle with such addition structure by $L_0$.
8901:
8902: Now to get a line bundle $L$ on $P_{\Ga}$ we have to concord every
8903: pair of line bundles $\Oh_{P_{v}}, \Oh_{P_{v'}}$
8904: at common point $p_{e}$ if $v, v' = \p e$. Under our trivializations such concordance is
8905: given by a multiplicative constant $ a (\vec e) \in \C^{*}$ such that under the orientation reversing involution $i_e$
8906: $$
8907: a(i_e(\vec e)) = a(\vec e)^{-1}
8908: $$
8909: (compare with (6.86)). Thus $L_0$ defines a map
8910: $$
8911: a \colon \vec E(\Ga) \to \C^*
8912: $$
8913: (compare with (6.85)).
8914:
8915: The changing of trivialization is given by a function
8916: $$
8917: \tilde{\la} \colon V(\Ga) \to \C^*
8918: $$
8919: (compare with (6.90)) which acts on the functions $a$ by the formula
8920: $$
8921: \tilde{\la}(a(\vec e)) = \tilde{\la}(v_s) \cdot a(\vec e) \cdot \tilde{\la}(v_t)^{-1}
8922: $$
8923: where $\p \vec e = v_s \bigcup v_t$ (compare with (6.91)).
8924:
8925: Let $\sA_{\C}$ be the space of trivialized topologicaly trivial line bundles that is the space of functions $a$
8926: and $\sG_{\C}$ be the group of trivializations.
8927: Then in terms of subsection 6.6 the space $\sA_{\C}$ is the space of flat $\C^*$-connections, $\sG_{\C}$ is the complexification of the unitar
8928: abelian gauge group and
8929: \begin{equation}
8930: Pic_0(P_\Ga) = \sA_{\C} / \sG_{\C}.
8931: \end{equation}
8932: Thus
8933: \begin{equation}
8934: Pic_0(P_\Ga) = Hom(\pi_1(\Ga), \C^*) = S^a_g
8935: \end{equation}
8936: is the abelian Schottky space (2.68).
8937:
8938:
8939: To describe a rk 2 bundle $E$ on $P_\Ga$ consider a collection of its trivializations on all components $P_v$ of $P_\Ga$ and denote the bundle with such addition structure by $E_0$.
8940:
8941: Now to get a rk 2 vector bundle $E$ on $P_{\Ga}$ we have to glue every
8942: pair of line bundles $\Oh_{P_{v}} \oplus \Oh_{P_{v}}, \Oh_{P_{v'}} \oplus \Oh_{P_{v'}}$
8943: at common point $p_{e}$ if $v, v' = \p e$. Under our trivializations such concordance is
8944: given by a matrix $ a (\vec e) \in SL(2, \C)$ with the standard
8945: property (6.57).
8946:
8947:
8948:
8949: Thus $E_0$ defines a flat $SL(2, \C)$-connection
8950: $$
8951: a \colon \vec E(\Ga) \to SL(2, \C)
8952: $$
8953: (compare with (6.56)).
8954:
8955: The changing of the trivialization is given by a function
8956: $$
8957: \tilde{g} \colon V(\Ga) \to SL(2, \C)
8958: $$
8959: (compare with (6.63)) which acts on the functions $a$ by the formula
8960: $$
8961: \tilde{g}(a(\vec e)) = \tilde{u}(v_s) \cdot a(\vec e) \cdot \tilde{g}(v_t)^{-1}
8962: $$
8963: where $\p \vec e = v_s \bigcup v_t$ (compare with (6.65)).
8964:
8965: Thus the space $\sA_{ \C}$ of trivialized topologicaly trivial rk 2 bundles coincides with the space of flat $SL(2, \C)$-connections
8966: and the group of trivializations $\sG_{ \C}$ coincides with the complex gauge group.
8967: Then in terms of subsection 6.5
8968: \begin{equation}
8969: M^{ss}_{vb}(P_\Ga) = \sA_{\C} / \sG_{\C}.
8970: \end{equation}
8971: Thus
8972: \begin{equation}
8973: M^{ss}_{vb}(P_\Ga) = \CLRep(\pi_1(\Ga), SL(2, \C)) = S_g
8974: \end{equation}
8975: is the Schottky space (see subsection 3.2).
