math0010059/sft.tex
1: 
2: \documentclass[leqno,10pt]
3: {article}
4: \usepackage{amsmath, amsfonts, amssymb}
5: \usepackage{psfig}
6: 
7:  \def\firstpage{1}\def\lastpage{1000}
8: %
9: 
10: %\makeatletter
11: %\def\magnification{\afterassignment\m@g\count@}
12: %\def\m@g{\mag=\count@\hsize6.5truein\vsize8.9truein\dimen\footins8truein}
13: %\makeatother
14: 
15: %%% Choose 10pt/12pt:
16: \oddsidemargin1.66cm\evensidemargin1.66cm\voffset1.2cm%10pt
17: %\magnification1200\oddsidemargin.41cm\evensidemargin.41cm\voffset-.75cm%12pt
18: 
19: \textwidth13.0cm\textheight19.5cm
20: 
21: 
22: \pagestyle{myheadings}
23: \pagenumbering{arabic}
24: \setcounter{page}{\firstpage}
25: 
26:  \def\sign{\mathrm {sign}}
27: \def\Fc{{\mathcal F}}
28: \def\Pc{{\mathcal P}}
29: \def\Cc{\mathcal{C}}
30: \def\a{\alpha}
31: \def\ora{\overrightarrow}
32: \def\ola{\overleftarrow}
33: \def\g{\gamma}
34: \def\d{\delta}
35: \newcommand{\up}{{}^}
36: \newcommand{\ointw}{\int\limits_{|w|=1}}
37: \newcommand\hook{\mathbin{\hbox{\vrule height .5pt width 3.5pt depth 0pt
38: \vrule height 6pt width .5pt depth 0pt}}}
39: \def\ThetaH{\Theta^{H_2}}
40: \newcommand{\qed}{\hfill {$\blacktriangleright$}}
41: \newcommand{\fW}{{\mathfrak W}}
42: \newcommand{\fA}{{\mathfrak A}}
43: \newcommand{\fP}{{\mathfrak P}}
44: \newcommand{\fK}{\mathfrak {K}}
45: \newcommand{\fF}{{\mathfrak F}}
46: \newcommand{\fD}{{\mathfrak D}}
47: \newcommand{\fL}{{\mathfrak L}}
48: \newcommand{\fH}{{\mathfrak H}}
49: \newcommand{\fE}{{\mathfrak E}}
50: \newcommand{\NOV}{{\mathrm {NOV}}}
51: \newcommand{\cont}{{\mathrm {cont}}}
52: \newcommand{\restr}{\mathrm {restr}}
53:  \newcommand{\bx}{{\mathbf x}}
54: \newcommand{\by}{{\mathbf y}}
55: \newcommand{\bz}{{\mathbf z}}
56: \newcommand{\bgg}{{\mathbf g}}
57: \newcommand{\bff}{{\mathbf f}}
58: \newcommand{\bg}{\boldsymbol{\gamma}}
59: \newcommand{\bo}{\boldsymbol{\omega}}
60: \newcommand{\bmu}{\boldsymbol{\mu}}
61: \newcommand{\bphi}{{\mathbf \phi}}
62: \newcommand{\bA}{{\mathbf A}}
63: \newcommand {\bE}{\mathbf{E}}
64: \newcommand{\bC}{{\mathbf C}}
65: \newcommand{\bh}{{\mathbf h}}
66: \newcommand{\bF}{{\mathbf F}}
67: \newcommand{\bG}{{\mathbf G}}
68: \newcommand{\bt}{{\mathbf t}}
69: \newcommand{\bH}{{\mathbf H}}
70: \newcommand{\bK}{{\mathbf K}}
71: \newcommand{\bk}{{\mathbf k}}
72: \newcommand{\bS}{{\mathbf S}}
73: \newcommand{\bV}{ \mathbf {V}}
74:  \newcommand{\CZ}{{\mathrm {CZ}}}
75:  \newcommand{\SFT}{{\mathrm {SFT}}}
76:  \newcommand{\RSFT}{{\mathrm {RSFT}}}
77: \newcommand{\bPhi}{\boldsymbol{\Phi}}
78: \newcommand{\bPsi}{\boldsymbol{\Psi}}
79: \newcommand{\bXi}{\boldsymbol{\Xi}}
80: \newcommand{\bTheta}{\boldsymbol{\Theta}}
81: \newcommand{\btheta}{\boldsymbol{\theta}}
82: \newcommand{\bd}{\boldsymbol{\delta}}
83: \newcommand{\wt}{\widetilde}
84: \newcommand{\cT}{{\mathcal T}}
85: \newcommand{\Mc}{\mathcal {M}}
86: \newcommand{\cM}{\mathcal {M}}
87: \newcommand{\cD}{\mathcal {D}}
88: \newcommand{\cS}{{\mathcal S}}
89: \newcommand{\cK}{{\mathcal K}}
90: \newcommand{\cP}{{\mathcal P}}
91: \newcommand{\cE}{{\mathcal E}}
92: \newcommand{\cF}{{\mathcal F}}
93: \newcommand{\ch}{{\mathcal h}}
94: 
95: \newcommand{\CR}{\mathcal{CR}}
96: \newcommand{\mc}{{\mathcal c}}
97: \newcommand{\ve}{\varepsilon}
98: \newcommand{\Ker}{{\mathrm {Ker}}}
99: %\newcommand{\Im}{{\rm Im}}
100: \newcommand{\cV}{{\mathcal V}}
101: \newcommand{\cN}{{\mathcal N}}
102: \newcommand{\cW}{{\mathcal W}}
103:  \newcommand{\dist}{{\mathrm {dist}}}
104: \newcommand{\Coker}{{\mathrm {Coker}}}
105: \newcommand{\Morse}{{\mathrm {Morse}}}
106: \newcommand{\Int}{{\mathrm {Int}}}
107: %\newcommand{\Ker}{{\mathrm {Ker}}}
108: \newcommand{\Id}{{\mathrm {Id}}}
109: \newcommand{\ind}{\mathrm{ind\,}}
110: 
111: \newcommand{\im}{\mathrm {Im}}
112: \newcommand{\comp}{\mathrm {comp}}
113: \newcommand{\D}{{\rm D}}
114: \newcommand{\TT}{T\Theta_+\Theta_-}
115: \newcommand{\TTH}{\TT^{H_2}}
116: \newcommand{\contr}{{\mathrm {contr}}}
117: \newcommand{\punct}{{\mathrm {punct}}}
118: \newcommand{\even}{{\mathrm {even}}}
119: \newcommand{\cl}{{\mathrm {closed}}}
120:  \newcommand{\proof}{{\sl Proof. }}
121: %
122: \newcommand{\1}{{{\mathchoice {\rm 1\mskip-4mu l} {\rm 1\mskip-4mu l}
123: {\rm 1\mskip-4.5mu l} {\rm 1\mskip-5mu l}}}}
124: \newcommand{\N}{{\mathbb N}}
125: \newcommand{\Z}{{\mathbb Z}}
126: \newcommand{\R}{{\mathbb R}}
127: \newcommand{\Q}{\mathbb {Q}\,}
128: \newcommand{\C}{\mathbb {C}\,}
129: \newcommand{\const}{{\mathrm const}}
130: %
131: \newcommand{\Om}{{\Omega}}
132: \newcommand{\om}{{\omega}}
133: \newcommand{\eps}{{\varepsilon}}
134: \newcommand{\si}{{\sigma}}
135: \newcommand{\ups}{{\upsilon}}
136: \renewcommand{\i}{{\iota}}
137: \newcommand{\p}{{\partial}}
138: \newcommand{\ba}{{\mathbf a}}
139: \newcommand{\bb}{{\mathbf b}}
140: \newcommand{\bc}{{\mathbf c}}
141:  \newcommand{\bX}{{\mathbf X}}
142:  \newcommand{\bY}{{\mathbf Y}}
143:  \newcommand{\DM}{{\overline{\cM}}}
144:    \newcommand{\pr}{\mathrm{pr}}
145:  \newcommand{\ct}{\mathrm{ct}}
146:  \newcommand{\ev}{\operatorname{ev}}
147:  \newcommand{\disk}{\mathrm{disk}}
148: \newcommand{\gt}{{\tau}}
149: \newcommand{\go}{{\omega}}
150: \newcommand{\gd}{{\delta}}
151: \newcommand{\gb}{{\beta}}
152: \newcommand{\w}{{\wedge}}
153: \newcommand{\bs}{ \mathbf {s}}
154: %
155: 
156: 
157:  \newtheorem{theorem}{Theorem}[subsection]
158: \newtheorem{corollary}[theorem]{Corollary}
159: \newtheorem{lemma}[theorem]{Lemma}
160: \newtheorem{proposition}[theorem]{Proposition}
161: \newtheorem{conjecture}[theorem]{Conjecture}
162: \newtheorem{question}[theorem]{Question}
163: \newtheorem{problem}[theorem]{Problem}
164: \newtheorem{criterion}[theorem]{Criterion}
165: \newtheorem{summary}[theorem]{Summary}
166: \newtheorem{remark}[theorem]{Remark}
167: \newtheorem{definition}[theorem]{Definition}
168: 
169: %\theoremstyle{definition}
170: 
171: 
172: \newtheorem{examples}[theorem]{Examples}
173: \newtheorem{example}[theorem]{Example}
174: \newtheorem{exercise}[theorem]{Exercise}
175: 
176: \newtheorem{generalization}[theorem]{Generalization}
177: %\numberwithin{section}{chapter}
178: %$\numberwithin{equation}{chapter}
179: %\numberwithin{figure}{chapter}
180: \begin{document}
181: %%%%% ------------- fill in your data below this line  -------------------
182: %%%%%    The following lines \Title ... \EndAddress must ALL be present
183: %%%%%    and in the given order.
184: \title{Introduction to Symplectic Field Theory}
185: \author{Yakov Eliashberg\\Stanford
186: University \and Alexander Givental \\  UC  Berkeley
187: \and Helmut Hofer\\New York University}
188: \date{October 2000}
189:  \maketitle
190: 
191:   %
192: \begin{abstract}
193: 
194: We sketch in this article  a new theory, which we call {\sl
195: Symplectic Field Theory } or SFT,  which provides an approach to
196: Gromov-Witten invariants of symplectic manifolds and their
197: Lagrangian submanifolds in the spirit of topological field theory,
198: and at the same time
199:  serves as a rich source
200: of new invariants of contact manifolds and their Legendrian
201: submanifolds. Moreover, we  hope that the applications of  SFT
202: go far beyond this framework.\footnote{The research is   partially
203: supported by the National Science Foundation.}
204: \end{abstract}
205: 
206:  \tableofcontents
207:  \bigskip
208: 
209:  \centerline{\bf Disclosure}
210:  \medskip
211: 
212: Despite its length, the current paper presents only a very sketchy
213: overview of  Symplectic Field Theory. It contains practically no proofs,
214: and in a few places where the proofs  are given
215: their role is    just to
216: illustrate the involved ideas,  rather than to give  complete rigorous
217: arguments.
218: 
219:  The ideas, the algebraic formalism, and some of the applications
220:  of this new theory were presented  and popularized
221:  by the authors at several conferences
222:  and seminars (e.g. \cite{Eliash-ICM}). As a result,
223:    currently there exists a  significant
224:  mathematical community which is in some form familiar with
225:    the subject.
226:   Moreover, there are
227:   many mathematicians, including several former
228:    and current students of the authors,
229:   who are actively working  on foundational aspects
230:   of the theory and its applications, and    even
231:   published papers on this subject. Their results
232:     show that already the simplest versions of the theory
233:   have some remarkable corollaries (cf. \cite{Ustilovsky}).
234:    We hope that   the present  paper will help
235:    attracting even more people
236:    to   SFT.
237: 
238:    Of course, our ideas give just a small new twist
239:     to many other active directions of research in
240:     Mathematics and Physics
241:     (Symplectic topology, Gromov-Witten invariants and quantum cohomology,
242:     Floer homology theory, String theory,
243:      just to mention few), pioneered by  V.I. Arnold,
244:      C.~Conley-E.~Zehnder,
245:       M.~Gromov, S.K. Donaldson, E.~Witten, A.~Floer,
246:         M.~Kontsevich and others
247:      (see \cite{Arnold-conjecture, Conley-Zehnder,
248:      Gromov-holomorphic, Donaldson-invariants, Floer, Ruan, Witten1, Witten2, Kontsevich-CP2,
249:      Konts-Manin}).
250:      Many people independently contributed results and ideas,
251:       which may be considered as  parts of  SFT.
252:       Let us just mention here the work of  Yu.Chekanov
253:       \cite{Chekanov},
254:        K.Fukaya--K.Ono--Y.-G.Oh--H.Ohta  \cite{FKO3},
255:        A.~Gathmann  \cite{Gathmann}, E.Ionel--T.Parker
256:       \cite{Ionel-Parker, Ionel}  , Y.Ruan--A.-M.Li  \cite{Ruan-Li}.
257:       It also  draws on  other results of   the  current authors
258:       and their coauthors (see \cite{EHS, EH-3ball, Givental, Givental-Maslov,
259:       Givental-toric, Hofer-Weinstein, Hofer-Wysocki-Zehnder, AbbasHofer,
260:       HoferWysockiZehnder:94d, HoferWysockiZehnder:99}). The contact-geometric
261:       ingredient of our work is greatly motivated by two
262:       outstanding conjectures in contact geometry: Weinstein's conjecture about periodic
263:       orbits of Reeb fields  \cite{Weinstein},
264:        and  Arnold's chord conjecture \cite{Arnold-steps}.
265: 
266: 
267:    Presently, we are working on a series of papers devoted
268:    to the foundations,
269:    applications,  and further development of  SFT.  Among the applications,
270:    some of which are mentioned in this paper, are new invariants of contact manifolds
271:    and Legendrian knots and links, new methods for computing Gromov-Witten invariants,
272:    new restrictions on the topology of Lagrangian submanifolds, new non-squeezing type
273:    theorems in contact geometry etc. We are expecting new links with the low-dimensional topology
274:    and, possibly, Physics. It seems,
275:    however,
276:    that what we see at the moment is just a tip of an iceberg.
277:    The main   body of Symplectic Field Theory  and its applications is yet to be discovered.
278: 
279:      \bigskip
280:      \noindent{\bf Guide for an impatient reader.}
281:      The paper consists of two parts. The first part, except Section \ref{sec:Floer}
282:      and the end of
283:        Section \ref{sec:orientation},
284:      contains some  necessary   background symplectic-geometric and analytic information.
285:      An impatient reader can try to begin reading with Section
286:      \ref{sec:Floer},and use
287:      the rest of the first part for  the references.
288:      The second part begins with its own introduction
289:      (Section \ref{sec:informal})  where we present a very rough sketch of SFT. At the end of Section
290:          \ref{sec:informal} we describe the plan of the remainder of the paper.
291: 
292:    \bigskip
293:    \noindent{\bf Acknowledgements.} The authors benefited a lot from discussions
294:     with many mathematicians, and from the ideas which they corresponded to
295:     us. We are especially grateful to C. Abbas,  P. Biran, F. Bourgeois, K. Cieliebak,
296:     T. Coates, T. Ekholm, K. Fukaya, E. Getzler,
297:     M. Hutchings, E.-N.
298:     Ionel, V.M. Kharlamov,
299:     K. Mohnke, L. Polterovich, D. Salamon, M. Schwarz, K. Wysocki and E. Zehnder.
300:       A  part of the paper
301:     was written,  when  the first author visited
302:     RIMS at Kyoto University.
303:      He wants  to thank  RIMS  and  K. Fukaya and K. Saito,
304:     the organizers of a special program
305:     in Geometry and String Theory,   for the  hospitality. He also thanks
306:     T. Tsuboi and J.-L. Brylinski for organizing   cycle of lectures on  SFT at
307:      the    University of
308:     Tokyo and the Pennsylvania State University. The second author is thankful to
309:      A. Kirillov, R. Donagi, and
310:      the Department of Mathematics of the University
311:     of Pennsylvania  for the possibility to present   SFT in a series of lectures.
312:      The third author acknowledges the hospitality of   FIM at the ETH Z\"urich, where   some
313:      of the work was
314: carried
315: out. Our special thanks to N.M. Mishachev for drawing the
316:     pictures and to J. Sabloff for
317:     the computer verification of the formula
318:     (\ref{eq:CP2-final}).
319: 
320: 
321: \section{Symplectic  and analytic setup}
322: 
323: 
324: \subsection{Contact preliminaries}\label{sec:prelim}
325: A $1$-form $\a$ on a $(2n-1)$-dimensional manifold $V$
326: is called {\it contact} if the
327: restriction of $d\a$ to  the    $(2n-2)$-dimensional tangent
328: distribution $\xi =
329: \{\a = 0\}$ is non-degenerate (and hence symplectic).
330: A  codimension $1$ tangent distribution $\xi$ on $V$ is called
331:  a {\it contact structure} if
332: it can be locally (and in the co-orientable case globally) defined
333: by  the Pfaffian equation $\a
334: = 0$ for some choice of a contact form $\a$.  The pair  $(V,\xi)$   is called
335: a {\it contact manifold}.  According to  Frobenius' theorem the contact
336: condition is a condition of maximal
337:  non-integrability of the tangent hyperplane field $\xi$. In particular, all integral
338: submanifolds of $\xi$ have
339: dimension $\le n-1$. On the other hand, $(n-1)$-dimensional integral submanifolds, called
340: {\it Legendrian}, always exist in abundance.    We will be dealing in this paper
341: only with co-orientable, and moreover co-oriented contact structures.
342: Any non-coorientable contact structure can be canonically double-covered
343:  by a coorientable one.
344: If a contact form $\a$ is fixed then one can associate with it the {\it Reeb vector field}
345: $R_\a$, which is transversal
346: to the contact structure $\xi=\{\alpha=0\}$.
347: The field $R_\a$ is uniquely determined by the equations
348: $R_\a\hook d\a=0;\,\a(R_\a)=1\,.$   The flow of $R_\a$ preserves the contact form $\a$.
349: 
350:  The  $2n$-dimensional manifold $M=(T(V)/\xi)^*  \setminus V$ ,  called the
351: {\it symplectization} of $(V,\xi)$, carries a natural
352: symplectic structure $\omega$ induced by an embedding $M\to  T^*(V)$
353:  which assigns to each linear form $T(V) / \xi \to \R$ the
354: corresponding form $T(V) \to T(V)/\xi \to \R$. A choice of a contact
355: form $\a$ (if $\xi$ is co-orientable) defines a splitting $M=V\times(\R\setminus 0)$. As     $\xi$ is
356:  assumed  to be co-oriented  we can
357:  pick
358:  the positive half  $V\times\R_+$ of  $M$, and call it symplectization
359: as well.  The symplectic structure $\omega$ can be written in terms of this splitting as
360: $d(\tau\a), \tau>0$. It will be more convenient for us, however, to use additive notation
361: and write $\omega$ as $d(e^t\a),\,t\in\R$, on $M=V\times\R$.
362: Notice that the vector field $T=\frac{\p}{\p t}$ is conformally symplectic:
363:  we have ${\cal L}_T\omega=\omega$, as well as ${\cal L}_T
364:  (e^t\a)=e^t\a$,
365:  where ${\cal L}_T$ denotes the Lie derivative along the vector
366:  field $T$.
367: All the notions of contact geometry can be formulated as the
368: corresponding symplectic notions, invariant or equivariant with
369: respect to this conformal action. For instance, any contact
370: diffeomorphism of $V$ lifts to an equivariant symplectomorphism of
371: $M$;  contact vector fields on $V$ (i.e. vector fields preserving
372: the contact structure) are   projections of $\R$-invariant
373: symplectic (and automatically Hamiltonian) vector fields on $M$;
374: Legendrian submanifolds in $M$ correspond to cylindrical (i.e.
375: invariant with respect to the $\R$-action) Lagrangian submanifolds
376: of $M$.
377: 
378:  Notice that the Hamiltonian vector field on $V\times\R$, defined by the Hamiltonian function
379: $H=e^t$ is invariant under translations $t\mapsto t+c$, and projects to the Reeb vector field
380: $R_\alpha$ under the projection $V\times\R\to \R$.
381: 
382: The symplectization of a contact manifold is an example of a
383: symplectic manifold with {\it cylindrical} (or rather conical)
384: ends. We mean by that    a possibly non-compact symplectic
385: manifold $(W,\omega)$ with    ends of the form
386: $E^+=V^+\times[0,\infty)$ and  $E^-=V^-\times(-\infty,0]$, such
387: that  $V^\pm$ are compact manifolds, and
388: $\omega|_{V^\pm}=d(e^t\alpha^\pm)$, where $\a^\pm$ are  contact
389: forms on $V^\pm$. In other words,  the ends  $E^\pm$ of
390: $(W,\omega)$ are symplectomorphic, respectively,
391:   to the positive or negative halves of the
392: symplectizations of    contact manifolds $(V^\pm,\xi^\pm=\{\a^\pm=0\})$.
393: We will consider   the splitting of the ends and the  the contact forms
394: $\a^\pm$ to be parts of the structure of a symplectic manifold with cylindrical ends.
395: We will also call
396: $(W,\omega)$ a {\it directed symplectic cobordism} between
397:  the contact
398: manifolds $(V^+,\xi^+)$ and $(V^-,\xi^-)$, and denote it  by $\ora{V^-V^+}$.
399: 
400: Sometimes we will have to consider the compact part
401: $W^0=W\setminus(\Int E^+\cup\Int E^-)$ of a directed symplectic
402: cobordism $\ora{V^-V^+}$.  If it is not clear from the context we
403: will refer to $W^0$   as a {\it compact}, and to $W$ as a {\it
404: completed} symplectic cobordism.
405: 
406: Let us point out that ``symplectic cobordism" {\it is not an
407: equivalence relation, but rather a partial order. }
408: Existence of a directed
409: symplectic cobordism $\ora{V^-V^+}$  does not imply
410: the existence of a directed symplectic cobordism   $\ora{V^+V^-}$,
411: even if one does not fix contact forms for the contact structures $\xi^\pm$. On the other
412: hand,
413: directed
414: symplectic cobordisms $\ora{V_0V_1}$ and $\ora{V_1V_2}$ can be glued, in an obvious
415: way, into a directed symplectic cobordism  $\ora{V_0V_2}=\ora{V_0V_1}
416: \circledcirc\ora{V_1V_2}$.
417: 
418: 
419: 
420: Contact  structures have no local invariants. Moreover, any contact form is locally
421: isomorphic to the
422: form $\alpha_0= dz-\sum\limits_1^{n-1}y_idx_i$ (Darboux' normal form).
423: The contact structure $\xi_0$ on $\R^{2n-1}$ given by the form $\a_0$ is called {\it
424: standard}.
425:  The standard contact structure on $S^{2n-1}$ is formed by complex tangent hyperplanes to the
426: unit sphere in $\C^n$. The standard contact structure on $S^{2n-1}$
427: is isomorphic in the complement of a point to the standard contact structure  on
428: $\R^{2n-1}$.
429: According to a theorem of J. Gray (see \cite{Gray}) contact structures
430:  on closed
431: manifolds   have  the following stability property:
432: {\it Given a family  $\xi_t$, $t\in[0,1]$, of contact  structures on  a closed manifold
433: $M$, there exists
434: an isotopy $f_t:M\to M$,
435: such that $df_t(\xi_0)=\xi_t; t\in [0,1]$.}
436: Notice that for contact {\it forms} the analogous statement is  wrong. For
437: instance,
438: the topology of the $1$-dimensional foliation determined by the  Reeb vector field $R_\a$
439: is very sensitive to deformations of the contact form $\a$.
440: 
441: The conformal class of the symplectic form $d\a|_\xi$ depends only on
442: the cooriented contact structure $\xi$ and not on the
443: choice of the contact form $\a$. In particular, one can associate with $\xi$
444:  an almost complex structure $J:\xi\to\xi$, compatible with $d\a$
445: which means that $d\alpha(X,JY); X,Y\in\xi,$ is an Hermitian metric on $\xi$.
446: The space of   almost complex structures $J$ with this property is contractible, and hence
447:  the choice of $J$ is homotopically canonical.
448: Thus a  co-oriented contact structure  $\xi$ defines on $M$ a {\it
449:  stable almost complex
450: structure}
451:  $\widetilde J=\widetilde J_\xi$,
452: i.e. a splitting of
453: the tangent bundle $T(V)$ into the Whitney sum of a complex bundle of (complex) dimension
454: $(n-1)$ and a trivial $1$-dimensional real bundle.
455: The existence of a
456:  stable almost complex
457: structure  is  necessary for  the existence of a contact structure on $V$. If $V$
458: is open (see \cite {Gro-PDR}) or ${\mathrm {dim}}\,V=3$ (see \cite{Martinet,Lutz})
459:  this property is also sufficient for  the existence of a contact structure in the
460: prescribed homotopy
461: class. It  is still unknown
462: whether this condition
463: is sufficient for  the existence of a contact structure on a closed manifold of dimension
464: $>3$. However,   a positive answer  to this question is extremely unlikely.
465: The homotopy class of $\widetilde J_\xi$, which we denote by
466: $[\xi]$ and call the {\it formal }homotopy class  of $\xi$,
467: serves as an
468: invariant of $\xi$.
469: For an open $V$ it is a complete invariant (see \cite {Gro-PDR})   up to   homotopy
470: of contact structures,
471: but not up to a contact diffeomorphism. For closed manifolds this is known to be false
472: in many, but not all dimensions. The theory discussed in this
473: paper serves as a rich source of contact invariants, both of
474: closed and open contact manifolds.
475: 
476: \subsection{Dynamics of Reeb vector fields}\label{sec:dynamics}
477: 
478: Let $(V,\xi)$ be a $(2n-1)$-dimensional manifold with  a
479: co-orientable contact structure with a fixed contact form $\alpha$.
480:  For a generic choice of $\a$ there are only countably many periodic trajectories
481:   of the
482: vector field $R_\a$.  Moreover, these trajectories can be assumed {\it non-degenerate}
483: in the sense  that the linearized Poincar\'e return map $A_\g$
484:  along any  closed
485:  trajectory $\g$, including multiples,
486: has no eigenvalues equal to $1$.  Let  us denote by $\Pc=\Pc_\a$
487:   the set of  all periodic trajectories of $R_\a$, including multiples.
488:   \footnote{As it is explained below in Section \ref{sec:orientation}
489:   the orientation issues require us to exclude certain multiple periodic orbits out
490:   of consideration. Namely, let us recall that real eigenvalues of symplectic matrices
491:   different from $\pm 1$ come in pairs $\lambda,\lambda^{-1}$.
492:  Let $\g\in \Pc$ be a simple periodic orbit and $A_\g$ its linearized Poincar\'e return map.
493: If
494:   the total multiplicity of   eigenvalues of $A_\g$ from the interval
495:   $(-1,0)$ is odd, then we   exclude from $\Pc$ all even
496:   multiples of $\g$.}
497: 
498: The reason for a such choice
499: is discussed in Section \ref{sec:orientation} below. We will also fix a point
500: $m_\gamma$ on each {\it simple} orbit from $\Pc$.
501: Non-degenerate trajectories can be divided into {\it odd} and {\it even} depending
502: on the sign of the Lefshetz number ${\mathrm {det}}  (I- A_\g)$.
503: Namely, we call $\g$ odd if $  {\mathrm {det}}  (I-  A_\g)<0$, and even otherwise.
504: The parity of  a periodic orbit $\g$  agrees   with the parity of a certain
505: integer grading  which is defined   if certain additional choices are made, as
506: it is described below.
507: %%%%
508: 
509: If $H_1(V)=0$
510:  then for each $\g\in\Pc$ we can choose  and
511:  fix a surface $F_{\g}$ spanning
512: the trajectory $\g$ in $V$.
513:   We will allow the case $H_1(V)\neq 0$,
514:    but will require in most of the paper that the torsion part is
515:    trivial
516:    \footnote{ The case when $H_1(V)$ has torsion elements is
517:    discussed in Section \ref{sec:torsion} below.}.
518: In this case   we choose a basis of $H_1(V)$,
519:   represent    it by oriented curves ${C}_1,\dots,{C}_K$, and    choose
520:   a symplectic trivialization of the bundle $\xi|_{{C}_i}$ for each chosen curve.
521:   We recall that the bundle $\xi$ is endowed with the symplectic
522:   form $d\alpha$ whose   conformal class depends only on $\xi$.
523:        For  any periodic orbit $\gamma\in\Pc$
524:    let us choose a surface $F_\gamma$ with $[\partial F_\gamma]=
525:    [\gamma]-\sum n_i[{C}_i]$.
526:   The coefficients $n_i$ are uniquely defined because of our  assumption that $H_1(V)$
527:   is
528:   torsion-free.
529: 
530: 
531: 
532: The above choices enable us to define the {\it Conley-Zehnder
533: index} $\CZ(\g)$ of  $\g$ as follows. Choose a homotopically
534: unique trivialization of the symplectic vector bundle $(\xi,d\a)$
535: over each trajectory $\g\in\Pc$ which extends to $\xi|_{F_{\g}}$
536: (and coincides with a chosen trivialization of $\xi|_{{C}_i}$
537: if ${C}_i$ is not homologically trivial). The linearized flow
538: of $R_\a$ along $\g$ defines then a path in the group
539: $Sp(2n-2,\R)$ of symplectic matrices, which begins at the unit
540: matrix and ends at a matrix with all eigenvalues different from
541: $1$. The Maslov index of this path (see \cite{Arnold-Maslov,
542: Robbin-Salamon}) is, by  the definition, the Conley-Zehnder index
543: $\CZ(\g)$ of the trajectory $\g$. See also
544:  \cite{HoferWysockiZehnder:94d}, Section 3, for an axiomatic
545:         description of the Conley-Zehnder index using our normalization
546:         conventions.
547: 
548: Notice that by changing the spanning surfaces for the trajectories
549: from $\Pc$ one can change Conley-Zehnder indices by  the value of
550: the cohomology class $2c_1(\xi)$, where $c_1(\xi)$ is the first
551: Chern class of the contact bundle $\xi$. In particular, ${\mathrm
552: {mod}}\,2$ indices can be defined independently of any spanning
553: surfaces, and even in the case when $H_1(V)\neq 0$. In fact,
554: $$(-1)^{\CZ(\g)}=(-1)^{n-1}\sign \left({\mathrm
555: {det}}(I-A_\gamma)\right).$$
556: 
557: 
558: 
559: 
560: 
561: 
562:   \subsection{Splitting of a symplectic manifold along \\
563:   a contact submanifold}
564: \label{sec:splitting} Let $V$ be a hypersurface of contact type,
565: or in a different terminology, a symplectically convex
566: hypersurface  in a symplectic manifold $(W,\omega)$. This means
567: that $\omega$ is exact, $\omega=d\beta$, near $V$, and the
568: restriction $\alpha=\beta|_V$ is a contact form on $V$.
569: Equivalently, one can say that the conformally symplectic vector
570: field $X$, $\omega$-dual to $\beta$, is transversal to $V$. Let us
571: assume that $V$ divides $W,\, W=W_+\cup W_-$, where
572:  the notation of the parts are chosen in such a way that   $X$
573: serves as an inward transversal for $W_+$, and an outward
574: transversal for $W_-$. The manifolds $W_\pm$ can be viewed as
575: compact directed symplectic cobordisms such that $W_-$ has only positive
576: contact boundary $(V,\alpha)$, while the same contact manifold
577: serves as a negative boundary of $W_+$.
578: 
579: Let $$(W_-^\infty,\omega_-^\infty)=(W_-,\omega)\cup\left(V\times[0,\infty),d(e^t\alpha)\right)$$ and
580: $$(W_+^\infty,\omega_+^\infty)=\left(V\times(-\infty,0],d(e^t\alpha)\right)\cup(W_+,\omega)$$
581:  be the  completions, and
582: $$(W_-^\tau,\omega_-^\tau)=(W_-,\omega)\cup\left(V\times[0,\tau],d(e^t\alpha)\right)$$ and
583: $$(W_+^\tau,\omega_+^\tau)=\left(V\times[ -\tau,0],d(e^t\alpha)\right)\cup(W_+^0,\omega)$$
584:   {\it partial completions} of
585: $W_\pm$. Let us observe that the symplectic manifolds
586: $$(W_-,e^{-\tau}\omega_-^\tau),\quad (V\times [-\tau,\tau],
587: d(e^t\alpha))\quad\hbox{and}\quad (W_+,e^\tau\omega_+^\tau)$$ fit
588: together into a symplectic manifold $(W^\tau, \omega^\tau)$, so
589: that $W^0=W$. Hence when $\tau\to\infty$ the deformation
590: $(W^\tau,\omega^\tau)$ can be viewed as a decomposition of the
591: symplectic manifold $W=W^0$ into the union   of two completed
592: symplectic cobordisms  $W^\infty_+$ and $ W^\infty_-$ . We will
593: write $W=W_-\circledcirc W_+$ and also
594: $W^\infty=W_-^\infty\circledcirc W_+^\infty$.
595: %%%%%%%%%%%%%%%%
596: 
597: \begin{figure}
598: \centerline{\psfig{figure=sf3.eps,height=60mm}}
599:  \caption{\small Splitting of a closed symplectic manifold $W$
600:  into two completed symplectic cobordisms $W_-^\infty$ and $W^\infty_+$}
601: \label{fig:splitting}
602: \end{figure}
603: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
604:  Let us give here two important examples of the above
605:  splitting construction.
606: 
607:  \begin{example}\label{ex:cotangent}
608:  {\rm Suppose $L\subset W$ is a Lagrangian submanifold.
609:   Its neighborhood  is symplectomorphic to a
610:    neighborhood of the $0$-section  in the cotangent bundle
611:    $T^*(L)$. The boundary $V$ of an  appropriately chosen neighborhood
612:    has  contact type, and thus we can apply    along $V$
613:    the  above splitting construction. As the result we split
614:    $W$ into  $W_+^\infty$ symplectomorphic
615:    to $W\setminus L$, and $W_-^\infty$ symplectomorphic to $T^*(L)$.
616:     }
617:     \end{example}
618:     \begin{example}\label{ex:section}
619:         {\rm Let $M$ be a hyperplane section of a K\"ahler manifold
620:    $W$, or
621:    more generally a symplectic hyperplane section of a
622:    symplectic manifold
623:    $W$, in the sense of Donaldson (see \cite{Donaldson}). Then $M$ has a neighborhood
624:    with a contact boundary $V$. The affine part
625:    $W\setminus M$ is a Stein manifold in the K\"ahlerian case,
626:    and in any case
627:    has a structure of   a symplectic Weinstein
628:    manifold $\widetilde W$ (notice that the symplectic structure
629:    of $\widetilde W$ does not coincide with  the induced
630:    symplectic structure
631:    on $W\setminus M$ but contains $W\setminus M$ as an
632:    open symplectic
633:    submanifold).
634:    The Weinstein manifold $W\setminus M$ contains
635:    an isotropic  deformation retract  $\Delta$. The splitting of $W$ along $V$
636:    produces $W^\infty_-$ symplectomorphic to
637:     $\widetilde W,$ and $W^\infty_+$
638:     symplectomorphic to $W\setminus \Delta$.
639:     If $\Delta$ is a smooth Lagrangian submanifold, then
640:     we could get the same decomposition by
641:     splitting along the boundary of a tubular neighborhood
642:     of $L$, as in Example \ref{ex:cotangent}.  }
643:               \end{example}
644: 
645: \subsection{Compatible almost complex structures}\label{sec:almost}
646: 
647: According to M. Gromov (see \cite{Gromov-holomorphic}) an almost
648: complex structure $J$ is called {\it tamed} by a  symplectic form
649: $\omega$ if $\omega$ is positive on complex lines. If, in
650: addition,  one adds the calibrating condition that $\omega$ is
651: $J$-invariant, then $J$ is   said {\it compatible} with $\omega$.
652: For symplectic manifolds with cylindrical ends one needs further
653: compatibility conditions at infinity, as it is described below.
654: 
655: At each positive, or negative end $\left(V\times\R_\pm,d(e^t\alpha
656: )\right)$ we require   $J$ to be invariant with respect to
657: translations $t\mapsto t \pm c$, $c>0$ at least for sufficiently
658: large $t$. We also require   the contact structure
659: $\xi^\pm|_{V\times t}$ to be invariant under $J$, and define
660: $J\frac{\partial}{\partial t}=R_\alpha$, where
661:  $R_\alpha$ is
662:  the Reeb vector field (see \ref{sec:dynamics} above)
663:   of the contact form $\alpha$. In the case when
664: $W=V\times\R$ is the symplectization of a manifold $V$, i.e. $W$ is    a
665: cylindrical manifold, we additionally  require   $J$ to be globally
666: invariant under all translations along the second factor.
667: 
668: To define a compatible almost complex structure $J$ in the above
669: Examples \ref{ex:cotangent} and \ref{ex:section}
670:  one needs to specify a contact form $\alpha$ on the contact manifold
671:  $V$. In the case of the boundary of a tubular neighborhood    of a Lagrangian
672:  submanifold $L$ a natural choice of a contact form is provided by
673:  a  Riemannian metric on $L$. The Reeb vector field for such a
674:  form $\alpha$ generates on $V$ the geodesic flow of the metric.
675: 
676:  When $V$ is the boundary of a neighborhood of a hyperplane section  $M$
677:  then there exists another good choice of a contact form. It is
678:  a $S^1$-invariant connection form $\a$ on the principal $S^1$-bundle
679:  $V\to M$, whose curvature equals the symplectic form $\omega|_M$.
680:  The contact manifold $(V,\xi=\{\a=0\})$ is called the {\it
681:  pre-quantization} of the symplectic  manifold $(M,\omega)$.
682:  Orbits of the  Reeb field $R_\a$ are all closed and coincide
683:  with the fibers of the fibration, or their multiples. Notice that though
684:  the Reeb flow in
685:  this case looks extremely nice and simple, all its  periodic orbits are
686:  highly degenerate, see Section \ref{sec:Bott} below.
687:  Notice that  the symplectization $W$  of $V$
688:  can be viewed  as the total space of a complex  line bundle
689:  $L$ associated with
690: the $S^1$-fibration $V\to M$ with the zero-section removed.
691: It is possible
692:  and convenient to choose
693: $J$  compatible with the structure of this bundle,
694: and in such a way that the projection   $W\to M$
695: becomes holomorphic with respect to a certain almost
696: complex structure $J_M$ on $M$
697: compatible with $\omega$.
698: 
699: 
700:    Let us describe now what the symplectic splitting
701:    construction from  Section \ref{sec:splitting}
702:    looks like from the point of view of  a compatible almost
703:    complex structure.
704: 
705:    First, we assume that the original almost complex structure
706:    $J$ on $W$ is chosen in such a way
707:    that     the contact structure $\xi=\{\alpha=0\}$ on $V$ consists
708:    of complex tangencies to $V$, and that $JX=R_\alpha$, where
709:    $X$ is a conformally symplectic vector field, $\omega$-dual
710:    to $\alpha$, and $R_\alpha$ is the Reeb vector field of $\alpha$.
711:    Next we define an almost complex structure
712:    $J^\tau$ on $W^\tau=W_-\cup V\times[-\tau,\tau]\cup W_+$ by setting
713:    $J^\tau|_{W_\pm}=J$ and requiring  $J^\tau$ to be independent
714:    of $t\in[-\tau,\tau]$ on  $V\times[-\tau,\tau]$.
715:    When $\tau\to\infty$  the almost complex structure
716:    $J_-^\tau$ on
717:    $W_-^\tau=W_- \cup V\times[-\tau,\tau]$   converges
718:     to an almost complex
719:    structure $ J^\infty_-$ on $W^\infty_- $ compatible   with
720:    $\omega^\infty_-$, and  $J_+^\tau$ on
721:    $W_+^\tau=V\times[-\tau,\tau]\cup W_+$   converges
722:     to an almost complex
723:    structure $ J^\infty_+$ on $W^\infty_+$ compatible   with
724:    $\omega^\infty_+$.
725: 
726:    \subsection{Holomorphic curves in symplectic
727:    cobordisms}\label{sec:holomorphic}
728: 
729: 
730:   Let $(V,\alpha)$ be a contact manifold with a fixed contact form
731: and $(W=V\times\R,\omega=d(e^t\alpha))$ its symplectization. Let
732: us denote by $\pi_\R$ and $\pi_V$ the projections $W\to\R$ and $W\to
733: V$, respectively. For a map $f:X\to W$ we write $ f_\R$ and $f_V$
734: instead of $\pi_\R\circ f$ and $\pi_V\circ f$.
735: 
736: Notice that given a trajectory $\gamma$ of the Reeb field
737: $R_\a$,
738: the cylinder $\R\times\gamma\subset W$ is a $J$-holomorphic curve.
739:  Let us also observe that
740: \begin{proposition}\label{prop:positivity}
741:   For a $J$-holomorphic curve  $C\subset W$
742:  the restriction $d\a|_C$ is non-negative, and if $d\a|_C\equiv 0$
743:  then $C$ is a (part of a) cylinder $\R\times\gamma$ over a
744:  trajectory $\gamma$ of the Reeb field $R_\a$.
745:   \end{proposition}
746: 
747:  Given a $J$-holomorphic map $f$ of a punctured disk $D^2\setminus
748:  0\to W$ we say that  the map $f$ is {\it asymptotically
749:  cylindrical } over a periodic orbit $\g$ of the Reeb field
750: $R_\a$ at $+\infty$ (resp. at $-\infty$) if
751: $\mathop{\lim}\limits_{r\to 0} f_\R(re^{i\theta})=+\infty$ (resp.
752: $=-\infty$), and $\mathop{\lim}\limits_{r\to
753: 0}f_V(re^{i\theta})=\bar f(\theta)$, where the map $\bar
754: f:[0,2\pi]\to V$ parameterizes the trajectory $\gamma$.
755: 
756:  %%%%%
757:  \bigskip
758: 
759:  The almost complex manifold $(W,J)$ is bad from the point of view of
760: the theory of
761: holomorphic curves: it has a {\it pseudo-concave} end
762: $V\times(-\infty,0)$, or using Gromov's terminology its geometry
763: at this end is not bounded. However, it was shown in
764: \cite{Hofer-Weinstein} that   Gromov compactness theorem can be
765: modified to accommodate this situation, see Theorems
766: \ref{thm:comp1} and \ref{thm:comp2}
767: below. We will mention in this
768: section only the following fact related to compactness, which motivates the usage
769: of holomorphic curves asymptotically cylindrical over orbits from
770: $\Pc_\a$.
771: \begin{proposition}
772:  Suppose that all periodic orbits of the Reeb field $R_\alpha$ are
773: non-degenerate.
774:  Let  $C$ be  a non-compact Riemann surface  without boundary and $f:C\to W$
775:    a proper $J$-holomorphic curve. Suppose that there
776: exists a constant $K>0$ such that
777: $\int\limits_{C}f^*d\a  < K$.
778: Then
779:  $C$ is conformally equivalent to a compact Riemann surface $S_g$
780:  of genus $g$ with $s^++s^-$ punctures
781:  $$ x^+_1,\dots,
782:  x^+_{s^+},x^-_1\dots,x^-_{s^-}\in S_g,$$
783:   such that near the punctures ${\bx^+}=(x^+_1,\dots,x^+_{s^+})$
784:   the map
785:  $f$ is asymptotically cylindrical over periodic orbits
786:   ${\Gamma^+}=\{\g^+_1,\dots,\g^+_{s^+}\}$ at $+\infty$, and near
787:  the punctures ${\bx^-}=\{x^-_1,\dots,x^-_{s^-}\}$ the map
788:  $f$ is asymptotically cylindrical over periodic orbits
789:   ${\Gamma^-}=\{\g^-_1,\dots,\g^-_{s^-}\}$ at $-\infty$.
790: \end{proposition}
791: 
792: 
793:  %%%%%
794:  Thus holomorphic maps of punctured Riemann surfaces,
795:    asymptotically cylindrical over periodic
796: orbits of the Reeb vector field $R_\alpha$, form a natural class of
797: holomorphic curves to consider in symplectizations
798: as well as  more general symplectic manifolds with cylindrical ends.  We
799: will define now moduli spaces of such curves.
800: 
801: Let $W =\ora{V^-V^+}$ be a (completed) directed cobordism,
802: $\alpha^\pm$    corresponding contact forms on $V^-$ and $V^+$,
803:  $\Pc^\pm$  the sets of all
804: periodic orbits (including multiple ones) of the Reeb vector
805: fields
806: $R_{\alpha^\pm}$. We assume that $\alpha^\pm$ satisfies
807: the genericity assumptions from Section \ref{sec:dynamics}.
808: Choose a compatible almost complex  structure  $J$   on $W$.
809:  Let $\Gamma^\pm$ be   ordered sets of trajectories from
810:  $\Pc^\pm$ of cardinality $s^\pm$.
811:  We also assume that every {\it simple} periodic orbit $\gamma$ from
812:  $\Pc^\pm$ comes with a fixed marker
813:  $m_\gamma\in\g$.
814: 
815:  Let $S=S_g$ be a compact Riemann surface of genus $g$ with a conformal
816:  structure $j$, with $s^+$ punctures $\bx^+=\{x^+_1,\dots,x^+_{s^+}\}$, called
817:  positive,
818:  $s^-$ punctures $\bx^-=\{x^-_1,\dots,x^-_{s^-}\}$, called negative, and
819:  $r$ marked points $\by=\{y_1,\dots,y_r\}$.
820:   We will also   fix an {\it asymptotic marker} at each
821:    puncture. We mean by that  a ray originating at each
822:    puncture.  Alternatively, if one
823:    takes the cylinder $S^1\times[0,\infty)$ as a conformal model of
824:    the punctured disk $D^2\setminus 0$ then an asymptotic marker
825:    can be viewed as a point on the circle at infinity.
826:    If a holomorphic map $f:D^2\setminus 0\to V^\pm\times\R_\pm$ is asymptotically
827:    cylindrical over a periodic orbit $\g$,  we say that
828:     a marker $\mu=\{\theta=\theta_0\}$
829:     is mapped by $f$ to the
830:    marker $m_\gamma\in\overline\gamma$, where $\overline\g$ is the simple
831:      orbit which  underlines $\g$, if $\mathop{\lim}\limits_{r\to
832:    0}f_{V^\pm}(re^{i\theta_0})=m_\g$.
833:   %%%%%
834: Let us recall (see Section \ref{sec:dynamics} above) that we
835: provided each periodic orbit from $\Pc^\pm$ with a ``capping"
836: surface. This surface bounds $\g\in\Pc^\pm$ in $V_\pm$ if $\g$ is
837: homologically trivial, or realizes a homology    between $\g$ and
838: the corresponding linear combination of  basic curves
839: ${C}_i^\pm$. We will  continue to rule out torsion elements in
840: the first homology
841:  (see the discussion of torsion in Section \ref{sec:torsion} below)   and choose
842:   curves ${C}_k\subset W$ which represent a basis of
843: the image $H_1(V^-\cup V^+)\to H(W)$ and for each curve
844: ${C}^\pm_i$ fix a surface $G^\pm_i$ which realizes a homology
845: in $W$
846: between ${C}^\pm_i$ and the corresponding linear combination of
847: curves ${C}_k$.
848: All the choices enable us to associate with a  relative
849:  homology class $A'\in H_2(W,\Gamma^-\cup\Gamma^+)$,
850:  $\Gamma^\pm\subset\Pc^\pm$,
851:  an absolute integral  class
852:  $A\in H_2(W)$, which we will view as an element of $H_2(W;\C)$.
853: 
854:   %%%%%
855: Let us denote by $\Mc^A_{g,r}(\Gamma^-,\Gamma^+;W,J)$ the moduli
856: space of $(j,J)$-holomorphic curves
857: $S_g\setminus(\bx^-\cup\bx^+)\to W$ with $r$ marked points, which
858: are asymptotically cylindrical over the periodic orbit $\g^+_i$
859: from $\Gamma^+$ at    the positive end at the puncture $x^+_i$,
860: and asymptotically cylindrical over the periodic orbit $\g^-_i$
861: from $\Gamma^-$ at    the negative end  at the puncture $x^-_i$,
862: and which send asymptotic markers to the
863: markers on the
864: corresponding periodic orbits. The curves from
865: $\Mc^A_{g,r}(\Gamma^-,\Gamma^+;W,J)$ are additionally required to
866: satisfy a stability condition, discussed in the next section. We
867: write $\Mc^A_{g}(\Gamma^-,\Gamma^+;W,J)$ instead of
868: $\Mc^A_{g,0}(\Gamma^-,\Gamma^+;W,J)$, and
869: $\Mc^A_{g,r}(\Gamma^-,\Gamma^+)$ instead of
870: $\Mc^A_{g,r}(\Gamma^-,\Gamma^+;W,J)$ if   it is clear from the
871: context which target almost complex  manifold $(W,J)$ is
872: considered.
873: 
874:  Notice, that we are not fixing $j$, and the configurations of
875:  punctures, marked points  or asymptotic markers.
876:  Two maps are called equivalent if they differ by a conformal map
877:  $S_g\to S_g$ which preserves all punctures, marked points and
878:  asymptotic markers.
879:      When the manifold $W=V\times\R$ is cylindrical,
880:      and hence the almost complex structure      $J$ is invariant under
881:      translations along the second factor, then
882:      all the moduli spaces $\Mc^A_{g,r}(\Gamma^-,\Gamma^+;W,J)$ inherit the $\R$-action.
883:      We will denote the quotient moduli space  by
884:        $\Mc^{A}_{g,r}(\Gamma^-,\Gamma^+;W,J)/\R$, and by
885: $\Mc^A_{g,r,s^-,s^+}(W,J)$ the union $\bigcup
886: \Mc^A_{g,r}(\Gamma^-,\Gamma^+;W,J)$ taken over
887: all sets of periodic orbits $\Gamma^\pm\subset\Pc^\pm$   with the
888: prescribed numbers $s^\pm$ of elements.
889: We will also need to consider the moduli space of disconnected
890:   curves of  Euler characteristic $2-2g$, denoted by
891:   $\wt{\Mc}_{g,r}^A(\Gamma^-,\Gamma^+)$.
892: 
893: 
894: 
895: 
896: 
897: 
898: 
899: 
900: \subsection{Compactification of the moduli spaces
901: $\Mc_{g,r}^A(\Gamma^-,\Gamma^+)$}\label{sec:compact}
902: 
903: To describe the compactification we need an appropriate notion of
904: a {\it stable holomorphic curve}.
905: 
906: Given a completed symplectic cobordism $W=\ora{V^-V^+}$ we first
907: define a {\it   stable curve of height $1$}, or a {\it $1$-story stable curve }
908:  as a ``usual" stable curve in a sense of M. Kontsevich (see \cite{Kontsevich-CP2}),
909: i.e. a collection of of holomorphic curves $h_i:S_i\to
910: W$  from  moduli spaces
911: $\Mc_{g_i,r}^A(\Gamma^-_i,\Gamma^+_i)$ for various genera $g_i$
912:  which realize  homology classes
913: $A_i$, and collections of periodic orbits $\Gamma_i^\pm$, such
914: that certain pairs of marked points  (called special) on these
915: curves are required to be mapped to one point in $W$. The
916: stability condition means the absence of infinitesimal symmetries
917: of the moduli space. Let us point out, however, that in the case
918: when
919:  $W$ is a cylindrical cobordism, and in particular the almost complex structure
920:  $J$ is translationally invariant, we would need to consider along
921:  with the above moduli space its quotient under the $\R$-action.
922:  The stability for this new moduli space still means an absence of
923:  infinitesimal deformations, but it translates into an additional restriction
924:  on holomorphic curves.
925:   Namely, in the first case the stability
926: condition means that each constant curve has, after   removal
927: of the marked points, a negative Euler characteristic. In the second
928: case it additionally requires  that when {\it all} connected
929: components of the curve are straight cylinders $\g\times\R,
930: \g\in\Pc$ then  at least one of these cylinders should have a
931: marked point.
932: 
933:   One can define an arithmetic genus $g$ of the
934: resulting curve, the total sets $\bx^\pm$ and $\by$ (equal to the
935: union of sets $\bx_i^\pm$ and $\by_i$ for the individual curves of
936: the collection), and the total  absolute  homology class  $A\in
937: H_2(W )$ (see the discussion in Section \ref{sec:holomorphic}
938: above), realized by the union of all curves of the collection.
939: 
940: Moduli of stable curves of height $1$, denoted by
941: ${}_1\Mc_{g,r}^A(\Gamma^-,\Gamma^+)$, form  a part of the
942: compactification of the moduli space
943: $\Mc_{g,r}^A(\Gamma^-,\Gamma^+)$.
944:   However, unlike the case of  closed symplectic manifolds, the
945:   stable curves of height $1$ are not   sufficient  to describe  the
946:  compactification of the moduli space $\Mc_{g,r}^A(\Gamma^-,\Gamma^+)$.
947: 
948: A finite sequence  $(W_1,\dots, W_k)$ of symplectic manifolds with
949: cylindrical ends is called a {\it chain} if  the positive end of
950: $W_i$ matches with the negative end of $W_{i+1}$, $i=1,\dots,k-1$.
951: This means that all data, assigned to an end, i.e. a contact
952: form, marking of periodic orbits, and an almost complex structure,
953: are the same for the matching ends.
954: 
955: Let us first suppose that
956:  none of the cobordisms  which form a chain $(W_1,\dots, W_k)$
957: is cylindrical.    Then
958: a {\it stable curve of height
959: $k$}, or a $k$-story stable curve  in the
960: chain $(W_1,\dots, W_k)$ is a $k$-tuple
961: $f=(f_1,\dots,f_k)$, where $f_{
962: i } \in
963: {}_1\wt{\Mc}_{g_i,r_i}^{A_i}(\Gamma_{i}^-,\Gamma_{i}^+;W_i,J_i)$,
964:  such that the  boundary data  of the curve $f_i$ at the positive end
965: match  the boundary data of $f_{i+1}$ on the negative one.
966:  One
967: also needs to impose the following additional equivalence relation regarding
968: the asymptotic markers on multiple orbits.
969:                     Suppose that $\g $ is
970:    a $k$-multiple periodic orbit, so that the holomorphic curve $f_i$ is
971:    asymptotically cylindrical over $\g$ at the positive end at a puncture $x^+ $, and
972:    $f_{i+1}$ is asymptotically cylindrical over $\g$ at  the negative end  at a puncture $x^-$.
973:     There are $k$ possible  positions $\mu^+_1,\dots,\mu^+_k$
974:     and $\mu^-_1,\dots,\mu^-_k$
975:   of   asymptotic  markers at  each of the punctures $x^\pm$.
976:    We assume here that the markers are numbered cyclically
977:   with respect to the orientation defined by the Reeb vector field
978:   at each of the punctures, and that the markers $\mu^+_1$ and
979:   $\mu^-_1$ are chosen for the curves $f_i$ and
980:   $f_{i+1}$.         Then we identify
981:   $f=\{\dots,f_i,     f_{i+1},\dots\}$
982:    with $(k-1)$ other stable curves of height $k$
983:   obtained by simultaneous cyclic  shift of
984:    the asymptotic markers at the punctures $x^+$ and $x^-$.
985: 
986:    The curves $f_i$, which form  a $k$-story stable curve $f=(f_1,\dots,f_k)$ are called
987: {\it floors},  or {\it levels} of $f$.
988: \medskip
989: 
990:    If some of the cobordisms which form the chain   $(W_1,\dots, W_k)$, say $W_{i_1},\dots,W_{i_l}$,
991: are cylindrical  then we will assume that the corresponding floors
992: of a $k$-story curve in $W=(W_1,\dots,W_k)$ are defined {\it only
993: up to translation.} In other words, if $W_i$ is cylindrical for some $i=1,\dots, k$
994: (i.e. $W_i=V_i\times\R$ and $J_i$ is translationally invariant) then
995: $f_i$ should be viewed as an element of
996: ${}_1\wt{\Mc}_{g_i,r_i}^{A_i}(\Gamma_{i}^-,\Gamma_{i}^+;W_i,J_i)/\R$,
997: rather than ${}_1\wt{\Mc}_{g_i,r_i}^{A_i}(\Gamma_{i}^-,\Gamma_{i}^+;W_i,J_i)$.
998:  It will be convenient for us, however, to introduce the following convention.
999:  When speaking about stable holomorphic  curves  in chains which contain
1000:  cylindrical cobordisms, we will treat the corresponding floors as curves representing
1001:  their equivalence classes from
1002:   ${}_1\wt{\Mc}_{g_i,r_i}^{A_i}(\Gamma_{i}^-,\Gamma_{i}^+;W_i,J_i)/\R$. Any statement about such
1003:   curves should be understood in the sense, that {\it there exist} representatives for
1004:   which the statement is true.
1005: 
1006: 
1007: \bigskip
1008: Let us define now the meaning of convergence of a
1009: sequence of holomorphic curves to a stable curve of height $l$.
1010: For $l=1$ this is   Gromov's standard  definition (see \cite{Gromov-holomorphic}). Namely, with
1011: each stable curve
1012:  $h=\{S_i,h_i\}\in {}_1\Mc_{g,r}^A(\Gamma^-,\Gamma^+)$   of height $1$ we associate a nodal
1013:    surface $\widehat S$
1014: obtained by identifying special pairs of marked points on $\coprod S_i$.  The maps $h_i$ fits together
1015: to a continuous map $\widehat S\to W$ for which we will keep the notation
1016: $h$.
1017:  Let us  consider
1018: also a smooth surface $S$ obtained by smoothing the nodes
1019: of $\widehat S$. There exist a partitioning of  $S$
1020:  by circles
1021:  into  open parts   diffeomorphic to  surfaces $S_i$    with removed special points, and a map
1022:   $g:S\to\widehat S$  which is a diffeomorphism from the complement $\tilde S$
1023:   of the dividing circles
1024:   in $S$ to the complement of  the double points
1025:     in $\hat  S$, and  which collapses the partitioning circles to
1026:   double points.
1027:   A sequence of holomorphic  $\varphi_l:(S,j_l)\to (W,J)$ is said to converge
1028:    to a stable curve
1029:  $h$
1030:     if
1031:       the sequence $\varphi_l|_{\tilde S}$ converges
1032:       to $h_i\circ g|_{\tilde S}$, and
1033:       $j_l$ converges to $g^*(j)$  uniformly on compact sets,
1034:       where $j$ is the conformal structure on the stable curve.
1035:       Of course, we also require  convergence of marked points and asymptotic markers.
1036:     A sequence of stable curves
1037:     $h^j=\{S^j_i,h^j_i\}_{i=1,\dots,k}\in {}_1\Mc_{g,r}^A(\Gamma^-,\Gamma^+), j=1,\dots,$  is said to converge
1038:     to a stable curve $h$, if $h$ can be presented as a collection
1039:     of stable curves $h_i$, $i=1,\dots, k$, such that  $h^j_i$ converges to $  h_i$ in the above
1040:     sense for each $i=1,\dots,k$.
1041:     \medskip
1042: 
1043:  The convergence of
1044:   a sequence of smooth curves to a stable curve of height  $l>1$  is understood
1045:   in a similar sense.
1046:   Let us  assume here     for simplicity that $l=2$ and that the   floors
1047:     $f_{1} :     S_1\to W_1$ and $f_2 :
1048:   S_2\to W_2$  of  a stable curve $f$ in a chain $(W_1,W_2)$
1049:   are smooth, i.e.
1050:   have no  special marked points.
1051:   As in the height $1$ case let us consider
1052:       \begin{description}
1053:       \item{-} the smooth surface $S$
1054:       partitioned  according to the combinatorics of our stable curve
1055:       by circles into two open  (possibly disconnected) parts $U_1$ and $U_2$  diffeomorphic to
1056:       the punctured surfaces $S_1$
1057:        and $S_2$,
1058:        \item{-} the surface
1059:         $\hat S$ with double points obtained by collapsing
1060:         these circles to points, and
1061:         \item{-} the projection $g:S\to\hat S$.
1062:       \end{description}
1063:        Let $(W,J)=(W_1,J_1)\circledcirc (W_2,J_2)$
1064:    be the composition of (completed) directed symplectic    cobordisms $W_1$ and $W_2$
1065:    with compatible almost complex structures $J_1$ and $J_2$. This means that
1066:    \begin{description}
1067:    \item{-}    there exists a contact    hypersurface $V\subset W$ which splits
1068:       $W$ into two cobordisms $W^0_1$ and $W^0_2$;
1069:       \item{-} $W_1=W^0_1\cup V\times[0,\infty),\;W_2=V\times(-\infty,0]\cup W^0_2$;
1070:       \item{-} $J|_{W^0_j}=J_1|_{W^0_j},\;\; j=1,2$;
1071:       \item{-} $J_1$ and $J_2$ are translationally invariant at the ends
1072:       $V\times[0,\infty)$ and $(-\infty,0]\times V$.
1073:       \end{description}
1074: 
1075:       We denote  by
1076:             $W^k $, $k=1,\dots,$ the quotient space of the disjoint union
1077:             $$W^0_1\coprod V\times[-k,k]\coprod W^0_2$$ obtained by identifying $V=\partial W^0_1$
1078:             with  and $V\times( -k)$ and $V=\partial W^0_2$
1079:             with  and $V\times k$,  and extend the   almost
1080:       complex  structures $J_1|_{W^0_1}$ and     $J_2|_{W^0_2}$
1081:       to the unique almost complex structure  $J^k$ on $W^k$ which is translationally
1082:        invariant on $V\times[-k,k]$. We also consider $W^k_1$ obtained
1083:        by gluing $W^0_1$ and $V\times[0,k]$ along $V=\partial
1084:        S^0_1=V\times 0$,
1085:        and $W^k_2$ glued in a similar way from $V\times[-k,0]$ and $W^0_2$.
1086:         We have $W_j=\bigcup\limits_{k=0}^\infty
1087:        W^k_j$, $j=1,2$. On the other hand, $W^k_1$ and $W^k_2$
1088:         can be viewed as submanifolds
1089:        of $W^k$.
1090: 
1091: 
1092:       \begin{definition}\label{def:convergence}
1093:       {\rm
1094:                 Suppose that we are  given a sequence $j^k$
1095:                  of conformal structures on the surface $S$
1096:        and a sequence  of  $1$-story $(j^k,J^k)$-holomorphic curves
1097:        $f^k:S\to W^k$.
1098:        We say that this sequence converges to a stable curve $f=(f_1,
1099:        f_2)$
1100:        of height   $2$
1101:        in $(W_1,W_2)$ if     there     exist two sequences of domains
1102:        $U_1^1\subset\dots \subset U_1^i\subset \dots\subset U_1$
1103:        and $U_2^1\subset\dots \subset U_2^i\subset \dots\subset U_2$,
1104:         such that
1105:         \begin{description}
1106:         \item{} $\bigcup\limits_{k=1}^\infty U^k_i=U_i,\; i=1,2$ ;
1107:         \item{} $f^k(U^k_i)\subset W^k_i$ for $i=1,2,\;\;k=1,\dots$;
1108:         \item{}  for $i=1,2$  the  holomorphic curves
1109:         $f^k|_{U^k_i}$ converges to $f_i\circ g: U_i\to W_i$,
1110:                 and the conformal structures   $j^k|_{U^k_1}$ converge to $g^*{j_i} $
1111:                 when $k\to\infty$
1112:          uniformly on
1113:         compact sets.
1114: As in the case of stable curves of level $1$
1115: we also require
1116: convergence of marked points and asymptotic markers.
1117:                 \end{description} }
1118:   \end{definition}
1119:   Let us emphasize that  when some of the cobordisms  are cylindrical
1120:   then according to  the convention which we introduced above
1121:     one is allowed to compose  the corresponding curves  with translations to satisfy
1122:   the above definition.
1123: 
1124:  \bigskip
1125:  Notice that if the  cobordism $W_2$ is cylindrical, i.e.
1126:  $W_2=V\times\R$ and $J_2$ is translationally invariant,
1127:  then $W_1\circledcirc W_2$ can be identified with $W_1$, and thus one can talk about
1128:  convergence of a sequence of curves
1129:   $f^k\in {}_1{\Mc}_{g,r}^{A}(\Gamma^-,\Gamma^+;W_,J_1)$
1130:   (where the almost complex structure
1131:   $J_1$ is fixed!)
1132:  to   a $2$-story curve $(f_1,f_2)$, where
1133:  $f_1 \in{}_1\wt{\Mc}_{g_1,r_1}^{A_1}(\Gamma^-,\Gamma;W_1,J_1)$,
1134:   $f_2\in{}_1\wt{\Mc}_{g_2,r_2}^{A_2}(\Gamma,\wt\Gamma^+;V\times\R,J_2)/\R$,
1135:   $g=g_1+g_2,\,r=r_1+r_2, A=A_1+A_2$, and $J_2$ is translationally invariant.
1136:  It is important to stress the point that the curve   $f_2$ is defined only up to translation.
1137: 
1138: 
1139: 
1140:        \begin{theorem}\label{thm:comp1}
1141:        Let $f_k\in {}_1\Mc_g^A(\Gamma^-,\Gamma^+)$, $k=1,\dots, $ be a
1142:        sequence of stable holomorphic curves in
1143:        a (complete)  directed symplectic cobordism $W$.
1144:         Then there exists a chain of  directed symplectic
1145:         cobordisms $$A_1,\dots, A_a,W,B_1,\dots,B_b,$$ where all
1146:         cobordisms $A_i$ and $B_i$ are cylindrical,
1147:         and a stable curve $f_\infty$ of height $a+b+1$ in this chain
1148:         such that a subsequence of $ \{f_i\}$  converges to
1149:         $f_\infty$. {\rm See Fig. \ref{fig:split1}}.
1150:         \end{theorem}
1151: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1152: 
1153: \begin{figure}
1154: \centerline{\psfig{figure=sf2.eps,height=100mm}}
1155: \caption{\small A possible splitting of a sequence of holomorphic curves in
1156:  a completed symplectic cobordism}
1157: \label{fig:split1}
1158: \end{figure}
1159: \begin{figure}
1160: \centerline{\psfig{figure=sf4.eps,height=80mm}} \caption{\small A
1161: possible splitting of a sequence of holomorphic curve   when
1162: $J_k\to J_\infty$ } \label{fig:split2}
1163: \end{figure}
1164: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1165:         \begin{theorem}\label{thm:comp2}
1166:         Let $W$ be a completed directed symplectic cobordism,
1167:         $V\subset W$ a contact hypersurface, and $J_k$    a
1168:         sequence of compatible almost complex structures on $W$
1169:         which realizes the splitting of $W$ along $V$ into two
1170:         directed symplectic cobordisms $W_-^\infty$ and
1171:         $W_+^\infty$
1172:         {\rm (see Section \ref{sec:almost} above)}.
1173:         Let $f_k$ be a sequence of stable $J_k$-holomorphic curves
1174:           from $ {}_1\Mc_g^A(\Gamma^-,\Gamma^+;W,J_k)$. Then there exists a
1175:         chain of directed symplectic cobordisms
1176:         $$A_1,\dots,A_a,W_-,B_1,\dots,B_b,W_+,C_1,\dots,C_c$$  where
1177:         all cobordisms $A_i, B_j, C_l$ are cylindrical, such that a
1178:         subsequence of
1179:         $\{f_i\}$   converges to a stable curve of height $a+b+c+2$
1180:         in
1181:          the chain $$A_1,\dots,A_a,W_-,B_1,\dots,B_b,W_+, C_1,\dots,
1182:          C_c.$$ {\rm See Fig. \ref{fig:split2}. The reader
1183:          may consult \cite{HoferWysockiZehnder:99}
1184:          for the analysis of splitting $\C P^2$
1185:          along the boundary of a tubular
1186:           neighborhood
1187:          of $\C P^1\subset\C P^2$.}
1188:         \end{theorem}
1189: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1190: 
1191: The definition of convergence can be extended in an obvious way to a sequences of stable curves
1192: of height $l>1$. Namely, we say that a sequence of $l$-story curves
1193: $f^k=(f^k_1,\dots,f^k_l)$, $k=1,\dots,\infty$, in a chain $(W_1,\dots,W_l)$ converge to a stable $L$-story,
1194: $L=m_1+\dots+m_l$, curve  $f=(f_{11},\dots,,f_{1m_1},\dots,f_{l_1}\dots f_{lm_l}) $ in a chain
1195:  $$(W_{11},\dots,W_{1m_1},\dots,W_{l_1}\dots W_{lm_l})$$
1196: if for each $i=1,\dots,l$  the cobordism $W_i$ splits into the composition
1197: $$W_i=W_{i1}\circledcirc\dots\circledcirc W_{im_i}$$
1198: and the sequence $f^k_i$, $k=1,\dots,\infty$, of stable curves of height $1$ converges
1199: to the $m_i$-story curve $f_i=(f_{i1},\dots,f_{im_i})$ in the chain $(W_{i1},\dots, W_{im_i})$ in the sense of Definition
1200: \ref{def:convergence}.
1201: \bigskip
1202: 
1203: %%%%%%%%%%%%%%%%%%%%%%%%%
1204: It is important to combine Theorems \ref{thm:comp1} and
1205: \ref{thm:comp2} with the following
1206: observation which is a
1207: corollary of  Stokes' theorem combined with Proposition \ref{prop:positivity}.
1208: 
1209: 
1210: \begin{proposition}\label{prop:one-end}
1211:  A holomorphic curve in an exact
1212:   directed symplectic cobordism
1213:  (and in particular in a cylindrical one)
1214:  must have at least one positive puncture.
1215:  \end{proposition}
1216: 
1217:  In particular,  we have
1218:  \begin{corollary}\label{cor:pos-puncture}
1219: Let $f^n\in {\Mc}_0(W,J)$  be a sequence of rational holomorphic curves with  one positive, and possibly
1220: several negative punctures.
1221:  Suppose that the sequence  converges to a stable
1222:  curve $$F=\{g_1,\dots,g_a,f,h_1,\dots,h_b\}$$ of height $a+b+1$ in a chain
1223:  $$A_1,\dots,A_a,W,B_1,\dots,B_b.$$ Then the  $W$-component   $f\in {\Mc}_0(W,J)$ of the stable curve
1224:  $F$ has precisely
1225:  one positive puncture as well.
1226:  \end{corollary}
1227: 
1228: 
1229:  \subsection{Dimension of the moduli spaces
1230: $\Mc^A_{g,r}(\Gamma^-,\Gamma^+)$}
1231: One has the following index formula for the corresponding
1232: $\bar\partial$-problem which compute the dimension
1233: of the moduli space $\Mc^A_{g,r}(\Gamma^-,\Gamma^+;W,J)$ for a {\it
1234: generic}
1235:  choice of $J$.
1236: \begin{proposition}
1237: 
1238:   \begin{equation}\label{eq:dim}
1239:   \begin{split}
1240:  & {\mathrm {dim}}\Mc^A_{g,r}(\Gamma^-,\Gamma^+;W,J)
1241:  =\sum\limits_1^{s^+}\CZ(\gamma^+_i)-
1242:  \sum\limits_1^{s^-}\CZ(\gamma^-_k)\\
1243:  &  +  (n-3)(2-2g-s^+-s^-)+2c_1(A)+2r,\\
1244:  \end{split}
1245:  \end{equation}
1246:   where
1247: $s^\pm$ are the cardinalities of the sets $\Gamma^\pm$, and $c_1\in
1248: H^2(W)$ is the first Chern class of the almost complex manifold
1249: $(W,J)$
1250: \end{proposition}
1251: 
1252:   Making the moduli spaces non-singular by picking
1253:    generic $J$ is needed for the purpose of curve counting but
1254:             does not always work properly. It is therefore crucial
1255:              that the moduli spaces of stable $J$-holomorphic
1256: curves are non-singular {\em virtually}. This means that for {\em
1257: any} $J$ the moduli spaces, being generally speaking singular, can
1258: be equipped with some canonical additional structure that make
1259: them {\em function in the theory the same way as if they were
1260: orbifolds with boundary and had the dimension prescribed by the
1261: Fredholm index}. In particular, the moduli spaces come equipped
1262: with rational fundamental cycles relative to the boundary (called
1263: {\em virtual fundamental cycles}) which admit pairing with
1264: suitable de Rham cochains and allow us to use the Stokes
1265: integration formula.
1266: 
1267: 
1268: 
1269: Technically the virtual smoothness is achieved by a
1270: finite-dimensional reduction of the following picture: a singular
1271: moduli space is the zero locus of a section defined by the
1272: Cauchy-Riemann operator in a suitable orbi-bundle over a moduli
1273: orbifold of stable $C^{\infty}$-maps. More general {\em virtual
1274: transverality} properties for families of
1275:  $J$'s also hold true
1276: (cf. \cite{Fukaya-Ono, Fukaya-Ono2, Liu-Tian, Li-Tian, Ruan-virtual, Siebert, McDuff} at al.).
1277: We
1278: are reluctant to provide in this quite informal exposition
1279:  precise
1280: formulations   because of numerous not entirely innocent
1281: subtleties this would entail.  Fortunately, what we intend to say in the rest of this
1282: paper does not depend much on the details we are omitting.
1283: 
1284: 
1285: \bigskip
1286: 
1287: 
1288: As it was explained in Section \ref{sec:compact} above the moduli
1289: space $\Mc^A_{g,r}(\Gamma^-,\Gamma^+;W,J)$  can be compactified by
1290: adding strata which consist of stable holomorphic curves of
1291: different height.  This compactification looks quite similar to
1292: the Gromov-Kontsevich compactification  of   moduli spaces of
1293: holomorphic curves in a closed symplectic manifold  with a
1294: compatible almost complex structure.
1295: There is, however, a major difference. In the case of a closed
1296: manifold   all the strata   which one needs to add to compactify the moduli space
1297: of smooth holomorphic curves have (modulo virtual cycle complications) codimension
1298: $\geq 2$. On the other hand  in our case  the  codimension one strata are
1299: present {\it generically}. Thus in this case the boundary of the compactified moduli space,
1300: rather than the moduli space  itself, carries  the fundamental
1301: cycle.
1302: 
1303:  In particular, this boundary
1304:   is tiled by codimension one
1305:  strata represented by stable curves $(f_-,f_+)$ of {\it height two}.
1306: Each such a stratum can be described by the constraint matching
1307: the positive ends of $f_-$ with the negative ends of $f_+$ in the
1308: Cartesian product of the moduli spaces $\Mc_{\pm}$ corresponding to   the
1309: curves
1310: $f_{\pm}$  separately.
1311: Proposition \ref{prop:boundary} below describes these top-dimensional boundary
1312:  strata more precisely
1313: in two important for our purposes situations.
1314: Let us point out that  \ref{prop:boundary} literally holds only under certain
1315:  transversality conditions.
1316: Otherwise it should be understood only virtually.
1317: 
1318: 
1319:  \begin{proposition}\label{prop:boundary}
1320:      \begin{description}
1321:      \item{1.} Let $(W=V\times\R,J)$ be a cylindrical cobordism.
1322:      Then any
1323:     top-dimensional stratum  $\cS$ on the boundary  of the compactified moduli space
1324:      $\overline{\Mc^A_{g,r}(\Gamma^-, \Gamma^+;W,J)/\R}$ consists
1325:      of
1326:          stable curves  $(f_-,f_+)$   of height   two,
1327:           $f_\pm\in\Mc_\pm/\R$, where
1328:           $$\Mc_-=\wt\Mc^{A_-}_{g_-,r_-}(\Gamma^-,\Gamma;W,J)/\R ,
1329:           \quad \Mc_+=\wt\Mc^{A_+}_{g_+,r_+}(\Gamma,\Gamma^+;W,J)/\R ,$$
1330:            $g=g_-+g_+, \;r=r_-+r_+,\;A=A_-+A_+$,   $\Gamma=\{\g_1,\dots,\g_l\}\subset\Pc$.
1331: 
1332:                  All  but one
1333:           connected components  of  each of the curves $f_-$ and $f_+$
1334:            are trivial cylinders
1335:           (i.e.  have the form   $\g\times\R, \g\in\Pc$) without
1336:           marked points.
1337:           \item{2.} Let $(W=\ora{V^-V^+},J)$  be any    cobordism,
1338:           and $(W_\pm,J_\pm)= (V^\pm\times\R,J_\pm)$
1339:           be the cylindrical cobordisms associated to its
1340:           boundary.
1341:           Then any  top-dimensional strata $\cS$ on the boundary  of the
1342:           compactified moduli space
1343:      $\overline{\Mc^A_{g,r}( \Gamma^-, \Gamma^+;W,J) }$ consists
1344:      of
1345:          stable curves  $(f_-,f_+)$   of height   two,
1346:           $f_\pm\in\Mc_\pm$, where either
1347:           $$\Mc_-=\wt\Mc^{A_-}_{g_-,r_-}(\Gamma^-,\Gamma;W_-,J_-)/\R,\;
1348:  \Mc_+=\wt\Mc^{A_+}_{ g_+,r_+}(\Gamma,\Gamma^+;W,J)\;\hbox{
1349:  and}\;
1350:   \Gamma\subset\Pc^-,$$ or
1351: $$\Mc_+=\wt\Mc^{A_+}_{g_-,r_-}(\Gamma ,\Gamma^+;W_+,J_+)/\R,\;
1352:  \Mc_-=\wt\Mc^{A_-}_{g_-,r_-}(\Gamma^-,\Gamma;W,J ) \;\hbox{
1353:  and}\;
1354:   \Gamma\subset\Pc^+.$$
1355:    In both cases  we have  $$g=g_-+g_+,
1356:  \;r=r_-+r_+,\;\;\hbox{and}\;\;A=A_-+A_+.$$
1357:   The part of the stable curve
1358:  $(f_-,f_+)$ which is  contained in in  $W_\pm$ must have
1359:  precisely one non-cylindrical connected component, while there
1360:  are no restrictions on the number and the character of connected components or the
1361:  other part of the stable curve.
1362:  \item{~~} In both cases the stratum
1363:          $\cS=\cS(\Gamma,g_-|g_+,r_-|r_+,A_-|A_+)$ is
1364:          diffeomorphic to a $\kappa$-multiple cover of the product
1365:          $\Mc_-\times\Mc_+$, where the multiplicity $\kappa$ is
1366:          determined by the multiplicities of periodic orbits from
1367:          $\Gamma$.
1368:  \end{description}
1369:               \end{proposition}
1370: 
1371: Proposition \ref{prop:boundary} is not quite sufficient for our
1372: purposes, as we also needs to know the structure of the boundary
1373: of   moduli spaces of $1$-parametric families of holomorphic
1374: curves. However, we are not  formulating the corresponding
1375: statement in this paper, because it  is   intertwined  in a much more serious way
1376: with
1377: the virtual cycle techniques and terminology. An algebraic description
1378: of this boundary is given in
1379:    Theorem
1380: \ref{thm:SFT-chain}
1381: below.
1382: 
1383: 
1384: 
1385: \bigskip
1386: 
1387: 
1388: Let us consider some special cases of the formula (\ref{eq:dim}).
1389: Suppose, for instance, that $W$ is the cotangent bundle of a manifold
1390: $L$. Then $W$ is a symplectic manifold which has  only a positive
1391: cylindrical end.
1392:  If $L$ is orientable then there is a canonical way to
1393: define Conley-Zehnder indices. Namely, one takes any
1394: trivialization along orbits, which is  tangent to vertical
1395: Lagrangian fibers. The resulting index is independent of a
1396: particular  trivialization. For this trivialization, and a choice
1397: of a contact form corresponding to  a metric on $L$ we have
1398:  \begin{proposition}
1399: Periodic orbits of the Reeb flow are lifts of closed geodesics,
1400: and if $L$ is orientable their Conley-Zehnder indices   are equal
1401: to Morse indices of the corresponding geodesics and we have
1402:  $${\mathrm {dim}}\Mc_g^A( \Gamma^+)= \sum\limits_i\Morse(\g^+_i)+ (n-3)(2-2g-s^+) .$$
1403: \end{proposition}\label{prop:Morse-Zehnder}
1404: Notice that for a metric on  $L$ of  non-positive curvature we have  $s^+>1$, because in this case
1405: there are no
1406: contractible geodesics. Moreover, if the metric has negative curvature then all geodesics
1407: have  Morse
1408: indices equal to $0$.  Hence, we get
1409: \begin{corollary}\label{cor:isolated} In  the cotangent bundle of a  negatively curved
1410: manifold  of dimension $>2$ there  could be only isolated
1411: holomorphic curves. If, in addition, $n\neq 3$ then these curves are spheres
1412: with two positive punctures. Each of these curves is asymptotically
1413: cylindrical at punctures over   lifts of the same geodesic with
1414: opposite orientations.
1415: \end{corollary}
1416: %CHANGE
1417: Let us point out that the orientability is not required in
1418: Corollary \ref{cor:isolated}. The corresponding result for a
1419: non-orientable manifold  follows from \ref{prop:Morse-Zehnder}
1420: applied to its orientable double  cover.
1421: 
1422: \subsubsection*{Absence of hyperbolic Lagrangian submanifolds  in uniruled manifolds}
1423: 
1424: As the first  application of the above compactness theorems let us
1425: prove here the following theorem of C. Viterbo. Let us recall that
1426: a  complex projective manifold $W$ is called uniruled, if there is a
1427: rational holomorphic curve
1428:   through each point of $W$. For instance,  according to
1429:   Y. Myaoka--S. Mori
1430:  \cite{Mori-Miayoka}  and J. Kollar  \cite{Kollar}  Fano manifolds are uniruled.
1431: 
1432: 
1433: 
1434: \begin{theorem} {\rm (C. Viterbo, \cite{Viterbo})}
1435: Let $W$ be a uniruled manifold of complex dimension $>2$,  $\omega$ its K\"ahler
1436: sympletic form, and $L\subset W$ an embedded Lagrangian
1437: submanifold. Then $L$ does not admit a Riemannian metric of
1438: negative sectional  curvature.
1439: \end{theorem}
1440: \proof J. Kollar  \cite{Kollar}  and in a more general case Y.
1441: Ruan   \cite{Ruan-virtual} proved  that there exists a homology
1442: class $A\in H_2(W)$, such that for any almost complex structure
1443: compatible with $\omega$ and any point $z\in W$ there exists
1444: $f\in\Mc_{0,1}^A(W,J)$ with $f(y)=z$, where $y$ is the marked
1445: point. Let us identify a neighborhood $U$ of $L$ in $W$ with  a
1446: neighborhood of the zero-section in $ T^*(L)$. Suppose $L$ admits
1447: a Riemannian metric of negative
1448:   curvature. We can assume that  $U$ is the round neighborhood of radius $1$ in
1449:   $T^*(L)$. Let us consider a sequence $J^m$ of almost complex structures
1450:   on $W$, which realizes the splitting along the contact  type hypersurface
1451:   $(V=\partial U,\alpha=pdq|_V)$.
1452:   (see Section \ref{sec:almost}). Then according to Example \ref{ex:cotangent}
1453:   $W$ splits into $W_-=T^*(L)$ and $W_+=W\setminus L$. The almost
1454:   complex structure $J_-$ on $T^*(L)$ is compatible at infinity  with the
1455:   contact $1$-form   $\alpha=pdq|_{V}$. According to Corollary
1456:   \ref{cor:isolated} for any choice $\Gamma=\{\g_1,\dots,\g_k\}$ and
1457:   any $g\geq 0$ the
1458:   moduli spaces $\Mc_g(\Gamma;W_-,J_-)$ are empty, or
1459:   $0$-dimensional. One the other hand, Theorem \ref{thm:comp2}
1460:   together with Ruan's theorem guarantee the existence of a
1461:   rational holomorphic curve  with punctures through every point
1462:   of $L$. This contradiction proves that
1463:   $L$ cannot admit a metric
1464:   of negative curvature.  \qed
1465: 
1466: %%%%%%%%%%%%%%%%%%%%%%%%
1467: \subsection{ Coherent orientation of the moduli spaces
1468:   of holomorphic curves}\label{sec:orientation}
1469: 
1470: 
1471:     To get started with the algebraic formalism,
1472:     one first needs to orient moduli spaces
1473:     $\Mc(\Gamma^-,\Gamma^+)$ of holomorphic curves with punctures.
1474:     This problem is much easier in the case
1475:     of moduli spaces of closed holomorphic curves, because in that
1476:     case moduli spaces are even-dimensional and carry a canonical almost
1477:      complex structure (see Section \ref{sec:coherent-closed} below).
1478:     In our case we have to adapt the philosophy of {\it coherent orientations}
1479:      of the moduli spaces
1480:     borrowed from Floer homology theory (see \cite{Floer-Hofer}).
1481:     We sketch this approach in this section.
1482: 
1483: 
1484: \subsubsection{Determinants}
1485: 
1486: In order to separate the problems of orientation and
1487: transversality we are going to orient the determinant line bundles
1488: of the linearized $\overline\partial$-operators, rather than  the moduli spaces themselves.
1489: 
1490:  For a linear Fredholm operator $F:A\rightarrow
1491: B$ between Banach spaces we can define its determinant line
1492: $\det(F)$ by $$ \det(F)= (\Lambda^{max} \Ker (F))\otimes
1493: (\Lambda^{max} \Coker (F))^{\ast}. $$ We note that for the trivial
1494: vector space $\{0\}$ we have $\Lambda^{max}\{0\}= {\R}$.  An
1495: orientation for $F$ is by definition an orientation for the line
1496: $\det(F)$. In particular, given an isomorphism $F$ we can define a canonical
1497: orientation given through the vector $1\otimes 1^{\ast}\in {\R
1498: }\otimes {\R}^{\ast}$.
1499: 
1500: Given a continuous family $F=\{F_y\}_{y\in Y}$ of Fredholm
1501: operators, parameterized by a topological space $Y$,  the
1502: determinants of operators $F_y$ form   a line bundle
1503: $\det(F)\rightarrow Y$. The fact that this is a line bundle in a
1504: natural way might be surprising since the dimensions of kernel and
1505: cokernel vary in general. This is however a standard fact, see for
1506: example \cite{Floer-Hofer}.
1507: 
1508: \subsubsection{Cauchy-Riemann Type Operators on Closed Riemann Surfaces}
1509: \label{sec:coherent-closed} Let $(S,j)$ be a closed, not
1510: necessarily connected Riemann surface and $E\rightarrow S$ a
1511: complex vector bundle. Denote by $X_E\rightarrow S$ the complex
1512: $n$-dimensional vector bundle whose fiber over $z\in S$ consists
1513: of all complex ant-linear maps $$ \phi:T_zS\rightarrow E_z,\ z\in
1514: S,\ \hbox{i. e.}\ J\circ\phi+\phi\circ j=0, $$ where $J$ is the
1515: complex structure on $E$. Fixing a connection $\nabla$ and a
1516: smooth $a\in\hbox{Hom}_{\mathbb R}(E,X_E)$ we can define a
1517: Cauchy-Riemann type operator $$L:C^{\infty}(E)\rightarrow
1518: C^{\infty}(X_E)$$ by the formula $$
1519: (Lh)(X)=\nabla_Xh+J\nabla_{jX}h +(ah)(X), $$ where $X$ is an
1520: arbitrary vector field on $S$. Since the space of connections is
1521: an affine space we  immediately see that the set ${\mathcal O}_E$
1522: of all Cauchy-Riemann type operators on $E$ is convex. For a
1523: proper functional analytic set-up, where we may chose H\"older or
1524: Sobolev spaces, the operator $L$ is Fredholm. By elliptic
1525: regularity theory the kernel and cokernel would be spanned always
1526: by the same smooth functions, regardless which choice we have
1527: made. The index of $L$ is given by the Riemann-Roch formula $$
1528: \hbox{ind}(L)= (1-g)  \hbox{dim}_{\mathbb R}(E) + 2 c(E), $$ where
1529: $c(E)$ the first Chern number $c_1(E)(S)$ of $E$. Here we assume
1530: $S$ to be a connected closed surface of genus $g=g(S)$.
1531: 
1532: 
1533:  Let   $\phi:(S,j)\rightarrow (T,i)$
1534:  be a biholomorphic map and $\Phi:E\rightarrow F$ a ${\mathbb
1535: C}$-vector bundle isomorphism covering $\phi$. Then $\Phi$ induces
1536: an isomorphism $$ { \Phi}_{\ast}:{\mathcal O}_E\rightarrow
1537: {\mathcal O}_F $$ in the obvious way. The operators   $(E,L)$ and
1538: $(F,K)$ are called isomorphic  if there exists ${
1539: \Phi}:E\rightarrow F$, so that ${\bf \Phi}_{\ast}(L)=K$. We will
1540: denote by $[E,L]$ the equivalence class  of an operator  $(E,L)$
1541: which consists of operators $(F,K)$, equivalent to $(E,L)$ under
1542: isomorphisms, {\it isotopic to the identity}, and by $[[E,L]]$ the
1543: equivalence class under the action of the full group of
1544: isomorphisms. The moduli space of equivalence classes $[[E,L]]$
1545: will be denoted by $\CR$, and the ``Teichmuller space" which
1546: consists of classes $[E,L]$ will be denoted by $\wt{\CR}$. An
1547: isomorphism $\Phi$ induces an isomorphism between the kernel
1548: (cokernel) of $L$ and ${ \Phi}_{\ast}L$ for every $L\in {\mathcal
1549: O}_E$, and hence one can canonically associate the determinant
1550: line to an isomorphism class, and thus define the {\it determinant
1551: line bundle} $\cV$ over the moduli space $\CR$.
1552:  Given an orientation $o$ for $L$ we obtain an induced
1553: orientation ${ \Phi}_{\ast}(o)$.  Let us note the following
1554: \begin{lemma}
1555:  The bundle $\cV$ is orientable.
1556: \end{lemma}
1557: \proof The lift $\wt\cV$ of the bundle $\cV$ to the Teichmuller
1558: space $\wt\CR$ is obviously orientable, because each connected
1559: component of the space $\wt\CR$ is contractible. However, one
1560: should check that an arbitrary isomorpism $\Phi:(E,L)\to (F,K)$
1561: preserves the orientation. This follows from the following
1562: observation. Any connected component of $\wt\CR$ contains an
1563: isomorphism class of a complex linear operator $(E,L_0)$,  and any
1564: two complex linear operators representing points in a given
1565: component of $\wt\CR$ are homotopic in the class of complex linear
1566: operators. The determinant of $(E,L_0)$ can be oriented
1567: canonically by observing that its kernel and cokernel are complex
1568: spaces.  Any isomorphism  maps a   complex linear operator to a
1569: complex linear operator and preserves its complex orientation.
1570: Hence, it preserves an orientation of the determinant line  of any
1571: operator $(E,L)$.
1572: \qed
1573: \medskip
1574: 
1575: We will call an orientation of $\cV$ {\it complex} if it coincides
1576: with    the complex orientation of   determinants   of complex
1577: linear operators.
1578: 
1579: 
1580: \bigskip
1581: 
1582: The  components of the space $\CR$ are
1583: parameterized by the topological type of the underlying surface
1584: $S$ and the isomorphism class of the bundle $E$.
1585: It turns out that the complex orientation of $\cV$ satisfies three
1586: {\it coherency} Axioms A1--A3 which we formulate below.
1587: They relate  orientations of $\cV$ over different components of
1588: $\CR$.
1589: Conversely, we will see that these axioms determine the
1590: orientation uniquely up to a certain normalization.
1591: 
1592: 
1593: 
1594: Given $(E,L)$ and $(F,K)$ over surfaces $\Sigma_0$ and $\Sigma_1$
1595:  we define a {\it disjoint union}
1596: $$ [E,L]\dot{\cup}[F,K] := [G,M]  $$ of $(E,L)$ and $(F,K)$ as a pair
1597: $(G,M)$, where $G$ is a bundle over     the disjoint union
1598: $\Sigma=\Sigma_0\coprod\Sigma_1$, so that $(G,M)|_{\Sigma_0}$ is
1599: isomorphic to $(E,L)$ and $(G ,M)_{\Sigma_1}$  is isomorphic to
1600: $(F,K)$. Clearly, the
1601: isomorphism class of a disjoint union is uniquely determined by the classes
1602: of $(E,L)$ and $(F,K)$.
1603: Thus, we have a well-defined construction called {\it disjoint
1604: union}:
1605:  The determinant $\det\Sigma$ is canonically isomorphic to $\det
1606:  L\otimes\det K$, and hence the orientations    $o_K$ and $o_L$
1607:  define
1608:  an orientation $o_K\otimes o_L$ of    $ \Sigma $.
1609:  Our first axiom  reads
1610:  \medskip
1611: 
1612: \noindent{\textsf{Axiom C1}}. For any disjoint union $
1613: [G,M]=[E,L]\dot{\cup}[F,K]$ the orientation $o_M$ equals
1614:  $o_K\otimes o_L$.
1615:  \bigskip
1616: 
1617: 
1618: 
1619: Given $(E,L)$ and $(F,K)$, where $E$ and $F$ are  bundles over $S$
1620: of possibly different rank, we can define an operator $(E\oplus
1621: F,L\oplus K)$. There is a canonical map $$
1622: \hbox{det}(L)\otimes\hbox{det}(K)\rightarrow \hbox{det}(L\oplus
1623: K),
1624: $$ and thus  given orientations $o_L$ and $o_K$ we obtain $o_L\oplus o_K$.
1625: 
1626: 
1627: \bigskip
1628: \noindent{\textsf{Axiom C2}}.$$o_{L\oplus K}=o_L\oplus o_K.$$
1629:  \bigskip
1630: 
1631: To formulate the third axiom, we need a construction, called {\it
1632: cutting and pasting}.
1633: 
1634: Let $(E,L)$ be given and assume that
1635: $\gamma_1,\gamma_2:S^1\rightarrow S$ be real analytic embeddings
1636: with mutually disjoint images. Assume that $\Phi:E|_{\gamma_1}
1637: \rightarrow E|_{\gamma_2}$ is a complex vector bundle isomorphism
1638: covering $\sigma=\gamma_2\circ\gamma_{1}^{-1}$. The maps
1639: $\gamma_1$ and $\gamma_2$   extends as holomorphic
1640: embeddings $\bar{\gamma}_j:[-\varepsilon,\varepsilon]\times S^1
1641: \rightarrow S$ for a suitable small $\varepsilon>0$, so that the
1642: images are still disjoint. Locally, near $\gamma_j$ we can
1643: distinguish the left and the right side of $\gamma_j$. These sides
1644: correspond to the left or the right part of the annulus
1645: $[-\varepsilon,\varepsilon]\times S^1$. Cutting $S$ along the
1646: curves $\gamma_j$ we obtain a compact Riemann surface $\bar{S}$
1647: with boundary. Its boundary components are $\gamma_{j}^{\pm}$,
1648: $j=1,2$, where $\gamma^{\pm}_{j}$ is canonically isomorphic to
1649: $\gamma_j$. The vector bundle $E$ induces a vector bundle
1650: $\bar{E}\rightarrow \bar{S}$. We define a space of smooth sections
1651: $\Gamma_{\Delta}(\bar{E})$ as follows. It consists of all
1652: smooth sections $\bar{h}$ with the property that $$
1653: \bar{h}|_{\gamma_{j}^{-} }= \bar{h}|_{\gamma_{j}^{+}}\ \hbox{for}\
1654: j=1,2. $$ Then $L$ induces an operator
1655: $\bar{L}:\Gamma_{\Delta}(\bar{E}) \rightarrow
1656: \Gamma(X_{\bar{E}})$. The operators  $L$ and $\bar{L}$
1657: have naturally isomorphic kernel and cokernel. So an orientation $o$ of $\hbox{det}(L)$
1658: induces one of $\hbox{det}(\bar{L})$. The boundary condition
1659: $\Delta$ can be written in the form
1660: \begin{eqnarray*}
1661: \left[\begin{array}{cc} 1&0\\ 0&1
1662: \end{array}\right]
1663: \cdot\left[\begin{array}{c}
1664: \Phi(\gamma_{1}(t)\bar{h}\circ\gamma_{1}^{-}(t)\\
1665: \bar{h}\circ\gamma_{2}^{-}(t)
1666: \end{array}\right]
1667: =\left[\begin{array}{c}
1668: \Phi(\gamma_1(t))\bar{h}\circ\gamma^{+}_{1}(t)\\
1669: \bar{h}\circ\gamma^{+}_{2}(t)
1670: \end{array}\right]
1671: \end{eqnarray*}
1672: We introduce a parameter depending boundary condition by
1673: \begin{eqnarray*}
1674: \left[\begin{array}{cc} cos(\tau)&sin(\tau)\\ -sin(\tau)&cos(\tau)
1675: \end{array}\right]
1676: \cdot\left[\begin{array}{c}
1677: \Phi(\gamma_{1}(t)\bar{h}\circ\gamma_{1}^{-}(t)\\
1678: \bar{h}\circ\gamma_{2}^{-}(t)
1679: \end{array}\right]
1680: =\left[\begin{array}{c}
1681: \Phi(\gamma_1(t))\bar{h}\circ\gamma^{+}_{1}(t)\\
1682: \bar{h}\circ\gamma^{+}_{2}(t)
1683: \end{array}\right]
1684: \end{eqnarray*}
1685: for $\tau\in [0,\frac{\pi}{2}]$. For all these boundary conditions
1686: $L$ induces an operator, which is again Fredholm of the same
1687: index. For every $\tau$ we obtain a Cauchy-Riemann type operator
1688: from $\Gamma_{\Delta_{\tau}}(\bar{E}) $ to $\Gamma(X_{\bar{E}})$.
1689: Note that for a section $h$ satisfying the boundary condition
1690: $\Delta_{\tau}$ the section $ih$ satisfies the same boundary condition.
1691: On the other hand for
1692: $\tau=\frac{\pi}{2}$ we obtain a Fredholm operator whose
1693: kernel and cokernel naturally isomorphic to the kernel on cokernel of
1694:  a   Fredholm operator
1695: on a new closed surface. Namely identify $\gamma^{+}_{1}$ with
1696: $\gamma^{-}_{2}$ and $\gamma^{+}_{2}$ with $\gamma^{-}_{1}$. For
1697: the bundle $\bar{E}$ we identify the part above $\gamma^{+}_{1}$
1698: via $\Phi$ with the part over $\gamma_{2}^{-}$ and we identify the
1699: part above $\gamma^{-}_{1}$ with $-\Phi$ to the part above
1700: $\gamma^{+}_{2}$. The latter surface and bundle we denote by
1701: $E_{\Phi}\rightarrow S_{\Phi}$ and the corresponding operator by
1702: $L_{\Phi}$. Letting the parameter run we obtain starting with an
1703: orientation $o$ for $L$ an orientation $o_{\Phi}$ for $L_{\Phi}$.
1704: If $o$ is the complex orientation it is easily verfied that
1705: $o_{\Phi}$ is the complex orientation as well. We  say that the operator
1706: $L_{\Phi}$  is an operator obtained
1707: from $L$ by cutting and pasting.  This operator $L_\Phi$ has the
1708: same index as $L$,  and the component of $[L,\Phi]$ in $\wt\CR$
1709: depends only the isotopy classes of the embeddings $\gamma_1$ and
1710: $\gamma_2$.
1711: 
1712: 
1713: 
1714: \bigskip
1715: 
1716: \noindent{\textsf{Axiom C3}}. $$o_{L_\Phi}=o_\Phi\, .$$
1717:  \bigskip
1718: 
1719:  Note that we have to require here that the parts of $L$ over the curves
1720:   $\gamma_1$ and $\gamma_2$ are isomorphic via the gluing data.
1721: It is straightforward to check that
1722: \begin{theorem}\label{thm:complex-coherent}
1723: The complex orientation of $\cV$ is coherent, i.e. it satisfies
1724: Axioms C1--C3.
1725: \end{theorem}
1726: 
1727: Let us point out a simple
1728:  \begin{lemma}
1729:   Let $(E,L)$ be  an isomorphism then the orientation by $1\otimes 1^{\ast}$
1730:   of the $\det
1731: L={\mathbb R}\otimes {\mathbb R}^{\ast}$ defines the complex
1732: orientation of $\cV$ over the component of $[E,L]$.
1733: \end{lemma}
1734:  The following
1735: theorem gives the converse of Theorem \ref{thm:complex-coherent}.
1736: 
1737: \begin{theorem}\label{thm:coherent-complex}
1738: Suppose that a coherent orientation of $\cV$  coincides with
1739: the complex orientation for  the trivial line bundle over $S^2$
1740: and for
1741: the line bundle over $S^2$ with Chern number $1$. Then the
1742: orientation is complex.
1743: \end{theorem}
1744: 
1745: \proof
1746: Let us first observe that according to Theorem
1747: \ref{thm:complex-coherent} the disjoint union, direct sum and
1748: cutting and pasting procedures preserve the class of complex
1749: orientations.  Consider the pair $(E_0,L_0)$, where  $E_0$
1750: is the trivial bundle $ S^2\times {\mathbb C}\to S^2$ and $L_0$ is the
1751: standard Cauchy-Riemann operator. Then the $\ind L_0=2$. Take
1752: small loops around north pole and south pole on $S^2$ and
1753:  identify the trivial bundles over these loops. Now
1754: apply the cutting and pasting procedure and Axiom C3 to obtain the
1755: disjoint union of the trivial bundle over the torus and the
1756: trivial bundle over $S^2$. Hence we can use Axiom C1 to obtain an
1757: induced orientation for the Cauchy-Riemann operator on the trivial
1758: bundle over $T^2$. Taking appropriate loops we obtain orientations
1759: for all trivial line bundles over Riemann surfaces of arbitrary
1760: genus. Using direct sums
1761:  and disjoint unions constructions, and applying Axioms C1 and C2  we see that  the orientation of
1762:   all trivial bundles
1763: of arbitrary dimensions over Riemann surfaces of arbitrary genus
1764: are complex. Let $E_1$ be the bundle over $S^2$ with Chern number
1765: $1$.  Then we can use C3 to glue  two copies of $(E_1,L_1)$ to
1766: obtain the complex orientation of the disjoint union of a Cauchy
1767: Riemann operator on the trivial bundle and one on the bundle with
1768: Chern number $2$. Now it is clear that  the given coherent
1769: orientation has to  be complex over all components of the moduli space
1770: $\CR$. \qed
1771: 
1772: In the next section we extend the coherent orientation from
1773: Cauchy-Riemann type operators over closed surfaces to
1774:   a  special class of Cauchy-Riemann type
1775: operators on Riemann surfaces with punctures.
1776: 
1777: \subsubsection{A special class of Cauchy-Riemann type
1778: operators on punctured Riemann surfaces}\label{sec:coherent-punct}
1779: 
1780: Let us view ${\mathbb C}^n$ as a real vector space equipped with
1781: the Euclidean  inner product which is the real part of the
1782: standard Hermitian inner product. We define a class of
1783: self-adjoint operators as follows. Their domain in
1784: $L^2(S^1,{\mathbb C}^n)$
1785:  is
1786: $H^{1,2}(S^1,{\mathbb C}^n)$ of  Sobolev maps $h:S^1\rightarrow
1787: {\mathbb C}^n$.  The operators have the form
1788: \begin{equation}\label{eq:asymptotic}
1789: (Ax)(t)=-i\frac{dx}{dt}-a(t)x,
1790: \end{equation}
1791:  where $a(t)$ is a smooth loop of
1792: real linear self-adjoint maps. We assume that $A$ is
1793: non-degenerate in the sense that $Ah=0$ only has the trivial
1794: solution, which just means that the time-one map $\psi(1)$ of the
1795: Hamiltonian flow
1796: \begin{equation}\label{eq:asympt-flow}
1797: \begin{split}
1798:  \dot{\psi}(t)=&i a(t)\psi(t),\\
1799:  \psi(0)=&\Id\\
1800:  \end{split}
1801:  \end{equation}
1802:  has no eigenvalues equal  to $1$.
1803:   In particular, $A:H^{1,2}\rightarrow L^2$ is an
1804: isomorphism.
1805: 
1806: Given a smooth vector bundle $E\rightarrow S^1$ we can define
1807: $H^{1,2}(E)$ and $L^2(E)$ and a class of operators $B$ by
1808: requiring that $A=\Phi B\Phi^{-1}$ for  an Hermitian
1809: trivialization $\Phi$ of the bundle $E$.  We shall call such
1810: operators {\it asymptotic}, for reasons which will become clear
1811: later.
1812: 
1813:  As it was defined in Section \ref{sec:holomorphic} above,
1814:   an asymptotically marked
1815: punctured Riemann surface is a triplet $(S, \bx , \mu )$, where
1816: $S=(S,j)$ is a closed Riemann surface,
1817:  $\bx =\{x _1,\dots,x _{s }\}$
1818: is the set  of  punctures, some of them called positive, some
1819: negative, and $\bmu=\{\mu_1,\dots,\mu_s\}$   is the set of
1820: asymptotic markers, i.e. tangent rays,
1821:  or equivalently oriented
1822: tangent lines at the punctures.
1823: 
1824: One can introduce near each puncture $x_k\in\bx$ a holomorphic
1825: parameterization, i.e. a holomorphic map $h_k:D\to S$ of the unit
1826: disk $D$ such that $h_k(0)=x_k$ and the asymptotic marker $\mu_k$
1827: is tangent to the ray $h_k(r),\;r\geq 0$. We assume that the
1828: coordinate neighborhoods $\cD_k=h_k(D)$ of all the punctures are
1829: disjoint. Then we define $\sigma_k:{\mathbb R}^+\times
1830: S^1\rightarrow {\mathcal D}\setminus\{0\}$ by $$
1831: \sigma_k(s,t)=h_k(e^{\pm{2\pi(s+it)}}), $$ where the sign $-$ is
1832: chosen if the  puncture $x_k$ is positive, and the sign $+$ for
1833: the negative puncture. We will refer to $\sigma_k$ as holomorphic
1834: polar coordinates adapted to $(x_k,\mu_k)$. Given two adapted
1835: polar coordinate systems $\sigma $ and $\sigma'$ near the same
1836: puncture $x\in\bx$
1837:   we observe that the transition map
1838: (defined for $R$ large enough) $$ \sigma
1839: ^{-1}\circ\sigma':[R,\infty)\times S^1 \rightarrow
1840: [0,\infty)\times S^1 $$ satisfies for every multi-index $\alpha$
1841: $$ D^{\alpha}[\sigma^{-1}\circ\sigma'(s,t)-(c+s,t)] \rightarrow 0
1842: $$ uniformly for $s\rightarrow\infty$, where $c$ is a suitable
1843: constant. The main point is the fact that there is no phase shift
1844: in the $t$-coordinate.