8976:
8977: Let us return to the subsection 3.2 and the handlebodies (6.9). Every such
8978: handlebody defines the exact sequence
8979: \begin{equation}
8980: 1 \to F_g \to \pi_1 (\Si_\Ga) \to \pi_1 (\Ga) \to 1
8981: \end{equation}
8982: where the normal subgroup is a free group with $g$ generators. Using this sequence we can choose the standard generators of $\pi_1 (\Si_\Ga)$
8983: $$
8984: \pi_1 = <a_1, ... ,a_g, b_1, ... , b_g \vert \prod_{i=1}^g [a_i, b_i] = 1>
8985: $$
8986: such that
8987: \begin{equation}
8988: F_g = <a_1, ... , a_g>
8989: \end{equation}
8990: $$
8991: \pi_1 (\Ga) = <b_1, ... , b_g>.
8992: $$
8993: Then we can consider the space of classes of $SL(2, \C)$-representations
8994: \begin{equation}
8995: \CLRep(\pi_1(\Ga), SL(2, \C)) = S_g
8996: \end{equation}
8997: as a family of flat vector bundles on $\Si_\Ga$. If we fix a complex structure
8998: $I$ on $\Si_\Ga$ we get the forgetful map
8999: \begin{equation}
9000: f_I \colon S_g \to M^{ss}(\Si_I)
9001: \end{equation}
9002: (3.44).
9003: The space of complex structures on $\Si_\Ga$ under fixing of the basis is
9004: the Teichmuller space $\tau_g$ (5.31) and we can consider the universal family
9005: of the forgetful maps
9006: \begin{equation}
9007: F \colon S_g \times \tau_g \to \sM
9008: \end{equation}
9009: where $u \colon \sM \to \tau_g$ is the universal family of moduli spaces of vector bundles such that $u^{-1}(I) = M^{ss}(\Si_I)$.
9010:
9011: Now it is easy to see that
9012: \begin{enumerate}
9013: \item the ll-curve $P_\Ga$ is contained by the Deligne-Mumford
9014: compactification of $\tau_g$,
9015: \item the moduli space $ \ov{M^{ss}(P_\Ga)}$ is contained by the compactification of
9016: $\sM$ corresponding to the Deligne-Mumford
9017: compactification of $\tau_g$,
9018: \item the specialization of the universal forgetful map F ( . ) on $P_\Ga$ coincides with the standard embedding
9019: \begin{equation}
9020: F_{P_\Ga} = i \colon S_g = M_{vb}^{ss}(P_\Ga) \to \ov{M^{ss}(P_\Ga)}
9021: \end{equation}
9022: described before.
9023: \end{enumerate}
9024:
9025: From this it easy to prove the following partial solution of the Narasimhan problem
9026:
9027: \begin{prop} For a general smooth algebraic curve $\Si_I$ of genus $g$ the image of the forgetful
9028: map $f_I : S_g \to M^{ss}(\Si)$ (3.44) is Zariski dense in this moduli space.
9029: \end{prop}
9030: Indeed, this fact is true for $P_\Ga \in \sM_g$ and this property is preserved
9031: by small deformations in $\sM_g$. (I thank Gerd Detloff for the simplification of my arguments.)
9032:
9033: This fact is quite productuve even in the classical abelian case (see subsection 2.4): the standard embedding
9034: \begin{equation}
9035: i \colon (\C^*)^g = S^a_g \to \ov{Pic_0(P_\Ga)}
9036: \end{equation}
9037: is a limit of the standard forgetful covers $f_I \colon (\C^*)^g \to J(\Si_I)$
9038: (see (2.68)- (2.73)). We can see what happens with the classical theta functions (2.73) under this limit procedure. In particular we can see why
9039: $\delta$-function (2.46) appears.