1845: 
1846: Given $(S, \bx , \mu )$ we associate to it a smooth surface
1847: $\bar{S}$ with boundary compactifying the punctured Riemann
1848: surface $S\setminus\bx$ by adjoining a circle for every puncture.
1849: Each circle has a distinguished point $0\in S^1={\mathbb
1850: R}/{\mathbb Z}$. Namely for each   positive puncture we compactify
1851: ${\mathbb R}^+\times S^1$ to $[0,\infty]\times S^1$, where
1852: $[0,\infty]$ has the smooth structure making the map $$
1853: [0,\infty]\rightarrow [0,1]:s\rightarrow s(1+s^2)^{-\frac{1}{2}},\
1854: \infty\rightarrow 1 $$ a diffeomorphism. We call
1855: $S^{+}_{k}=\{\infty\}\times S^1$ the circle at infinity associated
1856: to $(x_k,\mu_k)$. For negative punctures we compactify at
1857: $-\infty$ in a similar way.
1858: 
1859: \begin{definition}
1860: A smooth complex vector bundle $E\rightarrow  (S, \bx , \mu )$ is
1861: a smooth vector bundle over $\hat{S}$ together with Hermitian
1862: trivializations $$ \Phi_k:E|_{S_{k}}\rightarrow S^1\times{\mathbb
1863: C}^n. $$ An isomorphism between two bundles $E$ and $F$ over
1864: surfaces $S$ and $T$ is a a complex vector bundle isomorphism
1865: $\Psi:E\to F$ which covers   a biholomorphic map
1866: $\phi:(S,j)\rightarrow (T,i)$, preserves punctures and the
1867: asymptotic markers  (their numbering and signs)  and respects the
1868: asymptotic trivializations.
1869: \end{definition}
1870: 
1871: Define as in Section \ref{sec:coherent-closed} above   the bundle
1872: $X_E\rightarrow \bar{S}$. Set $\dot{S}=S\setminus\Gamma$. We
1873: introduce the Sobolev space $H^1(E)$ which  consists of all
1874: sections $h$ of $E\rightarrow \dot{S}$ of class $H^{1,2}_{loc}$
1875: with the following behavior  near   punctures.
1876:   Suppose, that $x$ is a positive puncture and   $\sigma$
1877: is an adapted system of  holomorphic polar coordinates. Pick a
1878: smooth   trivialization $\psi$ of $E\rightarrow \bar{S}$ over
1879: $[0,\infty]\times S^1$ (in local coordinates) compatible with the
1880: given asymptotic trivialization. Then the map $(s,t)\rightarrow
1881: \psi(s,t)h\circ\sigma(s,t)$ is assumed to belong to
1882: $H^{1,2}({\mathbb R}^+\times S^1,{\mathbb C}^n)$. A similar
1883: condition is required for   negative punctures. In a similar
1884: way we define  the space $L^2(X_E)$. Observe that defacto we use
1885: measures which are infinite on $\dot{S}$ and that the
1886: neighborhoods of punctures look  like half-cylinders.
1887: 
1888: A Cauchy-Riemann type operator $L$ on $E$ has the form $$ (Lh)X =
1889: \nabla_Xh+J\nabla_{jX}h +(ah)X, $$ where $X$ is a vector field on
1890: $S$.   We require, however a particular behaviour of $L$  near the
1891: punctures. Namely, regarding   $E$ as a trivial bundle
1892: $[0,\infty]\times \C^n$ with respect to the chosen polar
1893: coordinates and trivialization near say a positive  puncture we
1894: require that $$ (Lh)(  s,t )(\frac{\partial }{\partial s})
1895: =\frac{\partial h}{\partial s} - A(s)h, $$ where $A(s)\rightarrow
1896: A_{\infty}$ for an asymptotic operator $A_{\infty}$,   as  it was
1897: previously introduced.
1898: \begin{theorem}
1899: The operator $L$ is Fredholm.
1900: \end{theorem}
1901:  The index of $L$ can be computed in terms of Maslov indices of
1902:  the asympotic operators (and, of course, the first Chern class of
1903:  $E$ and the topology of $S$).
1904: 
1905: 
1906: 
1907: Similar to the case of closed surfaces we define the notion of
1908: isomorphic pairs $(E,L)$ and $(F,K)$, where we emphasize the
1909: importance of the compatibility of the asymptotic trivializations,
1910: define the moduli space $\CR_{\punct} \supset \CR_{\cl}$ and the
1911: Teichmuller spaces $\wt\CR_{\punct} \supset\wt\CR_{\cl}$, and
1912: extend the determinant line bundle $\cV$ to $\CR_{\punct}$ and
1913: $\wt\cV$ to $\wt\CR_{\punct}$. The bundle $\wt\cV$ is orientable
1914: by the same reason as in the
1915:   case of closed surfaces: each component of the space $\wt\CR_{\punct}$ is
1916: contractible. However, unlike the closed case,  there is no
1917: canonical (complex) orientation of $\wt\cV$. Still due to the
1918: requirement
1919:  that
1920: isomorphisms  preserves the end structure of the operators, one
1921: can deduce the fact that even isotopically non-trivial isomorphisms
1922: preserve the orientation of $\wt\cV$, which shows that the bundle
1923: $\cV$ over $\CR_{\punct}$ is orientable.
1924: 
1925: 
1926:  Let us review now Axioms C1--C3 for the line bundle
1927: $\cV$ over $\CR_\punct$. The formulation of Disjoint Union Axiom
1928: C1
1929:  should be appended by the  following requirement. Let $(E,L)$ and $(F,K)$ be
1930:  operators
1931:  over the punctured
1932:  Riemann surfaces $(S,\bx=\{x_1,\dots,x_s\})$ and $(T,\by=\{y_1,\dots,y_t\})$, respectively.
1933:  Then $(E,L)\dot{\cup}(F,K)$ is an operator over the surface
1934:  $S\coprod T$ with the set of punctures ${\mathbf z}=\{x_1,\dots,x_s,y_1,\dots,y_t\}$.
1935:  The disjoint union operation is associative, but not necessarily
1936:  commutative (unlike the case of closed surfaces).
1937:   Axioms C2 and C3  we formulate without any changes compared to the closed case.
1938:  By a coherent orientation of the bundle $\cV$ over $\CR_{\punct}$
1939:  we will mean any orientation of $\cV$ which satisfies Axioms
1940:  C1--C3.
1941: 
1942: Take  the trivial (and globally trivialized) line bundle $E_0=
1943: \C\times\C$ over the 1-punctured  Riemann sphere $\C=\C P^1\setminus \infty$.
1944: For any
1945: admissible asymptotic operator $A$ we choose a Cauchy-Riemann
1946: operator $L^\pm_A$ on $E_0$ which has $A$ as its asymptotics at $\infty$. The
1947: superscript $\pm$ refers to the choice of $\infty$ as the positive
1948: or negative puncture.
1949:  Note, that the   component  of $([E_0,L^\pm_A)$ in the  moduli spaces
1950:   $\CR$   is uniquely determined
1951:   by the  homotopy class $[A]$ of  the asymptotic operator $A$ in the space
1952:   of {\it non-degenerate} asymptotic operators.
1953: 
1954: The following theorem describes all possible coherent orientations
1955: of  the line bundle $\cV$ over $\CR$.
1956: 
1957: 
1958: \begin{theorem}\label{thm:coherent}
1959: Let us choose an  orientation $o^\pm_A$ of the operator
1960: $(E_0,L^\pm_A)$ for a representative $A$ of each homotopy class
1961: $[A]$ of non-degenerate  asymptotic operators. Then this choice
1962: extends to the unique coherent orientation of the bundle $\cV$
1963: over $\CR_{\punct}$, which coincide with the complex orientation
1964: over $\CR_{\cl}$.
1965: \end{theorem}
1966: Thus  there are infinitely many coherent orientations of $\cV$
1967: over $\CR_{\punct}$ unlike the  case of closed surfaces, when there are
1968: precisely four.
1969: 
1970: We sketch below  the proof of Theorem \ref{thm:coherent}.
1971:  First, similar to the case of closed surfaces, it is sufficient
1972:  to consider only operators on the trivial, and even globally
1973:  trivialized bundles.
1974:  Next take the disjoint union of
1975:  $(E_0,L^-_A)$ and $ (E_0,L^+_A)$, consider two circles $\g^\pm$ around the  punctures
1976:  in the two copies of $\C$ and apply the cutting/pasting
1977:  construction along these circles. As the result we get a disjoint
1978:  union of an operator $\wt L_A$ on the trivial line  bundle over the closed Riemann
1979:  sphere, and an operator $\overline L_A$ over  the cylinder $C=S^1\times\R$,
1980:  which we view as the Riemann sphere
1981:  with  two
1982:  punctures $x_1=\infty$ and $x_2=0$ and consider $x_1$ as a positive puncture
1983:  and $x_2$ as a negative one. The operator $\overline L_A$ has the
1984:  same asymptotic operator $A$ at both punctures.
1985:  Then Axioms C1 and C3 determine the orientation of $\overline
1986:  L_A$, because for  the operator $\widetilde L_A$ we have chosen
1987:  the complex orientation. Notice that if one glue $L_A^\pm$ in the opposite order,
1988:  then we get an operator $\overline L_A'$ which has the reverse
1989:   numbering of the punctures.
1990:   The orientation of $\overline L_A'$  determined by the gluing  may be the same, or opposite as
1991:   for the operator $\overline L_A$, depending on the parity of the Conley-Zehnder
1992:   index of the asymptotic operator $A$.\footnote{
1993:   The operator $\overline L_A$
1994:  is homotopic to an isomorphism, and thus has a canonical
1995:  orientation
1996:  $1\otimes 1^{\ast}$. If we insist on that normalization, than
1997:  our construction would determine the orientation of $L_A^-$ in
1998:  terms of $L_A^+$.}
1999: 
2000: 
2001: Consider now an arbitrary operator  $(E,L)$  acting on sections of a complex line bundle
2002: $E$ over   a punctured
2003: Riemann surface
2004:  $(S,\bx,\bmu)$ with $\bx=\{x_1,\dots,x_s\}$, $E=S\times\C^n$, and the  asymptotic operators
2005:  $A_1,\dots,A_s$ at the corresponding punctures. For each
2006:  $i=1,\dots,s$ consider an operator $(E_0,L_i=L_{A_i}^\pm)$, where
2007:  $E_0=\C\times\C^n$,  the sign $+$ is chosen if the puncture
2008:  $x_i$ is negative, and the sign $+$ is chosen  otherwise.
2009:  Using Axiom C1 we orient the operator $
2010:  (E,L)\dot{\cup}(E_0,L_s)$, and then choosing circles around the
2011:  puncture $x_s$ and $\infty$ apply the cutting/pasting procedure.
2012:  As the result we get the disjoint union of an operator $L'$ over
2013:  the Riemann surface with punctures $(x_1,\dots,x_{s-1})$ and the
2014:  operator $\overline L_A$, or $\overline L_A'$ depending on
2015:  whether the puncture $x_s$ was negative, or positive. Hence
2016:  Axioms C1 and C3 determine the orientation of $L'$ in terms of the orientation
2017:   of $L$. Repeating the
2018:  procedure for the punctures $x_{s-1},\dots,x_1$ we express
2019:   the orientation of $L$ in terms of the complex
2020:  orientation of an operator over the closed surface.
2021: 
2022:  It remains to observe that if $E$ is a trivial complex bundle of rank $r>1$,
2023:   then any asymptotic operator
2024:  $A$ can be deformed through non-degenerate asymptotic operators to an operator
2025:   $\widetilde A$ which is split into the direct
2026:  sum of asymptotic operators on the trivial complex line. Hence we can use the direct sum
2027:   axiom C2 to orient determinants of  operators  acting on bundles of arbitrary rank.
2028:  \bigskip
2029: 
2030: \subsubsection{Remark about  the coherent orientation for
2031:  asymptotic  operators with symmetries}
2032: Let $A$ be an asymptotic operator
2033: given by the formula (\ref{eq:asymptotic}),
2034: where the loop $a(t),\,t\in S^1=\R/\Z$, of symmetric matrices has a symmetry
2035: $a(t+1/2)=a(t),\;t\in\R/\Z$. Let $L$ be a Cauchy-Riemann type
2036: operator on a bundle $E\to S$, which has $A$ as its
2037: asymptotic
2038: operator at a puncture $x\in S$ with an asymptotic marker $\mu$.
2039: Let $L'$ be an operator which differs from $L$ by
2040: rotating by the angle $\pi$ the marker $\mu$ to a marker $\mu'$,
2041: with the corresponding change of the trivialization near the
2042: puncture.
2043:  Let $h:S\to S$ be  a diffeomorphism  which rotates
2044:  the polar coordinate
2045:  neighborhood $\cD$ of the punctures $x$ by $\pi$, and is fixed
2046:  outside a slightly larger neighborhood.
2047:  Then the operator $h_*L'$ has the same asymptotic data as $L$ and
2048:  the  isomorphism classes $[E,L]$ and $[E,h_*L']$ belongs to the
2049:  same component of the space $\wt\CR$. Given a coherent
2050:  orientation of $\cV$, do the orientations $o_L$ and $o_{h_*L}$
2051:  coincide? It turns out that
2052:  \begin{lemma}\label{lm:bad-orbits}
2053:  Let $\Psi$ be the time-one map of the linear Hamiltonian flow $\psi(t)$,
2054:  defined by the equation {\rm (\ref{eq:asympt-flow})}. The orientations
2055:  $o_L$ and $o_{h_*L}$ coincide if and only if the number of  real eigenvalues  of  $\Psi$
2056:  (counted with multiplicities)  from the interval $(-1,0)$ is
2057:  even.
2058:  \end{lemma}
2059: This lemma is the reason why we excluded certain periodic orbits
2060: from $\Pc$ in Section \ref{sec:dynamics} above. See also Remarks
2061: \ref{rem:bad-orbits1} and \ref{rem:bad-orbits2}.
2062: 
2063: 
2064: \subsubsection{Coherent orientations of moduli spaces}
2065: The moduli spaces of holomorphic curves which we need to orient are zero sets of nonlinear
2066: Cauchy-Riemann type operators, whose linearizations are related to
2067: operators of the kind we described (see below for more details).
2068: In general, the moduli spaces are neither manifolds nor orbifolds,
2069: due to the fact  that Fredholm sections cannot be made transversal
2070: to the zero section by changing natural parameters like the almost
2071: complex structure or the contact form. Such a transversality will
2072: only be achievable by making abstract perturbations, leading to
2073: virtual moduli spaces. Those virtual spaces will be the moduli
2074: spaces which will provide us with the data for our constructions.
2075: Nevertheless the Fredholm operators occurring in the description
2076: of the virtual moduli spaces will only be compact perturbations of
2077: the Cauchy-Riemann type operators, and hence the orientation scheme
2078: for these virtual moduli spaces does not differ from the case of moduli spaces
2079: of holomorphic curves.
2080: 
2081: 
2082: A moduli space  $\Mc(\Gamma^+,\Gamma^-;W,J)$ of  holomorphic
2083: curves in a directed symplectic cobordism $(W=\ora{V^-V^+},J)$ is
2084: a fiber bundle over the corresponding moduli space of Riemann
2085: surfaces. Its base is   a complex orbifold, and hence it is
2086:  canonically oriented, while the
2087: fiber over a point $S$, where $S$ is a Riemann surface with a
2088: fixed conformal structure
2089:  and positions
2090: of punctures,  can be viewed as the space  solutions of the $\overline
2091: \partial_J$-equation.
2092:  If the  transversality is achieved than the  tangent bundle  of a
2093:  moduli space  $\Mc(\Gamma^+,\Gamma^-;W,J)$  arise as  the kernel of the linearized
2094:  surjective  operator $\overline\partial_J$.
2095: The linearization  of $\overline
2096: \partial_J$
2097:   at a point $f\in \Mc(\Gamma^+,\Gamma^-;W,J)$
2098: is a Fredholm operator in a suitable functional analytic setting.
2099: This set-up involves Sobolev spaces with suitable asymptotic
2100: weights derived from the non-degeneracy properties of the periodic
2101: orbits. It is a crucial observation, again a corollary of the
2102: behaviour near the punctures, that up to a compact perturbation,
2103: the operator $L$ splits into two operators $L'$ and $L''$, where
2104: $L'$ is   a complex linear operator acting on the complex line
2105: bundle $T(S)$ of the Riemann surface $S$, and $L''$ is a
2106: Cauchy-Riemann  type on the the bundle $E$, such that $T(S)\oplus
2107: E=f^*(TM)$. This operator is usually only real linear, but most
2108: importantly it is of the kind we just described in our linear
2109: theory.
2110:   The trivialization of $E$ near the punctures
2111: is determined by the chosen in \ref{sec:holomorphic}
2112: trivialization
2113:  of the contact structure near
2114: periodic orbit of   the Reeb vector fields on $V^\pm$, and the
2115: asymptotic operators are determined by the linearized Reeb flow
2116: near the periodic orbits. We have $\det L=\det L'\otimes\det L''$.  But  $\det L'$ has a canonical complex
2117: orientation, and   hence the orientation for $\det L$ is  determined by   the orientation of
2118:  $\det L''$. Therefore, a choice of
2119:   a coherent orientation of $\cV$ over $\CR$ determines in the
2120: transversal case the orientation of  all the moduli spaces
2121: $\Mc(\Gamma^+,\Gamma^-;W,J)$.
2122: 
2123: 
2124: %%%%%%%%%%%%%%%%%%%%%%%
2125: 
2126: 
2127: 
2128: 
2129: 
2130: 
2131: 
2132: 
2133:    \subsection{First attempt of algebraization:
2134:    Contact Floer homology}  \label{sec:Floer}
2135:    \subsubsection{Recollection of finite-dimensional Floer theory} \label{sec:Morse}
2136:    Let us  first recall   the basic steps in defining a Floer homology theory in the simplest case of a
2137: Morse function $f$ on a finite-dimensional orientable closed manifold
2138: $M$. We refer the reader
2139: to  Floer's  original papers (see, for instance, \cite{Floer}), as well as an excellent exposition by D. Salamon  \cite{Salamon:Floer}
2140: for the general theory.
2141: 
2142:  First,  one forms a  graded complex $C(f,g)$ generated by
2143: critical points  $c_1,
2144: \dots,c_N$ of      $f$, where the grading is given by the Morse index of critical points.
2145:  Next, we  choose a  generic Riemannian metric $g$ on $M$  which satisfy the Morse-Smale condition of transversality
2146:  of stable and unstable varieties
2147:  of critical points. This enables us to define  a differential $d=d_{f,g}:C(f,g)\to C(f,g)$ by counting  gradient trajectories
2148:  connecting critical points   of neighboring indices:
2149:  $$d(c_i)=\sum L_i^jc_j,$$ where the sum is taken
2150:  over all critical points $c_j$  with   $\ind c_j=\ind c_j-1$. The coefficient  $L_i^j$
2151:  is the {\it  algebraic number}  of    trajectories    connecting $c_i$ and $c_j$.
2152:   This means that the trajectories are counted
2153:  with signs.   In the finite-dimensional case the signs could  be determined
2154:  as follows. For each critical point we orient arbitrarily its stable manifold. Together with
2155:  the orientation of $M$ this allows us to orient  all unstable manifolds,
2156:   as well as the intersections
2157:  of stable and unstable ones.   If $\ind c_j=\ind c_i-1$ then
2158:  the stable manifold of $c_i$ and the unstable manifold
2159:  of $c_j$ intersect along finitely many trajectories which   we want to count,
2160:   and hence each of these trajectories
2161:  gets an orientation. Comparing this orientation with the one given by the direction of the gradient $\nabla f$ we can associate
2162:  with every trajectory a sign. \footnote{The generalization of this procedure to an infinite-dimensional
2163:  case is not straightforward, because stable and unstable manifolds not only can become
2164:  infinite-dimensional, but in most interesting cases cannot  be defined at all.  On the other hand,
2165:   the moduli spaces of
2166:  gradient trajectories connecting pairs of critical points  (which in the finite-dimensional case
2167:  coincide with the intersection
2168:  of stable and unstable manifolds of the critical points) are often defined, and one can use the coherent
2169:  orientation scheme, similar to the one described in Section \ref{sec:orientation}
2170:  above for the moduli spaces of holomorphic curves,
2171:   to define their orientation.}
2172: 
2173:   To show that $d^2=0$, which then  would allow us to define the homology group
2174:   $H_*(C(f,g),d)$, we proceed as follows.
2175: Let us observe that  the coefficients $K_i^j$ in the expansion $d^2(c_i)=\sum K_i^jc_j$
2176:   count  the algebraic number of broken gradient trajectories  $(\d_{il},\d_{lj})$  passing through an intermidiate
2177:   critical point $c_l$, $l=1,\dots N$.
2178:  But each broken trajectory    $(\d_{il},\d_{lj})$, which connects critical points whose
2179:  indices differ by $2$, is a boundary point
2180:  of the $1$-dimensional manifold of smooth trajectories connecting $c_i$ and $c_j$. The  algebraic number of boundary points of a compact
2181:  $1$-dimensional  manifold is, of course, equal to $0$. Hence $K_i^j=0$, and thus $d^2=0$.
2182: 
2183: 
2184:  Next we want to show that the homology group $H_*(C(f,g),d)$ is  an invariant of the manifold $M$
2185:  (of course, in the case we consider it is just $H_*(M)$), i.e. it is independent of the choice of the function
2186:  $f$ and the Riemannian metric $g$.  The proof of the invariance consists of three steps.
2187: \medskip
2188: 
2189: \noindent \textsf{Step 1.}
2190: Let us show that given a homotopy  of functions $F=\{f_t\}_{t\in[0,1]}$,   and a homotopy  of Riemannian metrics
2191:  $G=\{g_t\}_{t\in[0,1]}$, one can define a homomorphism $\Phi=\Phi_{F,G}:C(f_1,g_1)\to C(f_0,g_0)$ which commutes with
2192:  the boundary homomorphisms $d_0=d_{f_0,g_0}$ and $d_1=d_{f_1,g_1}$, i.e.
2193:   \begin{equation}\label{eq:Morse-boundary}
2194:   \Phi\circ d_1-d_0\circ\Phi=0\,.
2195:   \end{equation}
2196:          To construct $\Phi$ we consider the product
2197:          $W=M\times\R $ and, assuming that the homotopies $\{f_t\}$ and $\{g_t\}$ are
2198:          extended to all $t\in\R$  as independent  of $t$ on
2199:          $(-\infty,-1]\cup[1,\infty)$, we
2200:           define on $W$
2201:         a function, still denoted by $F$,
2202:         by the formula
2203:          \begin{equation*}
2204:            F(x,t)=
2205:            \begin{cases}
2206:            f_0(x)+ct,&  \;\;t\in (\infty,0) ;\\
2207:            f_t(x)+ct,&  \;\;t\in[0,1]\;\;  ;\\
2208:            f_1(x)+ct,&  \;\;t\in(0,\infty) ,\\
2209:                         \end{cases}
2210:            \end{equation*}
2211: where the constant $c$ is chosen to ensure that $\frac{\partial
2212: F}{\partial t}>0$. Similarly, we use the family of Riemannian
2213: metrics $g_t$
2214: to define a metric $G$ on $W$  which is equal to $g_t$ on $M\times
2215: t$ for all $t\in\R$, and such that $\frac{\partial}{\partial t}$ is the unit vector field
2216: orthogonal to the slices $M\times t,t\in\R$.
2217: The gradient trajectories of $\nabla F$ converge to
2218: critical points of $f_1$ at $+\infty$ , and  to the critical points
2219: of $f_0$ at $-\infty$. For a generic choice of $G$ the moduli space of the
2220: (unparameterized)
2221: trajectories connecting two critical points, $c^1$ of $f_1$ and
2222: $c^0 $ of $f_0$, is a compact $k$-manifold with boundary with corners,
2223: where $k=\ind c^1-\ind c^0$. Hence,
2224: similarly
2225:  to the above definition of the differential $d$, we can define a homomorphism
2226: $\Phi:C(f_1,g_1)\to C(f_0,g_0)$ by  taking an algebraic count of gradient trajectories between
2227: the critical point of $f_1$ and $f_0$ of the same Morse index,
2228: i.e.
2229: $\Phi(c^1_j)=\sum \wt L^i_jc^0_j$. The identity
2230: (\ref{eq:Morse-boundary}) comes from the  description of the  boundary
2231: of the $1$-dimensional moduli spaces of trajectories of $\nabla
2232: F$. Notice that the function $F$ has no critical points, and hence
2233: a family of gradient trajectories cannot converge to a broken
2234: trajectory in a usual sense. However, this can happen {\it at
2235: infinity}. Let us recall that the function $F$ and the metric $G$
2236: are cylindrical outside of  $M\times[-1,1]$. Hence
2237: away from a compact set  a gradient
2238: trajectory of $F$  projects to a gradient
2239: trajectory of $f_0$ or $f_1$. When  the projection, say at $+\infty$,
2240:  of a sequence  $\d_n:\R\to W$ of trajectories of $\nabla F$   converges  to
2241: a broken trajectory  of    $\nabla f_1$ this can be interpreted as a
2242: splitting at $+\infty$. This phenomenon is very similar to  the one  described
2243: for the moduli spaces of holomorphic curves in Section
2244: \ref{sec:compact}. Namely, there exist    gradient  trajectories $\d:\R\to W $  of $\nabla F$,
2245: and $\d':\R\to M_1$ of $\nabla f_1$,  such that
2246: \begin{description}
2247: \item{$-\;$} $\d_n\to\d$ uniformly on $(-\infty,C]$ for all $C$;
2248: \item{$-\;$} there exists a sequence $C_n\to+\infty$ such that
2249: $\d'_n(t)=\d_n(t+C_n)$ converges to $(\d'(t),t)$ uniformly on all
2250: subsets $[-C,\infty)$.
2251:  \end{description}
2252: 
2253: In this sense broken trajectories of the form $(\d,\d')$ and
2254: $(\d'',\d)$, where $\d''$ is a   trajectory of $\nabla f_0$ form
2255: the boundary of the $1$-dimensional moduli spaces of trajectories
2256: of $\nabla F$ connecting critical points  $c^1$ of $f_1$ and
2257: $c^0 $ of $f_0$  with $\ind c^1-\ind c^0=1$. Therefore the
2258: algebraic number of these trajectories equals $0$. On the other hand,
2259: this number is equal   to $\Phi\circ d_1-d_0\circ\Phi $  which yields
2260: the identity (\ref{eq:Morse-boundary}).
2261: \medskip
2262: 
2263: \noindent\textsf{Step 2.} Our next goal is to check that  if
2264: $(F_u,G_u), u\in[0,1],$  is a homotopy of homotopies which is
2265: constant outside of a compact subset of $W$, then
2266:   the homomorphisms $\Phi_0=\Phi_{F_0,G_0}$ and $\Phi_1=\Phi_{F_1,G_1}$ are related via the
2267: chain homotopy formula
2268: \begin{equation}\label{eq:Morse-chain}
2269: \Phi_1-\Phi_0=K\circ d_1+d_0\circ K,
2270: \end{equation}
2271: for a homomorphism $K:C(f_1,g_1)\to C(f_0,g_0)$. The space of all homotopies $(F,G)$
2272: connecting given pairs $(f_0,g_0)$ and $(f_1,g_1)$ is contractible,
2273:  and hence (\ref{eq:Morse-chain}) implies
2274:   that the homomorphism $\Phi_*:H_*(C(f_1,g_1),d_1)\to
2275: H_*(C(f_0,g_0),d_0)$ is independent of the choice of a homotopy
2276: $(F,G)$.
2277: 
2278: To prove (\ref{eq:Morse-chain}) one studies moduli spaces  of
2279: gradient trajectories  of the whole $1$-parametric family of
2280: functions $F_u$. For a generic choice of the homotopy one has
2281: isolated critical values of the parameter $u$  when appear
2282: {\it handle-slides}, i.e. gradient connections between critical
2283: points with the index difference $-1$.  By  counting these
2284: trajectories
2285: one can then define a
2286: homomorphism $K:C(f_1,g_1)\to\C(f_0,g_0)$ in exactly the same way as the homomorphism $\Phi$
2287: was  defined in Step 1
2288:    by counting   trajectories   with the index difference  $0$.
2289: 
2290:  The identity (\ref{eq:Morse-chain}) expresses  the fact that the broken
2291:  trajectories  of the form $(\d,\d')$ and $(\d'',\d)$, where $\d$
2292:  is a handle-slide trajectory and $\d'$ is a trajectory of $\nabla
2293:  f_1$, form the boundary of the moduli space  of   index $0$
2294:  trajectories in the family $(F_u,G_u)$. The difference in signs
2295:  in formulas (\ref{eq:Morse-boundary}) and (\ref{eq:Morse-chain})
2296:  is a reflection of the fact that the homomorphism $K$ raises the grading by
2297:  $1$, while $\Phi$ leaves it unchanged.
2298: 
2299: 
2300: 
2301: 
2302: 
2303: \medskip
2304: \noindent\textsf{Step 3.} Finally we need to show that
2305: \begin{equation}
2306: \label{eq:Morse-composition}
2307: (\Phi_{F,G})_*=(\Phi_{F',G'})_*\circ(\Phi_{F'',G''})_*,
2308: \end{equation}
2309: if
2310:  $(F,G)=\{f_t,g_t\}_{t\in[0,2]}$ is the composition of homotopies
2311:  $(F'',G'')=\{f''_t,g''_t\}_{t\in[0,1]}$ and
2312:  $(F',G')=\{f'_t,g'_t\}_{t\in[1,2]}$. To prove this we  view,
2313:  as in Step 1, the  homotopy $(F,G)$ as a function and a metric
2314:  on the cylinder $W=M\times\R$. Consider a deformation $(F_T,G_T)$ of $F,G$, by
2315:  cutting $W$ open along $M\times 1$ and inserting a cylinder
2316:  $M\times[0,T]$ of growing height $T$ with the function and the
2317:  metric independent of the coordinate $t$. When $T\to+\infty$
2318:  the gradient trajectories of $F_T$ with respect to $G_T$  split  in
2319:  a  sense, similar to  the one explained in Step 2,
2320:  \footnote{See
2321:  also the discussion of  a similar phenomenon for the
2322:  moduli spaces of holomorphic curves   in Section \ref{sec:compact}
2323:  above.} into a ''broken trajectory"
2324:  $(\d'',\d')$, where $\d'$  (resp.  $\d"$) is a   trajectory of $\nabla_{G'}F'$
2325: ( resp. $\nabla_{G''}F''$).  Consider the $1$-dimensional moduli  space $\Mc$ of
2326: trajectories  of $\nabla_{G_T}F_T, T\in[0,\infty),$ connecting a
2327: fixed
2328: critical point  $c=c^2$ of $f_2$  with an arbitrary critical point
2329: $c^0 $ of $f_0$  with $\ind c-\ind c^0=1$. Then the boundary of
2330: $\Mc$ consists of
2331: \begin{description}
2332: \item{ a) } all the trajectories of
2333: $\nabla_{G_0}F_0=\nabla_G F$ connecting $c^0$ and $c$; they are
2334: given by the expression $\Phi(c)$;
2335: \item{ b) } all the broken trajectories $(\d'',\d')$ described
2336: above, such that $\d''$ begins at $c^0$ and ends  at a critical point
2337: $c^1$ of $f^1$ which is, necessarily, of the same Morse index as
2338: $c^0$ and $c^2$, $\d'$ begins at $c^1$ and ends at
2339: $c$; these broken trajectories are described by the
2340: expression
2341: $\Phi_{F',G'}\big(\Phi_{F'',G''}(c)\big)$;
2342: \item{ c) } broken trajectories defined according to Step 2 for the
2343: $1$-dimensional family $F_T, T\in[0,\infty)$; they are described
2344: by the expression $K(d_0(c))+d_2( K(c))$ for some homomorphism
2345: $K:C(f_2,g_2)\to C(f_0,g_0)$.
2346: \end{description}
2347:  Thus the sum (taken with appropriate signs) of the three
2348:  expressions defined in a)--c)  equals $0$, and thus we get
2349:  \begin{equation*}
2350: \Phi_{F',G'}\big(\Phi_{F'',G''}(c)\big)-\Phi(c )=K\circ d_0(c)+d_2\circ
2351: K(c),
2352: \end{equation*}
2353: i.e. the homomorphisms $\Phi$ and
2354: $\Phi_{F',G'}\circ\Phi_{F'',G''}$ are chain homotopic, which
2355: yields formula (\ref{eq:Morse-composition}).
2356: 
2357: We can finish now the proof that the homology group  $H_*(C(f,g),d)$
2358: is independent of the choice of $f$ and $g$ as follows.
2359:  Given two pairs   $(f_0,g_0)$ and  $(f_1,g_1)$ we
2360:  first take any homotopy $(F,G)$  connecting  $(f_0,g_0)$ with  $(f_1,g_1)$, and also take  the
2361:  inverse homotopy $(\overline F,\overline G)$ connecting  $(f_1,g_1)$ with
2362:  $(f_0,g_0)$.
2363:   The composition $(\wt F,\wt G)$ of  the  homotopies $(F,G)$ and  $(\overline F,\overline G)$
2364:   connects  the pair $(f_0,g_0)$ with itself. According to Step 3
2365:   we have
2366:   $(\Phi_{\wt F,\wt G})_*=(\Phi_{F,G})_*\circ(\Phi_{\overline
2367:   F,\overline
2368:   G})_*$. On the other hand, we have  shown in Step 3  that the
2369:   homomorphism $(\Phi_{\wt F,\wt G})_*$ is independent of the
2370:   choice of a homotopy, connecting $(f_0,g_0)$ with itself, and
2371:   hence it equals the identity. Therefore, we conclude that $(\Phi_{F,G})_*$ is
2372:  surjective, while $ (\Phi_{\overline
2373:   F,\overline
2374:   G})_*$ is injective. Taking the composition of homotopies  $(\overline F,\overline G)$
2375:   and $(F,G)$ in the opposite order we prove that both
2376:   homomorphisms are bijective.
2377:   \bigskip
2378: 
2379:  A. Floer  discovered  that the
2380:     finite-dimensional scheme  which we explained  in this section  works,
2381:      modulo some
2382:     analytic  complications, for several geometrically interesting functional on infinite-dimensional
2383:     spaces. For instance,
2384:   in the  symplectic Floer homology theory one deals with
2385:   critical points of the action functional. Its critical points are periodic orbits of
2386:   a Hamiltonian system, while for an appropriate choice of
2387:   a metric  and an almost complex structure the  gradient trajectories    can be interpreted
2388:   as holomorphic
2389:   cylinders   which connect these trajectories.
2390:   The role of broken trajectories is played here  by split holomorphic cylinders,
2391:    and    finite-dimensional  compactness theorems are replaced by  the  highly
2392:    non-trivial Gromov  compactness theorem for holomorphic curves.
2393: 
2394:    In the  rest of this section we explore  the  Floer-theoretic approach for
2395:    the problem of defining invariants of contact manifolds. We will see that this approach
2396:    works only  in a very special and restrictive   situation. However, the  general
2397:    algebraic formalism of SFT, though quite different, has  a distinctive flavor    of
2398:    a Floer homology theory.
2399: 
2400: 
2401: 
2402: \subsubsection{Floer homology for the Action functional}
2403:    Let us make an attempt to define invariants of contact manifolds
2404:    in the spirit of Floer homology theory. Let $(V,\xi)$     be a contact manifold
2405:          with a fixed contact form $\alpha$ and an  almost complex structure
2406:          $J:\xi\to\xi$,   compatible
2407:     with the symplectic form $d\alpha|_\xi$.  Then $J$ and $d\alpha$
2408:     define a Riemannian metric on the vector bundle $\xi$
2409:     by the formula $g(X,Y)=d\alpha(X,JY)$ for any vectors $X,Y\in\xi$.
2410:     We extend $g$ to the whole tangent bundle $T(V)$ by declaring the vector field $R_\alpha$ to be the unit normal
2411:     field to $\xi$.  Consider the free loop space
2412:     $$\Lambda(V)=\{u:S^1=\R/\Z\to V\},$$
2413: and define the {\it action functional}
2414:    \begin{equation}
2415:    S:\Lambda(V)\to\R\quad\hbox{  by the formula}\quad
2416:     S(\g) =\int\limits_\gamma \alpha .
2417:     \end{equation}
2418:    The least action principle tells us that the critical points of the functional $S$
2419:    are, up to parameterization,   the
2420:    periodic orbits of the Reeb field $R_\a$.
2421: 
2422: 
2423:      The metric $g$ on $T(V)$ defines a        metric on $\Lambda(V)$
2424:       and thus allows us to consider
2425:     gradient trajectories of the action functional connecting critical points of $V$.
2426:       The gradient direction $\nabla S(u)$, $u\in \Lambda(V)$, is given
2427: by the vector field $J\pi( \frac{du}{dt})$, where $\pi:T(V)\to\xi$
2428: is the projection along the Reeb direction, so that a gradient
2429: trajectory $u(t,s),\,t\in\R/\Z,s\in\R,$ is given by the equation
2430: \begin{equation}\label{eq:CR}
2431: \frac{\partial u}{\partial s}(t,s)=J\pi\big(\frac{\partial u}{\partial
2432: t}(t,s)\big)\,.
2433:  \end{equation}
2434: Equation (\ref{eq:CR}) has a flavor of a Cauchy-Riemann equation.
2435: We want to modify it into a genuine one.
2436: Namely, consider the Cauchy-Riemann equation
2437:  \begin{equation*}
2438:   \frac{\partial U}{\partial t}(t,s)=J\frac{\partial U}{\partial
2439:   s}(t,s)
2440:   \end{equation*}
2441:   for $U(s,t)=   \big(u(s,t),\varphi(s,t)\big)\in V\times\R$.
2442:   It  can be rewritten
2443:   as a system
2444: \begin{equation}\begin{split}\label{eq:CR2}
2445: \frac{\partial u}{\partial s}(t,s)&=J\pi\big(\frac{\partial u}{\partial
2446: t}(t,s)\big)+\frac{\partial \varphi}{\partial t}(t,s)R_\a(u(t,s)\\
2447:  \frac{\partial \varphi}{\partial s}(s,t)&=-\big\langle\frac{\partial
2448: u}{\partial t}(t,s),R_\a(u(t,s))\big\rangle\,.\\
2449: \end{split}
2450:  \end{equation}
2451: Notice that $dS(\nabla S+\psi R_\a)\geq 0$ for any function $\psi(t,s)$.
2452:  Hence,
2453:    the first
2454:   equation of the system (\ref{eq:CR2})  can be viewed as the flow equation of the
2455:   gradient-like vector-field $\nabla S+\frac{\partial \varphi}{\partial
2456:   t} R_\a$. Trajectories  of  this gradient like field   connecting critical points
2457:   $\gamma^-,\gamma^+$ of the action functional   correspond to
2458:    elements
2459:   of the moduli space  $\Mc_0(\gamma^-,\gamma^+;W,J)$, and
2460:   therefore the Floer homology philosophy (\cite{Floer}),
2461:   which we described above in the finite-dimensional
2462:   case, suggests the following construction.
2463: 
2464: 
2465: 
2466: 
2467: 
2468:   Let us associate a variable $q_\gamma$ with every periodic orbit $\gamma\in\Pc_\a$
2469:   and assign to it the grading $$\deg{ q_\gamma}=\CZ(\gamma)+(n-3).$$
2470:   The choice of the   constant $n-3$ is not important
2471:   for purposes of this definition, but
2472:  it will become important for generalizations considered in the second
2473:  part of this paper.
2474: 
2475:   Let  $A$ be  the group algebra $ \C\,[H_2(V)]$. We will fix a basis
2476:   $A_1,\dots, A_N$ of $H_2(V;\C\,)$ and identify each homology class $\sum d_iA_i$ with its {\it degree}
2477:   $d=(d_1,\dots, d_N)$. Thus  we can view the algebra $A$ as
2478:   the  algebra of  Laurent polynomials
2479:    of $N$ variables $z_1,\dots,z_N$ with complex coefficients, and write its elements
2480:    in the form $\sum a_dz^d,$ where $z^d=z_1^{d_1}\cdots
2481:    z_N^{d_N}$. The variables $z_i$ are also considered graded,
2482:    $\deg\,z_i=-2c_1(A_i),\,i=1,\dots,N$.
2483:    Consider a complex $\fF$ generated by the (infinitely many) graded variables $q_\gamma$ with coefficients
2484:   in the graded algebra $A$,
2485:   and define  a differential $\partial: \fF\to\fF $ by the formula:
2486:   \begin{equation} \label{eq:Floer1}
2487:   \partial q_\gamma= \sum\limits_{\g',d}\frac{n_{\gamma,\gamma',d}}{\kappa_{\g'}}z^d  q_{\gamma'},
2488:   \end{equation}
2489:   where $\kappa_{\gamma'}$ denotes the multiplicity of the orbit
2490:   $\gamma'$,
2491:    the sum is taken over all trajectories
2492:   $\g'\in\Pc_\a $ and $d=(d_1,\dots,d_N)$ with $$\CZ(\g')=\CZ(\g)+2\langle
2493:   c_1,d\rangle-1,$$
2494:    and the coefficient $n_{\gamma,\gamma',d}$ counts
2495:   the algebraic number of components of the $0$-dimensional moduli space
2496:   $\Mc^d_0(\gamma',\gamma;W,J)/\R
2497:   $.\footnote{
2498:   Let us recall that  according to our definition of the moduli
2499:   space $\Mc^d_0(\gamma',\gamma;W,J)/\R
2500:   $ the coefficient $n_{\gamma,\gamma',d}$ counts
2501:   equivalence classes of holomorphic curves {\it with asymptotic
2502:   markers}, and hence     each holomorphic cylinder connecting $\g$ and
2503:   $\g'$ is counted $\kappa_\g\kappa_{\g'}$ times, unless the
2504:   cylinder itself is
2505:   multiply covered. The role of the denominators $\kappa_{\g'}$
2506:   in formula (\ref{eq:Floer1}), as well in a similar
2507:   formula (\ref{eq:Floer2}) below, is to correct this ``over-counting". }
2508:   Notice that the Liouville flow of the vector field  $\frac{\partial}{\partial t}$
2509:     defines  a $\R$-action on
2510:   the moduli spaces    $\Mc^d_0(\gamma',\gamma;W,J)$, which makes the
2511:    $1$-dimensional components of the moduli
2512:   spaces
2513:     canonically oriented.
2514:    Comparing this orientation
2515:   with the coherent orientation we produce  signs   which we use in the formula
2516:   (\ref{eq:Floer1}).
2517: 
2518: To simplify the  assumptions in the propositions which we formulate below
2519: we will  assume for the rest of this section that $c_1|_{\pi_2(V)}=0$.
2520:  This assumption allows us to define for any
2521:  contractible periodic orbit $\g$ the Conley-Zehnder index
2522:  $\CZ_{\disk}(\g)$ computed with respect to {\it any disk}
2523:  $\Delta$ spanned by $\g$ in $V$. We denote
2524:  $\deg_{\disk}(\g)=\CZ_{\disk}(\g)+n-3$.
2525: 
2526:   \begin{proposition}\label{prop:Floer1}
2527:   If for a  contact form $\alpha$ the Reeb
2528:   field $R_\alpha$ has no contractible periodic orbits $\g\in\Pc_\a$  with
2529:   $\deg_{\disk}(\g)=1$, then
2530:   $\partial^2=0$.
2531:   \end{proposition}
2532: 
2533: {\sl Sketch of the proof}. Similarly to the finite-dimensional
2534: case considered in Section \ref{sec:Morse} above the identity
2535: $\partial^2=0$ in Floer homology  is equivalent to the fact that
2536: the codimension $1$ stratum of the compactified moduli spaces
2537: $\Mc^d_0(\gamma',\gamma)$   consists of broken
2538: trajectories, which in our case are represented by the height $2$
2539: stable curves $(f_1,f_2)$, $f_1\in\Mc^{d'}_0(\gamma',\gamma'')/\R$, $f_2\in
2540: \Mc^{d''}_0(\gamma'',\gamma)/\R$, where $d=d'+d''$. However, in  the general case a sequence of
2541: holomorphic cylinders in $\Mc^d_0(\gamma',\gamma)$ can split into curves
2542: different from cylinders, as it is stated in Proposition \ref{prop:boundary} and  Corollary
2543: \ref{cor:pos-puncture}. But
2544:  if this happens then the  first-floor curve
2545: $f_1$ must have  a component which is conformally equivalent to $\C\,$ and
2546: asymptotically cylindrical over a  contractible orbit at
2547: $+\infty$. Moreover, if $(f_1,f_2)$ belongs to a top-dimensional
2548: stratum of the boundary of the moduli space
2549: $\Mc^d_0(\gamma',\gamma)$, then $\deg_{\disk}(\g)=1$, which
2550: contradicts our assumption.
2551: 
2552: \begin{remark}\label{rem:bad-orbits1}{\rm
2553: Let us recall that we excluded from $\Pc$ certain ``bad" periodic
2554: orbits (see  the footnote in Section \ref{sec:dynamics}).
2555: However on the boundary of the moduli space
2556: $\Mc^d_0(\gamma',\gamma)$  there could be a stratum
2557:  which consists of height $2$
2558: stable curves $(f_1,f_2)$, $f_1\in\Mc^{d'}_0(\gamma',\gamma'')$, $f_2\in
2559: \Mc^{d''}_0(\gamma'',\gamma')$, where the orbit $\gamma''$ is
2560: one of  the bad orbits which we excluded from $\Pc$. The
2561: orbit $\g''$ has even multiplicity $2k$, and hence on the boundary
2562: of   $\Mc^d_0(\gamma',\gamma)$ there are $2k$ strata which
2563: correspond
2564: to
2565: $2k$ different possible positions of the asymptotic marker at
2566: the punctures mapped to $\g''$. The Poincar\'e return map of the Reeb flow
2567: along the orbit $\g''$ has an odd number of eigenvalues in the interval $(-1,0)$,
2568: and hence  according to Lemma \ref{lm:bad-orbits}  the coherent orientation
2569: will automatically assign to these orbits opposite signs, which means that these strata
2570: will not contribute to the sum  (\ref{eq:Floer1}). This explains why the exclusion of bad orbits
2571: is {\em possible}. Remark \ref{rem:bad-orbits2} below explains why this
2572: exclusion is {\em necessary}.}
2573: \end{remark}
2574: 
2575: Now we follow Steps 1--3 in Section \ref{sec:Morse} above to show the independence
2576: of   the homology group   $$\oplus H_k(\fF ,\partial )= \Ker\partial /{\mathrm {Im}}\partial, $$
2577:     graded by the degree $k$,  of the choice of
2578:   a  nice contact form $\a$ and  a compatible almost complex
2579:   structure  $J$.
2580: 
2581: Suppose now that we have a  directed symplectic cobordism
2582: $W=\ora{V^-V^+}$,  and $J$  is a compatible almost complex structure on $W$.
2583: Suppose that the inclusions $V^\pm\hookrightarrow W$ induce
2584: isomorphisms on $2$-dimensional homology.  Then we can define
2585:   a homomorphism $\Phi=\Phi_W:\fF^+\to\fF^-$ by the formula
2586: \begin{equation}\label{eq:Floer2}
2587: \Phi( q_\gamma)= \sum\limits_{\g',d} \frac{1}{\kappa_{\gamma' }}n_{\gamma,\gamma',d}z^dq_{\gamma'},
2588: \end{equation}
2589:   where the sum is taken over all trajectories $\g'\in\Pc^- $  and $d$
2590:   with
2591:   $\CZ(\g')=\CZ(\g)+2\langle c_1,d\rangle$, and
2592:   the coefficient $n_{\gamma,\gamma',d}$ counts
2593:   the algebraic number of points of the compact $0$-dimensional moduli space $\Mc^d_0(\gamma',\gamma;W,J)
2594:   $.
2595:    If the   condition on the second homology is not satisfied then the above construction
2596:   gives us only a correspondence, rather than
2597:   a homomorphism. See Section \ref{sec:composition}  for the discussion of a
2598:    more general case.
2599: 
2600: \begin{proposition}\label{prop:Floer2}
2601: Suppose that the contact forms   $\a^\pm$ associated to the
2602: ends satisfy the condition  $\deg_{\disk}(\g)\neq 0,1$ for any {\em contractible in $W$} periodic orbit
2603:  $\g\in\Pc^\pm$. Then the homomorphism $\Phi_W$ commutes
2604: with $\partial $.
2605: \end{proposition}
2606: \begin{proposition}\label{prop:Floer3}
2607:  Let $J_t,t\in[0,1] $, be a family of  almost complex structures compatible with
2608:  the directed symplectic cobordism
2609: $W=\ora{V^-V^+}$. Suppose  that the forms   $\a^\pm$ associated to the
2610: ends satisfies the condition  $\deg_{\disk}(\g)\neq -1,0,1$ for any  {\em contractible in $W$} periodic orbit
2611:  $\g\in\Pc^\pm$. Then the  homomorphisms
2612:  $\Phi_0=\Phi_{W,J_0}$ and  $\Phi_1=\Phi_{W,J_1}$  are chain
2613:  homotopic, i.e there exists a homomorphism $\Delta:\fF^+\to\fF^-$
2614:  such that $\Phi_1-\Phi_0=\partial \Delta +\Delta\partial$.
2615:  \end{proposition}
2616: \begin{proposition}\label{prop:Floer4}
2617: Given two cobordisms $W_1$ and $W_2$, and a compatible almost complex structure
2618: $J$ on the composition $W_1\circledcirc W_2$, the homomorphism
2619: $\Phi_{W_1\circledcirc W_2}$ is chain-homotopic to
2620: $\Phi_{W_1}\circ\Phi_{W_2}$.
2621: \end{proposition}
2622: 
2623: Together with an obvious remark that for the cylindrical cobordism   $W_0$
2624: the homomorphism $\Phi_{W_0}$ is the identity,
2625:   Propositions \ref{prop:Floer1}--\ref{prop:Floer3}
2626:   imply that if a contact structure $\xi$ on $V$ admits a {\it
2627:   nice} contact form, i.e. a form without contractible periodic
2628:   orbits of index $-1,0$ and $1$, then the {\it contact homology
2629:   group}
2630:   $$\oplus HC_k(V,\xi)=\oplus H_k(\fF ,\partial ) $$
2631:    is well defined and
2632:     independent of the choice of
2633:   a  nice contact form and  a compatible almost complex
2634:   structure
2635:   (however, if $H_2(V)\neq 0$ and/or $H_1(V)\neq0$ it depends on a
2636:   choice of spanning surfaces $F_\g$ and the  framing of the bundle
2637:   $\xi$ over basic loops).
2638:    Similarly to what was explained in the sketch of
2639:      the proof  Proposition
2640:       \ref{prop:Floer1} the  ``niceness" assumptions   guarantees
2641:       that the  top codimension strata   on the boundary of the involved
2642:       moduli spaces consist of height 2 cylindrical curves,
2643:       and thus         the proofs of Propositions  \ref{prop:Floer2}--\ref{prop:Floer4}
2644:       may precisely follow the standard scheme of the Floer
2645:       theory (see \cite{Floer, Salamon:Floer}).
2646: \begin{remark}\label{rem:bad-orbits2}{\rm
2647: Similarly to what we explained in Remark \ref{rem:bad-orbits1} the
2648: coefficient $n_{\g,\g'}$ in the definition (\ref{eq:Floer2}) of
2649: $\Phi$ equals $0$ if at least one of the orbits $\g,\g'$ is ``bad''.
2650: Hence, in the presence of ``bad" orbits the
2651: homomorphism $\Phi$  could never be  equal to the  identity, even
2652: for the cylindrical cobordism. This explains why
2653: the exclusion of ``bad" periodic orbits is {\em necessary}.}
2654: \end{remark}
2655: 
2656: 
2657: 
2658: 
2659:     Besides the degree (or Conley-Zehnder)  grading,
2660:    the contact homology group is graded by elements of $H_1(V)$, because
2661:    the boundary operator preserves the homology class of a periodic orbit.  We will denote the part of  $HC_*(V,\xi)$
2662:    which correspond to a class $a\in H_1(V)$ by $HC_*(V,\xi|a)$.
2663:   One can  similarly construct a contact homology group
2664:   $HC^{\contr}_*(V,\xi)$,
2665:   generated  only by contractible periodic orbits,  which is another invariant
2666:   of
2667:   the contact manifold $(V,\xi)$.
2668: 
2669: Contact structures which admit nice contact forms do exist, as it
2670: is
2671: illustrated by examples in Section \ref{ex:Floer}  below. However, the
2672: condition of existence of a nice form is too restrictive. The
2673: general case leads to an algebraic formalism developed in
2674: Sections \ref{sec:af-contact}--\ref{sec:composition} below.
2675: 
2676: \subsubsection{Examples}\label{ex:Floer}
2677: 
2678: 
2679: 
2680: \noindent 1. \textsf { Contact homology of the standard contact sphere $S^{2n-1}$.}
2681: 
2682:  Take the $1$-form $\alpha=\frac 12\sum (x_idy_i-y_idx_i)$,
2683:  which is a primitive of the standard symplectic structure
2684:  in $\R^{2n}$. Its restriction  to a generic  ellipsoid
2685:  $$S=\{\sum\frac{x_i^2+y_i^2}{a_i^2}=1\}$$ is a  nice contact form for the standard contact structure
2686:  $\xi$ on the sphere $S=S^{2n-1}$. The form $\alpha|_S$ has
2687:  precisely one periodic orbit for each  Conley-Zehnder index
2688:  $n +2i-1$ for $i=1,\dots,$. Hence the  contact homology group
2689:  $HC_*(S,\xi)$ has one generator in each  dimension $2i, i\geq n-1$.
2690:  See also the discussion in Section \ref{sec:Bott} below.
2691:  \medskip
2692: 
2693: \noindent 2. \textsf{ Contact homology of Brieskorn spheres.}
2694: 
2695:  Ilya Ustilovsky  computed (\cite{Ustilovsky}) the
2696:  contact homology of certain Brieskorn spheres.
2697: 
2698:  Let us  consider the Brieskorn manifold
2699:  \begin{equation*}
2700:  \Sigma(p,\underbrace{2,\dots,2}_n)=
2701:  \{z_0^p+\sum\limits_1^n  z_j^2=0\}\cap \{\sum\limits_0^n|z_j|^2=1\}\subset \C^{n+1}.
2702:  \end{equation*}
2703:  $\Sigma(p,\underbrace{2,\dots,2}_n)$ carries a
2704:   canonical contact structure  as a strictly pseudo-convex hypersurface in a complex manifold.
2705: 
2706:  Suppose that $n=2m+1$ is odd, and $p\equiv 1\,\mod\,8.$
2707:  Under this assumption  $\Sigma(p,\underbrace{2,\dots,2}_n)$
2708:  is diffeomorphic to $S^{2n-1}$ (see \cite{Brieskorn}).
2709:   However, the following theorem of Ustilovsky implies that     the contact structures on
2710:    Brieskorn spheres
2711:  $\Sigma(p,\underbrace{2,\dots,2}_n)$ and $\Sigma(p',\underbrace{2,\dots,2}_n)$ are not isomorphic, unless $p=p'$.
2712: This result should be confronted with a computation of Morita
2713: (\cite{Morita} ), which implies
2714:  that  the formal homotopy class (see Section \ref{sec:prelim} above) of the contact structure on $\Sigma(p,\underbrace{2,\dots,2}_n)$
2715:   is   standard, provided $p\equiv 1\,\mod\, 2(2m!)$. Hence, Ustilovsky's theorem provides infinitely many non-isomorphic contact
2716:   structures on $S^{4m+1}$ in the standard formal homotopy class.
2717: 
2718:   \begin{theorem} {\rm (I. Ustilovsky, \cite{Ustilovsky})}
2719:   The contact homology $$HC_*\left(\Sigma(p,\underbrace{2,\dots,2}_n)\right)$$ is defined, and
2720:   the dimension  $$c_k=\dim\,HC_k\left( \Sigma(p,\underbrace{2,\dots,2}_n)\right) $$
2721:   is given by the formula
2722:   \begin{equation*}
2723:   c_k  =
2724:   \begin{cases}
2725:   0,& k \;\;\hbox {is odd or}\;\;k<2n-4,\\
2726:   2,& k=2\left[ \frac{2N}{p}\right] +2(N+1)(n-2),\;\hbox{for  }\;\;
2727:   N\geq 1,\,2N+1\notin p\Z,\\
2728:   1,& \hbox{in all other cases.}\\
2729:   \end{cases}
2730:   \end{equation*}
2731:   \end{theorem}
2732: 
2733: \medskip
2734: 
2735: \noindent 3. \textsf{ Contact homology of boundaries
2736: of subcritical Stein manifolds.}
2737: 
2738:  A co-oriented contact manifold $(V,\xi)$ is called Stein-fillable if it can be realized as a strictly pseudoconvex
2739:  boundary of a complex manifold $W$,
2740:  whose interior is Stein, and if the co-oriented contact structure $\xi$ coincides with the canonical contact structure
2741:  of a strictly pseudo-convex hypersurface.   We say that $(V,\xi)$ admits a
2742:  subcritical Stein filling if the corresponding Stein manifold $\Int W$ admits an exhausting plurisubharmonic function
2743:  without critical points of dimension $\dim_\C(W)$. If $\dim V>3$ then one can equivalently require
2744:  that $W$ deformation  retracts to   a CW-complex of dimension $<\dim_\C W$ (see \cite{Eliash-Stein}).
2745: 
2746:  Mei-Lin Yau studied in her  PhD thesis \cite{MLYau} contact homology of contact manifolds
2747:  admitting a subcritical Stein filling.
2748:  Here is her result.
2749:  \begin{theorem}\label{prop:MLYau} {\rm (Mei-Lin Yau,  \cite{MLYau})}
2750:  Let $(V,\xi)$ be a  contact manifold of dimension $2n-1$ which admits a subcritical Stein filling $W$.
2751:  Suppose that $c_1(V)=0$ and $H_1(V)=0$. Let $c_1,\dots, c_k$ be generators of $H_*(W)$.
2752:   Then  the contact homology $HC_*(V)$ is defined  and generated
2753:   by elements
2754:   $ q_{i,j}$ of  degree $\deg q_{i,j}= 2(n+i-2)-\dim c_j$, where  $j=1,\dots, k$, and
2755:   $i\geq 1$.
2756:   \end{theorem}
2757: 
2758: \medskip
2759: 
2760: \noindent 4. \textsf{ Contact homology of $T^3$ and its coverings.}
2761: \medskip
2762: 
2763:  Set $\a_n=\cos 2\pi nz\,dx+\sin 2\pi nz\,dy$. This contact form descend
2764:  to the $3$-torus $T^3=\R^3/\Z^3$ and defines there a contact structure $\xi_n$.
2765:  The structure $\xi_1$ is  just the canonical contact structure on $T^3$ as the space
2766:  of co-oriented contact elements of $T^2$. The form  $\a_n$ for $n>1$  is
2767:  equal to the pull-back $\pi^*_n (\a_1)$, where $\pi_n:T^3\to T^3$
2768:  is the covering
2769:  $(x,y,z)\mapsto(x,y,nz)$. Notice that all structures $\xi_n$
2770:  are homotopic as plane field to the foliation $dz=0$.
2771: 
2772:  \begin{theorem}\label{thm:T3}
2773:  The contact homology group $HC_*(T^3,\xi_n|w)$, where $w$ is the homology class $(p,q,0)\in H_1(T^3)$,
2774:  is isomorphic to $\Z^{2n}$.
2775:  \end{theorem}
2776:  In particular we get as a corollary a theorem of E. Giroux:
2777:  \begin{corollary} {\rm (E.  Giroux, \cite{Giroux})}
2778:  The contact structures $\xi_n$, $n=1,\dots,$ are pairwise non-isomorphic.
2779:  \end{corollary}
2780: 
2781:   The contact manifold $(T^3, \xi_1)$
2782:  is foliated by pre-Lagrangian
2783:  tori $L_{p,q}$, indexed by simple homology classes $ (p,q)\in H_1(T^2)$. Each torus
2784:  $L_{p,q}$ is foliated by  the $S^1$-family of lifts of
2785:   closed geodesics which represent the class $(p,q)$.
2786:   Thus for any given $ (p,q)\in H_1(T^2)$ (even when  $(p,q)$ have common divisors)
2787:    the set of closed orbits
2788:   in  $\Pc_{\a_1}$ which represent the class $(p,q,0)\in H_1(T^3)$
2789:    is  a circle $S_{p,q}$, and for any $n\geq 1$ the set of closed orbits
2790:   in  $\Pc_{\a_n}$ which represent the class $(p,q,0)\in H_1(T^3)$
2791:   consists of $n$ copies $S^1_{p,q},\dots,  S^n_{p,q}$ of such circles.
2792:    The forms $\a_n$ have no contractible periodic orbits, but of course, they
2793:     are degenerate. To compute the contact homology groups, one can   either work directly with these
2794:     degenerate forms, as it is explained in Section \ref{sec:Bott} below, and
2795:     show that  $HC_*(T^3,\xi_n|w)=H_*(\bigcup\limits_1^n S^i_{p,q})=\Z^{2n}$, or  first  perturb the form $\alpha_1$,
2796:   and respectively all its covering forms $\a_n= \pi_n^*(\a_1)$, in order to substitute each circle $S^i_{p,q}$
2797:   by two non-degenerate periodic orbits, and then show that the orbits from  each of these pairs  are connected by precisely
2798:   two holomorphic cylinders, which cancel each other in the formula for the boundary operator $\partial$.
2799: 
2800: 
2801: \subsubsection {Relative contact homology and contact non-squeezing theorems}
2802: \label{sec:squeezing}
2803: Let us observe that  the  complex $(\fF,\partial)$  is filtrated
2804: by the values of the action  functional $S$, $\fF=\mathop{\bigcup}\limits_{a\in\R}\fF^a$, where
2805: the complex $\fF^a$
2806: is generated  by variables $q_\gamma$ with $S(\g)\leq a$. The differential
2807: $\partial$ respects this filtration, and hence descends to $\fF^b/\fF^a, a<b$.
2808:  Hence, one can define  the homology
2809: $H_*^{(a,b]}(\fF,\partial)=H_*(\fF^b/\fF^a,\partial)$ in the window $(a,b]\subset\R$. Of
2810: course, $H_*^{(a,b]}$ depends on a choice of a particular nice form
2811: $\alpha$. If $\alpha>\beta$ then we  have a map $\Phi_*:H_*^{(a,b]}(\fF,\partial;\alpha)\to
2812: H_*^{(a,b]}(\fF,\partial;\beta)$. We write $H^a$ instead of $H^{(-\infty,a]}$.
2813: 
2814: Consider now a contact manifold $(V,\xi)$ which is
2815:  either closed, or satisfies
2816: the following
2817:   {\it pseudo-convexity}  condition at infinity.
2818: A contact manifold $(V,\xi=\{\a=0\})$  with a fixed contact form $a$
2819: is called pseudo-convex at infinity
2820: if there exists a compatible almost complex structure
2821:  $J$ on the symplectization $V\times\R$ for which $V$ can be exhausted
2822:  by compact domains $V_i$ with smooth  pseudo-convex
2823:  boundary. A sufficient condition for pseudo-convexity is existence of
2824:   an exhaustion $V=\bigcup V_i$, such that for each $i=1,\dots, $
2825:  trajectories of the Reeb field
2826:  $R_\a|_{V_i}$ do not have interior tangency points with $\partial  V_i$.  For instance, for
2827:  the standard contact form  $\a=dz-\sum y_idx_i$
2828:  on $\R^{2n+1}$ the latter condition is satisfied for an exhaustion of $\R^{2n+1}$
2829:  by round balls, and  ence the standard contact form on $\R^{2n+1}$ is pseudo-convex at
2830:   infinity.
2831: 
2832: 
2833: 
2834: 
2835: Our goal is to define  a relative contact homology group $HC_*(V,U,\xi)$
2836:  for a relatively compact open subset
2837: $U\subset V$, so that  this group would be invariant    under a contact isotopy of $U$ in $V$.
2838: 
2839: Let us fix a contact form $\alpha$ on $V$  which satisfies the above pseudo-convexity
2840:  condition.
2841: Let us denote by $\Fc_{U,\a}$ the set of  $C^\infty$-functions $f:V\to[0,\infty)$
2842: which  are $\leq 1$ on   $U$,   and for which  the contact form
2843: $f\alpha$ is nice and pseudo-convex at infinity.\footnote{ Of course, the set
2844:  $\Fc_{U,\a}$ may be
2845: empty, because the niceness condition is very restrictive.
2846: In this case one needs to employ
2847: a more general construction from Section \ref{sec:3algebras}.}
2848:  Take a strictly increasing sequence
2849: of functions
2850: $f_i\in \Fc_{U,\a}$, such that
2851: \begin{description}
2852: \item{a)}
2853: $\max\limits_K f_i \mathop{\to}\limits_{i\to\infty}\infty$ for   each compact
2854:  set $K\subset (V\setminus
2855: \overline{U})$;
2856: \item{b)} $f_i|_U \mathop{\to}\limits_{i\to\infty} 1$ uniformly on
2857: compact sets.
2858: \end{description}
2859: 
2860: \begin{proposition}\label{prop:rel-homol}
2861: The limit
2862: \begin{equation}
2863: HC_*(V,U,\xi)=\lim\limits_{a\to+\infty}
2864: \lim\limits_{\leftarrow}HC^a_*(V,f_i\alpha)
2865: \end{equation}
2866: is independent of $\alpha$,
2867:  and
2868: thus it is  an invariant of  the contact pair $(V,U)$.
2869: A contact isotopy  $f_t:V\to V$ induces a  family of
2870: isomorphisms $(f_t)_*:H_*(V,U)\to H_*(V, f_t(U))$.
2871: An inclusion $i:U_1\mapsto U_2$ induces a homomorphism
2872: \begin{equation*}
2873: i_*: HC_*(V,U_1,\xi)\to HC_*(V,U_2,\xi).
2874: \end{equation*}
2875: \end{proposition}
2876: 
2877: 
2878: 
2879: \bigskip
2880: 
2881: One of the most celebrated results in Symplectic topology is
2882: Gromov's non-squeezing theorem which  states that one cannot
2883: symplectically embed a  $2n$-ball of radius $1$ into
2884: $\D^2_r\times\R^{2n-2}$ for $r<1$. Here $D^2_r$ denotes a
2885:  $2$-disk of radius $r$ and $\D^2_r\times\R^{2n-2}$ is endowed
2886:  with the product of standard symplectic structures.
2887:   Because of the conformal character of contact geometry one
2888:   cannot expect as strong non-squeezing results
2889:   for
2890:   contact manifolds: one can embed any domain in the standard
2891:   $\R^{2n-1}$ in an arbitrary small ball. However, it turns out
2892:   that it is not always possible  to realize a contact squeezing via a contact
2893:   isotopy inside a manifold with  a non-trivial first Betti number.
2894: 
2895: As an example, consider the $1$-jet bundle $$V=J^1(\R^n,S^1)=T^*(\R^n)\times S^1$$
2896:  of
2897: $S^1$-valued functions with its standard contact structure $\xi$,
2898: given by the contact form $\alpha=dz-\sum y_idx_i$, $(x,y)\in
2899: \R^{2n}=T^*(\R^{n}),\, z\in S^1 =\R/\Z$.
2900: The contact form $\alpha$ satisfies the condition
2901: of  pseudo-convexity   at infinity
2902: and it is nice: the Reeb field equals $\frac{\partial}{\partial z}$,
2903: and thus it has no contractible periodic orbits.
2904:  Let us consider the class $\cN$ of domains $\Omega\subset V$
2905:  which are images of the split domains $U\times S^1\subset V$
2906:  under a contact isotopy of $V$, where $U$ is connected.
2907: For any $\Omega\in\cN$ the relative contact  homology
2908: group $HC_*(V,\Omega)$ is  well defined
2909:  because   for any function $f:\R^{2n}\to\R$
2910:  the form $f(x,y)\alpha$ is nice.
2911: 
2912: Let us denote by $\cE_r(\Omega)$, $\Omega\in\cN$,
2913: the space of contact embeddings $ D_r\times S^1\to
2914:  \Omega\times   S^1$, contact
2915:  isotopic in $V$  to the inclusion
2916:  $$D_r\times S^1\hookrightarrow \R^{2n}\times S^1=V\,.$$
2917: Notice, that for any two embeddings $f,g\in\cE_r(\Omega)$
2918: there exists  a positive $\rho\leq r$, such that the
2919: restrictions
2920: $f|_{D_\rho\times S^1}$ and $g|_{D_\rho\times S^1}$
2921: are isotopic       via a {\it contact} isotopy.
2922: 
2923: 
2924: 
2925: Given a  contact embedding $f\in\cE_r(\Omega)$
2926:  there is defined a homomorphism
2927: \begin{equation*}
2928: f_*: HC_*(V,D_r\times S^1,\xi)\to HC_*(V,\Omega,\xi)\,.
2929: \end{equation*}
2930:   Let us choose a
2931:     symplectic trivialization
2932:   of the contact bundle $\xi$ induced by the projection
2933:   $V\to\R^{2n}$. We will assume that indices of periodic orbits,
2934:   and hence the grading of contact homology groups,
2935:   are associated with this trivialization.
2936: 
2937: For each homology class $k\in \Z=H_1(D_r\times S^1)$ let us
2938: consider the maximal $l=l(f,k)$ such that
2939: $$\Ker\big( f_*|_{HC_l(V,D_r\times S^1,\xi|k)}\}\neq 0 ,,$$
2940:     and define an invariant $w_{\cont}(V,\Omega)$, called
2941:     the relative {\it
2942: contact width} by the formula
2943: \begin{equation}
2944: w_{\cont}(V,\Omega)=\sup \limits_{k,r>0,f\in
2945: \cE_r(U)}\frac{2 k}{l(f,k)}\;.
2946: \end{equation}
2947: 
2948: 
2949:   S.-S. Kim has computed $w_{\cont}(V,\Omega)$ for
2950:   certain domains $\Omega$. In particular,
2951:    she proved
2952: \begin{proposition}\label{prop:Kim}
2953: \begin{eqnarray*}
2954: w_{\cont}(V,D^{2n}_r\times S^1)= \pi r^2;  \\
2955: w_{\cont}(V,D^2_r\times D^{2n-2}_R\times S^1) =\pi r^2,
2956: \end{eqnarray*}
2957: if $R\geq r$.
2958: \end{proposition}
2959:     The contact width is clearly a monotone invariant, i.e.
2960: $$ w_{\cont}(V,U_1\times S^1)\leq w_\cont(V,U_2\times S^1)$$
2961: if $U_1\subset U_2$. Hence Proposition \ref{prop:Kim} implies
2962: 
2963:   \begin{corollary}\label{cor:non-squeezing}
2964:   Suppose that $r<\min(r',R).$ Then there is no  contact isotopy
2965:    $f_t:D^{2n}_{r'}\times S^1\to V$ such that $f_0$
2966:    is the inclusion, and
2967:   $$f_1( D^{2n}_{r'}\times S^1)\subset D^2_{r}\times D_R^{2n-2}\times S^1 .$$
2968:   \end{corollary}
2969: 
2970: 
2971: 
2972: 
2973:      \begin{problem}  Suppose there exists a contact isotopy
2974:       $f_t:V=\R^{2n}\times S^1\to V$ with
2975:      $f_0=\Id$ and $f_1(U_1\times S^1)\subset U_2\times S^1$. Does
2976:      there exist
2977:      a Hamiltonian isotopy $g_t:\R^{2n}\to\R^{2n}$ such that
2978:       $g_0=\Id$ and $g_1(U_1)\subset U_2$?
2979:       \end{problem}
2980:       Notice that the converse is obviously true.
2981: 
2982: \section{Algebraic formalism}\label{sec:algebraic}
2983: 
2984: \subsection{Informal introduction}\label{sec:informal}
2985: 
2986: The Floer-theoretic formalism described in Section \ref{sec:Floer}
2987:  is applicable only to a very limited class
2988: of contact manifolds. As it follows from Theorem \ref{thm:comp1}
2989: the boundary of the moduli space   of holomorphic cylinders in the
2990: symplectization may   consist of stable curves, different from broken
2991: cylinders; for instance,  it may contain    height 2 stable curves
2992: which consist  of a pair of pants on the upper level, and a copy
2993: of $\C$ plus a trivial cylinder at the bottom one. Hence the
2994: minimal class of holomorphic curves
2995:  in symplectizations which
2996: has the property that the  stable curves  of height $>1$ on the boundary
2997: of the corresponding moduli space  are made of curves from the same class,
2998: must contain all  rational curves with one positive   and an arbitrary
2999:  number of  negative punctures.
3000:  The counting of curves  with one positive and arbitrary number of negative punctures
3001:  can still be interpreted as a differential, but this time defined
3002:  not on the {\it vector space} generated by    periodic trajectories but on the   graded {\it algebra}
3003:   which they
3004:  generate. Thus
3005:  this  leads to a straightforward generalization of the Floer type formalism considered
3006:    in    Section \ref{sec:Floer}
3007:  when instead of  the additive  Floer complex   $\fF$  generated by the
3008:  variables $q_\g$,
3009:  one considers a graded commutative algebra  $\fA$
3010:  generated by these variables, and when  instead of the  formula (\ref{eq:Floer1})
3011:   the differential  $\partial q_\g$ is defined
3012:  as a polynomial of  a higher degree, rather than a linear
3013:  expression as in the Floer homology case. Namely, we define
3014:  \begin{equation}\label{eq:d}
3015: \partial q_\g= \sum \frac{n_{\Gamma,I,d}}{k!\prod^k_1
3016: i_j!\kappa_{\g_j}^{i_j}}q_{\g_1}^{i_1}\dots q_{\g_k}^{i_k}z^d,
3017:    \end{equation}
3018:    where the sum is taken over all ordered
3019:    \footnote{The coefficient $\frac{1}{k!}$ is the price we pay
3020:    for taking {\it ordered} sets of periodic orbits.} sets of
3021:    different periodic orbits $\Gamma=\{\g_1,\dots,\g_k\} $,   multi-indices
3022: $d=(d_1,\dots, d_N)$ and $I=(i_1,\dots,i_k), i_j\geq 0, $
3023:  and where the coefficient $n_{\Gamma,I,d}$  counts
3024: the algebraic number of elements of the moduli space
3025: $$\Mc^d_0(\g; \underbrace{\g_1,\dots,\g_1}_{i_1},\dots,\underbrace{\g_k,\dots,\g_k}_{i_k})/\R,$$ if
3026: this space is $0$-dimensional, and equals $0$ otherwise.
3027: The differential $\partial$ extends to the algebra $\fF$ according
3028: to the graded Leibnitz rule.
3029: Roughly speaking, $\partial q_\g$ is a polynomial,
3030: whose monomomials $q_{\g_1}\dots q_{\g_l}$ are in 1-1
3031: correspondence with rigid, up to translation,   rational
3032: holomorphic   curves with one positive  cylindrical end
3033: over $\g$ and $l$ negative cylindrical ends over
3034: trajectories $\g_1,\dots,\g_l$.
3035: 
3036: 
3037:  It turns out that the  quasi-isomorphism class
3038:  of the differential algebra $(\fA,\partial)$
3039:  is  independent of all extra choices (see Section \ref{sec:3algebras} below).
3040:  In particular, the {\it contact homology algebra} $ H_*(\fA,\partial)$
3041:   is an invariant of the contact manifold
3042:  $(M,\xi)$.
3043: 
3044:  Having included into the picture the moduli spaces of rational
3045:  curves with one positive and several negative punctures, one may
3046:  wonder, what is the role of rational curves with an arbitrary
3047:  number of positive and negative punctures. One can try to
3048:  interpret the
3049:  counting of rational holomorphic curves with
3050: fixed number of  positive and an arbitrary number of negative
3051: punctures  as a sequence of bracket type  operations on the
3052: algebra $\fA$. These operations  satisfy an infinite system of
3053: indentities, which remind the formalism of homotopy Lie algebras.
3054: 
3055:  However,  there is a more adequate algebraic formalism
3056:  for this picture. Let us associate   with  each periodic
3057:   orbit $\g$ two graded variables $p_\g$ and $q_\g$ of the
3058:   same parity (but   of
3059:   different integer grading, as we will see in Section
3060:   \ref{sec:correlators} below),
3061:   and consider an algebra $\fP$ of formal power series
3062:   $\sum f_\Gamma(q)p^\Gamma$, where $f_\Gamma(q)$
3063:     are polynomials of $q=\{q_\g\}$ with coefficients
3064:   in (a completion of) the group algebra  of $H_2(V)$.
3065:   It is useful to think about the algebra $\fP$ as the graded
3066:   Poisson
3067:    algebra of functions
3068:   on the     infinite-dimensional symplectic super-space $\bf{V}$ with the even
3069:   symplectic form $\sum\limits_{\g\in\Pc} {\kappa_\g}^{-1} dp_\g\wedge dq_\g$, or rather on
3070:   its formal analog along the $0$-section $\{p=\{p_\g\}=0\}$.
3071:   With each $0$-dimensional moduli space
3072:   $\Mc^d_0(\Gamma^-,\Gamma^+)/\R$, $$\Gamma^\pm=\{\underbrace{\g^\pm_1,\dots\g^\pm_1}_{i_1^\pm},\dots,\underbrace{\g^\pm_{s^\pm}\dots
3073:   \g^\pm_{s^\pm}}_{i^\pm_{s^\pm}}\},$$
3074:   we associate a monomial
3075:   $$\frac{n_{\Gamma^-,\Gamma^+,d}}{s^-!s^+!(i_1^-)!\dots(i_{s^-}^-)!(i_1^+)!\dots(i_{s^+}^+)!}
3076:     q_{\g^-}^{I^-}p_{\g^+}^{I^+}z^d,$$
3077:   where $q_{\g^-}^{I^-}=(q_{\g^-_{1}})^{i^-_{1}}\dots(q_{\g^-_{s^-}})^{i^-_{s^-}}$, $p_{\g^+}^{I^+}=(p_{\g^+_1})^{i^+_1}\cdots
3078:   (p_{\g^+_{s^+}})^{i^+_{s^+}}$,  and $n_{\Gamma^-,\Gamma^+,d}$ is
3079:    the algebraic number of elements of the moduli
3080:   space     $\Mc^d_0(\Gamma^-,\Gamma^+)/\R$.
3081: 
3082: The sum of all these monomials over all $1$-dimensional moduli
3083: spaces    $\Mc^d_0(\Gamma^-,\Gamma^+) $ for   all ordered sets
3084:     $\Gamma^-,\Gamma^+$ of periodic orbits
3085: is  an {\it odd} element $\bh\in\fP$. All the operations on the
3086: algebra $\fA$ which we mentioned above appear as the expansion
3087: terms of $\bh$ with respect to $p$-variables. It turns out that
3088: the   infinite system of  identities  for
3089: operations on $\fA$, which we mentioned above, and which is
3090:  defined by counting holomorphic curves with  a
3091: certain fixed number of positive punctures, can be encoded into a single
3092: equation $\{\bh,\bh\}=0$. Then the differentiation with respect to
3093: the Hamiltonian vector field, defined by the Hamiltonian $\bh$:
3094: \begin{equation*}
3095: d^{\bh}(g)=\{\bh,g\},\;g\in\fP,
3096: \end{equation*}
3097: defines a differential $d=d^{\bh}:\fP\to\fP$, which satisfies the
3098: equation $d^2=0$. Thus one can define the homology $H_*(\fP,d^{\bh})$
3099: which inherits the structure of a graded Poisson algebra.
3100: 
3101: 
3102:  The identities, like  $d^2=0$ and $\partial^2=0$,
3103:  encode in algebraic terms information about the structure of the  top-dimensional strata on the boundary of
3104:  compactified moduli spaces of holomorphic curves,
3105: as it is described in Proposition \ref{prop:boundary} above. For
3106: instance, the codimension $1$ strata on the boundary of the moduli
3107: space $\Mc^d_0(\Gamma^-,\Gamma^+)/\R$ consists of height two stable
3108: rational curves $(f_1,f_2)$.  Each floor of
3109:  this curve  may be disconnected,
3110: but  precisely one of its components differs from the straight cylinder.
3111:   Each   connected component of $f_1$  can be glued with a component of $f_1$ only
3112: along one of their ends. One can easily see that the combinatorics  of  such gluing   precisely corresponds to
3113: the Poisson bracket formalism and that the algebraic sum of the monomials associated to all
3114: stable curves of height two equals $\{\bh,\bh\}$. On the other hand,
3115:  the algebraic number of such height 2 curves equals
3116:  $0$ because they form   the  boundary
3117: of the a compactified  $1$-dimensional moduli space of holomorphic curves.
3118: Hence we
3119:   get the identity $\{\bh,\bh\}=0$. The identity is
3120: not tautological due to   the fact that $\bh$ is odd.   In view of the super-Jacobi identity
3121: it is equivalent to the identity $(d^{\bh})^2=0$.
3122: 
3123: 
3124: One can go further and include into the picture  moduli spaces of punctured holomorphic
3125: curves
3126: of higher genus.  Introducing a new variable, denoted $\hbar$, to keep track
3127: of the genus, one can associate with each $0$-dimensional moduli
3128: space
3129:  $\Mc^d_g(\Gamma^-,\Gamma^+)/\R$
3130:    a monomial  $$\frac{n_{\Gamma^-,\Gamma^+,d,g}}{s^-!s^+!(i_1^-)!\dots(i_{s^-}^-)!(i_1^+)!\dots(i_{s^+}^+)!}
3131:     q_{\g^-}^{I^-}p_{\g^+}^{I^+}\hbar^{g-1}z^d,$$
3132:   and form a generating function $\bH=\hbar^{-1}\sum\limits_{g=0}^\infty\bH_g\hbar^{g}$
3133:   counting all rigid holomorphic curves of arbitrary
3134:   genus, whose term $\bH_0$ coincides with $\bh$.    Again,
3135:   the codimension $1$ strata of the boundary of the moduli spaces
3136:   $\Mc_g(\Gamma^-,\Gamma^+)$ consists of height 2 stable curves, but unlike the case of  rational
3137:   curves, two connected components
3138:   on different levels can be glued together along an arbitrary number of ends.
3139:    The combinatorics of such gluing
3140:   can be described by the formalism of algebra of higher order differential operators.
3141:   Fig. \ref{fig:combinatorics}
3142:   illustrates how the composition
3143: of
3144:   differential operators can be interpreted via gluing of Riemann
3145:   surfaces with punctures. A letter $p_i$ in the picture represents a differential
3146:   operator
3147:              $\hbar\frac{\partial}{\partial q_i}$, and  a surface of genus $g$ with upper punctures
3148:     $p_{i_1},\dots,p_{i_k}$ and lower punctures $q_{j_1},\dots,q_{j_l}$ represents
3149:     a differential operator
3150:         $$\hbar^{g-1}q_{j_1}\dots q_{j_l}p_{i_1}\dots p_{i_k}=\hbar^{g-1}q_{j_1}\dots q_{j_l}
3151:         \big(\hbar\frac{\partial}{\partial q_{i_1}}\big)
3152:         \dots \big(\hbar\frac{\partial}{\partial q_{i_k}}\big).$$
3153:                  Thus
3154:          we are led to consider $\bH$ as an element of the Weyl super-algebra
3155:   $\fW$. This algebra should be
3156:    viewed
3157:   as a quantization of the Poisson algebra $\fP$, so that the description of the boundary of the
3158:   moduli spaces is given by the equation $[\bH,\bH]=0$, where   $[\;,\;]$ denotes the commutator in $\fW$.
3159:   As in  the rational case, this  identity is equivalent
3160:   to the identity $D^2_{\bH}=0$  for the differential
3161:   $D^\bH(f)=[H,f]$.
3162:   Hence we can define the
3163:   homology algebra $H_*(\fW,D^{\bH})$, which   also turns out to be an invariant of the contact manifold
3164:   $(V,\xi)$. Similarly to the standard Gromov-Witten theory for closed
3165:   symplectic manifolds one can develop an even more general
3166:   formalism  by encoding in $\bH$ information about higher-dimensional moduli spaces
3167:   of holomorphic curves. This leads to a deformation of the differential
3168:   algebra $(\fW,D^{\bH})$ along the space of closed
3169:   forms on $V$. The corresponding family of homology algebras   is
3170:   then  parameterized by $H^*(V)$.
3171: 
3172: %%%%%%%%%%%%%%%%%%%%
3173: 
3174: \begin{figure}
3175: \centerline{\psfig{figure=sf1.eps,height=170mm}} \caption{\small
3176:          There are  four different way to glue the lower and upper surfaces
3177:     on the picture along their matching ends, i.e. the ends denoted by $p$'s
3178:     and $q$'s with the same index.  These 4 ways correspond to  4
3179:     terms in the composition formula for differential operators:
3180:                              $(\hbar^{-1}p_1 p_2 p_3) \circ(\hbar^{-1}q_1 q_2 p_1)=p_1 p_3+
3181:                                 \hbar^{-2} q_1 q_2 p_1^2 p_2 p_3 +
3182:                 \hbar^{-1}q_1 p_1^2 p_3+\hbar^{-1}q_2 p_1 p_2 p_3 \;.
3183:                $  We  are ignoring here the sign issues and assuming all  the boundary components
3184:                  to be simple orbits.
3185:                                  }
3186:         \label{fig:combinatorics}
3187: \end{figure}
3188: %\end{document}
3189: %%%%%%%%%%%%%%%%%%%%%%
3190:   After going that far it is natural to  make the above algebraic structure
3191:   for contact manifolds a part
3192:   of a formalism in the spirit of topological field theory, which we call
3193:    {\it Symplectic Field Theory},
3194:   and which also includes the theory of Gromov-Witten invariants of closed manifolds.
3195:   To do that one considers moduli spaces of holomorphic curves with cylindrical ends
3196:   in directed symplectic cobordisms $W=\ora{V^-V^+}$. The generating function
3197:   counting rational holomorphic curves
3198:   in $W$ can be naturally written as a function  $\bff(q^-,p^+ )$ of $p^+$-variables
3199:   associated with the positive end $V^+$, and
3200:   $q^-$-variables associated with the negative end $V^-$ of the cobordism $W$.
3201:   It turns out that the Lagrangian submanifold
3202:   in $(-\mathbf{V}^-)\times\mathbf{V}^+$ generated by the function $\bff$
3203:   defines a {\it Lagrangian correspondence}
3204: $  \mathbf{L}_W\subset(-\mathbf{V}^-)\times(\mathbf{V}^+) $
3205:   which transforms the Hamiltonian
3206:   functions $\bh^+$ and  $\bh^-$ to each other, i.e.
3207:   \begin{equation}\label{eq:HJ-informal}
3208:   \left(\bh^-(p^-,q^-)-\bh^+(p^+,q^+)\right)|_{\mathbf{L}_W}=0,
3209:   \end{equation}
3210:   where
3211:   \begin{equation*}
3212:   L_W=
3213:   \begin{cases}
3214:   q^+_{\g^+}=&\kappa_{\g^+}\frac{\partial \bff}{\partial p^+_{\g^+}}(q^-,p^+));\\
3215: p^-_{\g^-}=&\kappa_{\g^-}\frac{\partial \bff}{\partial
3216: q^-_{\g^-}}(q^-,p^+)).\\
3217: \end{cases}
3218: \end{equation*}
3219:    We recall that $\kappa_{\g^\pm}$ denotes the multiplicity of the orbit $\g^\pm$.
3220: 
3221:   The composition of symplectic cobordisms produces
3222:   the composition of Lagrangian correspondences,
3223:   so that  if one  consider a ``Heegard splitting" of  a closed symplectic
3224:   manifold $W$ along a contact hypersurface $V$, then  the computation of Gromov-Witten invariants
3225:    of $W$ can be viewed
3226:   as a Lagrangian intersection problem in the   symplectic super-space
3227:   $\mathbf{V}$ associated to the contact manifold $V$.
3228: 
3229:   After what was said it should not come as a surprise that in
3230:    the quantized picture Lagrangian correspondences
3231:   are being replaced by Fourier integral operators, and the
3232:    composition of correspondences by the convolution
3233:   of the corresponding operators.
3234: 
3235:  We describe below the SFT-formalism with more details. We treat
3236: contact manifolds in Section \ref{sec:af-contact} and  symplectic
3237: cobordisms in Section \ref{sec:af-cobordism}. Section
3238: \ref{sec:chain}  is devoted to  the   SFT-version of the  chain  homotopy statement
3239:  in Floer homology theory.   In Section  \ref{sec:composition}  we introduce the composition
3240:  formula for the SFT-invariants of symplectic cobordisms. In Section \ref{sec:contact-invariants}
3241:  we discuss    how the   introduced algebraic structures of contact manifolds
3242:  depend on extra choices.  Section \ref{sec:master}  is devoted
3243:  to a differential equation for the  potential $\bF$ of a directed symplectic
3244:  cobordism    with  a {\it non-empty } boundary.  Together with the gluing formula from Section
3245:  \ref{sec:composition} this  equation  provides an effective tool for  computing
3246:   Gromov-Witten invariants.
3247: The remainder
3248: of the paper has even  more sketchy character than the rest of
3249: the paper.
3250:  Section \ref{sec:Legendrian}   is
3251: devoted to invariants of Legendrian submanifolds via SFT. Section
3252: \ref{sec:remarks} is devoted to various examples and possible
3253: generalizations of  SFT. In particular, in Section
3254: \ref{sec:Bott} we discuss   how  one can adapt the theory to
3255: include an important for applications, though non-generic, case of
3256: contact forms with continuous families of periodic orbits. In
3257: Section \ref{sec:computing} we describe a  new recursive
3258: procedure for computing rational  Gromov-Witten invariants of $\C
3259: P^n$. Finally, in Section \ref{sec:satellites} we  just touch the
3260: wealth of other  invariants which exist in  Symplectic Field
3261: Theory.
3262: 
3263:  \subsection{Contact manifolds}\label{sec:af-contact}
3264: 
3265:  \subsubsection{Evaluation maps}\label{sec:contact-eval}
3266:  Let $(V,\xi)$ be a contact manifold with a fixed contact form
3267:  $\alpha$, $(W=V\times\R,d(e^t\alpha))$ the symplectization of
3268:  $(V,\xi)$, and $J$  a compatible almost complex structure.  As in Section \ref{sec:holomorphic}
3269:   we denote  by $f_V$ and $f_\R$
3270:   the $V $- and $\R$-components of  a
3271:  $J$-holomorphic curve $f$ in $W$, and
3272:  by $\Mc_{g,r,s^-,s^+}(W,J)$ the  disjoint union
3273:    $$\bigcup\Mc_{g,r}^A(\Gamma^-, \Gamma^+),$$
3274:    which is taken over all $A\in H_2(V)$, and all
3275:     sets $ \Gamma^-,\Gamma^+\subset\Pc_\alpha$ of cardinalities $s^\pm$.
3276: 
3277:    %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3278: Let us view the set  $\Pc=\Pc_\a$ of periodic orbits of  the Reeb fields $ R_{\alpha }$ as a discrete topological space.
3279:  It naturally splits into the   disjoint
3280: union
3281: $$\coprod\limits_{k=1}^{\infty}\Pc _k,$$ of identical subspaces, where $\Pc_k$ is the space of
3282:   periodic orbits of
3283: multiplicity $k$.
3284: 
3285:            Consider now three sets of evaluation maps:
3286:             $$ev^0_i: \Mc_{g,r,s^-,s^+}/\R \to V,\;i=1,\dots,r,$$
3287:             $$ev^+_j:\Mc_{g,r,s^-,s^+ }/\R    \to \Pc ,\;j=1,\dots      ,s^+,$$
3288:             and
3289:               $$ev^-_k:\Mc_{g,r,s^-,s^+ }/\R    \to \Pc ,\;k=1,\dots      ,s^-,$$
3290:               where $ev_i^0$ is the  evaluation map $f_V(y_i)$
3291:                at the $i$-th marked point $y_i$, while
3292:                             $ev^\pm_j$   are the evaluation maps
3293:                             at asymptotic markers $\mu^{\bx^\pm}_j$.
3294:                More precise,
3295:                let
3296:                            $$\overline{f}=(f,j,\bx^-,\bx^+,\by,\mu^{\bx^-},\mu^{\bx^+})\in
3297:                             \Mc_{g,r,s^-,s^+ },$$
3298:               and $f$ be asymptotically cylindrical over a
3299:                             periodic orbit $\gamma^\pm_j\in \Pc $ at $\pm\infty$ at the
3300:               puncture   $x^\pm_j$. Then $ev_j^\pm(\overline{f})$ is a point of $
3301:               \Pc $ representing the orbit $\g^\pm_j$ (comp. Section \ref{sec:Bott} below).
3302: 
3303:               All the above evaluation maps can be combined into a map
3304:               $$ev:\Mc_{g,r,s^-,s^+ }/\R\to V^r\times (\Pc^-)^{s^-}\times (\Pc^+)^{s^+}\,,$$
3305: which extends to the compactified moduli space
3306: $\overline{\Mc_{g,r,s^-,s^+ }/\R}$.
3307:    %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3308: \subsubsection{Correlators}\label{sec:correlators}
3309: %%%%%
3310:           Now we are ready to define
3311:   correlators.
3312:  Given $r$    differential
3313: forms $\theta_1,\dots,\theta_r$ on $V$ and $s^\pm$ ($0$-dimensional) cohomology classes $\alpha^\pm_1,\dots,
3314: \alpha^\pm_{s^\pm}\in H^*(\Pc )$  we define the    degree $-1$, or contact {\it
3315: correlator}
3316: \begin{eqnarray}\up{-1}\langle\,\theta_1,\dots,\theta_r;\alpha^-_1,\dots,
3317: \alpha^-_{s^-};\alpha^+_1,\dots,
3318: \alpha^+_{s^+} \rangle^A_{g}=\\
3319:  \nonumber \int\limits_{\overline{ {\Mc_{g,r,s^-,s^+ }^A}/\R}}ev^*(\theta_1\otimes\dots
3320: \otimes \theta_r\otimes\alpha^-_1\otimes\dots\otimes\alpha^-_{s^-}
3321: \otimes\alpha^+_1\otimes\dots\otimes\alpha^+_{s^+} ).
3322: \end{eqnarray}
3323: 
3324: Usually we will assume that the forms $\theta_1,\dots,\theta_r$
3325: are closed, but even in this case the above  correlator  depends
3326: on the actual forms, and not just their cohomology classes,
3327: because the domain of integration may have a boundary.  As we will see
3328: below the
3329: superscript $-1$ in $\up{-1}\langle\dots\rangle $ corresponds
3330: to the grading of the generating function for these
3331: correlators. It also refers to the
3332: enumerative meaning of the correlators: they count
3333: components of $1$-dimensional moduli spaces of holomorphic curves.
3334: We will consider below also correlators $\up0\langle\dots\rangle $
3335: and $\up1\langle\dots\rangle$, counting $0$-dimensional and
3336: $-1$-dimensional (i.e. appearing in 1-dimensional families) moduli
3337: spaces of holomorphic curves.
3338: 
3339: 
3340:  If we are given $K$ linearly independent differential forms
3341:   $\Theta_1,\dots,\Theta_K$,
3342:   then it is convenient
3343:   to introduce a ``general form" $t=\sum\limits_1^Kt_i\Theta_i$
3344:   from the space $L=L(\Theta_1,\dots,\Theta_K)$
3345:   generated by the  chosen forms, and view $t_i$ as graded variables
3346:   with  $\deg(t_i) =\deg(\Theta_i)-2$. In particular, all terms in
3347:   the sum $\sum\limits_1^Kt_i\Theta_i$ have even degrees.
3348: 
3349: 
3350:   Let us   consider two copies
3351:  $\Pc^+$ and $\Pc^-$ of the $0$-dimensional space
3352:  $\Pc$, one   associated with the positive end of $W$, the other with
3353:   the negative one. Cohomology classes  in $\Pc_+$ we will denote
3354:   by $p$, and in $\Pc_-$ by $q$, and will write
3355:   $$p=\sum_{\g\in\Pc}\frac{1}{\kappa_\g}p_\g[\g],\quad q=
3356:   \sum_{\g\in\Pc}\frac{1}{\kappa_\g}q_\g[\g],$$
3357:    where   $\kappa_\g$ is
3358:  the multiplicity of $\g$, and the cohomology classes $[\g]$
3359:   form the canonical basis of $H^*(\Pc)$, dual to points in $\Pc$.
3360:   Of course, speaking about cohomology classes of a discrete space may sound
3361:   somewhat ridiculous. However, this point of view   is useful, especially
3362:    in preparation for a more general  case when
3363:     some periodic orbits may be degenerate and  thus the spaces  $\Pc^\pm$
3364:       need not to be  anymore discrete, see Section \ref{sec:Bott} below.
3365:  We will also fix a basis $A_1,\dots,A_N$ of
3366:  $H_2(V)$. The coordinate vector  $d=(d_1,\dots,d_N)$ of a class $A$
3367:  is called the degree.   Here $d_j$ are integers, while we   consider
3368:  $t,p,q $ as graded variables,
3369:   where the  degrees of the variables $p,q$ are   defined by the  formulas
3370:   $$\deg (p_\gamma )  =-\CZ(\gamma)+(n-3),\, $$
3371:  $$ \deg (q_\gamma )  =+\CZ(\gamma)+(n-3).$$
3372:  The choice
3373:   of grading,
3374:    somewhat strange at the first glance,  is explained by Proposition \ref{thm:finiteness-of-H} below.
3375: 
3376:      The correlators
3377:   $$ \langle \underbrace{t,\dots,t}_r;\underbrace{q,\dots,q  }_{s^-}
3378:      ;\underbrace{p,\dots,p}_{s^+}\rangle^d_g$$ with  different
3379:      $r,d,g$ determine all the correlators involving forms from the
3380:      space $L$.
3381: 
3382: 
3383: \subsubsection{Three differential algebras}\label{sec:3algebras}
3384: 
3385:    Similar to the theory of Gromov-Witten invariants of   closed symplectic manifolds we
3386:    will organize all correlators
3387:       into a generating function, called {\it Hamiltonian},
3388:    $${\bH}=\frac{1}{\hbar}\sum\limits_{g=0}^\infty{\bH}_g\hbar^g,$$
3389:    where,
3390: \begin{equation}
3391:  {\bH}_g=\sum\limits_d\sum\limits_{r,s^\pm=0}^\infty\frac{1}{r!s^-!s^+!}
3392:      \up{-1}\langle \underbrace{t,\dots,t}_r;\underbrace{q,\dots,q  }_{s^-}
3393:      ;\underbrace{p,\dots,p}_{s^+}\rangle^d_g z^d,
3394:      \end{equation}
3395: and  $t=\sum_1^Kt_i\Theta_i$. We will assume throughout the paper,
3396: that
3397: all forms $\Theta_1,\dots,\Theta_K$ are closed (see,
3398: however, Remarks \ref{rem:non-closed1} and \ref{rem:non-closed2}, and
3399: Section \ref{sec:master} below).
3400:  The variables $\hbar$ and $z=(z_1,\dots,z_N)$ are also considered as graded
3401:  with $\deg \hbar= 2(n-3)$ and $\deg (z_i)=- 2c_1(A_i)$, where $c_1$ is the first
3402:  Chern class of  the almost complex structure $J$.
3403: 
3404: 
3405: 
3406: \begin{proposition} \label{thm:finiteness-of-H}
3407: \begin{description}
3408:   \item{a)}  For each $g=0,\dots$  the series $\bH_g$
3409:    can be viewed as  formal power series in variables $p_\g$   with coefficients
3410:   which are   polynomials   of variables $q_\g$ and formal power series of $t_i$ \footnote{
3411:   In fact, $\bH_g$ depends polynomially on all variables $t_i$ of degree $\neq 0$.
3412:   The degree $0$ variables, i.e. the variables
3413:   associated with $2$-forms, enter into the constant part of $\bH_g$ (i.e.
3414:   the part describing constant holomorphic curves) polynomially, while
3415:    the non-constant part    of $\bH_g$ depends
3416:    polynomially on $e^{t_i}$. This fact is similar to the standard Gromov-Witten theory
3417:   and  will not discussed  in the present paper.}
3418:   with coefficients
3419:   in the group algebra $\C[H_2(V)]$
3420:     {\rm (which we identify with the algebra
3421:   of Laurent polynomials of $z$ with complex coefficients)};
3422:   \item{b)} All terms of $\bH$  have degree $-1$;
3423:   \item{c)} $\bH\big|_{p=0}=\bH_{\const}$, where
3424:    $$\bH_{\const}=\hbar^{-1}\sum\limits_{g,r=0}^\infty
3425:    \frac{1}{r!}\up{-1}
3426:    \langle\underbrace{t,\dots,t}_r\rangle^0_g\hbar^g$$
3427:    accounts for the contribution of constant holomorphic curves.
3428:    In particular, $\bH\big|_{p=0}$ is independent of $q$ and $z$.
3429:   \end{description}
3430:   \end{proposition}
3431: The polynomial dependence of $\bH_g$  on variables $q_\g$ and $z$   in a  geometric language just means that
3432: the union $\overline{\Mc^d_g(\Gamma^+)}$, $\Gamma^+=\g_1\dots\g_{s^+}$,
3433: of the  compactified moduli spaces of holomorphic curves  of a fixed genus of any degree
3434:  with prescribed
3435: positive ends  $\g^+_1,\dots,  \g_{s^+}$
3436: is compact, and in particular that there are only finitely many possibilities for the degrees
3437: and the negative
3438: ends of these curves.
3439:  This follows from the fact  that  for each curve $C\in \Mc_g(\Gamma^-,\Gamma^+)$ we have
3440:  \begin{equation}\label{eq:pos-ends}
3441:  0\leq \int\limits_C d\alpha= \sum\limits_{\g_i\in\Gamma^+}\int\limits_{\g_i}\alpha-
3442:  \sum\limits_{\g_j\in\Gamma^-}\int\limits_{\g_j}\alpha\leq
3443:   \sum\limits_{\g_i\in\Gamma^+}\int\limits_{\g_i}\alpha,
3444:   \end{equation}
3445:   the fact   that there exists a constant $m>0$ such that $\int\limits_{\g}\a>m$ for
3446:   any periodic   orbit $\g\in\Pc_\a$ and Theorem \ref{thm:comp1}  above.
3447:      Proposition \ref{thm:finiteness-of-H}b)   follows from the formula
3448:   (\ref{eq:dim}) for the dimension of the moduli spaces of holomorphic curves, our degree
3449:   convention   and the fact that a correlator
3450:       $\up{-1}\langle\,\theta_1,\dots,\theta_r;\g^-_1,\dots,
3451: \g^-_{s^-};\g^+_1,\dots,
3452: \g^+_{s^+} \rangle^A_{g}$  may be different from $0$ only if the total dimension of the forms
3453: $\theta_1,\dots\theta_r$ equals the dimension of the moduli  space   $\Mc^A_{g,r}( \g^-_1,\dots,
3454: \g^-_{s^-};\g^+_1,\dots,
3455: \g^+_{s^+})/\R $.
3456:   Proposition \ref{thm:finiteness-of-H}c) just means that every non-constant holomorphic curves should have at least
3457:   one positive end, which follows from   inequality  (\ref{eq:pos-ends}), or alternatively the
3458:   maximum principle for holomorphic curves.
3459:   \medskip
3460: 
3461:  Let us consider the {\it Weyl super-algebra}
3462:   $\fW=\{\sum\limits_{\Gamma,g} f_{\Gamma,g }(q,t )p^\Gamma\hbar^g \}$,
3463:   where
3464:  $$\Gamma=(\g_1,\dots,\g_a),\gamma_i\in\Pc,\,a=1,\dots,\, p^{\Gamma}=p_{\g_1}\dots p_{\g_a},$$
3465:   and        $f_{\Gamma,g }(q,t)$ are polynomials of variables
3466:   $q_\gamma$ and formal power series of $t_i$\footnote{See the previous footnote.}. Proposition
3467:   \ref{thm:finiteness-of-H}a) states that $\bH\in\hbar^{-1}\fW$.
3468: 
3469:   The product operation $F\circ G$ in $\fW$ is associative and satisfies the following commutation
3470:   relations: all variables
3471:   are super-commute (i.e. commute or anti-commute according to their grading), except
3472:   $p_\gamma$ and $q_\gamma$ which correspond to the same periodic orbit $\gamma$. For these pairs of variables
3473:    we have the following commutation relation:
3474:  \begin{equation}
3475:  [p_\gamma,q_\gamma]=p_\g\circ q_\g-(-1)^{\deg p_\g\deg q_\g}q_\g\circ p_\g=
3476:  {\kappa_\gamma}\hbar \,
3477:  \end{equation}
3478:   where $\kappa_\g$ is the multiplicity of the orbit $\g$.
3479:   The algebra $\fW$ can be represented as an  algebra of formal differential   operators  with respect
3480:   to
3481:   $q$-variables acting  {\it on the left} on the space of polynomials
3482:     $  f(q,z,\hbar) $, by setting
3483:     $$p_\g= {\kappa_\g}\hbar\ora{\frac{\partial}{\partial
3484:     q_\g}}.$$
3485:     Alternatively by setting
3486:     $$q_\g={\kappa_\g}\hbar\ola{\frac{\partial}{\partial p_\g}}$$ we can
3487:     represent
3488:     $\fW$   as an algebra  of  polynomial differential operators  acting {\it on the right} on the algebra $\{\sum
3489:     \limits_{\Gamma,g}f_{\Gamma,g}(q,z)\hbar^gp^\Gamma\}$ of formal power series
3490:     of $\hbar$ and the $p$-variables.
3491: 
3492:   Notice that  the commutator $[F,G]$ of  two homogeneous elements $F,G\in\fW$
3493:   equals $F\circ G-(-1)^{\deg F\deg G}G\circ F$, and hence if $F\in\fW$ is an odd element (i.e. all its
3494:   summands are odd) then
3495:      $[F,F] =2F\circ F,$
3496:      and $[F,F]=0$ if $F$ is even. For any two elements
3497:      $F,G\in\fW$ the commutator $[F,G]$ belongs to the ideal
3498:      $\hbar\fW$.