9040:
9041: Moreover, the theory of classical theta functions can be constructed by CST-method from subsection 2.3 on $Pic_0(P_\Ga)$ and can be extended to any smooth curve using
9042: the classical holomorphic flat connection perfectly described by Welters in
9043: \cite{W}. We use the same strategy as well as in the non-abelian case:
9044: CST-constructions from section 6 can be interpreted as the theory of non-abelian theta functions on $ M_{vb}^{ss}(P_\Ga)$. After this theory
9045: can be extended to any smooth curve using the Hitchin projective flat connection. In this situation to avoid the meromorphicy of the forgetful map
9046: we apply the geometry of Lagrangian and Legandrian cycles and Bohr-Sommerfeld
9047: quantization.
9048:
9049: Here we use existence of an algebraic-geometrical compactification
9050: of the spaces of vector bundles on ll-curves. Now we describe these compactifications precisely.
9051:
9052:
9053: \subsection{Compactifications and AG-theory of theta functions
9054: for ll-curves}
9055:
9056: Both of our spaces $Pic_0(P_\Ga)$ and $M^{ss}_{vb}(P_\Ga)$ are
9057: affine varieties admitting canonical compactifications.
9058: The compactification in the abelian case will be called {\it jacobian}:
9059: \begin{equation}
9060: J(P_{\Ga}) = \ov{Pic_{0}(P_{\Ga})}.
9061: \end{equation}
9062:
9063: Every rk 1
9064: torsion free sheaf $F$ is given by a function
9065: $a$ as before but we have to
9066: compactify the target space by two points $0$ and $\infty$. So the new target
9067: space is the projective line $\PP^{1}_{\C^{*}}$ with fixed points
9068: $0$ and $\infty$ or fixed standard $\C^{*}$-action
9069: (distinguishing fixed points $0$ and $\infty$).
9070:
9071: So the compactification is
9072: \begin{equation}
9073: J(P_{\Ga}) = (\PP^{1}_{\C^{*}})^{3g-3} / \sG_{\C^{*}}.
9074: \end{equation}
9075: The boundary is the union of $6g-6$ divisors
9076: \begin{equation}
9077: D_{\Ga} = J(P_{\Ga}) - Pic_{0}(\Ga) = \bigcup_{e \in E(\Ga)}
9078: (D_{+}(e) \bigcup D_{-}(e))
9079: \end{equation}
9080: where
9081: \begin{equation}
9082: D_{+}(e) = \{ a \vert a (\vec e) = 0\},
9083: \end{equation}
9084: \begin{equation}
9085: D_{-}(e) = \{ a \vert a (\vec e) = \infty\}
9086: \end{equation}
9087: and
9088: \begin{equation}
9089: D_{+}(e) \bigcup D_{-} (e)= \{F \vert p_{e} \subset Sing F \}.
9090: \end{equation}
9091:
9092: Now consider an abelian spin network $\Ga_w$ (6.99) on our graph $\Ga$
9093: given by a function $w$ (6.99) subjecting to the conditions (6.100).
9094: Then we have divisor
9095: \begin{equation}
9096: D_w = \sum w(e)\cdot D_{sign (w(e))}
9097: \end{equation}
9098: with properties
9099: \begin{enumerate}
9100: \item this divisor is principal that is it is divisor of rational
9101: function $\theta_w$;
9102: \item this function is regular on $Pic_0(P_\Ga)$ being the limit of the ordinary theta
9103: functions for smooth curves;
9104: \item if $\Ga_w \in SNW^k_a$ that is $w $ is of level $ k$ (6.104)-(6.107) Then
9105: we can fix a pole to make all $w(e)$ non negative and get the canonical basis
9106: of theta functions with characteristics.
9107: \end{enumerate}
9108:
9109: Every rk 2
9110: torsion free sheaf $F$ is given by the same type function
9111: $a$ but we have to consider as the target space the compactification
9112: \begin{equation}
9113: SL(2, \C) \subset \PP^3 = \PP End(\C^2).
9114: \end{equation}
9115: with fixed quadric $Q$.
9116:
9117: So
9118: \begin{equation}
9119: M^{ss}(P_{\Ga}) = \ov{M^{ss}_{vb}}(P_\Ga) = (\PP^{3})^{E(\Ga)} / \sG_{\C}
9120: \end{equation}
9121: where the gauge group is the same (9.79).