3499:                     According to Proposition \ref{thm:finiteness-of-H}   the Hamiltonian
3500:      $\bH$ can be viewed as an element of $\frac{1}{\hbar}\fW$, and hence
3501:      the above remark shows that for $F\in\fW$
3502:      we have $[\bH,F]\in\fW$.
3503:   \begin{theorem} \label{thm:HH}
3504:  The Hamiltonian  $\bH$ satisfies the identity
3505:   \begin{equation}\label{eq:HH}
3506:  \bH\circ \bH=0.
3507:  \end{equation}
3508:  \end{theorem}
3509: 
3510: This theorem (as well as Theorems
3511: \ref{thm:SFT-cobordism}, \ref{thm:SFT-chain} and \ref{thm:SFT-composition}
3512: below) follows from the description of the boundary of the
3513: corresponding moduli spaces of holomorphic curves. As it was stated
3514: in Proposition \ref{prop:boundary}
3515:   this boundary
3516:   is tiled by codimension one
3517:  strata represented by stable curves of height $2$, so that the
3518: (virtual) fundamental cycles of the boundary of the compactified  moduli
3519: spaces
3520:   can be
3521: symbolically written as   $\partial [\Mc]=\kappa\sum [\Mc_{-}]\times [\Mc_{+}]$,
3522: where $[\Mc_\pm]$ are chains represented by the corresponding moduli spaces and  where
3523:  the coefficient
3524:  $\kappa$ depends on multiplicities of orbits along which the two levels of the corresponding
3525:  stable curve are glued. Together
3526: with the Stokes formula $\int\limits_{[\Mc]} d\omega = \int\limits_{\partial [\Mc]} \omega
3527: $, and the fact that the integrand is a closed form, we obtain
3528: identity
3529: (\ref{eq:HH}).
3530: 
3531: \begin{remark}\label{rem:non-closed1}
3532: {\rm The same argument shows that when the forms $\Theta_i$
3533: generating the space $L$  are not necessarily closed
3534: we get the following  equation
3535: \begin{equation}\label{eq:non-closed1}
3536: d\bH+\frac12[\bH,\bH]=0,
3537:  \end{equation}
3538:  which
3539: generalizes (\ref{eq:HH}) and can be interpreted as the
3540: zero-curvature equation for the connection $d +[\bH,\cdot\,]$. We
3541: denote here by $d$ the de Rham differential, i.e.
3542:  \begin{equation}
3543:  \begin{split}
3544:  d\bH&=d\Big(\sum\limits_{d,g}\sum\limits_{r,s^\pm=0}^\infty\frac{1}{r!s^-!s^+!}\\
3545:    & \quad \big\langle \underbrace{\sum\limits_1^Kt_i\Theta_i,\dots,\sum\limits_1^Kt_i\Theta_i}_r;\underbrace{q,\dots,q  }_{s^-}
3546:      ;\underbrace{p,\dots,p}_{s^+}\big\rangle^d_g
3547:      z^d\hbar^{g-1}\Big)\\
3548:       \end{split}
3549:      \end{equation}
3550:  \begin{equation*}
3551:  \begin{split}
3552:      \;\;&=\sum\limits_{d,g}\mathop{\sum\limits_{s^\pm=0,}}\limits_{r=1}^\infty\frac{1}{(r-1)!s^-!s^+!}\\
3553:     &\quad \big\langle\sum\limits_1^K t_i
3554:      d\Theta_i, \underbrace{\sum\limits_1^Kt_i\Theta_i,
3555:      \dots,\sum\limits_1^Kt_i\Theta_i}_{r-1};\underbrace{q,\dots,q  }_{s^-}
3556:      ;\underbrace{p,\dots,p}_{s^+}\big\rangle^d_g
3557:      z^d\hbar^{g-1}\,.\\
3558:      \end{split}
3559:      \end{equation*}
3560:  }
3561: \end{remark}
3562: \bigskip
3563: 
3564:  The identity $\bH\circ\bH=0$
3565:  is equivalent to $[\bH,\bH]=0$, because $\bH$ is an odd element.
3566:  Let  us  define the differential
3567:  $D=D^{\bH}:\fW\to\fW$
3568:  by the formula
3569:  \begin{equation}
3570:  D^{\bH}(f)=[\bH,f]\quad\hbox{ for }\quad f\in\fW .
3571:  \end{equation}
3572: Then Theorem \ref{thm:HH} translates into the identity $D ^2=0$. The differential
3573:  $D^{\bH}$ satisfies the Leibnitz rule, and thus
3574:  $(\fW,D )$
3575:  is a        differential Weyl (super-)algebra.
3576: In particular, one can define the homology algebra $H_*(\fW,D)$,
3577: which inherits  its multiplication operation from  the Weyl
3578: algebra $\fW$.   The differential $D^{\bH}$ extends in an obvious way to the modules
3579: $\hbar^{-k}\fW$, $k=1,\dots.$
3580: 
3581: \begin{example}\label{ex:circle}
3582: {\rm Let
3583:  $V=S^1$. We have  in this case
3584:  \begin{equation}\label{eq:circle}
3585:  \bH=\hbar^{-1}\big(\frac{t_1t_0^2}{2}+ t_1\sum p_kq_k-\frac{t_1\hbar }{24}\big),
3586:  \end{equation}
3587:   where
3588: $ t=t_01+t_1d\phi$ is a general harmonic differential form on $S^1$,
3589: so that $\deg t_1=-1$,   $\deg t_0=-2$, $\deg\hbar=-4$ and  $\deg p_k,\deg q_k=-2$,
3590: which corresponds to the convention
3591: that the Maslov index of any path in the
3592: $1$-point group $Sp(0)$ equals 0.
3593: The term $t_1t_0^2/2=\int\limits_{S^1} t^{\wedge 3}/6$ is the contribution of the moduli
3594: space $S^1$ of constant maps $\C P^1\to \R\times S^1$ with $3$ marked points.
3595:  The term $-\frac{t_1\hbar}{24}$ is accounted for
3596: the contribution of constant curves of genus $1$ (see
3597: \cite{Witten2}), and the term $t_1p_kq_k$ represents the
3598: contribution $t_1=\int\limits_{S^1} t$ of trivial curves of multiplicity
3599: $k$ with one marking. All other curves do not contribute to $\bH$
3600: for dimensional reasons and because $t_1^2=0$.
3601: 
3602: Notice that if we organize all variables $p_k,q_k$   into
3603:  formal Fourier series  (comp. \cite{Getzler})
3604: \begin{equation}
3605: u(x)= \sum\limits_{k=1}^\infty(p_ke^{ixk}+q_ke^{-ixk}),
3606: \end{equation}
3607:  then the term accounting for
3608:   the contribution of rational curves in the formula (\ref{eq:circle})
3609:   takes the form
3610:   %CHANGE
3611:   \begin{equation}
3612:      \frac{t_1}{4\pi }\int\limits_0^{2\pi} (t_0+   u(x)) ^2  \ dx,
3613: \end{equation}
3614: see further discussion of this $u$-formalism in Section \ref{sec:Bott} below.
3615:  }
3616: \end{example}
3617: 
3618: 
3619:  We will  associate now with $(\fW,D)$    two other differential algebras, $(\fP,d)$ and $(\fA,\partial)$, which can be viewed
3620:  as {\it semi-classical} and {\it classical} approximations of the Weyl  differential algebra.
3621: 
3622: Let us denote  by
3623: \begin{description}
3624:  \item{-} $\fP$ -- a graded Poisson algebra  of formal power series in variables
3625:   $p_\g $ with coefficients which are polynomials of $q_\g, z_j,
3626:   z_j^{-1}$, and formal power series of $t_i$,
3627:   \footnote{See the first footnote in Section \ref{sec:3algebras}.} and by
3628: \item{-} $\fA $ --  a graded commutative algebra generated by variables
3629:  $q=\{q_\g\}_{ \g\in\Pc}$   with coefficients in   the algebra
3630:     $\C[H_2(V)][[t]]$.
3631:     \end{description}
3632:   The Poisson bracket on $\fP$ is defined by the formula
3633:   \begin{equation}\label{eq:Poisson}
3634:   \{h,g\}= \sum\limits_\g  {\kappa_\g}\left(\frac{\partial h}{\partial p_\g}
3635:   \frac{\partial g}{\partial q_\g}-(-1)^{{\deg h}{\deg g}}
3636:    \frac{\partial g}{\partial p_\g}\frac{\partial h}{\partial q_\g}
3637:  \right)    ,
3638:  \end{equation}
3639:  assuming that $h$ and $g$ are {\it monomials}.
3640:  When computing partial derivatives, like $\frac{\partial h}{\partial q_\g}$,
3641:   one should remember that we are working in the super-commutative environment,
3642:  and in particular the operator $\frac{\partial }{\partial q_\g}$
3643:  has the same parity as  the variable $q_\g$.
3644:  \begin{remark}\label{rem:u-notation}
3645:  {\rm
3646:   Notice, that if  similarly to  Example \ref{ex:circle} above
3647:  we organize the variables $p_k=p_{\g_k},q_k=q_{\g_k}$ corresponding to multiples
3648:  of each simple periodic orbit $\g=\g_1$ into  a
3649:  Fourier series
3650:  $$u_\g=\sum\limits_{k=1}^\infty(p_ke^{ixk}+q_ke^{-ixk}),$$
3651:  then the value of the Poisson tensor
3652:  (\ref{eq:Poisson}) on covectors $\delta u,\delta v$ takes      the
3653:   form
3654:  \begin{equation}
3655:  \label{eq:u-notation}
3656:  \frac{1}{2\pi i}\int\limits_0^{2\pi}  (\d u)'\d vdx  \,.
3657:  \end{equation}
3658:     }
3659:  \end{remark}
3660:  \bigskip
3661: 
3662:     In order to define differentials
3663:      on the algebras $\fA$ and $\fP$ let us  first     make
3664: the following observation.
3665: \begin{lemma} \label{lm:HH-Weyl}
3666: We have $$[\bH,\bH]=\frac{1}{\hbar}\{\bH_0,\bH_0\}+\dots\,,$$
3667: and for any $f=\frac{1}{\hbar}\sum f_g\hbar^g\in \frac{1}{\hbar}\fW$
3668:  we have
3669: $$D^\bH(f) = \frac{1}{\hbar}\{\bH_0,f_0\}+\dots\,.$$ In
3670: particular, $\bH_0$ satisfies the equation
3671: $\{\bH_0,\bH_0\}=0.$
3672: \end{lemma}
3673: 
3674: To cope with a growing number of indices we will rename $\bH_0$
3675: into $\bh$.
3676: %%%%%%%%%
3677:    Lemma \ref{lm:HH-Weyl} allows us to define the differential
3678:    $d=d^{\bh}:\fP\to\fP $ by the formula
3679:    \begin{equation}
3680:    dg=\{\bh,g\} \quad\hbox{ for }\quad g\in\fP .
3681: \end{equation}
3682: Theorem \ref{thm:HH} then implies
3683:    \begin{proposition}\label{prop:dd-Poisson} We have
3684:    $d^2=0$ and $d\{f,g\}=\{df,g\}+ (-1)^{\deg f}\{f,dg\}$ for any homogeneous element
3685:    $f\in\fP$. In other words,   $(\fP,d)$ is a
3686:    graded differential Poisson algebra with  unit.
3687:    \end{proposition}
3688:    Proposition \ref{prop:dd-Poisson}
3689:    enables us to define the homology $H_*(\fP,d)$ which inherits  from
3690:     $\fP$ the structure of a graded Poisson algebra with unit.
3691: 
3692:      Let us recall that according to \ref{thm:finiteness-of-H}
3693:      $\bh|_{p=0}=\bh_{\const}$,
3694:      where $\bh_{\const}$ accounts for constant rational
3695:      holomorphic curves, and thus it is independent of $q$-variables.
3696:      In fact, $$\bh_{\const}(t)=\frac16\sum\limits_{i,j,k=1}^K c^{ijk}t_it_jt_k,$$
3697:      where $c^{ijk}=\int\limits_V\Theta_i\wedge\Theta_j\wedge\Theta_k$ are
3698:      the structural
3699:      coefficients of the cup-product.
3700:           Hence,  $$\bh=\bh_{\const}+\sum h_{\g }(q,t,z)p_\g   +\dots,$$
3701:      where $\dots$ denote terms at least quadratic in $p_\g$.  Thus
3702:      we have
3703:       \begin{equation}
3704:      \{\bh,\bh\} = 2\sum\limits_{\g ,\g'\in\Pc }\kappa_{\g'}
3705:       h_{\g'  }(q,t)\frac {\partial h_{\g  }}{\partial q_{\g'}}(q,t)p_{\g}+ o(p)=0\,.
3706:       \end{equation}
3707:         Therefore,
3708:         \begin{equation}\label{eq:d2}
3709:          \sum\limits_{\g'  \in\Pc }\kappa_{\g'}
3710:       h_{\g' }(q,t)\frac{\partial h_{\g  }}{\partial
3711:       q_{\g'}}(q,t)=0
3712:       \end{equation}
3713:       for all $t$ and all $\g\in\Pc$.
3714: 
3715: 
3716:    Let us define a differential $\partial: \fA\to\fA$ by the
3717:       formula
3718: 
3719:       \begin{equation}
3720:       \partial f=\{\bh,f\}|_{\{p=0\}}=\sum\limits_{\g\in\Pc}\kappa_\g h_\g\frac{\partial
3721:       f}{\partial q_\g}.
3722:       \end{equation}
3723:       Then the equation (\ref{eq:d2}) is equivalent to
3724: 
3725:       \begin{proposition}\label{prop:zero-section}
3726:      $\partial^2=0$, and hence $(\fA,\partial)$ is a graded
3727:      commutative
3728:      differential algebra with  unit.
3729:      \end{proposition}
3730: 
3731:    The homology group $H_*(\fA,\partial)$
3732:      inherits the structure
3733:  of a graded commutative algebra with unit.
3734: 
3735: As it was already mentioned in Section \ref{sec:informal},
3736:  it is convenient to view the Poisson algebra $\fP$ as an algebra
3737: of functions on an infinite-dimensional symplectic super-space $\bV$ with the even symplectic form
3738: $\bo=\sum k_\g^{-1}dp_\g\wedge dq_\g$.
3739: Then  the differential $d^\bh$ is   the  Hamiltonian  vector field on $\bV$ generated by the  Hamiltonian
3740: function $\bh$. One should remember, however, that the $p$-variables are formal,
3741: so all that we have is the infinite jet of the symplectic space
3742: $\bV$ along the $0$-section.  The equation $\bh|_{p=0}=\bh_{\const}$ translates into
3743:  the fact that the vector field $d^\bh$
3744: is tangent to the $0$-section, and the differential $\partial$ is just the restriction of this vector field
3745: to the $0$-section.
3746:   The higher order terms in the expansion of $\bh$ with respect
3747:   to $p$-variables define a sequence  of (co-)homological operations on the algebra $\fA$.
3748:       \medskip
3749: 
3750:     Notice  also that the differentials $D,d$ and $\partial  $ do not involve any differentiation
3751:     with respect to $t$. Hence the  differential algebras
3752:       $(\fW,D^{\bH})$, $(\fP,d^{\bh})$ and
3753:        $(\fA,\partial)$ can be viewed
3754:              as {\it families} of differential algebras, parameterized
3755:       by $t\in H^*(V)$, and in particular,
3756:       one can compute the homology at any fixed $t\in H^*(V)$.   We will
3757:       sometimes denote
3758:        the corresponding
3759:        algebras and their homology
3760:        groups with the subscript $t$, i.e.
3761:         $(\fW,D)_t$,  $H_*(\fP,d)_t,$ etc., and call them
3762:        {\it specialization} at the point $t\in H^*(V)$.
3763:   We will also use the notation
3764:   $$H_*^{\SFT}(V,\xi|\,J,\a ),\;\;
3765:     H_*^{\RSFT}(V,\xi|\,J,\a  ), \quad\hbox{and }\quad
3766:   H_*^{\cont}(V,\xi|\,J,\a )$$
3767:      instead of
3768:    $H_*(\fW,\partial), H_*(\fP,\partial)$ and $H_*(\fA,\partial)$,
3769:    and
3770:    will usually omit the   extra data $J,\a$ from the notation:
3771:   as we will see in Section \ref{sec:composition} below all these
3772:   homology algebras are independent  of  $J,\a$ and other extra
3773:   choices, like
3774:   closed forms representing cohomology classes of $V$, a coherent orientation of the moduli spaces,
3775:   etc. The
3776:   abbreviation $\RSFT$ stands here for  {\it Rational Symplectic
3777:   Field Theory}.
3778: \begin{remark}\label{rem:grading}
3779: {\rm It is   important to observe that the algebras $\fW,\fP $ and
3780: $\fA$ have an additional grading by elements of $H_1(V)$ (comp. Section
3781: \ref{sec:Floer} above). This
3782: grading is also inherited by the corresponding homology algebras. However, this grading carries a
3783: non-trivial information only when we consider homology of algebras, specialized at points
3784: $t=\sum\limits_1^K t_i\Theta_i$ with  $t_i=0$ for at least some of the
3785:   coordinates $t_i$ corresponding
3786: $1$-dimensional forms.  Otherwise
3787: all cycles in these algebras are graded by  the $0$-class from $H_1(V)$.}
3788: \end{remark}
3789: 
3790: \subsection{  Symplectic cobordisms}\label{sec:af-cobordism}
3791: 
3792: 
3793: 
3794: \subsubsection{Evaluation maps and correlators}
3795:  \label{sec:sympl-eval}
3796: Let us now repeat the constructions of the previous section for a
3797: general directed symplectic cobordism $W=\ora{V^-V^+}$
3798:  between  two contact manifolds
3799:  $V^-$ and $V^+$ with fixed contact forms $\alpha^-$ and $\alpha^+$.
3800:  As in Section \ref{sec:contact-eval} we consider the sets $\Pc^\pm$
3801:   of periodic orbits of the Reeb fields $R^\pm=R_{\alpha^\pm}$ as   discrete topological spaces.
3802: 
3803: We denote by $\Mc^A_{g,r,s^-,s^+}( W,J)$ the  disjoint union
3804:    $$\bigcup\Mc_{g,r}^A(\Gamma^-, \Gamma^+;W,J),$$
3805:    where the union is taken over all sets $(\Gamma^-, \Gamma^+)$ of cardinality $(s^-,s^+) $, and
3806:   consider  three sets of evaluation maps:
3807:             $$ev^0_i: \Mc^A_{g,r,s^-,s^+}( W,J) \to W,\;i=1,\dots,r,$$
3808:             $$ev^\pm_j:\Mc^A_{g,r,s^-,s^+}(W,J)    \to \Pc^\pm,\;j=1,\dots      ,s^\pm,$$
3809:               where $ev_i^0$ is the  evaluation map $f (y_i)$ of
3810:                the map $f$  at the $i$-th marked point $y_i$, while
3811:                             $ev^\pm_j$   are the evaluation maps
3812:                             at asymptotic markers
3813:                             $\mu^{\bx^\pm}_j$, i.e.
3814:   $ev_j^\pm(f)$ is a point of $
3815:               \Pc^\pm$ representing the orbit $\g^\pm_j$, which contains the image of the
3816:               corresponding marker.
3817:                The evaluation maps $ev^0_i$ and $ev^\pm_j$ can be combined into a map
3818:               $$ev:\Mc^A_{g,r,s^-,s^+}(W,J)\to W^r\times (\Pc^+)^{s^+}\times (\Pc^-)^{s^-}\,.$$
3819: 
3820:             Now we are ready to define   degree $0$, or symplectic
3821:    correlators.  We will have to consider on $W$ differential
3822:   forms, which do not necessarily have compact support, but
3823:  have,   however, {\it cylindrical ends}. We say that a
3824:    differential  form  $\theta$ on $W$ is said to have cylindrical ends
3825:   if it satisfies the following
3826:   condition:
3827: \smallskip
3828: 
3829: \noindent there exists  $C>0$  such that
3830: $$\theta|_{V^-\times(-\infty,-C)}=(\pi^-)^*(\theta^-)\quad\hbox{ and}\quad
3831:  \theta|_{V^+\times(C,\infty)}=(\pi^+)^*(\theta^+),$$ where
3832: $\theta^\pm$ are forms on $V^\pm$, and $\pi^\pm$ are the
3833: projections of the corresponding ends to $V^{\pm}$.
3834:  We will denote the forms $\theta^\pm$ also by  $\restr^\pm(\theta)$, or $\theta|_{V^\pm}$.
3835:   In what follows we assume that all considered differential  forms on $W$ have   cylindrical ends.
3836:  \smallskip
3837: 
3838:   Given $r$    differential
3839: forms $\theta_1,\dots,\theta_r$ on $W$ and $s^\pm$   cohomology
3840: classes $$\alpha^\pm_1,\dots, \alpha^\pm_{s^\pm}\in
3841: H^*(\Pc^\pm)=H^*_0(\Pc^\pm)$$  we define the degree $0$
3842: correlator
3843: \begin{eqnarray}
3844: \up{0}\langle\,\theta_1,\dots,\theta_r;\alpha^-_1,\dots,
3845: \alpha^-_{s^-};\alpha^+_1,\dots,
3846: \alpha^+_{s^+} \rangle^A_{g}= \hfill \\
3847: \nonumber  \int\limits_{\overline{\Mc_{g,r,s^-,s^+ }^A}}ev^*(\theta_1\otimes\dots
3848: \otimes \theta_r\otimes\alpha^-_1\otimes\dots\otimes\alpha^-_{s^-}
3849: \otimes\alpha^+_1\otimes\dots\otimes\alpha^+_{s^+} ).
3850: \end{eqnarray}
3851: 
3852: Similar to
3853:   Section \ref{sec:contact-eval} above, we denote  the cohomology classes in
3854:   $H^*(\Pc^+)= H^0(\Pc^+)$
3855:   (resp. in   $H^*(\Pc^-) = H^0(\Pc^-)$) by
3856:   $p^+$ (resp.   $q^-$), and write $$p^+=\sum\limits_{\g\in\Pc^+}k_g^{-1}p^+_\g[\g]\quad
3857:   \hbox{\big(resp.}\quad q^-=\sum\limits_{\g\in\Pc^-}k_g^{-1}q^-_\g[\g]\,\big).$$
3858:    We will also fix a basis $A_1,\dots,A_N$ of
3859:  $H_2(W)$ and denote
3860:  by $d=(d_1,\dots,d_N)$ the degree of $A$ in this basis.
3861: %CHANGE
3862: 
3863: Let us call a system of linearly independent
3864:  closed forms  $\theta_1,\dots,\theta_m $
3865:   on $W$ with cylindrical ends
3866:  {\it basic}, if
3867:            \begin{equation}\label{eq:basic}
3868:  \begin{split}
3869:  &\hbox{a) the image}\quad
3870:     \restr^\pm(L(\theta_1,\dots,\theta_m ))\\
3871:     &\quad\hbox{ generates }\quad \im (H^*(W)\to H^*(V^\pm)); \\
3872:     & \hbox{b) the homomorphism}
3873:  \quad \Ker\big((\restr^+\oplus \restr^-)|_L\big)\to
3874:  H_{\mathrm{comp}}^*(W)\\
3875:  &\quad\;\hbox {is bijective.}\\
3876:  \end{split}
3877:  \end{equation}
3878:         Here we denote by $L(\theta_1,\dots,\theta_m )$ the subspace
3879:  generated by the forms $\theta_1,\dots,\theta_m $, and by
3880:  $H_{\mathrm{comp}}^*(W)$ the cohomology with compact support.
3881:  Equivalently, one can say that a basic system of forms consists
3882:  of a basis of $H^*(W)$ together with a basis of
3883:  $\Ker\big(H^*_{\comp}(W)\to H^*(W)\big)$.
3884: 
3885: A general point     $t\in L(\theta_1,\dots,\theta_m )$  we will
3886:    write in the form
3887:     $t=\sum\limits_1^mt_i\theta_i.$
3888:    The grading of the variables
3889:  $t,p^+,q^- $  is defined   as in Section
3890:  \ref{sec:correlators}:
3891:  \begin{equation}
3892:  \begin{split}
3893:  \deg(t_i) =&\;\deg(\theta_i)-2;\,\\
3894:   \deg (p_\gamma^+) =&-\CZ(\gamma^+)+(n-3),\\
3895:    \deg (q_\gamma^-) =&\; \CZ(\gamma^-)+(n-3)\\
3896:    \end{split}
3897:    \end{equation}
3898: 
3899: 
3900: 
3901: 
3902: 
3903: 
3904: \subsubsection{Potentials of symplectic cobordisms}\label{sec:relative}
3905:   Let us now organize the  correlators
3906:    into a generating function, called the { \it potential} of the
3907:    symplectic
3908:    cobordism $(W=\ora{V^-V^+},J,\a^\pm)$
3909:   \begin{equation}\label{eq:GW-potential}     \begin{split}
3910:   &{ { \bF}}=
3911:    {  \bF}_{W,J,\alpha^\pm} =\frac1\hbar
3912:    \sum\limits_{g=0}^\infty {\bF}_g\hbar^g,   \\
3913:    &\quad\hbox{  where }\\
3914:   &{\bF }_g=\sum\limits_d\sum\limits_{r  ,s^\pm=0}\frac{1}{r!s^+!s^-!}\up0\langle
3915:     \underbrace{t,\dots,t}_r;
3916:  \underbrace{q^-,\dots,q^-
3917:  }_{s^-};\underbrace{p^+,\dots,p^+}_{s^+}
3918:  \rangle^d_{g}z^d\,.\\
3919:  \end{split}
3920:  \end{equation}
3921: 
3922: When $W$ is a closed symplectic manifold, then the potential $F$
3923: is just the Gromov-Witten invariant of the symplectic manifold
3924: $W$. However, if $W$ is not closed, then $F$ itself is not an
3925: invariant. It depends on    particular forms
3926: $\theta_i$, rather than their cohomology classes, on $J$, on a
3927: coherent orientation, and several other choices. We will see,
3928: however, that the {\it homotopy class} of $F$, which we define in
3929: Section \ref{sec:chain} below, is independent of most of these
3930: choices.
3931: \medskip
3932: 
3933: In order to make sense out of the expression for
3934: $\bF$ let us consider a graded commutative
3935:  algebra $ \fD=\fD({W,\a^\pm })$  which consists of
3936: power series of the form
3937: \begin{equation}
3938: \sum\limits_{\Gamma,d,g}\varphi_{\Gamma,d,g} (q^-,t ) z^d(p^+)^{\Gamma}\hbar^g,
3939: \end{equation}
3940: where $\varphi_{\Gamma,d,g}$ are polynomials
3941:  of $q_\g$, formal power series of variables
3942:   $t_i$,\footnote{See the first footnote in Section \ref{sec:3algebras}.}
3943:  and where $\Gamma$ and
3944: $d$ satisfies the following Novikov type inequality:
3945: \begin{equation}\label{eq:Novikov}
3946: [\omega](d)=\sum
3947: d_i\int\limits_{A_i}\omega>-\left|\Gamma\right|=-\sum\limits_{i=1}^k|\gamma_i|,
3948: \end{equation}
3949: where
3950: $\Gamma=\{\g_1,\dots,\g_k\}$, and
3951: $|\gamma_i|=\int\limits_{\gamma_i}\alpha^+$ is the period of the
3952: periodic orbit $\g_i\in\Pc^+$, or equivalently its action. Recall
3953: that $(p^+)^\Gamma=p^+_{\g_1}\dots p^+_{\g_k}$.
3954: 
3955: 
3956: 
3957:  \begin{proposition}
3958: We have
3959: $$\bF_{W,J,\alpha^\pm}
3960: \in\frac1{\hbar}\fD({W,\alpha^\pm}).$$
3961: \end{proposition}
3962: Let us also consider a bigger algebra $\fD\fD$ which consists
3963: of series
3964: \begin{equation}
3965: \sum\limits_{\Gamma,d}\varphi_{\Gamma,d} (q^-,t,\hbar ) z^d(p^+)^{\Gamma},
3966: \end{equation}
3967: where $\varphi_{\Gamma,d}$ are polynomials of $q^-_\g$, formal power series of $t_i$ and formal {\it Laurent
3968: series}
3969: of $\hbar$, while $\Gamma$ and $d$  still satisfy
3970:  condition
3971: (\ref{eq:Novikov}).
3972: For instance, for any element $f\in \hbar^{-1}\fD$ we have $e^f\in\fD\fD$.
3973: 
3974: The algebra $\fD\fD=\fD\fD(W,J,\alpha^\pm)$ has a structure
3975: of a {\it left $D$-module} over the Weyl algebra
3976: $\fW^-=\fW(V^-,J,\alpha^-)$, and of a {\it right $D$-module} over the Weyl algebra
3977: $\fW^+=\fW(V^+,J,\alpha^+)$.
3978: Indeed,  we first associate with an element
3979: $$\Delta^-= \sum\limits_{\Gamma=\{\g_1,\dots,\g_m\},\Gamma',d,g }
3980:   {\delta}^-_{\Gamma',\Gamma,d,g  }( t )
3981:    (q^-)^{\Gamma'}(p^-)^{\Gamma}z^d\hbar^g\in\fW^-$$
3982:    a differential operator
3983: 
3984: \begin{equation}
3985: \sum\limits_{\Gamma=\{\g_1,\dots,\g_m\},\Gamma',d,g }
3986:  {\delta}^-_{\Gamma,\Gamma',d,g }(t )
3987:    (q^-)^{\Gamma'}{\hbar^{m+g}}\prod\limits_{i=1}^m
3988:    \kappa_{\g_i}
3989:    \overrightarrow{\frac{\partial}
3990: {\partial q^-_{\g_i}}}z^d,
3991: \end{equation}
3992: then  change   the coefficient ring via  the inclusion homomorphism
3993: $H_2(V^-)\to
3994: H_2(W)$,  and finally lift functions
3995: $ {\delta}^-_{\Gamma,\Gamma',d,g }(t )$ to the space of forms with cylindrical
3996: ends on $W$ via the restriction map $t\mapsto t|_{V^-}$.
3997:  We will   denote  the resulting operator by
3998: $\overrightarrow{\Delta^-}$. Similarly we associate with
3999: $\Delta^+\in\fW^+$ an operator $ \overleftarrow{\Delta^+}$ by
4000: first quantizing $q^+_\g\Rightarrow  \hbar
4001: \kappa_\g\overleftarrow{\frac{\partial}{\partial p^+_\g}}$  and then making an
4002: appropriate change of the coefficient ring.
4003:                                    It is straightforward to verify that
4004: for $f\in \fD$ we have $\overrightarrow{\Delta^-}f,\;f \overleftarrow{\Delta^+}\in \fD$,
4005: and for $F\in\fD\fD$ we have
4006: $\overrightarrow{\Delta^-}F,\; F\overleftarrow{\Delta^+}\in \fD\fD$.
4007: 
4008: \medskip
4009: 
4010: Let us denote the Hamiltonians (see Section \ref{sec:3algebras} above)  in
4011: $\fW^\pm$ by $\bH^\pm$ and define a map
4012: $D_W=D_W:\fD\fD\to\fD\fD$ by the formula
4013: 
4014: \begin{equation}\label{eq:DW}
4015: D_W(G)= \overrightarrow{
4016: \bH^-}G-(-1)^{\deg G}G\overleftarrow{\bH^+},\quad G\in\fD\fD,
4017: \end{equation}
4018: where we assume $G$ dimensionally homogeneous.
4019: Clearly, Theorem \ref{thm:HH} implies that $D_W^2=0$. However, the differential algebra
4020: $(\fD\fD,D_W)$ is too big
4021: and instead of considering its homology we will define a differential on the algebra $\fD$,
4022: or which is equivalent but more convenient, on the module $\hbar^{-1}\fD$.
4023: 
4024: For an {\it even} element $F\in\hbar^{-1}\fD$  let us define  a map
4025:  $D^F=T_FD_W:\hbar^{-1}\fD\to\hbar^{-1}\fD$
4026: by the formula
4027: \begin{equation} \label{eq:DF}
4028: D^F(g)=e^{-F}[D_W,g](e^F)=e^{-F}\big(D_W(ge^F)-(-1)^{\deg g}g D_W(e^F)\big),
4029: \quad g\in\hbar^{-1}\fD.
4030: \end{equation}
4031:       The map    $T_FD_W$ is         the linearization of  the map
4032:        $\wt D_W:\hbar^{-1}\fD\to\hbar^{-1}\fD$,
4033:       defined by the formula $$\wt D_W(F)=e^{-F}D_W(e^F),\quad F\in\hbar^{-1}\fD\,.$$
4034:             at the point $ F\in\hbar^{-1}\fD$.  Notice that   if $D_W(e^F)=0$ then
4035:             $D^F(g)=e^{-F}D_W(ge^F)$.
4036:       Let us first state a purely algebraic
4037:       \begin{proposition} \label {prop:SFT-cobordism}
4038:       Suppose that     for $F\in\hbar^{-1}\fD$ we have $D_W(e^F)=0$. Then
4039:    \begin{description}
4040:     \item{1.}   $(D^F)^2=0;$
4041: \item{2.}  The homology algebra         $H_*(\fD,D^{F})$
4042: inherits the structure of a left module over the homology
4043: algebra
4044: $ H_*(\fW^-,D^-)$, and
4045:  the structure of a right module over the homology
4046: algebra $ H_*(\fW^+,D^+)$;
4047:   \item{3.} The homomorphisms
4048: $F^\pm:  \fW^{\pm}  \to \fD $,
4049: defined by  the formulas
4050: \begin{equation} \label{eq:end-homomorphisms}
4051: \begin{split}
4052: &f\mapsto e^{-F}
4053: \overrightarrow{f}e^{F},\,f\in\fW^-,\quad\hbox{and}\\
4054: & f\mapsto  e^{F} \overleftarrow{f}e^{-F},\,f\in\fW^+\,,\\
4055: \end{split}
4056: \end{equation}
4057: commute with the  boundary maps of chain complexes, i.e.
4058: $$F^\pm\circ D^\pm= D^{F}\circ F^\pm,$$ and thus induce
4059: homomorphisms of homology
4060:  $$F^\pm_*:H_*(\fW^{\pm},D^\pm)
4061:  \to H_*(\fD,D^{F}),$$
4062:  as modules over $H_*(\fW^{\pm},D^\pm)$.
4063:  \end{description}
4064:  \end{proposition}
4065: 
4066: \begin{theorem}\label{thm:SFT-cobordism}
4067: The   potential $ \bF\in\hbar^{-1}\fD$  defined  above by the formula
4068: (\ref{eq:GW-potential})
4069: satisfies the equation
4070: \begin{equation}\label{eq:SFT-cycle}
4071: D_We^{\bF}=0 ,
4072: \end{equation} and hence all conclusions of Proposition \ref{prop:SFT-cobordism}
4073: hold for $\bF$.
4074:  \end{theorem}
4075: 
4076:   The appearance of $e^{\bF}$ in  equation
4077: (\ref{eq:SFT-cycle}) has the following reason.
4078:    Similar to Theorem
4079: \ref{thm:HH} above, equation (\ref{eq:SFT-cycle}) follows from the
4080: description of codimension $1$ strata on the boundary of the
4081: moduli space of holomorphic curves in the cobordism $W$, see
4082:  Proposition \ref{prop:boundary} above.   Notice
4083: that $e^{\bF}$ is the generating function  counting possibly disconnected
4084:  holomorphic curves in $W$. Thus the identity
4085: $$\overrightarrow{\bH^-} e^{\bF}- e^{\bF}\overleftarrow{\bH^+}=0$$
4086: asserts, in agreement with Proposition \ref{prop:boundary},
4087: that the  codimension $1$ strata on the  boundary of the moduli space
4088: $\wt\Mc(W)$
4089: of not necessarily connected curves
4090: in $W$ correspond
4091:   to  stable
4092: curves $(f_1,f_2)$ of height $2$, where one of the  curves $f_1,f_2$ belongs to $\wt\Mc(W)$,
4093:  while the second one
4094: is contained in the symplectization of $V^\pm$  and has
4095: precisely one component different from the   straight cylinder
4096: over a periodic orbit from $\Pc^\pm$.
4097: 
4098: 
4099: \medskip
4100: 
4101: \begin{remark}\label{rem:non-closed2}
4102: {\rm
4103: (Comp. Remark \ref{rem:non-closed1} above) The   potential
4104: $\bF$, extended to the space of all, not necessarily closed
4105:  differential forms
4106:  satisfies the   equation
4107: \begin{equation}\label{eq:non-closed2}
4108: d(e^\bF)=D_We^{\bF},
4109: \end{equation}
4110: where $d$ is the de Rham differential.
4111: This equation generalizes
4112: equation   (\ref{eq:SFT-cycle}).
4113: }
4114: \end{remark}
4115:  Following the scheme of Section \ref{sec:3algebras} above we will associate
4116:  now with the cobordism $W$ two other left-right modules, one over
4117:  the
4118:  Poisson algebras $\fP^\pm$, and another over the graded differential
4119:  algebras $\fA^\pm$.
4120: 
4121:  Consider the graded commutative algebra  $\fL=\fL({W,\a^\pm })$
4122:   of
4123: power series of the form
4124: \begin{equation}
4125: \sum\limits_{\Gamma,d}\varphi_{\Gamma,d} (q^-,t ) z^d(p^+)^{\Gamma},
4126: \end{equation}
4127: where $\varphi_{\Gamma,d}$ are polynomials of $q_\g^-$ and formal power seies of $t_i$, while $\Gamma$ and
4128: $d$  satisfies the  above inequality (\ref{eq:Novikov}).
4129: Let us also consider the larger graded commutative algebra
4130: \begin{equation}\label{eq:fL}
4131: \widehat{\fL}=\{\sum\limits_{\Gamma^+,\Gamma^-,d}
4132: \varphi_{\Gamma^+,\Gamma^-,d} (q^-,q^+,t )
4133: z^d(p^+)^{\Gamma}(p^-)^{\Gamma'}\},
4134: \end{equation}
4135: where the Novikov condition (\ref{eq:Novikov}) is satisfied for both pairs
4136: $(d,\Gamma)$ and $(d,\Gamma')$.
4137:    The algebra $\widehat{\fL}$
4138:  has a natural Poisson  bracket so that the homomorphisms
4139: $f\mapsto \widehat{f}$,
4140:  where we  denote by $ \widehat{f}$
4141: the image in $\widehat{\fL}$ of an element $f\in\fP^\pm$ under the
4142: coefficient homomorphism, are Poisson homomorphisms.  We set
4143: $\widehat{\bh}=\widehat{\bh^-}-\widehat{\bh^+}$, and   for any $f\in\fL$ denote
4144: by $L_{f}$ the ``Lagrangian variety''
4145:  $$\{p^-_\g=\kappa_\g\frac{\partial
4146: \widehat{f}}{\partial q^-_\g}, q^+_\g=\kappa_\g\frac{\partial
4147: {\widehat f }}{\partial p^+_\g}\}. $$ Strictly speaking $L_f$ is
4148: an ideal in the Poisson algebra $\widehat{L}$. However, it is
4149: useful to think about $L_f$ as a Lagrangian variety in the
4150: symplectic super-space $(\bV^-)\oplus\bV^+$ with the symplectic
4151: form $$\sum \kappa^{-1}_{\g^-}dp^-_{\g^-}\wedge
4152: dq^-_{\g^-}+\kappa^{-1}_{\g^+}dq^+_{\g^+}\wedge dp^+_{\g^+}\,,$$
4153: and with an appropriate change of the coefficient ring.
4154: 
4155: For any function $f\in\fL$, which satisfies the Hamilton-Jacobi
4156: equation
4157: \begin{equation}\label{eq:HJ1}
4158: \widehat\bh|_{L_f}=0
4159: \end{equation}
4160:  the Hamiltonian vector
4161: field defined by the Hamiltonian $\widehat{\bh}$ is tangent to
4162: $L_{f}$, and hence
4163:   the differential $d^f:\fL\to\fL$, defined by the formula
4164:   $$d^f(g)=\{\widehat\bh,g\}|_{L_f}$$ has the following meaning: we identify
4165:   $\fL$ with the space of functions on $L_f$ and differentiate
4166:   them along the Hamiltonian vector field determined by $\widehat\bh$.
4167: 
4168: Here is an analog of Proposition \ref{prop:RSFT-cobordism}
4169: for the algebra
4170: $\fL$.
4171: \begin{proposition}\label{prop:RSFT-cobordism}
4172:  Suppose that    $\widehat\bh|_{L_f}=0$.
4173:  Then
4174:  \begin{description}
4175:  \item{1.} $(d^f)^2=0;$
4176:  \item{2.} The maps $f^\pm:\fP^\pm\to\fL$, defined by the
4177: formula $g\mapsto {\widehat{g}}|_{L_{f}}$, are homomorphisms of
4178: chain complexes, i.e. $$d^{f}\circ f^\pm=f^\pm\circ
4179: d^\pm;$$
4180: \item{3.} If  $g_1,g_2\in\fP^\pm$ Poisson commute with $\bh^\pm$
4181:  or, in other words, if $g_1,g_2\in{\mathrm {Ker}}\,d^\pm$
4182: then
4183: \begin{equation*}
4184: \{f^\pm(g_1),f^\pm(g_2)\}=f^\pm(\{g_1,g_2\}).
4185: \end{equation*}
4186: where the left-side Poisson bracket is taken  in the algebra $\widehat{\fL}$.
4187: \end{description}
4188: \end{proposition}
4189: Let us recall that $\bF=\bF_W\in\fD$ has the form
4190: $\bF=\hbar^{-1}\sum\limits_{g=0}^\infty\bF_g\hbar^g$.
4191: Again, to simplify the notation we will write $\bff $ instead of
4192: $\bF_0$.
4193: The following theorem is the reduction of Theorem
4194: \ref{thm:SFT-cobordism} to the level of rational Gromov-Witten theory.
4195: \begin{theorem}\label{thm:RSFT-cobordism}
4196: The series $\bff(q^-,p^+,t)$ belongs to the algebra $\fL$ and satisfies
4197: the  equation
4198: \begin{equation}\label{eq:Lagr}
4199: \widehat{\bh}|_{L_{\bff}}=0.
4200: \end{equation}
4201: \end{theorem}
4202: In particular, all statements of  the above Proposition
4203: \ref{prop:RSFT-cobordism} hold for $\bff$, and  this allows
4204: us    to   define    the homology
4205: \begin{equation*}
4206: {}_{\bff}H^{\RSFT}_*(W|J,\a^\pm)=H_*(\fL,d^{\bff}).
4207: \end{equation*}
4208:   The chain maps $\bff^{\pm}$ induce homomorphism of Poisson homology
4209: algebras
4210: \begin{equation}
4211: (\bff^\pm)_*:H^{\RSFT}_*(V^\pm|J,\a^\pm)=H_*(\fP^\pm,d^\pm)\to
4212: {}_{\bff}H^{\RSFT}_*(W|J,\a^\pm)\,.
4213: \end{equation}
4214: 
4215: \bigskip
4216: For the rest of this section we assume that $W$ is a  rational
4217: homology
4218: cobordism, i.e. the restriction maps $$H^*(V^-;\R)\leftarrow
4219: H^*(W;\R)\rightarrow H^*(V^+,\R)$$ are isomorphisms.  Equivalently,
4220: this means that
4221:   the inclusions $V^\pm\to W$ induce isomorphisms of rational
4222:   homology groups.
4223: 
4224: 
4225: The potential $\bff\in\fL$ which we defined above can be written in the
4226:   form
4227: \begin{equation}
4228: \bff=\sum\limits_i\sum\limits_{|\Gamma^+|=i}f_{\Gamma}^i(q^-,t)(p^+)^{\Gamma^+}.
4229: \end{equation}
4230:                         Notice that the assumption that $W$
4231:                          is a homology cobordism implies
4232:   that $f^0(q^-,t)$ is independent of $q^-$.
4233:  Let us now define a homomorphism
4234:  $\bPsi : \fA^+\to \fA^-$
4235:  by the formula
4236:  \begin{equation}
4237:  \bPsi (q^+_{\g } )=f^1_{ \g }(q^-,t)\in\fA^-
4238:  \end{equation}
4239: on the generators $q^+_\g,\g\in\Pc^+$,
4240: of the algebra $\fA^+$ and then   extend by
4241: linearity.
4242: 
4243: \begin{theorem}   The homomorphism
4244: $\bPsi : \fA^+\to \fA^-$ commutes with the boundary operators
4245: $\partial^\pm$, i.e. $\partial^-\circ\bPsi =\bPsi
4246: \circ\partial^+$, and in particular defines a homomorphism of
4247: homology algebras
4248: \begin{equation*}
4249: (\bPsi )_*:H_*(\fA^+,\partial^+)\to H_*(\fA^-,\partial^-).
4250: \end{equation*}
4251: \end{theorem}
4252: 
4253: Without the assumption that $W$ is a homology cobordism
4254: one gets only a correspondence between the algebras $\fA^+$ and $\fA^-$, similar to the
4255: ``semi-classical" case considered above.
4256: 
4257: \subsection{Chain homotopy}\label{sec:chain}
4258: Let $W=\ora{V^-V^+}$ be a directed symplectic cobordism with fixed
4259: contact forms $\a^\pm$ on $V^\pm$. We will discuss in this section
4260: how the function $\bF=\bF_{W,J,\a^\pm}(p^+,q^-,t)$ and other
4261: associated structures change when one replaces  $J$ with another
4262: compatible almost complex structure $J'$ and replaces $t$ with a
4263: form $t'=t+d\theta$ where $\theta$ has   compact support in $W$.
4264: 
4265: Let us begin with some algebraic preliminaries. Two series
4266: $F_0,F_1\in\hbar^{-1}\fD$ are called {\it homotopic}, if they can
4267: be included into  a family $F_s\in\hbar^{-1}\fD,\,s\in[0,1]$,
4268: which satisfies the following differential equation
4269: \begin{equation} \label{eq:homotopy}
4270:     \frac{d F_s}{ds }=
4271:     e^{-F_s}
4272:     \left(\ora{[\bH^-,K_s]}e^{F_s}+
4273:      e^{F_s}\overleftarrow{[
4274:      K_s,\bH^+]}\right),\,\,s\in[0,1],
4275:     \end{equation}
4276:      for a family $K_s\in\hbar^{-1}\fD$.
4277:      Here  $[\bH^-,K_s]$ and $[K_s ,\bH^+]$ are commutators in the
4278: algebra $\hbar^{-1}\widehat{\fD}$, defined similar to
4279: $\widehat{\fL}$ in (\ref{eq:fL}) above, i.e.
4280: \begin{equation}
4281: \widehat{\fD}=\{\sum\limits_{\Gamma^+,\Gamma^-,d,g}
4282: f_{\Gamma^+,\Gamma^-,d} (q^-,q^+,t )
4283: z^d\hbar^g(p^+)^{\Gamma^+}(p^-)^{\Gamma^-}\},
4284: \end{equation}
4285: where the Novikov condition (\ref{eq:Novikov}) is satisfied for
4286: both pairs $(d,\Gamma^+)$ and $(d,\Gamma^-)$.
4287:       In other words, we view $K_s$ as an operator on
4288:        $\hbar^{-1}\fD$, acting by the multiplication
4289:     by the series $K_s$,
4290:     and  view $\bH^-$ and $ \bH^+  $
4291:     as left and right differential operators.
4292: 
4293: 
4294: Notice that the family $ K_s\in\hbar^{-1}\fD, s\in [0,1],$
4295: defines a flow $\Phi^s=\Phi_K^s: \fD\fD\to\fD \fD$,
4296:   by a differential equation
4297:    \begin{equation} \label{eq:chain-flow}
4298:     \frac{d \Phi^s(G) }{ds }=\cK_s\big(\Phi^s(G)\big),
4299:  \end{equation}
4300:  where we set
4301: \begin{equation*}
4302:     \cK_s(G)=\left(\ora{[\bH^-,K_s]}G+
4303:     G\overleftarrow{[ K_s,\bH^+]}\right),\,\,s\in[0,1].
4304:  \end{equation*}
4305: The linear operators $\Phi^s$ preserve  the ``submanifold" $\fE=e^{
4306: \hbar^{-1}\fD_{\even}}$, where $\fD_\even$ is the even part of
4307: $\fD$, and we have $$\Phi^s(e^{F_0})=e^{F_s},$$ where
4308: the family $F_s$ satisfies the equation (\ref{eq:homotopy}).
4309: 
4310: The tangent space to $\fE$ at a point $e^F$, $F\in \hbar^{-1}\fD_\even$, consists of
4311: series $fe^F$, $f\in\hbar^{-1}\fD_\even$, and thus it is naturally parameterized by
4312: $\hbar^{-1}\fD_\even$. With respect to this parameterization  the
4313: differential  of the flow $\Phi^s|_\fE$ defines a family of maps
4314: $  T_F^s:\hbar^{-1}\fD_\even\to\hbar^{-1}\fD_\even$, $F\in \hbar^{-1}\fD_\even$, by the formula
4315:  \begin{equation}\label{eq:lin-flow}
4316:     T_F^s(f)=e^{-F_s}\Phi^s(fe^F),\;\;\hbox{ where
4317:     }\;\;F_s=\Phi^s(F);
4318:   \end{equation}
4319:   We extend   $T_F^s$ to the whole
4320:    $\hbar^{-1}\fD$ by the same formula
4321:    (\ref{eq:lin-flow}).
4322: Let us list some properties of the flows $\Phi^s$ and $T^s_F$
4323: \begin{proposition}\label{prop:SFT-chain}
4324: Suppose that    for an  element  $F\in\hbar^{-1}\fD_\even$ we
4325:  have $$D_W(e^F)=\ora{\bH^-}e^F -e^F\ola{\bH^+}=0.$$
4326:     Then
4327: \begin{description}
4328: \item{1.}
4329:        The   flow  $T^s_F :\hbar^{-1}\fD\to\hbar^{-1}\fD$
4330: satisfies the equation
4331: \begin{equation}
4332:  T_F^s\circ D^{F}=D^{F_s}\circ T_F^s.
4333:   \end{equation}
4334:   for all $s\in[0,1].$
4335:   In particular,    $D_W(F_s)=0$  for all $s\in[0,1]$, and $T_F^s$
4336:        defines a family of isomorphisms
4337: $ H_*(\fD,D^{F})\to
4338:  H_*(\fD,D^{F_s})$.
4339: \item{2.} The homology class
4340: $[e^{F_s}]\in H_*(\fD \fD,D_W)$ is independent of  $s$.
4341:                     \item{3.} The diagram
4342: \bigskip\bigskip
4343: 
4344: \centerline{$ \mathop{\fD\qquad}\limits_{\;
4345:  \; \nwarrow \, F^\pm}
4346: \mathop{\longrightarrow}\limits^{T_F^s}
4347: \mathop{\fD}\limits_{\nearrow \, F_s^\pm}$}
4348: 
4349: \centerline{$\fW^\pm$}
4350: 
4351: homotopically commutes, i.e. there exist operators
4352: $A_s^\pm:\fW^\pm\to\fD$,
4353: such that
4354: \begin{equation}\label{eq:SFT-diagram}
4355: (T_F^s)^{-1}\circ F_s^\pm- F^\pm=D^{F}\circ A_s^\pm+ A_s^\pm\circ
4356: D^\pm\,.
4357: \end{equation}
4358: In particular,
4359: this diagram commutes on the level of homology algebras.
4360:   \end{description}
4361:   \end{proposition}
4362:   %%%%%%%%%%%%%%%%%%%%%%%%%
4363: The proof of this proposition is a  straightforward computation by
4364: differentiating  the corresponding equations.
4365: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4366: To illustrate the argument, let us   verify
4367: (\ref{eq:SFT-diagram})
4368: in \ref{prop:SFT-chain}.
4369: 
4370: Take, for instance, $f\in\fW^-$ and
4371: set
4372:  \begin{equation*}
4373: \begin{split}
4374: G_s(f)=&(T_F^s)^{-1}(e^{-F_s}\ora fe^{F_s}))-e^{-F}\ora fe^F\\&=
4375: e^{-F}(\Phi^s)^{-1}\left(\ora fe^{F_s}\right)-e^{-F}\ora fe^F.\\
4376: \end{split}
4377: \end{equation*}
4378: 
4379: 
4380: 
4381: Then we have $G_0(f)=0 $ and
4382: 
4383: \begin{equation*}
4384: \begin{split}
4385: \frac{dG_s(f)}{ds}
4386: &= -e^{-F}\left((\Phi^s)^{-1}
4387:  (  \cK_s \ora{f}  e^{F_s }  )\right)
4388: +  e^{-F}
4389: (\Phi^s)^{-1}\left(
4390:   \overrightarrow{f} \cK_s\left( e^{F_s }\right)\right)
4391:  \\
4392: & =           e^{-F}(\Phi^s)^{-1}\left(
4393:            \left[\cK_s,\overrightarrow{f}\right]
4394:             e^{F} \right).  \\
4395: \end{split}
4396: \end{equation*}
4397: Now remember that $\cK_s=[\widehat{\bH}  ,K_s]$,
4398:  where
4399: $\widehat{\bH}=\overrightarrow{\bH^-}-\overleftarrow
4400: {\bH^+ }$, and using the Jacobi identity we get
4401: \begin{equation*}
4402: \begin{split}
4403:   \frac{dG_s(f)}{ds}
4404:   &=  e^{-F} (\Phi^s)^{-1}\left(
4405:            \left[[\widehat{\bH}  ,K_s],
4406:            \overrightarrow{f}\right]
4407:             e^{F_s}\right) \\
4408:       & =
4409:  e^{-F}(\Phi^s)^{-1}\left(
4410:            \left[[K,\overrightarrow{f}],
4411:            \widehat{\bH}  \right]
4412:             e^{F_s }  \right)\\
4413:             &+
4414:               e^{-F} (\Phi^s)^{-1}\left(
4415:            \left[[\overrightarrow{f},
4416:            \widehat{\bH}  ],K_s\right]
4417:              e^{F_s } \right). \\
4418:                            \end{split}
4419:          \end{equation*}
4420:          Let us define now a linear operator
4421:          $B_s:\hbar^{-1}\fW_-\to\hbar^{-1}\fD $ by the formula
4422:          \begin{equation}
4423:          B_s(g)=   e^{-F}(\Phi^s)^{-1}\left(
4424:          \overrightarrow{[g,K_s] }
4425:           e^{F_s}\right).
4426:          \end{equation}
4427:           for $g\in \hbar^{-1}\fW_-$.
4428:         Recall that
4429:          $D_W(e^F)=\widehat{\bH}e^{F}=0$, and observe that
4430:          $\widehat{\bH}$ commutes with
4431:          $\cK_s=[\widehat{\bH},K_s]$ because $\widehat{\bH}\circ
4432:           \widehat{\bH}=0$.  We also have
4433:                     $[\widehat{\bH},f]= D^-f$ for
4434:                      $f\in\hbar^{-1}\fW_-$.
4435:           Hence
4436:            $ \frac{dG_s(f)}{ds}$ can be rewritten as
4437:           \begin{equation}\label{eq:derivative}
4438:           \frac{dG_s(f)}{ds}= D^{F}(B_s(f))    +
4439:             B_s(D^-(f))
4440:           \end{equation}
4441: 
4442:         Finally we integrate $B_s$ into the required linear operator
4443:          $A^-:\hbar^{-1}\fW^-\to\hbar^{-1}\fD$:
4444:          \begin{equation}
4445:         A_s^-(g)=\int\limits_0^s B_s(g)ds,
4446:         \quad\hbox{for}\quad g\in\hbar^{-1}\fW^-.
4447:         \end{equation}
4448: 
4449:          In view of   equation
4450:          (\ref{eq:derivative})
4451:          we get
4452:   \begin{equation}
4453: (\Phi^s)^{-1}\circ(F^s)^- -(F)^-=D^F\circ A_s^-+
4454: A_s^-\circ D^-\,.
4455: \end{equation}
4456:  \qed
4457: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4458: 
4459: \bigskip
4460:  Let us consider now a generic family $J^\tau$, $\tau\in[0,1]$,     of
4461: compatible almost complex structures on $W$ connecting $J^0=J$
4462: with $J^1=J'$. We  assume that the deformation $J^\tau$ is
4463: fixed outside of a compact subset of $W$.
4464: 
4465: \medskip
4466: Set
4467: \begin{equation}
4468: \Mc_{g,r,s^+,s^-}(W,\{J^\tau\}) =\bigcup\limits_{\tau\in[0,1]}
4469: \Mc_{g,r,s^+,s^-}(W,J^\tau).
4470: \end{equation}
4471: 
4472: The evaluation maps   defined for each $\tau$ can be
4473: combined into the evaluation map
4474: \begin{equation}
4475: ev:\Mc_{g,r,s^+,s^-}(W,\{J^\tau\})\to
4476: (W\times I)^r\times\left(\Pc^-\right)^{s^-}\times\left(\Pc^+\right)^{s^+}.
4477: \end{equation}
4478: 
4479: 
4480: Consider  closed forms
4481: $\widehat {\theta}_1,\dots \widehat{\theta}_r$    on
4482:   $W\times I$, such that
4483:   $\widehat{\theta}_i=\wt{\theta}_i+d\beta_i$, $i=1,\dots, r$, where
4484:   $\wt{\theta}_i$ is the pull-back of  a form $\theta_i$ on $W$
4485:   with cylindrical ends, and
4486:   $\beta_i$ has   compact support in $W\times \Int I$.
4487: 
4488: 
4489: 
4490: Similarly to correlators  of degree $-1$ and $0$
4491: (see \ref{sec:correlators} and \ref{sec:sympl-eval})
4492: we  can define correlators of degree $1$, or
4493: $1$-parametric correlators by the formula
4494: \begin{eqnarray}
4495:  \up1\langle\widehat {\theta}_1,\dots \widehat{\theta}_r\,;\theta^-_1,\dots,
4496: \theta^-_{s^-};\theta^+_1,\dots,
4497: \theta^+_{s^+} \rangle^A_{ g}  =
4498:   \int\limits_{\overline{\Mc_{g,r,s^+,s^-}(W,\{J^\tau\})} }\\
4499: \nonumber   ev^*(\widehat{\theta}_1\otimes\dots
4500: \otimes \widehat{\theta}_r
4501: \otimes\theta^-_1\otimes\dots\otimes\theta^-_{s^-}
4502: \otimes\theta^+_1\otimes\dots\otimes\theta^+_{s^+} ),
4503: \end{eqnarray}
4504:   for a homology class $A\in H_2(W)$, and cohomology
4505: classes $$\theta^\pm_i\in H^*(\Pc^\pm),\; i=1,\dots,s^\pm.$$
4506: 
4507: 
4508: 
4509: Consider a closed form $T=\widetilde{t}+d\beta$,
4510:  where  the notation $\wt t$ and $\beta$
4511:   have the same meaning as
4512:  above, i.e. $\wt t$ is the pull-back of a form $t$
4513:  on $W$ with cylindrical ends, and $\beta$ has
4514:    compact support in $ W\times\Int I$.
4515: We can organize the correlators
4516:  $$\up1\langle T,\dots,T\,; q^-,\dots,q^-,p^+,\dots, p^+\rangle^A_{
4517:  g}$$
4518:     into a generating function
4519: $\bK=\frac 1{\hbar}\sum\limits_{g=0}^\infty\bK_g\hbar^g\in\frac 1{\hbar}\fD$, defined
4520: by the formula
4521: \begin{equation}
4522: \bK=\sum\limits_d\sum\limits_{g,r
4523: ,s^\pm=0}^\infty
4524: \frac{1}{r!s^+!s^-!}\up1\langle \underbrace{T,\dots,T}_r;
4525:  \underbrace{q^-,\dots,q^-
4526:  }_{s^-};\underbrace{p^+,\dots,p^+}_{s^+}\rangle^d_{ g}\hbar^gz^d\,
4527:  \end{equation}
4528: 
4529: 
4530:   \medskip
4531:  Let us define an operator $\cK:\fD\fD\to\fD\fD$
4532:  by the formula
4533:  \begin{equation}
4534:  \cK(G)=  \overrightarrow{ [\bH^+,\bK]}G +G
4535: \overleftarrow{[\bK ,\bH^-]},\; G\in\fD\fD .
4536:   \end{equation}
4537: 
4538: Similar to Theorems \ref{thm:HH} and \ref{thm:SFT-cobordism}
4539: the next theorem  can be viewed
4540: as an algebraic    description of the boundary of the compactified
4541: moduli
4542: space        $\overline{\Mc_{g,r,s^+,s^-}(W,\{J^\tau\})}$.
4543: \begin{theorem}\label{thm:SFT-chain}
4544: For a generic family $J_\tau,\tau\in[0,1]$, of compatible almost complex
4545: structures on $W$ we have
4546: \begin{equation}\label{eq:SFT-chain}
4547: e^{\bF^1}= e^{\cK }(e^{\bF^0}),
4548: \end{equation}
4549: where
4550: $\bF^0=\bF_{W,J_0}(T|_{W\times 0}),\,\bF^1 =\bF_{W,J_1}(T|_{W\times 1}),$ and
4551: $\cK=\cK(T).$
4552: \end{theorem}
4553:  Notice that if we define $\bF^s$,
4554:  $s\in [0,1]$, by the formula
4555: \begin{equation}
4556: e^{\bF^s}= e^{s\cK }(e^{\bF^0}),
4557: \end{equation}
4558: then the flow $\Phi^s(F)=F^s$   satisfies the differential
4559: equation (\ref{eq:chain-flow})
4560:  with $K(s)\equiv \bK$. Hence $\bPhi^0$ and $\bPhi^1$
4561:  are homotopic, and therefore
4562:   Theorem
4563: \ref{thm:SFT-chain} and
4564: Proposition \ref{prop:SFT-chain} imply
4565: \begin{corollary}\label{cor:SFT-chain}
4566: \begin{description}
4567: \item{1.} The homology class $[e^{\bF}]\in H_*(\fD
4568: \fD,D_W)$ is
4569: independent of the choice of a compatible almost complex structure
4570: $J$ on $W$ and of the choice of
4571: $t\,{\mathrm {mod}}\left(d\left(\Omega_{\mathrm
4572: {comp}}(W)\right)\right)$, where $ {\Omega}_{\mathrm
4573: {comp}}(W)$ is the space of forms with compact support.
4574: \item{2.} For a generic compatible deformation $J^\tau$, $\tau\in[0,1]$,
4575: the isomorphism  $T :\fD\to\fD$
4576:    defined by the formula
4577: \begin{equation}
4578:     T (f)= e^{-\bF_1}
4579: e^{\cK}(fe^{\bF^0 } )
4580: \end{equation}
4581:  satisfies the equation
4582: \begin{equation}
4583:  T \circ D^{\bF^0}=D^{\bF^1}\circ T  ,
4584:  \end{equation}
4585:  and thus defines an isomorphism
4586: $ H_*(\fD,D^{\bF^0})\to
4587:  H_*(\fD,D^{\bF^1})$.
4588: \item{3.} The diagram
4589: \bigskip\bigskip
4590: 
4591: \centerline{$ \mathop{\fD\qquad}\limits_{\;
4592:  \; \nwarrow \, \left(\bF^0\right)^\pm}
4593: \mathop{\longrightarrow}\limits^{T }
4594: \mathop{\fD}\limits_{\nearrow \, \left(\bF^1\right)^\pm}$}
4595: 
4596: \centerline{$\fW^\pm$}
4597: 
4598: homotopically commutes, i.e. there exist operators
4599: $A^\pm:\fW^\pm\to\fD$,
4600: such that
4601: \begin{equation}
4602: T^{-1} \circ\left(\bF^1\right)^\pm-
4603: \left(\bF^0\right)^\pm=D^{\bF^0}\circ A^\pm+
4604: A^\pm\circ D^\pm\,.
4605: \end{equation}
4606:   \end{description}
4607: \end{corollary}
4608: 
4609: \bigskip
4610: Consider now the equivalence relation for
4611:  rational   potentials.
4612: 
4613:  Two series
4614:  $ f_0,f_1\in\fL$ are called {\it homotopic}
4615: if there exist families
4616: $f_s, k_s\in\fL$, $s\in[0,1]$, so that the family
4617:  $f_s$ connects $f_0$ and $f_1$ and satisfies
4618: the following Hamilton-Jacobi differential
4619: equation
4620: \begin{equation}\label{eq:HJ}
4621:  \frac{\partial f_s( p^+,q^-,t)}{\partial s}=
4622:  \bG(p^+,\frac{\partial f_s ( p^+,q^-,t)}
4623:  {\partial p^+},
4624:  \frac{\partial f_s ( p^+,q^-,t)}
4625:  {\partial q^-} ,q^-),
4626: \end{equation}
4627:  where
4628:   \begin{eqnarray}
4629:  \nonumber \bG(p^+,q^+,p^-,q^-,t)=\{\bh^+-\bh^-,k_s\}=\\
4630:  \sum\limits_{
4631:  \g^+\in\Pc^+,\g^-\in\Pc^-} \kappa_{\g^-}\frac{\partial \bh^-(p^-,q^-,t)}
4632:  {\partial p^-_{\g^-}}
4633:  \frac {\partial k_s( p^+,q^-,t)}{\partial  q^-_{\g^-}}+\\
4634:   \nonumber \kappa_{\g^+}\frac{\partial k_s( p^+,q^-,t)}{
4635:   \partial  p^+_{\g^+}}\frac {\partial \bh^+(p^+,q^+,t)}
4636:  {\partial q^+_{\g^+}} \,.
4637:  \end{eqnarray}
4638:   Here $\kappa_\g$ denotes, as usual,  the multiplicity of
4639:   $\g$.
4640: 
4641: 
4642: We can view  the correspondence
4643:   \begin{equation*}
4644:   f ( p^+,q^-,t)\mapsto f_s( p^+,q^-,t),
4645:   \end{equation*}
4646:   where $f_s$ is the solution of
4647:   the above equation (\ref{eq:HJ}) with the initial data
4648:   $f_0=f$,
4649:   as a non-linear operator $ S^s:\fL\to\fL$.
4650:   Let us denote by
4651:   $T^s_f $ the linearization of $ S^s$ at a point
4652:   $f$.
4653:   The next proposition is a rational version of Proposition
4654:   \ref{prop:SFT-chain}. It can be either deduced from
4655: \ref{prop:SFT-chain}, or similarly  proven by    differentiation
4656: with respect to the parameter $s$.
4657:   Denote by $\cS$ the subspace of $\fL$ which consists of solutions of
4658:     the Hamilton-Jacobi equation
4659: (\ref{eq:HJ1}), i.e.
4660:  \begin{equation*}
4661:  \widehat\bh|_{L_f}=0,\quad\hbox{  where}\quad
4662:   L_f=\{p^-_\g=\kappa_\g\frac{\partial
4663:  {f}}{\partial q^-_\g}, q^+_\g=\kappa_\g\frac{\partial
4664: {  f }}{\partial p^+_\g}\}.
4665: \end{equation*}
4666: 
4667: 
4668: \begin{proposition}\label{prop:RSFT-chain}
4669:   \begin{description}
4670:   \item{1.} The subspace  $\cS\subset\fL$ is invariant
4671:     under the flow $S^s$.
4672:   \item{2.}   For $f\in \cS$
4673:   the isomorphism  $T^s_f:\fL\to\fL$
4674:  satisfies the equation
4675: \begin{equation}
4676:  T^s_f\circ d^{f }=d^{S^s(f)}\circ T^s,
4677:  \end{equation}
4678:  and thus defines an isomorphism
4679: $H_*(\fL,d^{f})\to
4680: H^*(\fL,d^{S^s(f)} )$.
4681:   \item{3.}  For $f\in\cS$ the diagram
4682:    \bigskip\bigskip
4683: 
4684: \centerline{$ \mathop{\fL\qquad}\limits_{\;
4685:  \; \nwarrow \, \left(f\right)^\pm}
4686: \mathop{\longrightarrow}\limits^{T^s_f}
4687: \mathop{\fL}\limits_{\nearrow \, \left(S^s(f)\right)^\pm}$}
4688: 
4689: \centerline{$\fP^\pm\;$}
4690: 
4691: \bigskip
4692: 
4693:   homotopically commutes.
4694:   \end{description}
4695: \end{proposition}
4696: 
4697:  Theorem     \ref{thm:SFT-chain} reduces on the level of
4698:  rational curves
4699:  to the following
4700: 
4701: \begin{theorem}\label{thm:RSFT-chain}
4702: Let $\bF^0$, $\bF^1$ and $\bK$ be as in Theorem
4703: \ref{thm:SFT-chain}. Set $\bff^0=\bF^0_0,\;\bff^1=\bF^1_0,\;
4704: \bk=\bK_0.$ Then
4705: $\bff_0$ and $\bff_1$ are homotopic, i.e. they can be included into a family
4706: $\bff_s,\,s\in [0,1]$, such that the Hamilton-Jacobi equation
4707: (\ref{eq:HJ}) holds with $f_s=\bff_s$ and  $k_s\equiv\bk$.
4708: \end{theorem}
4709: 
4710: Hence, Proposition \ref{prop:RSFT-chain} implies
4711: \begin{corollary}\label{cor:RSFT-chain}
4712:  For a generic compatible deformation $J_s$,
4713: $s\in[0,1]$, we have
4714: \begin{description}
4715: \item{1.}
4716:  The operator  $S^1:\fL\to\fL$ defines an automorphism of
4717:  the space  of solutions of (\ref{eq:HJ1});
4718:     \item{2.} The linearization $T^1 =T^1_{\bff^0}$ of $S^1$ at the point
4719:     $\bff_0$ satisfies the
4720:   equation
4721: \begin{equation}
4722:  T^1\circ d^{\bff^0}=d^{\bff^1}\circ T^1,
4723:  \end{equation}
4724:  and thus defines an isomorphism
4725: $H_*(\fL,d^{\bff^0})\to
4726: H^*(\fL,d^{\bff^1})$.
4727:   \item{2.} The diagram
4728:    \bigskip\bigskip
4729: 
4730: \centerline{$ \mathop{\fL\qquad}\limits_{\;
4731:  \; \nwarrow \, \left(\bff^0\right)^\pm}
4732: \mathop{\longrightarrow}\limits^{T^1}
4733: \mathop{\fL}\limits_{\nearrow \, \left(\bff^1\right)^\pm}$}
4734: 
4735: \centerline{$\fP^\pm$}
4736: 
4737: \bigskip
4738: 
4739:   homotopically commutes.
4740:   \end{description}
4741: \end{corollary}
4742: 
4743: 
4744: To formulate the ``classical level'' corollary of Theorem
4745: \ref{thm:RSFT-chain}
4746: we assume, as usual, that $W$ is a homology cobordism.
4747: \begin{theorem}\label{thm:classical-chain}
4748: The homomorphisms  $\Psi^1_{J_0} ,\Psi^1_{J_1}: \fA^+\to\fA^-$ associated
4749: to two compatible almost complex structures $J_0$ and $J_1$ are
4750: homotopic, i.e. there exists a map $\Delta:\fA^+\to\fA^-$ such
4751: that
4752: $$\Psi^1_{J_1} -\Psi^1_{J_1} =\partial^-\circ\Delta+\Delta\circ\partial^+.$$
4753:       \end{theorem}
4754: The map $\Delta$ can be
4755: expressed through the function
4756:   $\bk\in\fL$. However, unlike the case of usual Floer homology
4757:   theory, $\Delta$ and $\bk$ are related via a first order non-linear PDE
4758:   (which can be deduced from the equation (\ref{eq:HJ})),
4759:      and thus  one cannot write a general explicit formula  relating
4760:      $\Delta$ and $\bk$.
4761: 
4762: 
4763:      \subsection{Composition of cobordisms}\label{sec:composition}
4764: In this section we   study the behavior of
4765: potentials and associated algebraic structures under
4766: the operation of composition of directed symplectic cobordisms.
4767: 
4768:     Let    us recall (see Section \ref{sec:splitting}) that   given
4769:     a dividing contact type hypersurface $V$ in a
4770:     directed symplectic cobordism $W=\ora{V^-V^+}$ one can split $W$
4771:     into a composition $W=W_-\circledcirc W_+$ of cobordisms $W_-=\ora{V^-V}$ and
4772:     $W_+=\ora{VV^+}$. From the point of view of an almost
4773:      complex structure        the process of splitting consists of deforming
4774:      an original almost complex structure $J=J^0$ to an
4775:      almost complex structure $J^\infty$, such that the restrictions
4776:      $J_\pm=J^\infty|_{W_\pm}$   are compatible  with the structure
4777:      of (completed) directed symplectic cobordisms $W_\pm$.
4778: 
4779: Conversely,  directed symplectic cobordisms
4780:  $W_-=\ora{V^-V}$ and
4781:     $W_+=\ora{VV^+}$ with matching data
4782:     on the common boundary    can be glued into
4783:     a cobordism         $W=\ora{V^-V^+}$ in the following sense:
4784:     there exists a family  $J^\tau$ of almost complex structures
4785:     on $W$ which in the limit  splits $W$ into the
4786:        composition of cobordisms $W_-=\ora{V^-V}$ and
4787:     $W_+=\ora{VV^+}$.
4788: 
4789: 
4790: 
4791:     \bigskip
4792: %%%%CONT
4793: In order to write the formula relating   the potentials
4794: of $W$ and $W_\pm$ we  first need to make more explicit the
4795: relation  between $2$-dimensional homology classes realized  by
4796: holomorphic curves in $W_\pm$ and $W$. We will keep
4797:  assuming that there are no torsion
4798: elements in $H_1$.
4799: Let us recall (see Sections \ref{sec:dynamics} and
4800: \ref{sec:holomorphic} above) that  we have chosen curves ${C}^i_-\subset
4801: W_-,i=1,\dots, m_-,\;\;{C}^j_+\subset W_+,j=1,\dots,m_+,$
4802: and ${C}^k
4803: \subset W,k=1,\dots m,$ which represent bases of first homology of
4804: the respective cobordisms. We also have chosen
4805: for every periodic orbit $\g\in
4806: \Pc_\a$ of the Reeb field $R_\a$ on $V$ a surface
4807: $F^\g_\pm$ realizing homology in $W_\pm$ between $\g$ and a linear
4808: combination of basic curves ${C}^i_\pm$.
4809: For our current purposes we have to make one extra choice: for
4810: each curve $ {C}^i_\pm$ we choose a surface $S^i_\pm$ which
4811: realizes homology in $W$ between ${C}^i_\pm$ and the
4812: corresponding linear combination of the curves ${C}^1,\dots,
4813: {C}^m$.
4814: All the choices enable us to associate with every orbit
4815: $\g\in\Pc_\a$ a homology class $C^\g$ which is realized by the chain
4816:  $$F^\g_+-F^\g_-+\sum\limits_1^{m_+} n^+_jS^j_+-\sum\limits_1^{m_-}
4817: n^-_iS^i_-\,,$$
4818: where $\partial F^\g_\pm=[\gamma]-\sum\limits_1^{m_\pm}
4819: n^{\pm}_j{C}^j_\pm$. We will denote by $d^\g$ the degree of
4820: $C^\g$, i.e. the string of its coordinates in the
4821: chosen basis
4822: $A_1,\dots,A_N\in H_2(W)$.
4823: 
4824: 
4825: 
4826:     Let us define an operation
4827:  \begin{equation}
4828:  \star:\fD\fD_-\otimes \fD\fD_+ \to \fD\fD,
4829:  \end{equation}
4830:   where   $\fD\fD_\pm=\fD\fD_{W_\pm}$ and
4831: $\fD\fD= \fD\fD_{W} $.
4832:   For
4833:    $F=\sum\limits_{\Gamma} f_{\Gamma}(t^-,q^-,\hbar,z^-)
4834:    p^{\Gamma}\in\fD\fD_-$ and
4835:     $
4836:   G=\sum\limits_{\Gamma^+} g_{\Gamma^+}
4837:  (t^+,q ,\hbar,z^+)(p^+)^{\Gamma^+} \in\fD\fD_+$ we set
4838:   \begin{equation}
4839:  \begin{split}
4840:   &F  {\star} G(t,q^-,p^+,\hbar,z)=\\
4841:   &\Big(
4842:  \sum\limits_{\Gamma }\wt f_\Gamma(t,q^-,\hbar,z)
4843:  \hbar^s  \prod\limits_{i=1}^s
4844:  \kappa_{\g_{i }}z^{d^{\g_i}}
4845:  \overrightarrow{\frac{\partial}{\partial
4846:  q_{\g_i}}}
4847: \sum\limits_{\Gamma^+} \wt g_{\Gamma^+}
4848: (t,q ,\hbar,z)(p^+)^{\Gamma^+}\Big)\Big|_{q=0}\,. \\
4849: \end{split}
4850: \end{equation}
4851: Here we denote by $\wt f_\Gamma$ and $\wt
4852: g_{\Gamma^+}$ the images
4853: of $ f_\Gamma$ and $ g_{\Gamma^+}$ under the coefficient
4854: homomorphisms $H_2(W_\pm)\to H_2(W)$.
4855: Let us explain what happens with the
4856:  variables $t$ and $z$ in more details.  Let
4857:   $A_1^\pm,\dots, A^\pm_{N^\pm},$
4858:   and
4859:    $A_1,\dots, A_{N},$
4860:     be   the chosen bases in     $H_2(W_\pm)$ and $ H_2(W)$.
4861:                    Then we have $$i^\pm_*(A^\pm_k)=\sum\limits_{j=1}^{N^\pm}n^\pm_{kj} A_j,$$
4862:    where $k=1,\dots, N$,  $(n^\pm_{kj})$ are integer matrices,
4863:    and $i^\pm:W_\pm\to W$ the inclusion maps.
4864: 
4865: We have
4866:  \begin{equation*}\begin{split}
4867:   & f_{\Gamma}(t^-,q^-,\hbar,z^-)
4868:    = \sum\limits_{d=(d_1,\dots, d_{N^-})}f_{\Gamma,d}(t^-,q^-,\hbar)(z_1^-)^{d_1}\dots
4869:    (z_{N^-}^-)^{d_{N^-}},\\
4870:   &g_{\Gamma^+}
4871:  (t^+,q ,\hbar,z^+) = \sum\limits_{d=(d_1,\dots, d_{N^+})}g_{\Gamma^+,d}(t^+,q,\hbar)(z_1^+)^{d_1}\dots
4872:    (z_{N^+}^+)^{d_{N^+}}  ,\\
4873:  \end{split}
4874:  \end{equation*} where
4875:     we denote by $z_\pm$ the ``$z$-variables" in $W_\pm$.
4876:     Then
4877:     \begin{equation*}
4878:     \begin{split}
4879:    & \wt f_\Gamma(t,q^-,\hbar,z)=
4880:           \sum\limits_{d=(d_1,\dots,d_{N_-})} \wt f_{\Gamma,d}(t|_{W_-},q^-,\hbar)z_1^{M_1^-}
4881:          \dots z_N^{M^-_{N}}, \\
4882:      & \wt g_{\Gamma^+}(t,q,\hbar,z)=
4883:           \sum\limits_{d=(d_1,\dots,d_{N_+})} \wt g_{\Gamma^+,d}(t|_{W_+},q^-,\hbar)z_1^{M_1^+}
4884:          \dots z_N^{M^+_{N}}, \\
4885:           \end{split}
4886:           \end{equation*}
4887:         where $M^\pm_j=   \sum\limits_{k=1}^{N^\pm}  n^\pm_{kj}d_k$, $j=1,\dots,N$.
4888: 
4889:    \bigskip
4890:   Let us  observe
4891:   \begin{lemma}\label{lm:star}
4892:   \begin{description}
4893:   \item{1.} The operation $\star$ is associative.
4894:   \item{2.}   For $F\in\hbar^{-1 }\fD_-,G\in\hbar^{-1 }\fD_+$ there exists a unique function $H\in \hbar^{-1 }\fD$,
4895:   such that
4896:  $$
4897:   e^H=e^F\star e^G.
4898:  $$
4899:  \end{description}
4900:   \end{lemma}
4901:   We will denote this $H$ by $F\lozenge G$, so that we have
4902:   $e^{F\lozenge G}=e^F\star e^G$.
4903: We will also consider the maps
4904: $$\lozenge G:\hbar^{-1}\fD_-\to \hbar^{-1}\fD,\;\;\lozenge
4905: G(F)=F\lozenge G,\,F\in\fD_-,$$
4906: and
4907: $$F\lozenge:\hbar^{-1}\fD_+\to \hbar^{-1}\fD,\;\;F\lozenge
4908:  (G)=F\lozenge G,\,G\in\fD_+,$$
4909:  and for even $F,G$ their linearizations:
4910:               \begin{equation*}
4911:  \begin{split}
4912: & T_F(\lozenge G):\hbar^{-1}\fD_-\to \hbar^{-1}\fD,\\
4913:  &T_F(\lozenge G)(f)=\frac{d\big(F+\varepsilon
4914:  f)\lozenge G\big)}{d\varepsilon}\big|_{\varepsilon=0}=e^{-F\lozenge
4915:  G}\big((fe^F)\star e^G\big),\;\; f\in\hbar^{-1} \fD_-,\\
4916:  \end{split}
4917:  \end{equation*}
4918: and
4919: \begin{equation*}
4920:  \begin{split}
4921: & T_G(F\lozenge  ):\hbar^{-1}\fD_+\to \hbar^{-1}\fD,\\
4922:  &T_G(F\lozenge  )(g)=\frac{d\big(F \lozenge (G+\varepsilon
4923:  g)\big)}{d\varepsilon}\big|_{\varepsilon=0}=e^{-F\lozenge
4924:  G}\big( e^F \star (ge^G)\big),\;\; g\in\hbar^{-1} \fD_+,\\
4925:  \end{split}
4926:  \end{equation*}
4927: 
4928: Let us first formulate an algebraic
4929: \begin{proposition}\label{prop:SFT-composition}
4930: Suppose that $F\in \hbar^{-1}\fD_-$ and $G\in \hbar^{-1}\fD_+$ are
4931: even elements, which
4932: satisfy the equations
4933: $$D_{W_-}(e^F)=\overrightarrow{\bH^-}e^F-e^F\overleftarrow{\bH}=0$$  and $$D_{W_+}(e^G)=
4934: \overrightarrow{\bH}e^F-e^F\overleftarrow{\bH^+}=0,$$
4935: where $\bH^\pm=\bH_{V^\pm},\,\bH=\bH_V$.
4936: Then we have
4937: \begin{description}
4938: \item{1.} $D_W(e^{F\lozenge G})=\ora{\bH^-}\, e^{F\lozenge G}-e^{F\lozenge
4939: G}\,\ola{\bH^+}=0$.
4940: \item{2.} The homomorphisms
4941: $ T_G(F\lozenge  ):\hbar^{-1}\fD_+\to \hbar^{-1}\fD$ and
4942: $T_F(\lozenge G):\hbar^{-1}\fD_-\to \hbar^{-1}\fD$ satisfy
4943: the equations $$T_G(F\lozenge  )\circ
4944: D^{G}=D^{F\lozenge G}\circ T_G(F\lozenge )$$ and $$T_F(
4945: \lozenge G )\circ D^{F}=D^{F\lozenge G}\circ T_F(
4946: \lozenge G),$$ and in particular they define homomorphisms
4947: of the corresponding homology algebras:
4948: $$ \left( T_G(F\lozenge  )\right)_*:H_*( \fD_+,D^G)\to H_*(  \fD,D^
4949:  {F\lozenge G})$$ and
4950: $$ \left( T_F( \lozenge G  )\right)_*:H_*( \fD_-,D^G)\to H_*(  \fD,D^
4951:  {F\lozenge G}).$$
4952: \item{3.} $$ T_F( \lozenge G)\circ F^-   =
4953: (F \lozenge G)^-$$ and $$  T_G( F\lozenge)
4954: \circ G^+= (F \lozenge G)^+.$$
4955: \item{4.}
4956: Suppose we are given three cobordisms $W_1,W_2,W_3$ with matching ends
4957:  so that we can  form  the  composition
4958: $W_{123}=W_1\circledcirc W_2 \circledcirc W_3$, and series
4959: $F_i\in\hbar^{-1}\fD_i=\hbar^{-1}\fD_{W_i},\,i=1,2,3$, such that
4960: $D_{W_i}e^{F_i}=0$, $i=1,2,3$.
4961: Then
4962: \begin{equation}
4963: T_{F_1\lozenge F_2}(\lozenge F_3)\circ T_{F_1}(\lozenge
4964: F_2)=T_{F_1}(\lozenge (F_2\lozenge F_3)).
4965: \end{equation}
4966: \end{description}
4967: \end{proposition}
4968: The proof of this proposition is immediate from the definition of
4969: the corresponding operations. For instance, to prove
4970: \ref{prop:SFT-composition}.1 we write
4971: \begin{equation*}
4972: \begin{split}
4973:  \ora{\bH^-}\,\big(e^{F\lozenge G}\big)-\big(e^{F\lozenge
4974: G}\big)\,\ola{\bH^+}&=\ora{\bH^-}\,e^F\star e^G -e^F\star e^G\ola
4975: {\bH^+}\\
4976: &=\left(\ora{\bH^-}\,e^F\right)\star e^G -e^F\star\left( e^G\ola
4977: {\bH^+}\right)\\
4978: &=\left(e^F\,\ola{\bH}\right)\star e^G
4979: -e^F\star\left( \ora{\bH}e^G\right).\\
4980: \end{split}
4981: \end{equation*}
4982: To finish the argument
4983: let us consider a cylindrical cobordism $W_0=V\times\R,$
4984: take the function
4985: $I=\sum \kappa_\g^{-1}p_\g q_\g$. Taking into account associativity
4986: of $\star$ (see \ref{lm:star}) we have
4987: \begin{equation}
4988: \ora {f}e^G=(fe^I)\star e^G,\quad\hbox{and}\quad
4989:  e^F\ola{f}=e^F\star (fe^I)\,.
4990:  \end{equation}
4991:  Hence, we have
4992:  \begin{equation*}
4993:  \begin{split}
4994:  & \left(e^F\,\ola{\bH}\right)\star e^G
4995: -e^F\star\left( \ora{\bH}e^G\right)\\
4996: &=e^F\star(\bH e^I)\star e^G-  e^F\star(\bH e^I)\star e^G=0.\\
4997:  \end{split}
4998: \end{equation*}
4999: \bigskip
5000: 
5001: Any cohomology class from $H^*(W)$
5002: can be represented by a form $t$ which splits into the sum of forms
5003: $t_\pm$  with cylindrical ends on $W_\pm$,
5004:     so that we have $t_\pm|_{V}=t_V$.
5005:      Let us define now a series
5006:       $\bF^\infty\in\hbar^{-1}\fD$
5007:       by the formula
5008:     \begin{equation}\label{eq:SFT-gluing}
5009:  {\bF^\infty(q^-,p^+,t)}=  {\bF_-(q^-,p,t_-)}
5010:  \lozenge
5011:    {\bF_+(q,p^+,t_+)} ,
5012:    \end{equation}
5013:     where $p,q$      are variable associated to the space $H^*(\Pc)$
5014:     of   periodic orbits of the Reeb vector field
5015:     $R_\a$ of the contact form $\alpha$ on $V$.
5016: 
5017:      The following theorem is the main claim of this section, and
5018:      similar to Theorems \ref{thm:HH} and \ref{thm:SFT-cobordism}
5019:      and \ref{thm:SFT-chain} it  is a statement about the
5020:      boundary of an appropriate moduli space of holomorphic curves.
5021:      This time we deal with limits of $J^s$-holomorphic curves
5022:      in $W$ when $s\to\infty$,
5023:      i.e. when the family $J^s$ realizes the splitting
5024:      of $W$ into the composition $W_-\circledcirc W_+$,
5025:       see Theorem \ref{thm:comp2} above.
5026: 
5027: 
5028:     \begin{theorem}\label{thm:SFT-composition}
5029: The element $\bF^\infty$ is homotopic to the
5030:    potential
5031:    $\bF=\bF_{W,J ,\a^\pm} $
5032:    for any generic  compatible almost complex structure $J$ on $W$.
5033:    \end{theorem}
5034:   \bigskip
5035:   Let us now describe the above results on the level of rational potentials.
5036: Let  $W_-=\ora{V^-V }$, $W_+=\ora{VV^+}$ and
5037:  $W=W_-\circledcirc W_+=\ora{V^-V^+}$ be as above. Set
5038: $\bh^\pm=\bh_{V^\pm},\bh=\bh_V$,
5039: $\widehat\bh_-=\bh^- -\bh$, $\widehat\bh_+=\bh  -\bh^+$,
5040: $\widehat\bh=\bh^+-\bh^-$,
5041: $\fL_\pm=\fL_{W_{\pm}}$ and $ \fL=\fL_W$.
5042: 
5043:  The operation
5044: $\lozenge:\hbar^{-1}\fD_-\times\hbar^{-1}\fD_+\to\hbar^{-1}\fD$ defined above reduces on the
5045: rational level to the
5046:    operation $$\sharp:\fL_-\times\fL_+\to\fL,$$ defined as follows.
5047:    For $ f_\pm\in\fL_\pm$ we set
5048: \begin{equation}
5049: f_-\sharp f_+(q^-,p^+)=\big(f_-(q^-,p)+f_+(q,p^+)
5050: -\sum_{\g\in\Pc}\kappa_\g^{-1}z^{-d_{\g}}q_\g p_\g\big)\big|_{L},
5051: \end{equation}
5052: where
5053: \begin{equation*}
5054: L=
5055: \begin{cases}
5056: q_\g=&\kappa_\g z^{d_\g}\frac{\partial f_-}{\partial p_\g};\\
5057: p_\g=&\kappa_\g z^{d_\g}\frac{\partial f_+}{\partial q_\g}.\\
5058: \end{cases}
5059: \end{equation*}
5060: Notice that given series $$F_-=\hbar^{-1}\sum\limits_0^\infty
5061: (F_-)_g\hbar^g\in\hbar^{-1}\fD^-\quad\hbox{ and}\quad
5062:  F_+=\hbar^{-1}\sum\limits_0^\infty (F_+)_g\hbar^g\in\hbar^{-1}\fD^+$$  with
5063: $F_-\lozenge F_+= \hbar^{-1}\sum\limits_0^\infty (F_-\lozenge
5064: F_+)_g\hbar^g\in\hbar^{-1}\fD $ then  $$(F_-\lozenge F_+)_0=(F_-)_0\sharp
5065: (F_+)_0.$$
5066:  We will also consider the operations
5067: \begin{equation*}
5068: \begin{split}
5069: &\sharp f_+:\fL_-\to\fL,\;\; \sharp f_+(f_-)=f_-\sharp
5070: f_+\;\;\hbox{and}\;\;
5071: \sharp f_-:\fL_+\to\fL,\;\; \sharp f_-(f_+)=f_-\sharp
5072: f_+,\\
5073: \end{split}
5074: \end{equation*}
5075: and their linearizations
5076: \begin{equation*}
5077: \begin{split}
5078: &T_{f_-}(\sharp f_+):\fL_-\to\fL,\;\; T_{f_-}(\sharp f_+)(g)=
5079: (g|_{L_+})\sharp f_+
5080: \quad\hbox{and}\\
5081: &T_{f_+}(\sharp f_-):\fL_+\to\fL,\;\; T_{f_+}(\sharp f_-)(g)=f_-\sharp
5082: (g|_{L_-})\,.\\
5083: \end{split}
5084: \end{equation*}
5085: Here we view $g|_{L_+}$ (resp. $g|_{L_-}$) as an element of
5086: $\fL_+$  which depends on variables $q^-$ as parameters (resp. an element of
5087: $\fL_-$, which depends on $p^+$ as parameters).
5088: %\end{proposition}
5089:   We have the following rational version
5090: of Theorem \ref{prop:SFT-composition}.
5091: \begin{proposition}\label{prop:RSFT-composition}
5092: Suppose that even elements $f_\pm\in\fL_\pm=\fL_{W_{\pm}}$ satisfy
5093: equation (\ref{eq:HJ1}), i.e.
5094: \begin{equation*}
5095: \widehat\bh_\pm|_{L_{f_\pm}}= 0 ,
5096: \end{equation*}
5097: where
5098: \begin{equation*}
5099: L_{f_-}=
5100: \begin{cases}
5101: q_\g&=k_\g\frac{\partial f_-}{\partial p_\g};\;\;\g\in\Pc\\
5102: p^-_{\g^-}& =\kappa_{\g^-}\frac{\partial f_-}{\partial q^-_{\g^-}},\;\;\g^-\in\Pc^-,\\
5103: \end{cases}
5104: \end{equation*}
5105: and
5106: \begin{equation*}
5107: L_{f_+}=
5108: \begin{cases}
5109: p_\g&=k_\g\frac{\partial f_+}{\partial q_\g};\;\;\g\in\Pc\\
5110: q^+_{\g^+}&=\kappa_{\g^+}\frac{\partial f_+}{\partial p^+_{\g^+}},\;\;\g^+\in\Pc^+.\\
5111: \end{cases}
5112: \end{equation*}
5113: Then
5114: \begin{description}
5115: \item{1.} The function $f_-\sharp f_+$ satisfies the Hamilton-Jacobi
5116: equation
5117: $$\widehat\bh_{L|_{f_-\sharp f_+}}=0;$$
5118: \item{2.} The homomorphisms $T_{f_-}(\sharp f_+):\fL_-\to\fL$ and
5119: $T_{f_+}( f_-\sharp):\fL_+\to\fL$ satisfy the equations
5120: \begin{equation*}
5121: \begin{split}
5122: T_{f_-}(\sharp f_+)\circ d^{f_-}=&d^{f_-\sharp f_+}\circ T_{f_-}(\sharp
5123: f_+),\\
5124: T_{f_+}( f_-\sharp)\circ d^{f_+}=&d^{f_-\sharp f_+}\circ T_{f_+}(\sharp
5125: f_-)\,, \\
5126: \end{split}
5127: \end{equation*}
5128: and hence define homomorphisms of the corresponding
5129: homology algebras:
5130: \begin{equation*}
5131: \begin{split}
5132: \big(T_{f_-}(\sharp f_+)\big)_*:&H_*(\fL_-,d^{f_-})\to H_*(\fL,d^{f_-\sharp f_+}),\\
5133: \big(T_{f_+}(f_-\sharp )\big)_*:&H_*(\fL_-,d^{f_-})\to H_*(\fL,d^{f_-\sharp f_+});\\
5134: \end{split}
5135: \end{equation*}
5136: \item{3.}
5137: \begin{equation*}
5138: \begin{split}
5139: T_{f_-}(\sharp f_+)\circ (f_-)^-=& (
5140: f_-\sharp f_+)^-\,, \\
5141: T_{f_+}(\sharp f_-)\circ (f_+)^-=& (
5142: f_-\sharp f_+)^+
5143: \,; \\
5144: \end{split}
5145: \end{equation*}
5146: \item{4.}
5147: Suppose we are given three cobordisms $W_1,W_2,W_3$ with matching ends
5148:  so that we can  form  the  composition
5149: $W_{123}=W_1\circledcirc W_2 \circledcirc W_3$, and series
5150: $f_i\in\hbar^{-1}\fL_i= \fL_{W_i}$  which satisfy
5151: Hamilton-Jacobi equations
5152: $\widehat{\bh}_{W_i}|_{L_{f_i}}=0,\,i=1,2,3$.
5153: Then
5154: \begin{equation*}
5155: T_{f_2}(\sharp f_3)\circ T_{f_1}(\sharp
5156: f_2)=T_{f_1}(\sharp (f_2\sharp f_3)).
5157: \end{equation*}
5158: \end{description}
5159: \end{proposition}
5160: 
5161: Let $t$ be a closed form on $W$ which is split into two forms $t_\pm$ on $W_\pm$ with
5162: cylindrical ends. Set $\bff_\pm=\bff_{W_\pm}$ and
5163: \begin{equation}
5164: \bff^\infty(q^-,p^+,t)=\bff_-(q^-,p,t_-)\sharp \bff_+(q,p^+,t_+).
5165: \end{equation}
5166: Alternatively $\bff^\infty$ can be defined as
5167: the first term  in the expansion
5168: $\bF^\infty=\hbar^{-1}\sum\limits_0^\infty\bF^\infty_g\hbar^g$.
5169:  The following theorem is a rational analog, and a direct
5170:  corollary of Theorem \ref{thm:SFT-composition}.
5171: \begin{theorem}
5172: \label{thm:RSFT-composition}
5173:   The series $\bff^\infty(q^-,p^+,t)$  and  $\bff_{W,J}(q,p,t)$
5174:   are  homotopic  for any generic  compatible almost complex structure
5175:   $J$ on $W$.
5176: \end{theorem}
5177: 
5178: Coming down to the   ``classical" level, let us assume that
5179:  $W,W_-$ and $W_+$ are homology cobordisms (see \ref{sec:3algebras}
5180:  above). Thus there are   defined  the homomorphisms  $\bPsi  :\fA^+\to\fA^-$,
5181:  $\bPsi_+ :\fA^+\to\fA $ and $\bPsi_- :\fA \to\fA^-$,
5182:     see Section \ref{sec:relative} above. Set
5183:  $\bPsi_\infty=\bPsi_+\circ\bPsi_-$.
5184:  Then we have
5185:  \begin{theorem}\label{thm:classical-composition}
5186:  For any generic compatible almost complex structure $J$ on $W$
5187:  homomorphisms
5188:  $\bPsi_1=\bPsi_J,\bPsi_\infty:\fA^+\to\fA^-$ are chain
5189:  homotopic.
5190: \end{theorem}
5191: 
5192: \subsection{Invariants of contact
5193: manifolds}\label{sec:contact-invariants}
5194: 
5195: 
5196: 
5197:  Theorem \ref{thm:SFT-composition}  allows us to define
5198:  SFT-invariants of contact manifolds.
5199:  Let $(V,\xi)$ be a contact manifold,  and
5200:  $\alpha^+$ and $\alpha^-$
5201:  two contact forms for $\xi$,
5202:  such that $\alpha^+>\alpha^-$, i.e.
5203:  $\alpha^+=f\alpha^-$, for a  function $f>1$.
5204:    Then for an appropriately chosen function $\zeta:V\times\R\to(0,\infty)$
5205:    the form
5206:  $\omega=d(\zeta\alpha^-)$ on $W=V\times\R$ is symplectic, and
5207:  $(W,\omega)$ is a directed symplectic cobordism between $(V,\alpha^-)$
5208:  and $(V,\alpha^+)$.  Let  $t^\pm$ be two cohomologous forms on $V$, and $t$ be a  closed
5209:  form  on $W$ with cylindrical ends which restricts to $t^\pm$ on $V^\pm$.
5210:  Suppose we are also given almost complex structures $J^\pm$ on $V$,
5211:   compatible with $\alpha^\pm$, which
5212:  are extended to  a compatible almost complex structure $J$ on $W$.
5213:  We will call a  directed symplectic cobordism  $(W,J,t)$,
5214:   chosen in the  above way, a {\it concordance} between
5215:   the data on its boundary. Notice, that concordance becomes   an
5216:   equivalence relation if we identify contact forms proportional
5217:   with a constant factor. A concordance $(W,J,t)$ is called {\it
5218:   trivial} if  $W=V\times\R$, the almost complex structure
5219:    $J$ is translationally  invariant, and
5220:   $t$ is the pull-back of a form $t_+$ under the projection $W\to
5221:   V$.
5222: 
5223: 
5224:  Let us denote, as usual,   by $(\fW^\pm,D^\pm)$  the differential Weyl algebras
5225:   associated to the data at the
5226:  ends of the cobordism $W$, by $(\fD,D_W)$ the $D$-module
5227:  $\fD(W,J,\a^\pm)$,  by $\bF\in\fW$
5228:  the    potential of the cobordism $W$, and by
5229:  $\bF^\pm:(\fW^\pm,D^\pm)\to(\fD,D_W)$ the
5230:  corresponding homomorphisms
5231:  of differential algebras defined in (\ref{eq:end-homomorphisms}).
5232: 
5233:  \begin{theorem}\label{thm:SFT-invariance}
5234:   For any concordance $(W,J,t)$ the  homomorphisms $$\bF^\pm:(\fW^\pm,d^\pm)\to(\fD,D_W)$$
5235:     are quasi-isomorphisms
5236:   of differential algebras. In particular, the homology algebras
5237:   $H_*(\fW^-,D^-)$ and
5238:   $H_*(\fW^+,D^+)$ are isomorphic.
5239:   \end{theorem}
5240: 
5241:  \noindent\proof We will prove   \ref{thm:SFT-invariance}  in three
5242:   steps.
5243: 
5244: \noindent Step 1. Let us begin with the trivial concordance $(W ,J
5245: ,t  )$. In this case $\fD$ can be identified with $\fW^\pm$ and we
5246: have $\bF(q,p,t)=\hbar^{-1}\sum\kappa_\g^{-1}q_\g p_\g$. Hence
5247: \begin{equation*}
5248: \bF^-(f)=e^{-\bF}\ora{f}e^{\bF}=
5249: f=e^{-\bF}\left(
5250: e^{\bF}\ola{f}\right)=\bF^+(f).
5251: \end{equation*}
5252: 
5253:          \noindent Step 2. If we add now to $t$ a form $d\theta$, where
5254: $\theta$ has a compact support, and change $J$ in a compact part
5255: of $W$ then according to Theorem \ref{thm:SFT-chain}
5256:  the potential $\bF_{W,J}(t+d\theta)$ remains the same
5257:  up to homotopy, and hence Corollary \ref{cor:SFT-chain} implies
5258: that the homomorphisms  $\bF^\pm_*$   induced on homology remain
5259:  unchanged.
5260: 
5261: 
5262: 
5263: \noindent Step 3. Now assume that $(W,J,t)=(W^1,J^1,t^1)$ is a general
5264: concordance. Consider the reversed concordance $(W^2,J^2,t^2)$, so
5265: that the compositions $$(W^{12}=W^1\circledcirc W^2,J^{12}=J^1\circledcirc
5266: J^2,t^{12}=
5267: t_1\circledcirc t_2)$$ and
5268:   $$(W^{21}=W^2\circledcirc W^1,J^{21}=J^2\circledcirc J^1,t^{21}=
5269: t_2\circledcirc t_1)$$ of concordances $(W^1 ,J^1 ,t^1 )$
5270: and $(W^2 ,J^2 ,t^2 )$   are   as in Step 2.
5271:  Then according to Theorem
5272: \ref{thm:SFT-composition}
5273: $\bF_{W^{12},J^{12}}( t^{12})$ is homotopic to
5274: $$\bF_{W^1,J^1}(t^1)\lozenge\bF_{W^2,J^2}(t^2)=
5275: \bF^1\lozenge\bF^2$$ and
5276:  $\bF^{21}=\bF_{W^{21},J^{21}}( t^{21})$ is homotopic to
5277: $$\bF_{W^2,J^2}(t^2)\lozenge\bF_{W^1,J^1}(t^1)=\bF^2\lozenge\bF^1.$$ Hence
5278: Proposition \ref{prop:SFT-composition}
5279: implies
5280: \begin{equation*}
5281: \Id=\left(\bF_{W^1\circledcirc W^2 }\right)^-_*
5282: =\left(T_{\bF^1}(\lozenge\bF^2) \right)_*\circ (\bF^1)^-_*
5283:   \end{equation*}
5284:   and
5285:   \begin{equation*}
5286:   \begin{split}
5287: & \Id= \left(T_{\bF^1}\left(\lozenge(\bF^2\lozenge\bF^1) \right)
5288: \right)_* \\
5289: &= \left(T_{\bF^1\lozenge\bF^2}(\lozenge \bF^1)\right)_*
5290: \circ \left(T_{\bF^1} (\lozenge  \bF^2)\right)_* .\\
5291: \end{split}
5292: \end{equation*}
5293: Hence
5294: $\left(T_{\bF^1}(\lozenge\bF^2) \right)_*$,  $\;(\bF^1)^-_*$,
5295: and similarly    $(\bF_1)^+_* $
5296: are isomorphisms. \qed
5297: 
5298: \bigskip
5299: 
5300: 
5301: The following rational and classical analogs of Theorem
5302: \ref{thm:SFT-invariance}
5303: can be either deduced directly from Theorem
5304: \ref{thm:SFT-invariance}, or alternatively can be proven similarly
5305: to \ref{thm:SFT-invariance}  using \ref{thm:RSFT-chain} (resp. \ref{thm:classical-chain}) and
5306: \ref{thm:RSFT-composition} (resp. \ref{thm:classical-composition}).
5307: 
5308: \begin{theorem}\label{thm:RSFT-invariance}
5309:     For any concordance $(W,J,t)$ the  homomorphisms $$\bff^\pm:(\fP^\pm,D^\pm)\to(\fL,D_W)$$
5310:      are quasi-isomorphisms
5311:   of differential algebras. In particular,  Poisson homology
5312:    algebras $H_*(\fP^\pm,d^\pm)$  are isomorphic.
5313:   \end{theorem}
5314: 
5315: 
5316: 
5317: \begin{theorem}\label{cor:classical-invariance}
5318:  For any concordance $(W,J,t)$ the  homomorphism  $$\bPsi:(\fA^+,\partial^+)\to
5319:  (\fA^-,\partial^-)$$
5320:   is a quasi-isomorphism
5321:   of differential algebras.
5322: \end{theorem}
5323: \bigskip
5324: 
5325: The definition of the differential algebras $(\fW,D),\;(\fP,d)$ and $(\fA,\partial)$
5326:  depends on the choice
5327: of a coherent orientation (see Section \ref{sec:orientation}), and
5328:   of spanning surfaces and framings
5329: of periodic orbits
5330: (see Section \ref{sec:dynamics}).
5331: As it is stated in Theorem \ref{thm:coherent}  a coherent orientation
5332: is determined by a choice of asymptotic operators associated
5333: with each periodic orbit $\g$. Let $\bH'$ be the new  Hamiltonian which
5334: one gets by changing the orientation of the asymptotic operator
5335: associated with a fixed  periodic orbit $\g $. One can then check that the
5336: change of variables  $(p_\g, q_\g)\mapsto(-p_\g,-q_\g)$ is an
5337: isomorphism  between the differential algebras $(\fW,D^{\bH})$ and
5338: $(\fW,D^{\bH'})$.