9122:
9123: The boundary $D = M^{ss}(P_{\Ga}) - M^{ss}_{vb}(P_{\Ga}) $ is the union of $3g-3$ divisors
9124: \begin{equation}
9125: D = \bigcup_{e \in E(\Ga)}
9126: D(e).
9127: \end{equation}
9128: Recall (see the text after formula (6.33) and Fig.12) that we can remove the edge $e$ from $\Ga$ and forget the vertices $v_1, v_2 = \p e$. We get a new graph
9129: $\Ga_e$ of genus $g-1$. Then the corresponding divisor can be expressed
9130: \begin{equation}
9131: D(e) = Q_e \times M^{ss}(P_{\Ga_e}).
9132: \end{equation}
9133:
9134: For every spin network $\Ga_w$ of level 1 the divisor
9135: \begin{equation}
9136: \Theta_w = \sum_{e \in E(\Ga)} w(e) \cdot D(e)
9137: \end{equation}
9138: is the limit of effective theta divisors on smooth curves and the collection of these divisors
9139: \begin{equation}
9140: \{ \Theta_w\} \quad , w \in SNW_\Ga^1 = W^1(\Ga)
9141: \end{equation}
9142: gives the Mumford basis (2.13) which coincides with the Bohr-Sommerfeld basis (4.39) in the wave function space (see Definition 4) for
9143: ll-curve $P_\Ga$.
9144:
9145: Let us fix a spin network of level 1 $w_0 \in W^1(\Ga)$. Then one has
9146: \begin{enumerate}
9147: \item the divisor $k \Theta_{w_0}$;
9148: \item the collection of divisors
9149: \begin{equation}
9150: \{ \l.c.m.(\Theta_w, k \Theta_{w_0})\} \quad , w \in SNW_\Ga^k = W^k(\Ga)
9151: \end{equation}
9152: \item this collection of effective divisors from the complete linear system $| k \Theta_{w_0}|$ gives the Bohr-Sommerfeld basis (4.39) in the wave function space for
9153: ll-curve $P_\Ga$ and the polarization of level $k$.
9154: \end{enumerate}
9155: By the restriction to $uS_g$ we can check that the extension of this basis to smooth curves due to the Hitchin connection coincides with the Bohr-Sommerfeld basis given by BPU-construction.
9156:
9157:
9158:
9159:
9160:
9161: \begin{thebibliography}{99}
9162:
9163: \bibitem[A1]{A1} Atiyah M. ``Reflections on Geometry and
9164: Physics", p. 423
9165:
9166: \bibitem[A2]{A2} Atiyah M. "Complex analytic connections in
9167: fibre bundles", Trans.Amer. Math. Soc., 85, 1957, 181-207.
9168:
9169: \bibitem[A3]{A3} Atiyah M. ``The path integral formulation'', Oxford
9170: seminar on Jones--Witten theory, Michaelmas Term 1988. Seminar 6,
9171: p.~106.
9172:
9173: \bibitem[AB]{AB} Atiyah M. and Bott R. "The Yang-Mills equations over Riemann
9174: surfaces.", Phil. Trans. R. Soc. Lond. A 308, 1982, 523-615.
9175:
9176: \bibitem[An]{An} J. E. Andersen "Asymptotic faithfulness of quantum $SU(n)$
9177: representations of the mapping class groups", QA/0204084.
9178:
9179: \bibitem[Ar]{Ar} I. V. Artamkin "Algebraic geometry of ll-curves." appears in
9180: Izv. RAN,ser.Math.
9181:
9182: \bibitem[B]{B} Baker H. "Abel's theorem and the allied theory includinng the
9183: theory of theta functions.",
9184: Cambridge Univ. Press, 1897.
9185:
9186: \bibitem[Be]{Be} Beauville A."Vector Bundles on Curves and
9187: Generalized Theta Functions: Recent Results and Open Problems"
9188: Complex Algebraic Geometry, MSRI Publications, vol.28, 1995,
9189: 17-33.
9190:
9191:
9192: \bibitem[BGH]{BGH} E. D. Bolker, V. Guillemin, T. S. Holm,
9193: "How is a graph like a manifold?", CO/0206103.