5339:   Different choices for spanning   surfaces and framings of periodic orbits do not affect
5340: $\mod\;2$ grading but change the integer grading of the
5341: differential algebras.
5342: 
5343: \begin{remark}\label{rem:other-choices}
5344: {\rm
5345:   An accurate introduction of  virtual cycle techniques
5346: would reveal that even more choices should be made.   However,  an independence
5347: of  all these extra choices can be also established following the scheme of this section.
5348: }
5349: \end{remark}
5350: 
5351: \subsection{A differential equation for potentials of symplectic cobordisms} \label{sec:master}
5352: 
5353: In this section we describe  differential equations for
5354: the  potentials $\bF_W$ and $\bff_W$ of a  directed symplectic
5355: cobordism with a non-empty boundary. These equations completely
5356: determine  the potentials, and
5357:  in
5358: combination with gluing Theorems  \ref{thm:SFT-composition}  and
5359: \ref{thm:RSFT-composition} they provide in many cases   an effective
5360:  recursive
5361: procedure for computing     potentials $\bF_W$ and $\bff_W$,
5362:  and even  Gromov-Witten invariants of closed symplectic manifolds  $W$
5363: (see some examples in Section \ref{sec:computing} below).
5364: 
5365: \bigskip
5366: Let us assume for simplicity that $W$ has only a positive end
5367: $E^+=V\times(0,\infty)$, and choose a basic system
5368: $\Delta_1,\dots,\Delta_k$, $\Theta_1,\dots,\Theta_m$ of closed
5369: forms so that the   following conditions are satisfied:
5370: \begin{description}
5371: \item{a)} $\Delta_1,\dots,\Delta_k$ form a basis of $H^*(W)$, and the restrictions
5372: $\delta_i=\Delta_i|_V$, $i=1,\dots,l$ for $l\leq k$ form a basis
5373: of    $\im\big(H^*(W)\to H^*(V)\big)$;
5374: \item{b)}  $\Theta_1,\dots,\Theta_m$ are compactly
5375: supported and    represent a basis of
5376: $\Ker\big(H^*_{\comp}(W)\to H^*(W)\big)$,
5377: \item{c)} there exist  forms $\theta_1,\dots,\theta_m$ on $V$ and
5378: a compactly supported $1$-form $\rho$ on $(0,+\infty)$, such that
5379: $\Theta_j=e_*\big(\rho\wedge \pi^*(\theta_j)\big),\;j=1,\dots, m$, where
5380: 
5381: $\pi$ is the projection $E=V\times(0,\infty)\to V$ and $e:E\hookrightarrow W$ is the inclusion.
5382: In  other words, $\Theta_j$ equals $\rho\wedge \pi^*(\theta_j)$
5383: viewed as a form on $W$.
5384: \end{description}
5385: \medskip
5386: 
5387: \begin{theorem}\label{thm:SFT-master}
5388: Let $\bH=\bH_{V,\a,J}$ be the Hamiltonian associated with the
5389: contcat manifold $V$.
5390: Set $$\bH^j(t_1,\dots,t_l,q,p)=\Big(
5391: \frac{\partial \bH}{\partial s_j}(\sum\limits_{i=1}^l
5392: t_i\delta_i+s_j\theta_j,q,p)\Big)\Big|_{s_j=0}
5393: ,\;j=1,\dots m,$$
5394: $$\bF^0(t_1,\dots,t_k, p)=\bF_{W,J}(\sum
5395:  t_i\Delta_i,p),$$
5396: and define
5397:  $\bF(t_1,\dots,t_k,\tau_1,\dots,\tau_m, p)$
5398:  by the formula
5399:  \begin{equation}
5400:  e^{\bF(t_1,\dots,t_k,\tau_1,\dots,\tau_m,
5401:  p)}=e^{\bF^0(t_1,\dots,t_k,p)}
5402:  \overleftarrow\bG(t_1,\dots,t_l,\tau_1,\dots,\tau_m,p),
5403:  \end{equation}  where we denote by
5404:  $\overleftarrow\bG$ the operator obtained from
5405:  \begin{equation}
5406:  %\begin{split}
5407:   \bG(t_1,\dots,t_l,\tau_1,\dots,\tau_m,q,p)
5408:   = e^{\tau_m\bH^m(t_1,\dots,t_l,q,p)}\dots
5409:  e^{\tau_1\bH^1(t_1,\dots,t_l,q,p)}
5410:  %\end{split}
5411:  \end{equation}
5412:  by quantizing   $q_\g=\kappa_\g
5413:  \hbar\overleftarrow{\frac{\partial }{\partial
5414:  p_\g}}$.
5415:  Then  $\bF(t_1,\dots,t_k,\tau_1,\dots,\tau_m, p)$  is
5416:   homotopic to  $$\bF_{W,J}(\sum\limits_{i=1}^k
5417:  t_i\Delta_i+\sum\limits_{r=1}^m\tau_r\Theta_r,p).$$
5418:   \end{theorem}
5419:   \proof
5420:  Set  $$T^j(s)={\bF_{W,J}(\sum
5421:  t_i\Delta_i+\sum\limits_{r=1}^{j-1}\tau_r\Theta_r+s\tau_j\Theta_j,p) }.$$
5422:   We have $$T^j(1)=T^{j+1}(0)\quad\hbox{
5423:  for}\;\;j=1,\dots,m-1 ,     $$
5424:   $$
5425:   T^m(1)= {\bF_{W,J}(\sum\limits_{i=1}^k
5426:  t_i\Delta_i+\sum\limits_{r=1}^m\tau_r\Theta_r,p)},$$
5427:  and
5428:  $$T^1(0)={\bF_{W,J}(\sum
5429:  t_i\Delta_i,p) }={\bF^0(t_1,\dots,t_k, p)}.$$
5430:  Let $S^j\in\hbar^{-1}\fD$ be defined from the equation
5431:  $$e^{S^j}=e^{T^j(0)}e^{\tau_j\ola{\bH^j(t_1,\dots,t_l,q,p)} }.$$
5432:  It is enough to prove that
5433:  $  T^j(1)$ is homotopic    to $S^j$ for $j=1,\dots,m$.
5434:  We have
5435:  \begin{equation*}
5436:  \begin{split}
5437:  &\frac{\partial e^{T^j(s)}}{\partial s}=
5438:   \frac{\partial T^j(s)}{\partial s} e^{T^j(s)}  \\
5439:   &=  e^{T^j(s)}\sum\limits_d\sum\limits_{g,u  ,v
5440:   =0}^\infty     \frac{1}{u!v ! }
5441:    \up0\langle \tau_j\Theta_j,
5442:     \underbrace{\sum\limits_{i=1}^k
5443:  t_i\Delta_i+\sum\limits_{r=1}^{j-1}\tau_r\Theta_r+s\tau_j\Theta_j}_u;
5444:  \underbrace{p ,\dots,p }_{v }
5445:  \rangle^d_{g}z^d \hbar^{g-1}.\\
5446:  \end{split}
5447:  \end{equation*}
5448:   The compactly supported form $\Theta_j$ is exact in $W$,
5449:   $$\Theta_j=d\wt\Theta_j,$$
5450: where $\wt\Theta_j$ is closed at infinity, has a cylindrical    end,
5451: and $\wt\Theta_j|_V=\theta_j$.
5452: Hence,
5453:  \begin{equation*}
5454:  % \begin{split}
5455:  \frac{\partial T^j(s)}{\partial s}  =
5456: \ \sum\limits_d\sum\limits_{g,u  ,v =0}^\infty
5457: \frac{1}{u!v !}
5458:   \up0\langle \tau_jd\wt \Theta_j,
5459:     \underbrace{\sum\limits_{i=1}^k
5460:  t_i\Delta_i+\sum\limits_{r=1}^{j-1}\tau_r\Theta_r+s\Theta_j}_u;
5461:   \underbrace{p ,\dots,p }_{v }
5462:  \rangle^d_{g}z^d \hbar^{g-1}
5463: % \end{split}
5464:  \end{equation*}
5465:   \begin{equation*}
5466:  \begin{split}
5467:  &= d\big( \sum\limits_d\sum\limits_{g,u  ,v=0}^\infty  \frac{1}{u!v
5468:  !}
5469:   \up0\langle \tau_j\wt \Theta_j,
5470:     \underbrace{\sum\limits_{i=1}^k
5471:  t_i\Delta_i+\sum\limits_{r=1}^{j-1}\tau_r\Theta_r+s\tau_j\Theta_j}_u;
5472:   \underbrace{p ,\dots,p }_{v }
5473:  \rangle^d_{g}z^d \hbar^{g-1}\big)\\
5474:   &= \frac{\partial}{\partial u}
5475:   \Big( d\big( \sum
5476:   \limits_d\sum\limits_{g,u  ,v=0}  \frac{1}{u!v!}\\
5477:  &\up0\langle
5478:     \underbrace{\sum\limits_{i=1}^k
5479:  t_i\Delta_i+\sum\limits_{r=1}^{j-1}\tau_r\Theta_r+s\tau_j\Theta_j
5480:  +u\tau_j\wt \Theta_j}_u;
5481:  \underbrace{p ,\dots,p }_{v}
5482:  \rangle^d_{g}z^d \hbar^{g-1}\big)\Big)\Big|_{u=0}\\
5483: & =  \frac{\partial}{\partial u}\Big( d\big(
5484: \bF_{W,J}(\sum\limits_{i=1}^k
5485:  t_i\Delta_i+\sum\limits_{r=1}^{j-1}\tau_r\Theta_r+
5486:  s\tau_j\Theta_j+u\tau_j\wt \Theta_j,p)\big)\Big)\Big|_{u=0} ,
5487:     \end{split}
5488:  \end{equation*}
5489:    where $d$ denotes the de Rham differential.
5490:    Using  equation (\ref{eq:non-closed2}) we get
5491:     \begin{equation*}
5492:  \begin{split}
5493: &  d\big(
5494: \bF_{W,J}(\sum\limits_{i=1}^k
5495:  t_i\Delta_i+\sum\limits_{r=1}^{j-1}\tau_r\Theta_r+
5496:  s\tau_j\Theta_j+u\tau_j\wt \Theta_j,p)\big) =\\
5497:  &e^{-\bF_{W,J}}\big( e^{\bF_{W,J}}\ola\bH(\sum\limits_{i=1}^l t_i\Delta_i +
5498:  u\tau_j\theta_j,p) \big),\\
5499:      \end{split}
5500:  \end{equation*}
5501:  and hence
5502:  \begin{equation*}
5503:  \begin{split}
5504:    &\frac{\partial T^j(s)}{\partial s}=
5505:    \frac{\partial}{\partial u}\Big( e^{-\bF_{W,J}}
5506:    \big( e^{T^j(s)}\ola\bH(\sum\limits_{i=1}^l t_i\Delta_i +
5507:  u\tau_j\theta_j,q,p) \big)\Big)\Big|_{u=0}\\
5508:  &= \tau_j e^{-T^j(s)}\Big(-\bF^j  (t_1,\dots,t_k,\tau_1,\dots,\tau_j,s,p)
5509:   \big(e^{T^j(s) }\ola\bH(\sum\limits_{i=1}^l t_i\Delta_i ,q,p)\big)\\
5510:  &+ \big( e^{T^j(s)} \bF^j  (t_1,\dots,t_k,\tau_1,\dots,\tau_j,s,p)\big)
5511:   \ola\bH(\sum\limits_{i=1}^l t_i\Delta_i ,q,p)\\
5512:  & +e^{T^j(s)}
5513:  \ola\bH^j(t_1,\dots,t_l,q,p)\Big) =
5514:  \tau_j e^{-T^j(s)}\Big(\big( e^{T^j(s)} [\bF^j , \ola\bH]
5515:  +e^{T^j(s)}
5516:  \ola\bH^j\Big),  \\
5517:     \end{split}
5518:  \end{equation*}
5519:     where
5520:     \begin{equation*}
5521:     \begin{split}
5522:     &\bF^j  (t_1,\dots,t_k,\tau_1,\dots,\tau_j,s,p) \\
5523:  &=\frac{ \partial   }{\partial u} \big(
5524:  \bF(\sum
5525:  t_i\Delta_i+\sum\limits_{r=1}^{j-1}\tau_r\Theta_r+
5526:  s\tau_j\Theta_j+u\wt\Theta_j,p)\big)\big|_{u=0}
5527:  \end{split}
5528:  \end{equation*}
5529:  and
5530:  $$\bH^j( t_1,\dots,t_l,q,p)=   \frac
5531:  { \partial   }{\partial u} \big(
5532:  \bH(\sum t_i\Delta_i +
5533:  u\theta_j,q,p)   \big)\big|_{u=0}\,.$$
5534:  Therefore,
5535:  \begin{equation*}
5536: \frac{\partial e^{T^j(s)}}{\partial s}=  e^{T^j(s)}
5537:   \frac{\partial T^j(s)}{\partial s}
5538:   =\tau_je^{T^j(s)}\big([\bF^j , \ola\bH]+\ola\bH^j\big)\,.
5539:  \end{equation*}
5540: 
5541: Let us define now a family
5542:  $U^j(s)\in \hbar^{-1}\fD,s\in [0,1]$
5543:  by the formula
5544:  \begin{equation}
5545:  e^{U^j(s)}=e^{T^j(s)}e^{(1-s)\tau_j\ola
5546:  {\bH^j}(t_1,\dots,t_l,q,p)}\,.
5547:      \end{equation}
5548:      Then $U^j(s)$ is a homotopy
5549:      between $S^j$ and $T^j(1)$. Indeed,   we have
5550:  $U^j(0)=S^j$ and $U^j(1)=T^j(1)$.
5551:  On the other hand we get an equation
5552:  \begin{equation*}
5553: % \begin{split}
5554: \frac {\partial e^{U^j(s)}}{\partial s} =\tau_je^{U^j(s)}
5555:  \big(-\ola\bH^j+ [\bF^j
5556: , \ola\bH]+\ola\bH^j\big) =e^{U^j(s)}[\tau_j\bF^j , \ola\bH]\;,
5557: %\end{split}
5558: \end{equation*}
5559: which is the definition of homotopy (see Section
5560: \ref{sec:chain} above).\qed
5561: 
5562: \bigskip
5563: 
5564: 
5565: 
5566: 
5567: We formulate now a version of Theorem \ref{thm:SFT-master} for
5568: rational potentials.
5569: 
5570: Set $$\bh^j(t_1,\dots,t_l,q,p)=\frac{\partial \bh}{\partial
5571: s_j}(\sum t_i\delta_i+s_j\theta_j,q,p) ,\;j=1,\dots m,$$ and for
5572: any $\bgg\in \fL$ we denote by $L_{\bgg}$ the Lagrangian variety of
5573: $\bV$, defined by the equation
5574: \begin{equation}
5575: L_\bgg=\{q_\g= \kappa_\g\frac{\partial\bgg}
5576: {\partial p_\g}\}.
5577: \end{equation}
5578: \begin{theorem}\label{thm:RSFT-master}
5579:  Let $\bff(t_1,\dots,t_k,\tau_1,\dots,\tau_m, p)$
5580: be the solution of the   Hamilton-Jacobi equation
5581: \begin{equation} \label{eq:RSFT-master}
5582: \frac{\partial \bff}{\partial \tau_j} (t_1,\dots,t_k,
5583: \tau_1,\dots,\tau_m)=
5584: \bh^j(t_0,\dots,t_l,q,p)|_{L_{\bff}}
5585: \end{equation}
5586: with the initial condition
5587:  $$\bff|_{\tau_j=0}=\bff_{W,J}(\sum
5588:  t_i\Delta_i+\sum\limits_{r\neq j}\tau_r\Theta_r,p).$$
5589: Then $$\bff(t_1,\dots,t_k,\tau_1,\dots,\tau_m, p)$$ is
5590:   homotopic to $$\bff_{W,J}(\sum
5591:  t_i\Delta_i+\sum\limits_{r=1}^m\tau_r\Theta_r,p).$$
5592:  \end{theorem}
5593: 
5594: 
5595: \subsection{Invariants    of Legendrian
5596: knots}\label{sec:Legendrian}
5597: Symplectic Field Theory can be extended to include Gromov-Witten
5598: invariants of pairs $(W,L)$, where $L$ is a Lagrangian
5599: submanifold of a symplectic manifold $W$.   The corresponding relative object is a
5600: pair $(W,L)$, where $W=\ora {V^-V^+}$ is a directed symplectic cobordism between
5601: contact manifolds $(V^\pm,\a^\pm)$, and $L$ is a Lagrangian cobordism between Legendrian
5602: submanifolds $\Lambda^\pm\subset V^\pm$.  More precisely, we assume that Lagrangian submanifold
5603: $L$ is cylindrical at infinity over $\Lambda^\pm$, i.e. there exists $C>0$, such that
5604: $\;L\cap V^-\times (-\infty, -C]=\Lambda^-\times (-\infty, -C]\;$ and
5605: $\; L\cap V^+\times (C,\infty]=
5606: \Lambda^+\times (C,\infty]\;$.
5607:                                   In other words, we require $L$
5608:                                    to coincide at infinity with  symplectizations
5609: of Legendrian submanifolds $\Lambda^\pm$.
5610: 
5611: The moduli space of holomorphic curves to be considered in this case
5612: consists of  equivalence classes of     holomorphic curves   with boundary which can have
5613: punctures of two types,
5614:  interior  and at the boundary.  The boundaries of holomorphic curves
5615: are required to be mapped to the Lagrangian submanifold $L$,   the holomorphic  curves
5616: should be cylindrical over periodic orbits from $\Pc^\pm$ at interior punctures,
5617:  while at boundary punctures
5618: we require them to be asymptotically cylindrical over Reeb chords connecting points  of the Legendrian submanifolds
5619: $\Lambda^\pm\subset V^\pm$.  A  more precise definition is given below.
5620: The algebraic structure arising from the stratification of boundaries of these moduli spaces is more complicated
5621: than in the closed case. First of all, unlike the interior punctures the    punctures
5622:  at the boundary are cyclically ordered, which leads to associative, rather than graded commutative algebras.
5623:  Second, the ``usual" cusp degenerations of curves with boundary
5624:  at boundary points (see \cite{Gromov-holomorphic}) has
5625:  in this case codimension
5626:  $1$, rather than $2$ as in the  closed case, and hence the  combinatorics of such
5627:   degenerations  should also be a part of the
5628:  algebraic formalism.
5629: 
5630:  We will sketch in this paper only the simplest of three cases of SFT, namely the ``classical case",
5631:  which corresponds to the theory of moduli spaces of holomorphic
5632:  disks with only $1$ positive puncture at the boundary.
5633:  \bigskip
5634: 
5635:          Let $(V,\xi=\{\a=0\})$ be a contact  manifold with a fixed
5636:          contact form $\a$, $W=V\times \R$ its symplectization
5637:          with a compatible almost complex structure $J$,
5638:          $\Lambda\subset V$ a  compact Legendrian submanifold,
5639:          and $L=\Lambda\times\R\subset W$ the symplectization of $\Lambda$, i.e.
5640:          the corresponding Lagrangian cylinder in $W$.   We  assume that all periodic orbits
5641:          of  the Reeb   vector field $R_\a$ are non-degenerate
5642:            and fix a marker
5643:          on every periodic orbit. We also consider the set $\Cc$   of Reeb chords
5644:           connecting points on $\Lambda$, and impose an extra
5645:            non-degeneracy condition along the   chords from $\Cc$
5646:           by requiring that the linearized flow of $R_\a$  along
5647:                    a chord   $c\in\Cc$  connecting points $a,b\in\Lambda$
5648:                      sends  the tangent  space $T_a(\Lambda)$ to
5649: a space transversal to $T_b(\Lambda)$.
5650: We also require that none of
5651: the chords from $\Cc$ be a part of an orbit from
5652: $\Pc$.
5653:  Under these assumptions, the set $\Cc$ is finite: $\Cc=\{c_1,\dots,c_m\}$.
5654: 
5655:  We will restrict the consideration to the case when
5656:  \begin{equation}
5657:  \pi_1(V)=0,\; \pi_2(V,\Lambda)=0,\quad\hbox{ and}\quad w_2(\Lambda)=0.
5658:  \end{equation}
5659:  First two assumptions are    technical and can be removed (comp. Section \ref{sec:dynamics} above).
5660:   However, the third one
5661:  is essential for orientability of the involved moduli spaces of holomorphic curves. Moreover, the invariants
5662:   we define depends on a particular choice of a spin-structure on $\Lambda$.
5663:   \footnote{We thank K. Fukaya for pointing this out.}
5664: 
5665:  As in  Section \ref{sec:dynamics} we choose capping surfaces $F_\g$ for $\g\in\Pc$, and  for each chord
5666:  $c\in \Cc$ we  also choose   a surface $G_c$
5667: which is bounded by  a curve
5668:   $c\cup\delta_c, $ where $\delta_c\subset \Lambda$.
5669:    The choice of  surfaces $F_\g,\,\g\in\Pc,$  allows us to define
5670:       Conley-Zehnder indices of  periodic orbits
5671:    (see Section \ref{sec:dynamics} above). Similarly, surfaces $G_c$ enable us to define
5672:    {\it Maslov indices} $\mu(c) ,\,c\in\Cc$.   Namely,  let us consider a Lagrangian subbundle of $\xi|_{\partial G_c}$,
5673:    which consists of  the Lagrangian sub-bundle
5674:   $T\Lambda|_{ \delta_c}\subset \xi|_{\delta_c}$ over $\delta_c$, together with
5675:    the family  of Lagrangian  planes  $T_u\subset \xi_u,u\in c$, which are images of $T_a(\Lambda)$ under the linearized flow of
5676:    the Reeb field $R_\a$.  Choose a symplectic trivialization of $\xi|_{ \partial G_c}$ which extends
5677:    to $G_c$.   With respect to this trivialization the above sub-bundle can be viewed
5678:    as a path of Lagrangian planes in a
5679:    symplectic vector space.  The Maslov index of such path
5680:     is defined as in \cite{Robbin-Salamon}.
5681: 
5682:     \medskip
5683:    Consider a unit disk
5684:    $D\subset\C$ with punctures $$\left(\{z^+,z_1^-,\dots,z_\sigma^-\}\cup\{x^-_1,\dots, x^-_s\}\right),$$
5685:    where  $\bz=\{z^+,z_1^-,\dots,z_\sigma^-\}$, $0\leq\sigma\leq m$,
5686:    is  a   counter-clockwise ordered set of punctures on $\partial D$,
5687:    and  $\bx=\{x^-_1,\dots, x^-_s\}$ is an ordered set of interior punctures.
5688:     As usual  we provide  interior punctures with asymptotic markers.
5689: 
5690:                         Let us denote by
5691:                          $\Mc^A(\{c_{i_1},\dots,c_{i_\sigma}\},\{\g_1,\dots,\g_s\},c_i;W,\Lambda,J)$
5692:                          the moduli space of
5693: $J$-holomorphic maps
5694: \begin{equation*}
5695: (D\setminus \left(\bz\cup\bx\right),
5696: \partial\left(D\setminus \left(\bz\cup\bx\right)\right) \to (W,L),
5697: \end{equation*}
5698:  which
5699: are asymptotically cylindrical at    the negative end over the periodic orbit $\g^-_{k}$
5700:   at the puncture $x^-_{i_k}$, and over the chord $c_{i_k}$ at the puncture $z_k$,
5701: asymptotically cylindrical at    the positive  end over the chord $c_i$
5702:    at the puncture $z^+$, and which send
5703: asymptotic markers of interior punctures to
5704: the  markers on the  corresponding
5705: periodic orbits.
5706: 
5707:   Two maps are called equivalent if they differ by a conformal map
5708:  $D\to D$ which preserves all punctures, marked points and     asymptotic markers.
5709:     Each moduli space  $\Mc^A(c_i,\{c_{i_1},\dots,c_{i_\sigma}\},\{\g_1,\dots,\g_s\};W,\Lambda,J)$ is invariant under
5710:     translations  $V\times\R\to V\times\R$ along the factor  $\R$, and we
5711:    denote the corresponding quotient moduli space  by
5712:      $$\Mc^A(c_i,\{c_{i_1},\dots,c_{i_\sigma}\},\{\g_1,\dots,\g_s\};W,\Lambda,J)/\R\;.$$
5713: 
5714: Let $(\fA,\partial)=(\fA(V,\alpha),\partial_J)$ be the graded commutative differential algebra
5715: defined in Section
5716: \ref{sec:3algebras} above, or rather its specialization at the point $0$.
5717:  Consider a graded  associative  algebra $\fK=\fK(V,\Lambda,\alpha)$ generated
5718: by elements $c_i\in \Cc$ with coefficients in the algebra $\fA$.
5719: We define a differential $\partial_{\Lambda}=\partial_{\Lambda,J}$
5720: on $\fK$ first on the generators $c_i$ by the formula
5721: \begin{equation}\label{eq:d-leg}
5722:  \partial_{\Lambda}(c_i)=
5723:  \sum \frac{n_{\Gamma,I,d}}
5724:  {k!\prod_1^k \kappa_{\g_j}^{i_j}i_j!}
5725:   c_{j_1}\dots c_{j_\sigma}
5726:  q_{\g_1}^{i_1}\dots q_{\g_k}^{i_k}z^d,
5727:   \end{equation}
5728:    where the sum is taken over all $d\in H_2(V)$, all ordered sets of
5729:    different periodic orbits $\Gamma=\{\g_1,\dots,\g_k\} $,
5730:     all multi-indices  $J=(j_1,\dots,j_\sigma),\,1\leq j_i\leq m$,
5731: and $I=(i_1,\dots,i_k)$, $i_j\geq 0, $ and where the coefficient
5732: $n_{\Gamma,I,d}$  counts the algebraic number of elements of the
5733: moduli space $$\Mc^d\big(c_i,\{c_{j_1},\dots,c_{j_\sigma}\},
5734: \{\underbrace{\g_1,\dots,\g_1}_{i_1},\dots,
5735: \underbrace{\g_k,\dots,\g_k}_{i_k}\}\big)/\R,$$ if this space is
5736: $0$-dimensional, and equals $0$ otherwise. The differential
5737: extends to the whole  algebra $\fK$ by the graded Leibnitz rule.
5738: However, it does not treat coefficients as constants: we have
5739: $\partial_\Lambda(q_\g)=\partial(q_\gamma)$, where  $\partial$ is
5740: the differential defined on the algebra $\fA$.
5741: 
5742: Then we have
5743: \begin{proposition}
5744: $$  \partial_{\Lambda}^2=0.$$
5745: \end{proposition}
5746: Given a family   of contact forms $\Lambda_\tau,
5747: \alpha_\tau,
5748: J_\tau\;\tau\in[0,1]$
5749: %CHANGE
5750: of Legendrian submanifolds, contact forms, and compatible almost complex structures one can define, similar to the  case of closed
5751: contact manifolds (see Sections \ref{sec:relative}  and Section \ref{sec:chain} above)
5752: a   homomorphism of differential algebras
5753: $$\Psi_S:\fK(V, \Lambda_0,\alpha_0)\to\fK(V, \Lambda_1,\alpha_1),$$
5754: which is independent up to homotopy of the choice of a connecting
5755: homotopy.
5756: Composition of homotopies generates composition of homomorphisms, and hence one conclude
5757: \begin{proposition}
5758: The quasi-isomorphism type of the differential algebra
5759:   $$(\fK(V,\Lambda,\alpha),\partial_{\Lambda,J})$$ depends only on the contact structure $\xi$
5760:   and the Legendrian isotopy class of $\Lambda$. The
5761:   %CHANGE
5762:   {\em Legendrian contact homology} algebra
5763:   $H_*(\fK,\partial_\Lambda)$ has a structure of a module over
5764:   the contact homology algebra
5765:   $H^\cont_*(V,\xi)=H_*(\fA,d)$, and it is an invariant of the Legendrian knot (or link) $\Lambda$.
5766:   \end{proposition}
5767: 
5768: 
5769:     The theory looks especially simple when the contact structure $\xi$ on
5770:     $V$ admits a contact form $\alpha$ such that the Reeb vector
5771:     field $R_\a$ has no closed periodic orbits. If, in addition
5772:     the space of trajectories is a  manifold $M$ (e.g. when
5773:     $V=J^1(N)=T^*(N)\times\R$ with a contact form $dz+pdq$), then
5774:      $W$ is automatically symplectic, and the projection $\pi:W\to V$ sends the Legendrian submanifold
5775:     $\Lambda\subset V$ to an immersed {\it Lagrangian} submanifold
5776:     $L\looparrowright M$ with transverse self-intersection points. These points
5777:     correspond to Reeb chords  $c_i$
5778:     connecting points on $\Lambda$.
5779:     Hence,  the algebra $\fK$ in this case is just a free associative
5780:     algebra, generated over $\C$ (or $\C$) by the
5781:     self-intersection points of $L$. It is possible
5782:      to choose a compatible  almost complex structures
5783:     $J$ on  the symplectization  $W=V\times\R$ and  $J_M$ on $M$
5784:     to make
5785:     the projection $W\to M$ holomorphic (comp. Section \ref{sec:Bott} below). Then  punctured
5786:     holomorphic disks in $W$  from moduli spaces
5787:     $\Mc^A(c_i,\{c_{i_1},\dots,c_{i_\sigma}\} ;W,\Lambda,J)$
5788:     project to $J_M$-holomorphic disks in $M$ with boundary in the
5789:     immersed Lagrangian  manifold $L$. Conversely, one can
5790:     check that each such disk  lifts to a disk from  the
5791:    corresponding moduli space $\Mc^A(c_i,\{c_{i_1},\dots,c_{i_\sigma}\}
5792:    ;W,\Lambda,J)$, uniquely, up to translation along
5793:     the $\R$-factor in $W=V\times\R$. This is especially useful when $\dim M=2$.  In
5794:    this case $L$ is an immersed curve, and the holomorphic disks
5795:    are precisely immersed, or branched disks with their boundaries in $L$.
5796:    Moreover, branched disks are never rigid, because the branching
5797:    point may vary. Hence, the  differential
5798:    $\partial:\fK\to\fK$ can be defined in this case in a pure
5799:    combinatorial way, just   summing the terms  corresponding
5800:    to  all  appropriate immersed disks whose boundary consists of arcs of $L$, and
5801:     which are locally convex near their corner.
5802: 
5803:     Yu. Chekanov independently realized  (see \cite{Chekanov}) this program  for
5804:     Legendrian links in the standard contact $\R^3$.
5805:     He was also  motivated by the  hypothetical description of the
5806:     compactification of the  moduli spaces of holomorphic discs,
5807:     but has chosen to prove the  invariance of the quasi-isomorphism type
5808:     of the differential algebra $(\fK,\partial)$ in a pure
5809:     combinatorial way.  In fact, he proved   a potentially stronger    form  of equivalence
5810:     of
5811:     differential algebras of isotopic Legendrian links,
5812:     which
5813:     he called     stable tamed isomorphism.  Stable
5814:     tame
5815:     isomorphism implies quasi-isomorphism, but we do not know
5816:     whether it is indeed stronger. Let also note that
5817:      Chekanov considered a   $\Z_2$-version of the theory.
5818:     In some examples it works better
5819:       the  $\Q$-version, which is provided by our formalism.
5820:     J. Etnyre--J. Sabloff (\cite{Etnyre-Sabloff}) and L. Ng (\cite{Ng}) worked out
5821:      the combinatorial
5822:     meaning of signs dictated by the coherent orientation
5823:     theory (see Section \ref{sec:orientation} above), and proved
5824:     invariance of the
5825:     stable tame type of the differential algebra $(\fK,\partial)$
5826:     {\it defined over $ \Z$}. Note that Chekanov's paper \cite{Chekanov}
5827:     also contains   examples  which show that the  stable  tame
5828:     $\Z_2$-isomorphism type do distinguish some Legendrian knots,
5829:     which could not be  previously distinguished.
5830: 
5831: 
5832: 
5833: 
5834:  Similar to the absolute case of SFT, one can define   further invariants of Legendrian
5835:   submanifolds by including in the formalism higher-dimensional moduli spaces. For instance,  by introducing
5836:     marked points on the boundary
5837:   of the disk
5838:   one gets a non-commutative deformation of Legendrian  contact homology
5839:   along the homology of Legendrian
5840:   manifolds. This is useful, in particular, to define invariants of
5841:     Legendrian links with ordered components.
5842:   However, the full-scale generalization of Symplectic Field Theory  to
5843:   directed symplectic-Lagrangian cobordisms between pairs
5844:   of contact manifolds and their Legendrian submanifolds, which would formalize
5845:   information  about moduli spaces of   holomorphic curves of arbitrary genus and arbitrary number
5846:   of positive and negative punctures, is not straightforward due to existence of different type
5847:   of codimension $1$ components on the boundary of the corresponding moduli spaces.
5848:   We will discuss this theory
5849:   in one of our future papers.
5850: 
5851: 
5852: 
5853: 
5854: \subsection{Remarks, examples, and further algebraic constructions in
5855: SFT}\label{sec:remarks}
5856:  \subsubsection{Dealing with torsion elements in $H_1$} \label{sec:torsion}
5857: Let us discuss grading issues for a contact manifold
5858: $(V,\xi=\{\a=0\})$ in the case when
5859:   the torsion part  of $H_1(V)$ is
5860: non-trivial. As we will see it is impossible to assign in a coherent way
5861: an integer grading to torsion elements and still keep the property
5862: that the Hamiltonian $\bH$ has total grading $-1$.  We will
5863: deal with this problem by assigning to some elements fractional
5864: degrees, and thus obtain a rational grading, incompatible with the
5865: canonical $\Z_2$-grading. In fact the
5866: term ``grading" is  misleading in this case, and more appropriately one
5867: should talk about
5868:      an Euler vector field with rational coefficients.
5869: 
5870: Let us split  $H_1(V)$ as $T\oplus F$, where $T$  and $F$ are   the
5871: torsion and free parts, respectively. As in Section \ref{sec:dynamics}
5872: above let us  choose curves ${C}_1,\dots,{C}_k$ representing a basis
5873: of $F$, fix a trivialization of the bundles $\xi_{{C}_i}$,
5874: for any periodic orbit $\g\in\Pc_\a$ with $[\g]\in F$ choose a
5875: surface $F_\g$ which realizes the homology between $[\g]$ and a
5876: linear combinations $\sum n_i [{C}_i] $, and trivialize the
5877: bundle $\xi|_\g$ accordingly.
5878: For any other  periodic orbit $\g$ let $\g_l$ be its smallest multiple which
5879: belong to  $F$. In particular, the bundle $\xi_{\g_l}$ is
5880: already trivialized by a framing $f$. The problem is that in general there is no
5881: framing over $\g$ which would produce $f$ over $\g_l$. Choose then
5882: an arbitrary framing $g$  over $\g$ and denote by $g_l$ the
5883: resulting framing over $\g_l$.  Let
5884: $2m(g_l,f)\in\pi_1(Sp(2n-2,\R))=\Z$ be the Maslov class of the
5885: framing $g_l$ with respect to $f$. The Conley-Zehnder indices of
5886: $\g_l$ with respect to these two gradings are then
5887:  related  by the formula
5888: \begin{equation*}
5889: \CZ(\g|f)=\CZ(\g|g_l)+2m(g_l,f).
5890: \end{equation*}
5891: We then assign to $\g$ the fractional  degree
5892: \begin{equation}
5893: \deg \g=\CZ(\g|g)-\frac{2m(g_l,f)}{l}\,.
5894: \end{equation}
5895: With this modification SFT can be extended to the case of contact
5896: manifolds with no restrictions on $H_1$. However, the price we pay
5897: is that this grading, even if integer, may not be compatible with
5898: the universal $\Z_2$-grading which determines the sign rules.
5899: 
5900: 
5901: 
5902: \subsubsection{Morse-Bott formalism}\label{sec:Bott}
5903: 
5904: 
5905:  Our assumption that all periodic orbits from $\Pc_\a$ for the
5906:  considered contact forms $\a$
5907:  are non-degenerate, though generic,  but is very inconvenient for
5908:  computations: in many interesting examples periodic orbits
5909:  come in continuous families. Sometimes  the  Reeb
5910: flow is periodic, and it sounds quite stupid to destroy this beautiful
5911: symmetry.
5912: 
5913: In fact the above formalism can be adapted to this ``Morse-Bott"
5914: case. We sketch below how this could be done   for the periodic
5915: Reeb flow of an $S^1$-invariant form of a
5916:  pre-quantization space.
5917: We consider below only the ``semi-classical" case which concerns
5918:  moduli spaces of rational holomorphic curves.
5919: 
5920: 
5921: 
5922: Let $(M,\omega)$ be a symplectic manifold  of dimension $2n-2$
5923: with  an integral cohomology class $[\omega]\in H^2(M)$. We will assume for simplicity
5924: that $H_1(M)=0$.
5925:  The
5926: pre-quantization space $V$ is a circle bundle over $M$ with
5927: first Chern class equal to $[\omega]$. The fibration $\pi:V\to M$
5928: admits a   $S^1$-connection form $\alpha$ whose curvature is
5929: $\omega$. It   defines a $S^1$-invariant contact structure  $\xi$
5930: on $V$, transversal to the fibers  of the fibration. The Reeb flow
5931: of $R_\a$ is periodic, so all its trajectories
5932:   are closed and coincide with the fibers of the fibration $\pi$, or their multiples.
5933: 
5934: 
5935:   The fiber of the fibration $V$ is a torsion element
5936: in $H_1(V)$, and if $l$ is the greatest divisor of the class $[\omega]$ then
5937: the  $l$-multiple of the fiber is homological to $0$.
5938: 
5939: 
5940: Consider the cylindrical cobordism  (the symplectization)
5941: $W=V\times\R$ with an almost
5942:  complex structure $J$ compatible with $\alpha$
5943: and denote by $\Mc_{0,r}(s|W,J,\a)$ the moduli  space of rational
5944: holomorphic curves in $W$ with $s$ punctures and $r$ marked points.
5945: Near punctures the curves are required to be  asymptotically
5946: cylindrical over some fibers of $V$, or their multiples.
5947: However, we do not specify to which particular fiber they are
5948:  being asymptotic, or whether this fiber is
5949:  considered
5950: on the positive, or negative end of $W$. We
5951:  do not equip   curves from $\Mc_{0,r}(s|W,J,\a)$
5952: with asymptotic markers of punctures, because   we  cannot
5953:   fix  in a continuous way  points on each simple periodic orbit,
5954: as we did in the non-degenerate case.
5955: 
5956: As it was already mentioned in Section \ref{sec:almost} above,
5957:  $W$ can be viewed  as the total space of the complex  line bundle  $L$ associated with
5958: the $S^1$-fibration $V\to M$, with the zero-section removed, and
5959: the almost complex structure $J$ can be chosen
5960:   compatible with the structure of this bundle, so that the projection   $W\to M$
5961: becomes holomorphic with respect to a certain almost complex structure $J_M$ on $M$
5962: compatible with $\omega$.
5963: Then automatically the   bundle induced over  any holomorphic curve in the base
5964: has a structure of a holomorphic line bundle.
5965: With this choice of $J$ each holomorphic curve   $f\in \Mc_{0,r}(s|W,J,\a)$ projects
5966: to a $J_M$-holomorphic sphere $\overline{f}:
5967: \C P^1\to M$,   and  can be viewed as a meromorphic section of the   induced
5968: bundle $(\overline{f})^*L$ over $\C P^1$.  This bundle is ample, and hence  poles of its
5969: sections
5970: correspond to   the  {\it negative} ends of $f$, while zeroes  correspond to  the positive ones.
5971: Notice that although  the moduli spaces
5972:  $\Mc_{0,r}(s|W,J,\a)$ can be identified with the moduli spaces
5973: of closed  holomorphic curves in a $\C P^1$-bundle over $M$
5974: with prescribed tangencies to two divisors,
5975:  their compactifications are different, and in particular the
5976:  compactification of the first moduli space may have codimension one strata
5977:  on its boundary.
5978: 
5979: 
5980: 
5981: The correspondence $f\mapsto\overline{f}$ define a fibration
5982:  $$\pr:\Mc_{0,r}(s|W,J,\a)/\R\to \Mc_{0,r+s}(M,J_M).$$
5983:  The fiber $\pr^{-1}(\overline{f})$ is   the
5984:   union of  (an infinite number of) disjoint circles, which are indexed by
5985:  sequences of integers $(k_1,\dots,k_{s+r})$ with $\sum k_i=d_0
5986:  =\int\limits_{A}\omega$,
5987:  where $A\in H_2(M)$ is the homology class realized by $\overline{f}$,
5988:   and where  there are precisely
5989:  $s$  non-zero coefficients  $k_i$.
5990: 
5991: Let us  consider two copies $\Pc^\pm$ of  the space $\Pc=\Pc_\a$ of periodic orbits,
5992:   as we need to differentiate
5993: between positive and negative ends of holomorphic curves. We will write $\ddot{\Pc}=\Pc^+\cup\Pc^-$
5994:  and define  the   evaluation maps:
5995: \begin{equation}
5996: ev_0:\Mc_{0,r}(s|W,J,\a)/\R\to V^{\times r}\;\; \hbox{ and}\;\;
5997:  ev^+_-:\Mc_{0,r}(s|W,J,\a/\R)\to\ddot{\Pc}^{\times s}\,.
5998: \end{equation}
5999: 
6000: Here $ev^+_-$ associates with each puncture the corresponding
6001: point of $\ddot\Pc$.
6002: The space $\Pc^\pm$  can be presented as
6003: $    \coprod\limits_{k=1}^{\infty}\Pc^\pm_k, $
6004:     where each $\Pc^\pm_k$ is a copy   of $M$, associated
6005:     with $k$-multiple orbits.
6006: 
6007: We will denote forms on $\Pc^+$ by $p$, on $\Pc^-$ by $q$,
6008:   denote by $p_k, q_k$ their restrictions to $\Pc_k^\pm$, and organize them
6009:   into   Fourier series
6010:   $u=\sum\limits_{k=1}^\infty(p_ke^{ikx}+q_ke^{-ikx})$. If we are given a
6011:   basis of $H^*(M)$ represented by forms $\Delta_1,\dots,\Delta_a$
6012:   we will consider only forms from the space generated by this
6013:   basis, and write $p_k=\sum\limits_{i=1}^a p_{k,i}\Delta_i,
6014: q_k=\sum\limits_{i=1}^a q_{k,i}\Delta_i$ and denote by
6015: $u_i$ the  $\Delta_i$-component of $u$, i.e.
6016:  $$u_i=\sum\limits_{k=1}^\infty(p_{k,i}e^{ikx}+q_{k,i}e^{-ikx})\;\;\hbox{
6017:  and}\;\;
6018:  u=\sum\limits_1^a u_i\Delta_i.$$
6019: 
6020: 
6021: Given a  closed form $t$ on $V$ and  a class $A\in H_2(V)$
6022:  we define the correlator
6023:  \begin{equation}
6024:  \begin{split}
6025:  & \up{-1}\langle
6026: \underbrace{t,\dots,t}_r; \underbrace{u,\dots,u}_s\rangle^A_{0} =\\
6027: &  \int\limits_{\overline{\Mc^A_{0,r}(s|W,J,\a)/\R} } ev_0^*\big(t\otimes \dots
6028: \otimes t\big)\wedge (ev^+_-)^*\big(u \otimes\dots
6029: \otimes u\big) \bigg{|}_{x=0}.\\
6030: \end{split}
6031: \end{equation}
6032: 
6033: Let us  choose a basis $A_0,\dots, A_N$ in
6034:   $H_2(M)$ in such a way that $\int\limits_{A_0}\omega=l>0$ and
6035:   $\int\limits_{A_i}\omega= 0$ for $i=1,\dots, N$.
6036:   Then the classes $A_i,i\geq 1 ,$ lift to classes $\wt{A}_i\in H_2(V)$  which
6037:   %CHANGE
6038: under the assumption $H_1(M)=0$ form
6039:    a basis of $
6040:   H_2(V)$. The   degree
6041:   $d=(d_1,\dots,d_N )$ of a class
6042:   $A\in H_2(V)$ is a vector of its coordinates in this basis.
6043: 
6044: 
6045: To associate an absolute   homology
6046: class with a holomorphic curve
6047: we  pick   the $l$-multiple (recall that   $l$ denotes the greatest divisor of $\omega$)
6048: of the fiber   $\g$ over a point $x\in M$
6049:  and
6050: choose a lift of the surface representing the class
6051:  $A_0$ with $\int\limits_{A_0}\omega =l$ as a spanning surface
6052: $F_\g$. Any other $m$-multiple of  $\g$ we will cap with the chain
6053:  $\frac ml[F_\g]$. However, to fix a spanning surface
6054: for a    fiber over any other point $y\in  M$  or
6055: its multiples, one needs to make some extra choices, for instance   fix
6056:  a path connecting $x$ and $y$. The condition $H_1(M)=0$ guarantees
6057:   independence of the homology class of the resulting surface of
6058:  the choice of this connecting path.  Notice that   with this choice,
6059:  the degree of $f\in \Mc_{0,r}(s|W,J,\a)/\R$  equals $(d_1,\dots,d_N)$, if the degree of its projection
6060:  $\pr(f)\in\Mc_{0,r+s}(M,J_m)$ is equal to $(d_0,d_1,\dots,d_N)$.
6061: 
6062: 
6063: In this notation the rational Hamiltonian $\bh=\bh_{V,J,\a }$ is defined
6064:  by the formula
6065: \begin{equation}
6066: \bh(t,u)=\sum\limits_{d}\sum\limits_{r,s=0}^\infty\frac{1}{r!s!}
6067:     \up {-1} \langle \underbrace{t,\dots,t}_r;\underbrace{u,\dots,u  }_{s}
6068:      \rangle_{0}^{d }  z^d.
6069:      \end{equation}
6070: 
6071:    Suppose that a basis of $H^*(M)$, represented by
6072:      closed forms   $\Delta_1,\dots,\Delta_a$, is  chosen in such
6073:      a way that   for $b\leq
6074: a$  the system forms $\wt{\Delta}_j=\pi^*(\Delta_j),\,j=1,\dots,b$, generate
6075:  the image $\pi^*(H^*(M))\subset H^*(V)$, and the forms
6076:  $\wt{\Theta}_1,\dots,\wt{\Theta}_c$,
6077:   complete  it
6078:    to a basis of $H^*(V)$.
6079:     We will denote  (graded) coordinates in the space generated by the forms
6080:      $\wt{\Delta}_j,\,j=1,\dots, b$ and
6081:   $\wt{\Theta}_1,\dots,\wt{\Theta}_c$ by
6082:     $ (t,\tau)=(t_1,\dots,t_b,\tau_1,\dots, \tau_c)$.
6083: 
6084: 
6085: As usual, the Hamiltonian $\bh$ is viewed as an element  of  a  graded
6086:  commutative Poisson
6087:  algebra $\fP$,
6088:  which consists of formal power series of coordinates of vectors $p_k$
6089:  and  $T=(t,\tau)=
6090:  (t_1,\dots,t_b,\tau_1,\dots,\tau_c)$
6091:   with coefficients
6092:  which are polynomials of coordinates of vectors $q_k=(q_{k,1},\dots,q_{k,a})$.
6093:   The coefficients of these polynomials
6094:  belong to a certain completion (see condition (\ref{eq:Novikov}) above) of the group algebra
6095:  of $H_2(V)$.
6096:  All  the variables $p_{k,i},q_{k,i}$  have in this case
6097:  the same parity as forms $\Delta_i$, the parity of variables $t_i$ and $\tau_j$ is the same as the  degree of
6098:  the corresponding
6099:  forms $\wt \Delta_i$ and $\wt\Theta_j$.
6100:   If $l=1$, i.e. when $H_1(V)=0$, then  the integer grading of variables which corresponds
6101:   to the   choice of capping surfaces described above  is defined
6102:    as follows:
6103: \begin{equation}\label{eq:dim-Bott}
6104: \begin{split}
6105: \deg t_i&=\deg \wt{\Delta}_i-2;\\
6106: \deg \tau_i&=\deg\wt{\Theta}_i-2;\\
6107: \deg q_{k,i}&=\deg\Delta_i-2+2ck;\\
6108: \deg p_{k,i}&=\deg\Delta_i-2-2ck;\\
6109: \deg z_i&==2c_1(A_i)\\
6110: \end{split}
6111: \end{equation}
6112: where $c=c_1(A_0)$.
6113: As it was explained in Section \ref{sec:torsion}  if $l>1$ one can only define fractional degrees, given by the above
6114: formulas (\ref{eq:dim-Bott}) with $c=\frac{c_1(A_0)}{l}$.
6115: 
6116: The following proposition is useful for applications (see below the discussion
6117: of Biran-Cieliebak conjecture about subcritical symplectic mainfolds).
6118: It follows from the fact that all the  moduli spaces
6119: $\Mc_{g,r}(s|W,J,\a)$ which we defined above  are even-dimensional.
6120: 
6121: \begin{proposition}\label{prop:even-dim}
6122: Let $(V,\xi)$ be the contact pre-quantization space for a
6123: symplectic manifold $(M,\omega)$. Then all contact homology algebras
6124: $$H_*^{\SFT}(V,\xi)\big|_{t=0},H_*^{\RSFT}(V,\xi)\big|_{t=0},H_*^{\cont}(V,\xi)\big|_{t=0}$$
6125: specialized  at $0\in H^*(V)$
6126: are
6127: free graded, respectively  Weyl, Poisson, or commutative algebras,
6128: generated by  elements
6129: $$p_{k,i},q_{k,i},\;\;i=1,\dots,a,\;k=1,\dots\,.$$ In particular, the parts of all these homology
6130: algebras graded by the homology class $w\in H_1(V)$ ({\rm see \ref{rem:grading} above})
6131: are   non-trivial.
6132:  \end{proposition}
6133: 
6134: The Poisson tensor on $\fP$  is determined in the ``$u$-notation" by
6135: the  following generalization of the   formula
6136:  (\ref{eq:u-notation}):
6137: \begin{equation}\label{eq:u-vector}
6138: %CHANGE
6139:  \frac{1}{2\pi i}\int\limits_0^{2\pi}\langle(\delta u)',\delta
6140: v\rangle dx,
6141: \end{equation}
6142:  where $\langle\;,\;\rangle$ denotes Poincar\'e pairing
6143: on cohomology $H^*(M)$,  which is given in the basis $\Delta_1,\dots, \Delta_a$ by the matrix
6144: $$\eta_{ij}=\langle\Delta_i,\Delta_j\rangle=\int\limits_M\Delta_i\wedge\Delta_j.$$
6145: The Poisson tensor  can be written in
6146: $(p,q)$-coordinates as
6147: \begin{equation*}
6148: \sum\limits_{k=1}^\infty k\sum\limits_{i,j=1}^{a }\eta_{ij}\frac{\partial}{\partial
6149: p_{k,i}}\wedge\frac{\partial}{\partial
6150: q_{k,j}}\,.
6151: \end{equation*}
6152: It can be shown that the  above Hamiltonian $\bh$
6153: satisfies the identity $\{\bh,\bh\}=0,$  and  that the differential
6154: Poisson algebra $(\fP,d^{\bh})$ is quasi-isomorphic to the
6155: corresponding  differential Poisson algebra defined  in Section
6156: \ref{sec:3algebras} for any non-degenerate contact form for the
6157: same contact structure  $\xi$ on $V$.
6158: \bigskip
6159: 
6160:  The following formula (\ref{eq:Bourg}), which sometimes allows to compute
6161:  the Hamiltonian $\bh$  of $V$
6162:   in terms of the Gromov-Witten invariant $\bff=\bff_{M,J_M}$ of $M$, emerged   in a discussion  of
6163:   the authors with T. Coates and  F. Bourgeois.
6164: 
6165:     \begin{proposition}\label{prop:Coates}
6166:     Set
6167:     $$
6168:     \bh^{j}_{W,J}(t,q,p,z)=\frac{\partial\bh}
6169:     {\partial\tau_j}(\sum\limits_1^b
6170:     t_i\wt\Delta_i+\tau_j\wt\Theta_j,q,p,z)
6171:     \big|_{\tau_j=0}\;,
6172:     $$
6173:     and
6174:     $$\widehat\bff^j(t,z)=\frac{\partial\bff}{\partial s}\big(
6175:     \sum\limits_1^a t_i\Delta_i +s\pi_*\wt\Theta_j,z\big)\big|_{s=0}
6176:     \,,$$
6177:     for $j=1,\dots,c$
6178:           (comp. Theorem \ref{thm:RSFT-master}).
6179:    Then   we have
6180:    \begin{equation}
6181:       \label{eq:Bourg}
6182:       \begin{split}
6183:     &\bh^j_{W,J}(t_1,\dots,t_b,q,p,z)\\&
6184:     =\frac{1}{2\pi }\int\limits_0^{2\pi}
6185:     \widehat\bff^j(t_1+u_1(x),\dots,t_b+u_b(x), u_{b+1}(x),\dots,u_a(x),\tilde z)dx
6186:     \end{split}
6187:         \end{equation}
6188:      where $z=(z_1,\dots,z_N)$,
6189:       $\tilde z=(e^{-ilx},z_1,\dots,z_N)$ and
6190:       $l$ is the greatest divisor of $\omega$.