9194:
9195:
9196:
9197: \bibitem[BL]{BL} Beauville A. and Laszlo Y. , "Conformal blocks and
9198: Generalized Theta Functions", Commun.Math. Phys. 164, 1994,
9199: 385-419.
9200:
9201: \bibitem[BPU]{BPU} D. Borthwick, T. Paul and A. Uribe, Legendrian
9202: distributions with applications to the non-vanishing of
9203: Poincar\'e series of large weight, Invent. math, {\bf122} (1995),
9204: 359--402 or hep-th/9406036
9205:
9206:
9207:
9208: \bibitem[Ber]{Ber} Bertram A.," Moduli of rank 2 vector bundles, theta
9209: divisors and the geometry of curves in projective space", J.
9210: Diff. Geom., 35,1992, 429--469.
9211:
9212:
9213:
9214: \bibitem[BD]{BD}
9215: Bradlow S. and Daskalopoulos G.
9216: " Moduli of stable pair for holomorphic bundles over Riemann
9217: surfaces", Int. J. Math., 2, 1991, 477--513.
9218:
9219:
9220: \bibitem[C]{C} Coble A. "Algebraic geometry and theta functions",
9221: AMS Coll.Publ. 10, Providence, R. I., 1929 (3 ed., 1969).
9222:
9223:
9224: \bibitem[CK]{CK} Cox D. and Katz S. "Mirror symmetry and Algebraic
9225: Geometry",AMS, Math.Surveys and Monographs, v. 68, 1999.
9226:
9227: \bibitem[DSS]{DSS} Danilov V., Shokurov V. and Shafarevich I.
9228: "Algebraic curves, algebraic manifolds and schemes", Encyclopedia
9229: of math. scinces. v. 23, Springer, 1998.
9230:
9231:
9232:
9233: \bibitem[D]{D} Delzant T. , "Hamiltonians periodiques et images
9234: convexed de l'application moment", Bull.Soc. Math. France, 116,
9235: 1988,315-339.
9236:
9237: \bibitem[Do]{Do} S. Donaldson ``Gluing techniques in the cohomology of
9238: moduli spaces'', in Topological methods in modern mathematics
9239: (Stony Brook, NY, 1991), Publish or Perish, Houston, TX, 1993,
9240: pp. 137--170.
9241:
9242:
9243: \bibitem[DS]{DS} Drinfeld V., and Simpson C., "B-structures on G-bundles
9244: and local triviality", Math. Res.-Letters, 2, 1995, 823-829.
9245:
9246:
9247: \bibitem[DKN]{DKN} Dubrovin B., Krichever I. and Novikov S.
9248: " Integrable systems I", Encycl. Math. Sci., v. 4, Dynamical
9249: systems IV, Springer-Berlin, 1990, 173-280.
9250:
9251:
9252:
9253:
9254:
9255: \bibitem[F1]{F1} Fay J. "Theta functions and Riemann surfaces",
9256: LNM, v 352, Springer-verlag, 1973.
9257:
9258: \bibitem[F2]{F2} Fay J. "Kernel Functions, Analytic Torsion and
9259: Moduli Spaces", AMS, Memoirs, v.98, 464, 1992.
9260:
9261: \bibitem[F3]{F3} Fay J. "The Non-Abelian Szego Kernel and
9262: Theta-Diviser", Contemporary Mathematics, v. 136, 1992, 171-183.
9263:
9264: \bibitem[FS]{FS} Friedan D. and Shenker S. "The analytic
9265: geometry of two-dimensional conformal field thory", Nuccl. Phys.
9266: B 281, 1987, 509.
9267:
9268:
9269: \bibitem[FMN]{FMN} C. Florentino, J. Mourao, J. Nunes,
9270: ``Analytical aspects of the theory abelian theta functions.''
9271: Preprint
9272: 3/2001, February 2001, IST, Lisboa.
9273:
9274:
9275: \bibitem[FMNT]{FMNT} C. Florentino, J. Mourao, J. Nunes and A. Tyurin
9276: ``Analytical aspects of the theory non-Abelian theta functions.''
9277: Preprint IST 2000.