6191:      \end{proposition}
6192:      To prove    (\ref{eq:Bourg}) one just observes that
6193:       the correlator
6194:           \begin{equation*}
6195:  \up{-1}\langle
6196: \wt \Delta_{j_1},\dots,\wt\Delta_{j_r},\wt\Theta_j;
6197: u_{i_1}\Delta_{i_i},\dots,u_{i_v}\Delta_v\rangle^d_{0}
6198: \end{equation*}
6199: equals  the Fourier coefficient with $e^{ilx}$ of
6200: the       correlator
6201:    \begin{equation*}
6202:  \up{0}\langle
6203: {\Delta_{j_1},\dots,\Delta_{j_r},\pi_*\wt\Theta_j}, {
6204: u_{i_1}\Delta_{i_i},\dots,u_{i_v}\Delta_v}\rangle^{\tilde
6205: d}_{0}\,.
6206: \end{equation*}
6207: 
6208:       Notice that if $\wt\Theta_j$ is an odd form, then
6209:       $$\bh(\sum\limits_1^b
6210:     t_i\wt\Delta_i+\tau_j\wt\Theta_j,q,p,\tilde z)=
6211:     \tau_j\bh^j(t,q,p,z),$$
6212:     because all terms of $\bh$ must contain $\tau_j$.
6213:                                         In particular,
6214:      for $M=\C P^{n-1}$  the manifold $V$ is a rational homology sphere, and thus
6215:      a volume form $\Theta$  on $S^{2n-1}$
6216:      generates the odd part
6217:      of $H^*(V;\R)$.
6218:      Hence, the formula
6219:      (\ref{eq:Bourg}) completely determines $\bh$.
6220:          Namely, let
6221:          $\bff(t,z)$ be the Gromov-Witten invariant of $\C P^{n-1}$,
6222:          and let $\Delta_{2i},i=0,\dots,n-1$, be closed forms generating
6223:      $H^*(\C P^{n-1})$, so that $\Delta_{2n-2}=\pi_*(\wt\Theta)$.
6224:       Set $\wt\Delta_0=\pi^*(\Delta_0)$
6225:       and $$\widehat\bff(t,z)=\widehat\bff^{2n-2}(t,z)=\frac{\partial\bff(t,z)}{\partial t_{2n-2}}.$$
6226: 
6227:      Then we have
6228:      \begin{equation}\label{eq:Bourg2}
6229:   \bh(t_0\wt\Delta_0+\tau\Theta,q,p) =
6230:    \frac{\tau}{2\pi } \int\limits_0^{2\pi}
6231: {\widehat\bff}( t_0\Delta_0+u, e^{-ix})  dx .
6232:  \end{equation}
6233: 
6234: 
6235: Let us consider some applications of the formula (\ref{eq:Bourg2}).
6236: 
6237: %%%%%%%
6238: 
6239: 
6240: 
6241: 
6242: \bigskip
6243: 
6244: \noindent\textsf {Contact homology of the standard contact $3$-sphere}
6245: \medskip
6246: 
6247:   Let $\pi:V=S^3\to M=\C P^1$ be the Hopf
6248:  fibration. $V$ is the pre-quantization space for $(S^2,\omega)$ with
6249:  $\int\limits_{S^2}\omega=1$. The $0$-form $\Delta_0=1$ and
6250:    the  symplectic  $2$-form $\Delta_2=\omega$
6251:    generate $H^*(M)$, the $0$-form
6252:        $\wt{\Delta}_0=\pi^*(\Delta_0)$ and the volume form $\wt{\Theta}_3$
6253:         with $\pi_*(\wt{\Theta}_3)=\Delta_2$ on $S^3$ generate $H^*(S^3)$.
6254:          Thus we have  functional variables
6255:                  \begin{equation*}
6256:            u_j(w)=\sum\limits_{k=1}^\infty \left(p_{k,j}e^{ikx}+q_{k,j}e^{-ikx}\right),
6257:  \end{equation*}
6258: associated  to the classes $\Delta_j,\; j=0,2,$ and variables $t_0$ and $\tau$
6259:  associated to $\wt{\Delta}_0$ and $\wt{\Theta}_3$. According to (\ref{eq:dim-Bott}) we have
6260: \begin{equation*}
6261: \begin{split}
6262: &\deg q_{k,0}=-2+4k,\quad \deg q_{k,2}= 4k,
6263: \quad\deg p_{k,0}=-2-4k,\\
6264: &\quad\deg p_{k,2}=- 4k,\quad
6265:  \deg t_0=-2,\quad\deg \tau=1.\\
6266: \end{split}
6267: \end{equation*}
6268:  The  potential $\bff$ for $M=\C P^1$ can be written, as it well known (see
6269:  also Section \ref{sec:computing} below),
6270: as
6271: \begin{equation}\label{eq:CP2}
6272: \bff=\frac{t_0^2t_2}{2} +e^{t_2}z,
6273: \end{equation}
6274:   and hence
6275:   \begin{equation*}
6276: \widehat\bff=\frac{t_0^2}{2} +e^{t_2}z,
6277: \end{equation*}
6278: Thus applying (\ref{eq:Bourg}) we get the following expression for the rational Hamiltonian
6279: $\bh$ for $S^3$:
6280: \begin{equation}\label{eq:3-sphere}
6281: \begin{split}
6282:  \bh&=\frac{\tau}{2\pi i}\int\limits_0^{2\pi} \left(\frac{(t_0+u_0)^2}{2}+
6283: e^{u_2-ix}\right)dx = \tau\Big(\frac{t_0^2}{2}+\sum\limits_{k\geq 1} q_{k,0}p_{k,0}\\
6284: &+\sum\limits_{t,s\geq 0}\mathop{\mathop{\sum
6285: \limits_{i_1,\dots,i_s\geq 0}}\limits_{j_1,\dots j_t\geq 0}}
6286: \limits_{\sum\limits_1^s li_l-\sum\limits_1^tmj_m=1}\frac{q_{1,2}^{i_1}\dots
6287: q_{s,2}^{i_s}p_{1,2}^{j_1}\dots p_{t,2}^{i_t}}{i_1!\dots i_s!j_1!\dots
6288: j_t!}\Big)\,.\\
6289: \end{split}
6290: \end{equation}
6291: 
6292: 
6293: Let us use (\ref{eq:3-sphere})
6294: to compute the contact homology algebra $$H^{\cont}_*(S^3,\xi_0)=H_*(\fA(S^3,J,\alpha),\partial).$$
6295: The part of $\bh$  linear in the $p$-variables has the form
6296: \begin{equation*}
6297: \tau\sum\limits_1^\infty\left( p_{k,0}q_{k,0}+ p_{k,2}h_k(q_{1,2},\dots,q_{k-1,2})\right)\;,
6298: \end{equation*}
6299: so that the differential $\partial :\fA\to\fA$
6300: is given by the formulas
6301: \begin{equation*}
6302: \partial q_{k,2}=k\tau q_{k,0},\quad \partial q_{k,0}=k\tau h_k(q_{1,2},\dots,q_{k-1,2})\;.
6303: \end{equation*}
6304: %CHANGE
6305: Here are
6306: few first polynomials $h_k$:
6307:  \begin{equation*}
6308:  \begin{split}
6309: h_1=&1,\\
6310: h_2=&q_{1,2},\\
6311: h_3=&q_{2,2}+\frac12q_{1,2}^2,\\
6312: h_4=&q_{3,2}+  q_{2,2}q_{1,2}+\frac16q_{1,2}^3.\\
6313: \end{split}
6314: \end{equation*}
6315: Notice that $\im\partial$ coincides with the ideal $I(\tau)$
6316: generated by $\tau$. Hence,  the homology algebra
6317: $H_*(\fA,\partial)$  specialized over a point $t=(t_0,0)$ is a
6318: free graded commutative algebra $ \fA_0$ generated by variables
6319: $q_{k,0},q_{k,2},\;k=1,\dots,$  and in particular it has one
6320: generator in each even dimension. On the other hand, over any
6321: point $t=(t_0,\tau)$ with  $\tau\neq 0$ the algebra
6322: $H_*(\fA,\partial)$ is isomorphic to a  proper subalgebra $\fA_1$
6323: of $ \fA_0$. It       has its first non-trivial generator $g_1=
6324: q_{1,2}-\frac12 q_{1,0}^2 $ in dimension $4$.
6325: \bigskip
6326: 
6327: \begin{remark}\label{rem:lens-space}
6328: {\rm The contact homology of the Lens space $V=L(l,1)$ which is
6329:  the pre-quantization space for $(S^2,\omega)$ with
6330:  $\int\limits_{S^2}\omega=l$ can be computed similarly.
6331:  The variables   $p_{k,0}, q_{k,0}, p_{k,2}, q_{k,2}, t_0 $ and $\tau$ have the same
6332:  $\Z_2$-grading, as in the case of $S_3$, i.e. all of them, except  $\tau$ are even.
6333:  However, the grading assigned by the Euler field to $p_{k,0}, q_{k,0}, p_{k,2}, q_{k,2}$
6334:   is fractional in this case
6335:  and given by formulas
6336: \begin{equation*}
6337: \begin{split}
6338: &\deg q_{k,0}=-2+\frac{4k}{l},\quad \deg q_{k,2}= \frac{4k}{l},
6339: \quad\deg p_{k,0}=-2-\frac{4k}{l},\\
6340: &\quad\deg q_{k,0}=- \frac{4k}{l} .\\
6341: \end{split}
6342: \end{equation*}
6343: 
6344: The formula for $\bh$ takes the form
6345: \begin{equation}\label{eq:lens-space}
6346: \bh=\frac{\tau}{2\pi i}\int\limits_0^{2\pi} \left(\frac{(t_0+u_0)^2}{2}+
6347: e^{u_2-ilx}\right)dx
6348: \end{equation}
6349: 
6350: We will not carry on here the computation of the contact homology  of
6351: the Lens space $L(l,1)$, and only note, that  as in the case of
6352: $S^3$
6353:   the homology algebra $H_*(\fA,\partial)$  specialized over a point $t=0$ is a free
6354: graded commutative algebra $ \fA_0$ generated by variables
6355: $q_{k,0},q_{k,2},\;k=1,\dots,$
6356:  and over any  point $t \neq 0$  the
6357: algebra $H_*(\fA,\partial)$ is isomorphic to a  proper subalgebra $\fA_1$ of $ \fA_0$.
6358: In particular, {\it over any point the contact homology algebra $H_*(\fA,\partial)$
6359:   has no odd elements}.}
6360: \end{remark}
6361:  %%%%%%
6362: \noindent\textsf{ Distinguishing contact structures on
6363: pre-quantizations spaces}
6364:  \medskip
6365: 
6366:  Formula (\ref{eq:Bourg}) can be used
6367:    for distinguishing contact structures on pre-quantization
6368:    spaces
6369:    of certain symplectic manifolds, which have different
6370:    Gromov-Witten invariants.
6371:                          Here is an example.
6372:                          \begin{proposition} \label{prop:Barlow}
6373:                          Let $(M_1,\omega_1)$
6374:                          and $(M_2,\omega_2,J_2)$ be two
6375:                          symplectic $4$-manifolds with  integral
6376:                          cohomology classes of their symplectic forms.
6377:                          Suppose that for compatible
6378:                          almost complex structures $J_1$ on $M_1$ and $J_2$
6379:                          on $M_2$ there are no non-constant  $J_1$-holomorphic spheres
6380:                          in
6381:                         $ M_1 $,
6382:                          while $M_2$ contains an embedded
6383:                          $J_2$-holomorphic $(-1)$-sphere $S$.
6384:                           Then  the pre-quantization spaces $(V_1,\xi_1)$ and $(V_2,\xi_2)$
6385: are not contactomorphic.
6386: \footnote{
6387: Yongbin Ruan proved in \cite{Ruan-virtual}
6388: that under the assumptions of Proposition \ref{prop:Barlow}   the
6389: symplectic manifolds $(M_1,\omega_1)\times \C P^1$ and
6390:  $(M_2,\omega_2)\times \C P^1$ are not symplectomorphic
6391:  (and  not even deformationally equivalent), despite   the fact
6392:  that    for a certain choice of $M_1$ and $M_2$
6393:  (e.g. $M_1$ is the Barlow surface and $M_2=\C P^2\#8\overline{\C P^2}$), and for
6394:  appropriate symplectic forms $\omega_1$ and $\omega_2$
6395:  the underlying manifolds $V_1$ and $V_2$ are diffeomorphic. }
6396: \end{proposition}
6397: \begin{remark}
6398: {\rm Even when the manifolds $M_1$ and $M_2$ are homeomorphic,
6399: the prequantization spaces $V_1$ and $V_2$ are not diffeomorphic
6400: (even not homotopy equivalent!)  for most choices   of symplectic forms
6401: $\omega_1$ and $\omega_2$, and hence the statement of the theorem is trivial in these
6402: cases. However, one can show that
6403: for homeomorphic $M_1$ and $M_2$ the
6404: symplectic forms can always be deformed in the class of symplectic forms compatible with
6405: the chosen almost complex structures $J_1$ and $J_2$, in order
6406:   to make $V_1$ and $V_2$ diffeomorphic.}
6407: \end{remark}
6408: To prove Proposition \ref{prop:Barlow}   we will ahow that
6409:    the ``classical"   contact homology algebras
6410:  $H^{\cont}_*(V_1,\xi_1) $ and $H^{\cont}_*(V_2,\xi_2) $ are not isomorphic.
6411: 
6412: Let  $S_0=S,S_1,\dots, S_m$ be the exceptional $J_2$-holomorphic
6413: spheres in $M_2$. Then the cohomology classes $ S^*_0,\dots,
6414: S^*_m,[\omega]$ are linearly independent, where  we denote by
6415: $S^*_i$ the cohomology class Poincar\'e-dual to $[S_i]\in
6416: H_2(M_2)$. Hence there exists a class $X\in H^2(M_2)$, such that
6417: \begin{equation}
6418: XS^*_0=1, \;X[\omega]=0,\;\;\hbox{and}\;\;XS^*_i=0\;\;\hbox{for}\;\; 1\geq i\geq m.
6419: \end{equation}
6420: Then the  potential $\bff_{M_2,J_2}(tX)$
6421: coincides with $$\bff_{S,J_2|_S}(tX|_{S})=e^tz.$$
6422:  Let us choose a basis of closed forms $\Delta_i, i=0,\dots,a$,
6423: generating
6424:   generating $H^*(M_2)$, so that one of the forms, say $\Delta_1$
6425:   represents the class $X$. The form $\Delta_1$ then
6426:      lifts to a  form $\wt \Delta_1$ such that
6427: $\pi_*\wt \Delta_1=\Delta_1$.
6428:  According to the formula
6429: (\ref{eq:Bourg2}) we have
6430: \begin{equation}
6431: \bh_V( \tau_1\wt\Delta_1,u)= \frac{\tau_1}{2\pi } \int\limits_0^{2\pi}
6432: e^{u_1-ilx} dx ,
6433: \end{equation}
6434: where $u=\sum\limits_0^au_j\Delta_j,\,u_j=\sum\limits_1^\infty
6435:  (p_{k,j}e^{ikx}+q_{k,j}e^{-ikx}),\;l=\int\limits_{S}\omega.$
6436: 
6437:  Hence, the contact homology algebra $H^{\cont}_*(M_2,\xi_2)$,
6438:  specialized     at the point $\tau\wt\Delta$ for $\tau\neq 0$
6439:  is isomorphic to the contact homology algebra of the  standard
6440:  contact
6441:  Lens space $L(l,1)=\pi^{-1}(S)\subset V$, specialized    at the volume
6442: form $\tau\wt\Delta|_{L(l,1)}$. It follows then from  Remark \ref{rem:lens-space}
6443:    that $H^{\cont}_*(V_2,\xi_2)$,
6444:  specialized     at any point  has
6445:  no odd elements.   On the other hand, for any $2$-dimensional
6446:  cohomology class $t\in H_2(M_1)$ we have
6447:  $\bff_{M_1,J_1}(t)=0$, and hence for any $3$-form
6448:  $\wt\Delta$ on $V_1$,
6449:  the formula (\ref{eq:Bourg2}) implies that the Hamiltonian
6450:  $\bh_{V_1} (\wt\Delta, u )$ vanishes as well, and  therefore
6451:  the contact homology algebra  $H^{\cont}_*(M_1,\xi_1)$,
6452:  specialized     at the point $\tau\wt\Delta,\tau\neq 0$
6453: is a free graded commutative algebra generated by the odd variable
6454: $\tau$, and even variables $p'_{k,j},q'_{k,j }$, which correspond
6455: to even dimensional generators of $H^*(M_1)$. Hence the contact
6456: manifolds $(M_1,\xi_1)$ and $(M_2,\xi_2)$ are not isomorphic. \qed
6457: 
6458: 
6459: 
6460:   \bigskip
6461: 
6462:  \noindent\textsf{ Subcritical symplectic manifolds}
6463:   \medskip
6464: 
6465:    The content of this example is a result of our discussion
6466:    with P. Biran and K. Cieliebak.
6467: 
6468:   In \cite{Donaldson} S. Donaldson  generalized  the Kodaira embedding theorem by
6469:   proving that for any closed  symplectic manifold $(W,\omega)$
6470:    with an integral cohomology class of the symplectic form there
6471:    exists an integer $l>0$ such that the homology class dual to $l[\omega]$
6472:    can be represented by an embedded symplectic hypersurface $W_0$.
6473:   In fact, S. Donaldson proved a stronger result, which  together
6474:   with
6475:    an improvement by  Biran-Cieliebak   asserts that for a sufficiently large $l$
6476:    the hypersurface $W_0$ can be chosen in such a way that in the complement
6477:    $W\setminus W_0$ there exists a vector field $X$ with the following properties:
6478:    \begin{description}
6479:    \item{} $L_X\omega=-\omega$, where $L_X$ denotes the Lie derivative
6480:    along $X$; in other words, $X$ is conformally symplectic and   contracting;
6481:     \item{}  $X$ is forward integrable;
6482:     \item{} $X$ is gradient-like for a Morse function $\phi:W\setminus W_0\to \R_+$,
6483:     which coincides with $-\log\dist^2$ near $W$, where $\dist(x)$ is the     distance
6484:     function from a point $x$ to $W_0$ with respect to some Riemannian metric.
6485:     \end{description}
6486: 
6487:        The vector field $X$ retracts $W\setminus W_0$ to the Morse complex $K$
6488:        of the function $\phi$,
6489:        which is automatically isotropic for the symplectic form $\omega$
6490:        (see  \cite{Eliash-Gromov-convex}), and, in particular, $\dim K\leq n$.
6491:          Biran-Cieliebak call the pair $(W,W_0)$ {\it subcritical}
6492:            if $\dim K<n$.
6493:       They constructed  in \cite{Biran-Cieliebak}
6494:     several interesting  examples of subcritical pairs, and conjectured
6495:     that {\it if $(W,W_0)$ is subcritical, then $l=1$}.
6496:     We sketch below  the proof of this conjecture.
6497: 
6498: First, let us observe that the contact structure $\xi$, defined by
6499: the contact form $\a=
6500:  X\hook\omega$ on the boundary  $V$ of a  small tubular neighborhood of
6501: $W_0$, is equivalent to the contact structure which is defined on $V$ as the
6502: pre-quantization space   of the symplectic manifold
6503: $(W_0,l\omega)$. On the other hand, the condition, that the pair
6504: $(W,W_0)$ is sub-critical implies that the contact manifold $(V,\xi)$
6505: is itself  subcritical in the sense of Examlpe \ref{ex:Floer}.4 above,
6506: i.e. it is  isomorphic to the strictly pseudo-convex
6507: boundary  of a sub-critical Stein (or Weinstein) manifold with its canonical complex structure.
6508: Let us recall (see \ref{rem:grading} above) that all SFT-objects, in particular Floer contact
6509: homology $HC_*(V,\xi)$ and the contact homology algebra $H^{\cont}_*(V,\xi)$ are
6510: graded by elements of $H_1(V)$.
6511: Using  arguments  as in  the theorem of Mei-Lin Yau (see \ref{prop:MLYau} above) one can show
6512:  that all  non-trivial elements
6513:  in  the
6514: contact homology algebra $H^{\cont}_*(V,\xi)$ of a subcritical contact manifold
6515: $(V,\xi)$ may correspond only to $0\in H_1(V)$.
6516:   On the other hand,
6517:  it follows from Proposition \ref{prop:even-dim} above that $H^{\cont}_*(V,\xi)$ specialized at $0\in H^*(V)$ has
6518:  non-trivial elements which correspond to the homology class of       the fiber in
6519:  $H_1(V)$. Therefore, $l=1$.
6520: 
6521: \subsubsection{Computing  rational Gromov-Witten
6522: invariants of $\C P^n$}\label{sec:computing}
6523: We will  show  in this section how SFT  can be used for computing      rational
6524:  Gromov-Witten invariants of $\C P^n$. Our method differs from
6525:  traditional ways (see \cite{Kontsevich-CP2,Fulton-Pand,Graber-Pand,Vakil,Ruan-Tian})   for this computation.
6526:   We will be simultaneously computing the rational potential of $\C^n$ and the rational
6527: Gromov-Witten invariant of $\C P^n$ by a recursion
6528: using Theorem \ref{thm:RSFT-master}
6529: 
6530: Let us choose basic forms in $\C^n$ as in the previous section,
6531: i.e $\Delta =1$, and $\Theta$ is a volume form with  compact
6532: support in $C^n\setminus 0=S^{2n-1}\times(0,\infty)$
6533:   with $\int\limits_{\C^n}\Theta =1$. We denote by $\delta$ the restriction of
6534:   $\Delta$ to $S^{2n-1}$.
6535: We also assume that $\Theta$ splits into a product $\hat\theta\wedge\rho$, where $\hat\theta$
6536: is pull-back of a unit volume form $\theta$ on $S^{2n-1}$, and $\rho$ is a
6537: compactly supported form in $(0,\infty)$.
6538: Set
6539: \begin{equation}
6540: \begin{split}
6541: \bh^1(t_0,q,p))&=\frac{\partial \bh}{\partial
6542: \tau}(t_0\delta+\tau\theta,q,p)|_{\tau=0}\\ &=\frac{1}{2\pi
6543: }\int\limits_0^{2\pi}\widehat\bff_{\C P^{n-1}}(t_0+u_0(x),u_2,\dots,
6544: u_{2n-2},e^{-ix})dx,\\
6545: \end{split}
6546: \end{equation}
6547:  where  $$\widehat\bff_{\C P^{n-1}}(t_0,\dots,t_{2n-2},z)=
6548:  \frac{\partial\bff_{\C P^{n-1}}}{\partial t_{2n-2}}(t_0,\dots,t_{2n-2},z),$$
6549:  $\bff_{\C P^{n-1}}(t_0,\dots,t_{2n-2},z)$ is the rational
6550: Gromov-Witten invariant  of $\C P^{n-1}$, and
6551: $$u_{2j}(x)=\sum\limits_1^\infty
6552: p_{k,2j}\,e^{ikx}+q_{k,2j}\,e^{-ikx}.$$ Then the equation
6553: (\ref{eq:RSFT-master}), which determines $
6554: \bff(t_0,t_{2n},p)=\bff_{\C^n}(t_0\Delta+t_{2n}\Theta,p)$ takes
6555: the form
6556: \begin{equation}\label{eq:Cn}
6557: %\begin{split}
6558: \frac{\partial\bff}{\partial t_{2n}}(t_0,t_{2n},p) =\frac{1}{2\pi
6559: }\int\limits_0^{2\pi} \widehat\bff_{\C P^{n-1}}(t_0+u_0(x),u_2,\dots,
6560: u_{2n-2},e^{-ix})dx\big|_{L_\bff},
6561: %\end{split}
6562: \end{equation}
6563: where $$L_{\bff}=\{q_{k,2j}=k\frac{\partial \bff}{\partial
6564: p_{k,2n-2j-2}}(t_0,t_{2n},p)\}.$$
6565:  Together with the initial data
6566: \begin{equation}
6567: \bff(t_0,0,p)=
6568: \begin{cases}
6569: p_{1,0},&\hbox{if}\;\;n=1;\\
6570: 0,&\hbox {otherwise} \\
6571: \end{cases}
6572: \end{equation}
6573: the equation (\ref{eq:Cn}) provides a recursive procedure for
6574: computing coefficients $f_j(t_0,p)$ of the expansion
6575: $$\bff(t_0,t_{2n},p)=\sum\limits_0^\infty f_j(t_0,p)t_{2n}^j.$$
6576: 
6577: For instance for $n=1$ we have (see Example \ref{ex:circle})
6578: $\bh^1=\frac{t_0^2}{2}+\sum\limits_1^\infty p_kq_k$, where we
6579: write
6580: $p_k,q_k$ instead of $p_{k,0},q_{k,0}$, and hence the equation
6581: (\ref{eq:Cn}) takes the form
6582: \begin{equation}
6583: \frac{\partial\bff}{\partial t_{2 }}(t_0,t_{2 },p)=\frac{t_0^2}{2}
6584: + \sum\limits_0^\infty k p_k\frac{\partial\bff}{\partial
6585: p_k}(t_0,t_{2 },p)
6586: \end{equation}
6587: with the initial data $\bff(t_0,0,p)=p_1$. This linear first order
6588: PDE is straightforward to solve, and we get $$\bff(t_0,t_{2
6589: },p)=\frac{t_2t_0^2}{2}+e^{t_2}p_1.$$
6590: \bigskip
6591: 
6592: For $n=2$ the Hamiltonian $\bh$ is given by the formula
6593: (\ref{eq:3-sphere}), and we have
6594: $$\bh(t_0,\tau,p)=\tau\bh^1(t_0,p).$$ Thus the equation for the
6595: potential of $\C^2$ has the form
6596: \begin{equation}\label{eq:C2}
6597: \begin{split}
6598: &\frac{\partial\bff}{\partial t_{4}}(t_0,t_{4},p)
6599:  =\frac{t_0^2}{2}+\sum\limits_{k\geq 1}k
6600: \frac{\partial\bff}{\partial p_{k,2}}p_{k,0}\\
6601: &+\sum\limits_{t,s\geq 0}\mathop{\mathop{\sum
6602: \limits_{i_1,\dots,i_s\geq 0}}\limits_{j_1,\dots j_t\geq 0}}
6603: \limits_{\sum\limits_1^s li_l-\sum\limits_1^tmj_m=1} \frac{ \left(
6604: \frac{\partial\bff} {\partial p_{1,0}}\right)^{i_1}\dots
6605:  \left(s \frac{\partial\bff}{\partial p_{s,0}}\right)^{i_s}p_{1,2}^{j_1}\dots
6606:  p_{t,2}^{i_t}}{i_1!\dots i_s!j_1!\dots
6607: j_t!};\\
6608: &\bff(t_0,0,p)=0\,.\\
6609: \end{split}
6610: \end{equation}
6611: Hence, we get
6612: \begin{equation}
6613: \begin{split}
6614: \bff(t_0,t_4,p)=&t_4(\frac{t_0^2}{2}+p_{1,2})+\frac{t_4^2}{2!}p_{1,0}+
6615: \frac{t_4^3}{3!}(p_{2,2}+\frac12 p_{1,2}^2)\\
6616: &+\frac{t_4^4}{4!}(2p_{2,0}+p_{1,2}p_{1,0})+\dots\;.
6617: \end{split}
6618: \end{equation}
6619: 
6620: To compute $\bff_{\C^n}$ for $n>2$ we need to know $\bff_{\C
6621: P^{n-1}}$. So to complete the recursion we will explain now how to
6622: express the rational Gromov-Witten invariant $\bff_{ \C P^{n-1}}$
6623: through $\bff_{\C^n}$.
6624: 
6625: \bigskip
6626: First of all we split, as it is described in Example
6627: \ref{ex:section}
6628:  above, $\C P^n$ along the boundary of a tubular neighborhood
6629: of $\C P^{n-1}\subset \C P^n$ into two completed symplectic
6630: cobordism $W_1=\C^n$ and $W_2=\C P^{n }\setminus {x}$, where we
6631: introduce on $W_2$ a complex structure of the holomorphic line
6632: bundle over $\C P^{n-1}$ determined by the hyperplane section $\C
6633: P^{n-2}\subset\C P^{n-1}$. We will denote by $\bff_1$ and $\bff_2$
6634: the potentials for $W_1$ and $W_2$, respectively.
6635: 
6636: Let $\Delta_0,\dots,\Delta_{2 n-2}$ be closed forms representing
6637: the standard basis of $H^*(\C P^{n-1})$. We will keep the same
6638: notation for the pull-backs of these forms to $W_2$. Let
6639: $\Delta_{2n}$ be  a closed $2n$-form with a compact support, which
6640: generates $$\Ker\big(H^*_{\comp}(W_2)\to H^*(W_2)\big).$$
6641:  We are interested
6642: in the potential
6643: $\bff_2(t_0,\dots,t_{2n},q)=\bff_2(\sum\limits_{i=1}^{n }
6644: t_{2i}\Delta_{2i},q)$. First of all notice that by dimensional
6645: reasons the moduli spaces of holomorphic curves which project  to
6646: non-constant curves in $\C P^{n-1}$ do not contribute to the
6647: potential $$\bff_2(t_0,\dots,t_{2n-2},0,q),$$ and hence we have
6648: \begin{equation}
6649: \bff_2(\sum\limits_{i=1}^{n-1} t_{2i}\Delta_{2i},q)= z\sum
6650: \limits_{i=0}^{n-1}q_{1,2i} \mathop{\sum\limits_{(s_1,\dots,s_{n-1})}}\limits_{  \sum
6651: s_j(j-1)=n-i-1}\prod\limits_{j=1}^{n-1}\frac{t_{2j}^{s_j}}{s_j!}\;.
6652: \end{equation}
6653: In particular, for $n=2$ we get
6654: \begin{equation*}
6655: \bff_2(t_0\Delta_0+t_2\Delta_2,q)=ze^t_2q_{1,2}
6656: \end{equation*}
6657:  One can recover $\bff_2(t_0,\dots,t_{2n},q)$ for $t_{2n}\neq 0$
6658: using the equation (\ref{eq:RSFT-master}), as we did it above for
6659: $W_1=\C^n$. However, for the purpose of our  computation of
6660: Gromov-Witten invariant $\bff_{\C P^n}$ this is not necessary, as
6661: we can proceed as follows.
6662: 
6663: Notice that the above chosen  forms
6664: $\Delta_2, \Delta_{2n-2}$ extend to $\ CP^n$. On the other hand, we
6665: will choose a  volume form $\Delta_{2n}$ on $\C P^n$ to be
6666: supported in the affine part, so that the restriction $\Delta_{2n}|_{\C^n}$
6667: coincides with the form $\Theta$ introduced above. Then Theorem
6668: (\ref{thm:RSFT-composition}) implies that
6669: \begin{equation}\label{eq:CPn}
6670: \begin{split}
6671: &\bff_{\C P^n}(\sum\limits_{i=1}^{n } t_{2i}\Delta_{2i})=\\
6672:  &\big(\bff_1(t_0,t_{2n},p)+
6673:  \bff_2(t_0,\dots,t_{2n-2},0,q)-\sum\limits_{i+j=n-1}
6674: \sum\limits_1^\infty \frac1{k} p_{k,2i}q_{k,2j}\big)\big|_L\;\;,\\
6675: \end{split}
6676: \end{equation}
6677: where
6678: \begin{equation}\label{eq:Lag-CPn}
6679: L=
6680: \begin{cases}
6681: p_{1,2i}=& z  \mathop{\sum\limits_{(s_1,\dots,s_{n-1})}}\limits_{
6682: \sum
6683: s_j(j-1)=i}\prod\limits_{j=1}^{n-1}\frac{t_{2j}^{s_j}}{s_j!}\;;\\
6684: p_{k,2i}=&0,\;\;\hbox{if}\;\; k>1\;; \\
6685: q_{k,2i}=&k\frac{\partial\bff_1}{\partial
6686: p_{k,2(n-i-1)}}(t_0,t_{2n}, p)\,.\\
6687: \end{cases}
6688: \end{equation}
6689: 
6690: Plugging expressions from (\ref{eq:Lag-CPn}) into equation
6691: (\ref{eq:CPn}) we get
6692: \begin{equation}\label{eq:CPn-Cn}
6693: \bff_{\C
6694: P^n}(t_0,\dots,t_{2n})=\bff_{\C^n}(t_0,t_{2n},p)\big|_{L_1}\;,
6695: \end{equation}
6696: where
6697: \begin{equation*}
6698: L_1=
6699: \begin{cases} p_{1,2i}&= z  \mathop{\sum\limits_{(s_1,\dots,s_{n-1})}}\limits_{
6700: \sum
6701: s_j(j-1)=i}\prod\limits_{j=1}^{n-1}\frac{t_{2j}^{s_j}}{s_j!}\;;\\
6702: p_{k,2i}&=0\;\;\hbox{if}\;\;
6703: k>1\,.\\
6704: \end{cases}
6705: \end{equation*}
6706: Indeed, two last terms in the formula (\ref{eq:CPn}) cancel each
6707: other (as it always happens when $\bff_2$ is linear with respect
6708: to $q$-variables).
6709:  For instance, for $n=1$ we get
6710: \begin{equation}
6711: \bff_1(t_0,t_2,p)=\big(\frac{t_0^2t_2
6712: }2+e^{t_2}p_1\big)\big|_{p_1=z}=\frac{t_0^2t_2}2+e^{t_2}z\,.
6713: \end{equation}
6714: 
6715: 
6716: For $n=2$ we have
6717: \begin{equation}
6718: L_1=
6719: \begin{cases}
6720:  p_{1,0}&= ze^{t_2}\,;\\
6721: p_{k,i}&= 0,\quad\hbox{ for all other}\quad k,i\,,\\
6722: \end{cases}
6723: \end{equation}
6724: and hence
6725: \begin{equation}\label{eq:CP2-final}
6726: \bff_{\C P^2}(t_0,t_2,t_4)=\bff_{\C^n}(t_0,t_4,
6727:  ze^{t_2},0,\dots)\,.
6728: \end{equation}
6729: 
6730: 
6731: \begin{remark} {\rm The method which we used  above
6732:  for computing of the rational
6733: potential of $\C P^n$, when  applied to an arbitrary
6734: symplectic manifold $W$,   allows us to express
6735:   $\bff_W$ through the potential of the affine part $W\setminus
6736:   M$.
6737:   The latter computation seems tractable when  the Weinstein manifold
6738:   $W\setminus M$ is
6739:   subcritical (see Section \ref{sec:splitting} above), i.e. when
6740:   its isotropic skeleton does not have Lagrangian cells.
6741:   On the other hand, when Lagrangian cells are present
6742:   this problem is related to central questions of Symplectic
6743:   topology.}
6744:   \end{remark}
6745: 
6746: 
6747: 
6748: \subsubsection{Satellites}\label{sec:satellites}
6749: Let $(V,\xi=\{\a=0\})$ be a  contact manifold, $(W
6750: = V\times \R, d(e^t\a))$  its symplectization, and   $J$   a
6751: compatible translation-invariant almost complex structure on $W$.
6752:    In this section  we will show that the
6753:  homological Poisson super-algebra $H_*(\fP ,  d^{\bh}  )$ comes equipped with some
6754: additional structures, rather unfamiliar in abstract Poisson
6755: geometry. Namely, the counting  of genus $g$ curves with a fixed
6756: complex structure and with a fixed configuration of $n$ points
6757: gives rise to an odd $n$-linear totally symmetric poly-form
6758: $\bh^{g,n}$ on the Poisson super-space $\bV$ underlying $\fP $. The
6759: poly-form descends well to the homology and thus yields another
6760: invariant of the contact structure which we call the genus-$g$
6761: $n$-point {\em satellite} of the Poisson structure.
6762: 
6763: Let us denote by
6764: $\overline{\cM_{g,m}(V)/\R}$   the compactified moduli space of stable connected
6765: $J$-holomorphic curves in $W$ which are characterized by the
6766: arithmetical genus $g$ and by the total number $m$ of punctures and
6767: marked points numbered somehow by the indices $1,...,m$ (see Section \ref{sec:compact} above). We
6768: emphasize that the moduli space in question contains equivalence
6769: classes of all such curves, and in particular, may have infinitely
6770: many connected components corresponding to different homotopy
6771: types of curves in $W$ and different numbering of the $m$
6772: markings. Let $\DM_{g,n}$ be the Deligne-Mumford compactification
6773: of the moduli space of genus $g$ Riemann surfaces with $n$ marked
6774: points. For any $g,n$ with $2g-2+n >0$ and $l\geq 0$ there is a
6775: natural {\em contraction map} $\ct :\overline{\cM_{g,n+l}(V)/\R}\to \DM_{g,n}$
6776: defined by forgetting the map to $W$ and the last $l$ markings
6777: followed by the contraction of those components of the curve which
6778: have become unstable.  Given a differential form
6779: $\tau$ on $\DM_{g,n}$ we will denote by $\ct^*\gt$ its  pull-back
6780:  to $\overline{\cM_{g,m}(V)/\R}$.
6781: 
6782: Let $u =(p,q,t)$ denote a point in $\bV$, that is $p,q$ and $t$ are
6783: (closed) differential forms on $\cP^{-},\cP^{+}$ and $V$
6784: respectively. We will denote $\ev_i^*{u}, \ i=1,...,m$, the
6785: pull-back by the evaluation map $$\ev_i: \overline{\cM_{g,m}(V)/\R}\to
6786: (\cP^{-}\cup \cP^{+}\cup V)$$ at the $i$-th marking. Let us emphasize the point that we
6787: are
6788: treating here  the marked points and punctures on equal footing.
6789: 
6790: Let $\gd u\in \bV$ be a tangent vector to $\bV$ at a point $u\in \bV$.
6791: We introduce the formal function
6792:  \begin{equation}
6793:  %\begin{split}
6794: \bh^{g,n}_{\gt}:=\frac{1}{n!}
6795: \sum_{l=0}^{\infty}\frac{1}{l!}\int\limits_{\overline{\cM_{g,n+l}(V)/\R }}\
6796: \ct^*{\gt}\w \ev_1^*\gd u\w...\w\ev^*_n\gd
6797: u\w\ev_{n+1}^*u\w...\w\ev_{n+l}^*u .
6798: %\end{split}
6799: \end{equation}
6800:  It is a super-symmetric  $n$-linear
6801: form in $\gd u$ with coefficients depending on the application
6802: point $u$.
6803: 
6804: Let $d^{\bh}(f)$ denote the Lie derivative of a tensor field $f$
6805:  along the odd Hamiltonian
6806: vector field $d^{\bh}$ on $\bV$ with the Hamilton function $\bh$.
6807: 
6808: \begin{proposition}
6809: Let $\gt $ be a top degree form on $\DM_{g,n}$. Then
6810: $d^{\bh}(\bh_{\gt}^{g,n})=0$. If the top degree form $\gt=d\a$ is exact
6811: then $\bh^{\gt}_{g,n}=d^{\bh}( \bh^{\a}_{g,n})$. In particular, the
6812: tensor field $\bh^{g,n}_{\gt}$ descends to the homology    algebra
6813: $H_*(\fP, \p)$ into a satellite which depends only on the total
6814: volume of $\gt$.
6815: \end{proposition}
6816: 
6817: This follows from the Stokes formula applied to
6818: $\bh^{d\gt}_{g,n}=0$ and respectively  to $\bh^{d\a}_{g,n}$.
6819: Codimension $1$ boundary strata of the moduli space $\cM_{g,m}(V)/\R$
6820: are formed by  stable curves of height $2$. Most of the strata do not contribute to
6821: the Stokes formula since they are mapped by the contraction map to
6822: complex codimension $1$ strata of the Deligne-Mumford space
6823: $\DM_{g,n}$, where $\gt$ and   $\a$ restrict to $0$ for
6824: dimensional reasons. Exceptions to this rule occur only if one of
6825: the two curves which form the stable curve is to be contracted. It is therefore a
6826: sphere with  glued to the other level of the stable curve along at precisely
6827: one end, and which have at most one marked points or ends with the index $\leq n$,
6828:  and with any
6829: number of ends or marked points with indices $>n$. Contributions
6830: of such curves to the Stokes formula is expressed bi-linearly via
6831: the $1$-st or $2$-nd derivatives of the Hamilton function $\bh$
6832: and the satellite. It is easy to see that the whole expression is
6833: interpreted correctly as the  Lie derivative of the tensor field
6834: $\bh^{\gt}_{g,n}$ along the Hamiltonian vector field $d^{\bh}$. \qed
6835: 
6836: 
6837: \medskip
6838: 
6839: We will assume further on that $\gt $ is normalized to the total
6840: volume $1$ and will often drop it from the notation for the
6841: satellite $\bh^{g,n}$.
6842: 
6843: \medskip
6844: 
6845: Let us consider now a directed symplectic  cobordism
6846: $W=\ora{V_-V_+}$ between two contact boundaries $V_{\pm}$. Then we
6847: have the Hamilton function $\widehat\bh
6848: =\bh^{+}-\bh^{-}$ and the satellites
6849: $\widehat\bh^{g,n}=\big(\bh^{g,n}\big)^+-
6850: \big(\bh^{g,n}\big)^-$ defined  as elements of
6851: the algebra $\widehat \fL$, which in the case when the cobordism
6852: is a concordance just equal to the tensor product of the Poisson
6853: algebras $\fP_{\pm}$. Also, we have the potential $\bff
6854: (p_{-},q_{+},t) $ counting rational $J$-holomorphic curves in $W$
6855: which defines a Lagrangian correspondence between $\fP^{\pm}$
6856: invariant under the vector field $ d^{\widehat{\bh}}$ with the Hamilton function
6857: $\widehat\bh$. Finally, using the moduli spaces $\cM_{g,m}(W)$ of
6858: $J$-holomorphic curves in the cobordism, we can introduce the
6859: satellites $\bff^{g,n}_{\gt}$ as symmetric $n$-forms on the space
6860: $(p_{-},q_{+},t)$-space parameterizing the Lagrangian
6861: correspondence. Then the arguments similar to the above proof of
6862: the proposition, but applied this time to $\bff^{g,n}_{d\gt}=0$,
6863: show that {\em the restriction of $\widehat\bh_{\gt}^{g,n}$ to the
6864: Lagrangian correspondence defined by $\bff$ coincides with the Lie
6865: derivative of $\bff^{g,n}_{\gt}$ along the vector field $d^{\widehat\bh}$
6866: restricted to the Lagrangian correspondence}
6867: (comp. Theorem \ref{thm:RSFT-cobordism} above).
6868: In this sense the
6869: Lagrangian correspondences defined by symplectic cobordisms
6870: preserve the satellite structures defined by $\big(\bh^{g,n}\big)^{\pm}$ on
6871: the homology $H_*(\fP_{\pm},d^{\bh^{\pm}})$. In particular, {\em the
6872: satellite structures of a contact manifold $V$ on the homology
6873: $H_*(\fP,d^{\bh})$ depend only on the contact structure.}
6874: 
6875: \medskip
6876: 
6877: The following discussion is the first steps in the study of the
6878: geometric structure defined by the satellites.
6879: 
6880: Let $\bh^{g,n}_{\a_1,...,\a_n}$ denote components of the satellite
6881: tensors on $\fP$. Using the Poisson tensor $\pi^{\mu\nu}$ we can
6882: couple two satellites with respect to some indices:
6883: $$\bh^{g',n'+1}_{...\mu}\pi^{\mu\nu}\bh^{g'',n''+1}_{\nu ...}.$$
6884: Similarly, we can couple two indices in $\bh^{g-1,n+2}$ with two
6885: indices in the $2$-nd differential $\gd^2\bh $ of the Hamilton
6886: function $\bh$.
6887: 
6888: \begin{proposition} If $g=g'+g''>0$ then the
6889: coupling of $\bh^{g',n'}$ and $\bh^{g'',n''}$ is a Lie derivative
6890: along $\p$ and thus vanishes in the homology $H_*(\fP ,\p )$.
6891: Similarly, the coupling of $\bh^{g-1,n+2}$ with $\gd^2\bh$
6892: vanishes in the homology $H_*(\fP,\p)$. \end{proposition}
6893: 
6894: The proof is based on some famous but non-trivial property of the
6895: spaces $\DM_{g,n}$ with $g>0$ to have complex codimension one
6896: strata homologically independent. Such strata correspond to
6897: different ways of cutting a $(g,n)$-surface along one circle and
6898: can be identified either with $\DM_{g',n'+1}\times
6899: \DM_{g'',n''+1}$ where $g'+g''=g, n'+n''=n$ or with
6900: $\DM_{g-1,n+2}$. The independence property implies that a volume
6901: form $\gt$ on the stratum, say  $\gt'\otimes\gt''$ in the first
6902: case, can be obtained as the restriction of a closed codegree two
6903: form $\go$ on $\DM_{g,n}$ which have exact (or even zero, for
6904: suitable choices of $\gt$) restrictions to all other
6905: codimension-$1$ strata in $\DM_{g,n}$. Applying the Stokes formula
6906: to $0=\bh^{g,n}_{d\go}$ we find that the coupling of
6907: $\bh^{g',n'+1}_{\gt'}$ and $\bh^{\gt'',n''+1}_{\gt''}$ (or --- in
6908: the second case --- of $\bh^{g-1,n+2}_{\gt}$ and $\gd^2\bh$)
6909: equals $d^{\bh}( \bh^{g,n}_{\go})$.
6910: 
6911: \medskip
6912: 
6913: \begin{remark} {\em
6914: To the contrary, coupling $\bh^{0,3}$ with itself via one index is
6915: not, in general, a $d^{\bh}$-derivative, but instead the following
6916: triple sum is:
6917: % I am afraid, the signs might be wrong
6918: \begin{equation*}
6919: \begin{split}
6920: & \bh^{0,3}_{\a\gb\mu}\pi^{\mu\nu}\bh^{0,3}_{\nu\g\d}+
6921:    (-1)^{(\deg{\a}+\deg{\gb})\deg{\g}}
6922:    \bh^{0,3}_{\g\a\mu}\pi^{\mu\nu}\bh^{0,3}_{\nu\gb\d}+\\
6923:   & (-1)^{\deg{\a}(\deg{\gb}+\deg{\g})}
6924:    \bh^{0,3}_{\gb\g\mu}\pi^{\mu\nu}\bh^{0,3}_{\nu\a\d} \equiv 0
6925:   \, .\\
6926:    \end{split}
6927:    \end{equation*}
6928: This follows from the property of the $3$ boundary strata in
6929: $\DM_{0,4}$ to represent the same homology class (use the Stokes
6930: formula for $\go =1$). In fact $\bh^{0,3}$ coincides with the
6931: $3$-rd differential $\gd^3\bh /6$ of the Hamilton function, and
6932: the above Jacobi-like identity can be derived by $4$
6933: differentiations of $\{ \bh ,\bh \}=0$ in $\a,\gb,\g,\d$. One can
6934: interpret the integrability property $(d^{\bh})^2=0$ of the odd vector
6935: field $d^{\bh}$ on $\bV$ as a homotopy Lie super-algebra structure on
6936: $\Pi \bV^*$, the dual space with changed parity. The identity in
6937: question corresponds to the Jacobi identity for the remnant Lie
6938: super-algebra structure in homology. }
6939:  \end{remark}
6940: 
6941: 
6942: It is sometimes convenient to extend the definition of genus $0$
6943: satellites to unstable values of $n$ by $\bh^{0,n}=\gd^n\bh /n!$
6944: for $n=0,1,2$. Also, one can define the function $\bh^{1,0}$ at
6945: least locally as a potential for $\bh^{1,1}$, using the following
6946: 
6947: \begin{proposition} The differential $1$-form $\bh^{1,1}$ is closed.
6948: \end{proposition}
6949: 
6950: Indeed, the partial derivatives $\gd_{\mu}\bh^{1,1}_{\nu}$ and
6951: $\gd_{\nu}\bh^{1,1}_{\mu}$ are identified with the satellites
6952: $(\bh^{1,2}_{\go})_{\mu\nu}$ corresponding to the $2$-form $\go $
6953: on $\DM_{1,2}$ pulled-back from $\DM_{1,1}$ by forgetting the
6954: $1$-st and respectively the $2$-nd marked point. But the two maps
6955: $\DM_{1,2}\to \DM_{1,1}$ coincide.
6956: 
6957: \medskip
6958: 
6959: It would be interesting to study other general properties of
6960: satellites which may depend on more sophisticated geometry of
6961: Deligne-Mumford compactifications. For instance, what can be said
6962: about Poisson brackets among the functions $\bh^{g,0}$?
6963: 
6964: \medskip
6965: 
6966: We complete the section with the computation of the satellites in the
6967: example $V=S^1$. Let $t=t_01+t_1d\phi $ denote the general
6968: harmonic form on $S^1$, $\gd t=\gt_0 1+\gd t_1\d\phi $,
6969: $u(x)=t_0+\sum p_ke^{ikx}+ q_ke^{-ikx}$, $\gd u = \gd t_0+\sum \gd
6970: p_ke^{ikx}+\gd q_k e^{-ikx}$.
6971: 
6972: \begin{proposition}
6973: For $2g-2+n \geq 0$ we have
6974: \[ \bh^{g,n+1}\ =\ \frac{\gd t_1}{2\pi\ n!}\int\limits_0^{2\pi} \ (u_{xx})^g \
6975: (\gd u)^n \ dx .\]
6976: \end{proposition}
6977: 
6978: Let us begin with the remark that the formula does not (and
6979: cannot) contain $t=t_0+t_1\phi$ because $\deg t < 2$, and
6980: therefore pushing forward from $\cM_{g,m+1}(V)/\R \to
6981: \cM_{g,m}(V)/\R$ by forgetting the corresponding marked point
6982: would send $t$ to $0$. Exceptions to this rule could occur only if
6983: $\cM_{g,m}(V)$ were ill-defined, that is only in the case of
6984: constant maps with "unstable" indices, $2g-2+m\leq 0$, which has
6985: no effect on the satellites with "stable" indices. On the other
6986: hand the factor $\d t_1$ is (and must be) present in the formula
6987: since the dimension of the moduli spaces is odd. With this
6988: information in mind, the enumerative question equivalent to
6989: computation of the satellites can be described as follows. On a
6990: Riemann surface $\Sigma$ of genus $g$, we are given a divisor $D$
6991: of $n$ distinct points with (possibly zero) multiplicities
6992: $m_1,...,m_n$. We have to count the divisors $l_1P_1+...+l_gP_g$
6993: which in the sum with $D$ form the divisor of a rational function.
6994: (In particular, the degree $\sum m_i+\sum l_j$ of the total
6995: divisor must vanish.)   The answer to this question is equal to
6996: the degree of the Abel-Jacobi map $\Sigma^g\to J_{\Sigma}$ defined
6997: by integration of holomorphic differentials
6998: $\vec{\go}=(\go_1,...,\go_g)$ on $\Sigma$ as
6999: \[ (P_1,...,P_g)\mapsto l_1\int^{P_1}\vec{\go}+...+l_g \int^{P_g}\vec{\go}
7000: .\] When the multiplicities $(l_1,...,l_g)=(1,...,1)$, the degree
7001: equals $g!$ (it is well-known that $S^g\Sigma^g\to J_{\Sigma}$ is
7002: a bi-rational isomorphism). For arbitrary $(l_1,...l_g)$ the
7003: Abel-Jacobi map has the Jacobi matrix $[l_j \go_i (P_j)]$. Thus
7004: the degree equals $l_1^2...l_g^2 g!$. Taking these answers as the
7005: coefficients in the generating function on the variables $t_0,
7006: p_l, q_{-l}$ corresponding to $l=0,l>0$ and $l<0$ we arrive at the
7007: factor $u_{xx}^g$. The other factor $(\gd u)^n/n!$ is similarly
7008: accountable for all possible choices of multiplicities
7009: $m_1,...,m_n$ in the divisor $D$. The contour integration of the
7010: product couples the choices with $m_1+...+m_n+l_1+...+l_g = 0$.
7011: 
7012: 
7013: \begin{thebibliography}{99}
7014: 
7015: \bibitem{AbbasHofer}
7016: C.~Abbas and H.~Hofer,
7017:   Holomorphic curves and global questions in contact geometry,
7018:   to appear in Birkh\"auser.
7019: \bibitem{Arnold-conjecture} V.I. Arnold,        {Sur une propri\'et\'e topologique des
7020: applications globalement canoniques de la m\'echanique classique},
7021:       {\it C.\ R.\ Acad. Paris}, 261(1965), 3719--3722.
7022: \bibitem{Arnold-Maslov} V. I. Arnold, On a characteristic class entering in quantization
7023: conditions,
7024: {\it Funct. Anal. and Applic.}, 1(1967), 1--14.
7025: \bibitem{Arnold-steps}  V. I. Arnold,       {First steps in symplectic topology},
7026: {\it Russian Math. Surveys}, 41(1986), 1--21.
7027: 
7028: \bibitem{Bennequin} D. Bennequin,
7029: Entrelacements et \'equations de Pfaff, {\it Ast\'erisque}, 106--107(1983).
7030: \bibitem{Biran-Cieliebak} P. Biran and K. Cieliebak, Symplectic
7031: topology on subcritical manifolds