9278:
9279: \bibitem[GH]{GH} Griffits Ph, Harris J. "Principles of Algebraic Geometry",
9280: Wiley-Intersience, New York, 1978
9281:
9282:
9283:
9284:
9285: \bibitem[GT]{GT} A. Gorodentsev and A. Tyurin, "ALAG", Izv. Ross.
9286: Akad. Nauk Ser. Mat. {\bf65}, n. 3, (2001), 15--50; English
9287: transl. in Russian Acad. Sci. Izv. Math.{\bf65}:3 (2001), avalable
9288: as Max-Planck-Institute fur Mathematik, Preprint Series 1999
9289: (130). 1-34.
9290:
9291:
9292: \bibitem[Gu1]{Gu1} Guillemin V.,"Moment maps and combinatorial invariants
9293: of Hamiltonian $T^n$-spaces",Birkhauser (Progress in Mathematics 122), 1994.
9294:
9295: \bibitem[Gu2]{Gu2} V. Guillemin, Symplectic spinors and partial
9296: differential equations, Coll. Inter. C.N.R.S., Aix-en-Provence,
9297: 1974) (1975), 217--252
9298:
9299:
9300:
9301:
9302:
9303: \bibitem[GS1]{GS1} Guillemin V. and Sternberg S. " Geometric asymptotics",
9304: Mathematical surveys, Vol.14, Providence, RI: AMS, 1977.
9305:
9306:
9307: \bibitem[GS2]{GS2} Guillemin V. and Sternberg S. "The Gelfand-Cetlin System and
9308: Quantization of the Complex Flag Manifolds" Journal of func.
9309: analysis, 52, 1983, 106-128. Compos. Math.,49, 173-194.
9310:
9311: \bibitem[G1]{G1} W. Goldman, The symplectic nature of fundamental groups of
9312: surfaces, Adv. in Math. {\bf54} (1984) 200--225
9313:
9314: \bibitem[G2]{G2} Goldman W. "Invariant functions on
9315: Lie groups and Hamiltonian flows of surface group
9316: representations", Invent. Math., 85, 1986, 263-302.
9317:
9318:
9319: \bibitem[Ha]{Ha} B.C. Hall, ``The Segal-Bargmann coherent state
9320: transform for compact Lie groups'', J. Funct. Anal. {\bf 122}
9321: (1994), 103-151.
9322:
9323:
9324: \bibitem[HT]{HT} Hatcher A. and Thurston W. "A presentation for the
9325: mapping class group of closed oriented surface", Topology,
9326: 19 (3),1980,221-237.
9327:
9328: \bibitem[H1]{H1} Hitchin N. "Flat connections and Geometrical Quantization",
9329: Commun.Math.Phys. 131, 1990,347-380.
9330:
9331: \bibitem[H2]{H2} Hitchin N. "Stable bundles and integral systems." Duke
9332: Math. Jour. ,v.54, n.1, !987, 91-114.
9333:
9334: \bibitem[HS]{HS} N. Hitchin and J. Sawon, ``Curvature and characteristic
9335: numbers of hyper\-K\"ahler manifolds'', math.DG/9908114, 17 pp.
9336:
9337:
9338:
9339: \bibitem[JW1]{JW1} L. C. Jeffrey and J. Weitsman, Bohr--Sommerfeld orbits
9340: in the moduli space of flat connections and the Verlinde
9341: dimension formula, Commun. Math. Phys. {\bf150} (1992), 593--630
9342:
9343:
9344: \bibitem[JW2]{JW2} L. C. Jeffrey and J. Weitsman, Half density
9345: quantization of the moduli space of flat connections and Witten's
9346: semiclassical invariants, Topo\-logy {\bf32} (1993), 509--529
9347:
9348:
9349: \bibitem[K1]{K1} Kohno T., "Topological invariants for 3-manifolds using
9350: representation of mapping class group I", Topology 31, 1992,
9351: 203-230.
9352:
9353: \bibitem[K2]{K2} Kohno T., "Topological invariants for 3-manifolds using
9354: representation of mapping class group II. Estimation tunnel
9355: number of knots",
9356: Commentary mathematics 175, 1994, 193-217.
9357:
9358: \bibitem[KZ]{KZ} Kninik V. and Zamolodchikov A. "Current algebra
9359: and Wess-Zumino model in two dimensions", Nucl. Phys. B,247,
9360: 1984, 83.
9361:
9362: \bibitem[KR]{KR} Kirillov A. and Reshetikhin N. "Representation of the algebtra
9363: $U_q(sl(2, \C))$, q-orthogonal polynomials and invariants of links,
9364: Infinite dimensional Lie algebras and groups.", World Scientific, 1988,
9365: 285-342.
9366:
9367: \bibitem[L]{L} Laszlo Yves, "Hitchin's and WZW connections are the same",
9368: J. Differential Geometry, 49, 1998, 547-576.
9369:
9370: \bibitem[MS]{MS} Moore G., and Seiberg N., "Classical and quantum conformal
9371: field theory", Commun. Math. Phys. 123, 1989,177-254.
9372:
9373:
9374: \bibitem[Mum]{Mum} D. Mumford, Tata lectures on theta. I: Progr. Math
9375: {\bf28}, Birkh\"auser (1983). II. Jacobian theta functions and
9376: differential equations: Progr. Math {\bf43}, Birkh\"auser (1984).
9377: III. Progr. Math {\bf97}, Birkh\"auser (1991)
9378:
9379:
9380: \bibitem[M]{M} Margalit D. "The Automorphisms Group of the
9381: Complex of Pant Decompositions", GT/0201319.
9382:
9383:
9384:
9385: \bibitem[MM]{MM} Masur H. and Minsky Y.
9386: "Geometry of the complex of curves I. Hyperbolicity", Invent.
9387: Math., 138 (1), 1999,103-149.
9388:
9389: \bibitem[MP]{MP} M. Mulase, M. Penkava, "Ribbon graphs, quadratic
9390: differentials on Riemann surfaces and algebraic curves defined over $\ov{\Q}$",
9391: math-ph/9811024.
9392:
9393: \bibitem[Mu]{Mu} M. Mulase, "Lectures on the asymptotic expansion of a Hermitian matrix integral", math-ph/9811023.
9394:
9395:
9396: \bibitem[NN]{NN} Newlender, A
9397: and Nirenberg, L. Ann. of Math. 65 (1957), 391-404.
9398:
9399:
9400: \bibitem[Ne]{Ne} Newstead P.E. "Characteristic
9401: classes of stable bundles over an algebraic curve ", Trans. Am.
9402: Math. Soc., 169,1972, 337-345.
9403:
9404:
9405: \bibitem[NR]{NR} Narasimhan M. and Ramanan R. "Deformations of
9406: the moduli space of vector bundles over n algebraic curve", Ann.
9407: of Math., v.101, 1975, 391-417.
9408:
9409:
9410:
9411:
9412: \bibitem[P]{P} Penrose R. "An approach to combinatorial space-time"
9413: in "Quantum Theory and Beyond", ed. T. Bastin, Cambridge Univ.
9414: press, 1971, 151-180.
9415:
9416: \bibitem[RS]{RS} Rovelli C. and Smolin L. "Spin Networks and
9417: Quantum Gravity", gr-qu/9505006.
9418:
9419: \bibitem[R]{R}
9420: ~Rawnslay J. Coherent States And K\"ahler Manifolds, Qart. J. of
9421: Maths. 28, 1997, pp.~404--415.
9422:
9423:
9424:
9425: \bibitem[RSW]{RSW} T. R. Ramadas, L. M. Singer and J. Weitsman, Some comments
9426: on Chern--Simons gauge theory, Commun. Math. Phys. {\bf126}
9427: (1989) 409--420
9428:
9429: \bibitem[Re]{Re} Reid M. "What is a flip?", Preprint Uta, 1993.
9430:
9431: \bibitem[Se]{Se} Seshadri C.S. "Desingurisation des varietes
9432: de modules de fibres semi-stables.", Asterisque, 98, 1982, 110-130.
9433:
9434:
9435:
9436: \bibitem[S]{S} Sorger C., "La vormula de Verlinde", Sem. Bourbaki, 794, 1994.
9437:
9438:
9439: \bibitem[Sc]{Sc} Schlichenmaier M."Sugawara construction for higher
9440: genus Riemann surfaces", Mannheimer Manuskripte, 231, 1998,1-19,
9441: QA/9806032.
9442:
9443: \bibitem[Si]{Si} Singer J. "Three dimensional manifolds and their Heegaard
9444: diagrams", Trans. Am. Math. Soc., 35, 1933, 88-111.
9445:
9446:
9447: \bibitem[Th1]{Th1} Thaddeus M. "Conformal field theory and
9448: the cohomology of the moduli space
9449: of stable bundles", J. Diff. Geom., 35, 1992, 131--149
9450:
9451:
9452: \bibitem[Th2]{Th2} Thaddeus M.
9453: "Stable pairs,linear system and the Verlinde formula",
9454: Invent. Math., 117, 1994, 317--353.
9455:
9456:
9457:
9458: \bibitem[TUY]{TUY} Tsuchiya A., Ueno K. and Yamada Y., "Conformal field theory on iniversal family of stable curves with gauge symmetries", Adv. Stud. Pure Math. 19, 1989, 459-566.
9459:
9460:
9461: \bibitem[TK]{TK} Tsuchiya A. and Kanie, "Vertex operators in
9462: conformal field on $\PP^{1}$ and monodromy representation of
9463: braid group", Advanced Studies in Pure Math. 16, 1988,297-372,
9464: (Errata) ibid. 19, 1990,675-682.
9465:
9466:
9467:
9468: \bibitem[T1]{T1} Tyurin A. "The geometry of moduli of vector bundles,
9469: Uspekhi Mat.Nauk 29:6 (1974), 59-88. Russian Math. Surveys
9470: 29:6 (1974), 57-88."
9471:
9472: \bibitem[T2]{T2} Tyurin A, "Algebraic geometric aspects of
9473: smooth structures. I, Donaldson polynomials", Russian Math.
9474: Surveys, 44:3, 1989, 117-178
9475:
9476:
9477: \bibitem[T3]{T3} Tyurin A. "Six lectures on four manifolds",
9478: Lect.Notes in Math.,1646, 186-246.
9479:
9480: \bibitem[T4]{T4} A. N. Tyurin ``Special Langrangian geometry and slightly
9481: deformed algebraic geometry (geometric quantization and mirror
9482: symmetry).'', Izv. RAN,ser.Math. 64:2, 141-224,
9483: \newline math.AG/9806006, 1-45.
9484:
9485:
9486: \bibitem[T5]{T5} A. Tyurin, ``Three mathematical facets of $\SU(2)$-spin
9487: networks'',\newline math.DG/0011035, 1-20.
9488:
9489:
9490: \bibitem[T6]{T6} A. Tyurin, "On Bohr-Sommerfeld bases, Izv. Ross.
9491: Akad. Nauk Ser. Mat. {\bf64} (2000), 163--196; English transl. in
9492: Russian Acad. Sci. Izv. Math.{\bf64}:5 (2000), avalable as
9493: math.AG/9909084;
9494:
9495: \bibitem[T7]{T7} Andrei Tyurin, Complexification of
9496: Bohr-Sommerfeld conditions, Preprint series, Inst. of Math of
9497: university of Oslo, {\bf 15}, 1999, 1-32.
9498:
9499: \bibitem[T8]{T8} Andrei Tyurin, Delzant model of moduli spaces,
9500: AG/0105216, appear in Russian Acad. Sci. Izv. Math.{\bf65}:3 (2002)
9501:
9502: \bibitem[W]{W} Welters G. Polarized abelian varieties and the heat euqation.
9503: Compos. Math.,49, 173-194.
9504:
9505: \bibitem[Wa]{Wa} Wajnryb B. "A simple presentation for
9506: the mapping class group", Israel J. Math. 45. 1983, 157-174.
9507:
9508:
9509:
9510: \bibitem[We]{We} Weil A. "Generalization des functions abeliennes"
9511: J. de Math. P. et Appl.,17,no.9,1938, 47-87.
9512:
9513: \bibitem[Wi]{Wi} Wirtinger W. "Untersuchengen uber Thetafunctionen"
9514: Teubner, Leipzig, 1895
9515:
9516: \end{thebibliography}
9517: \end{document}
9518: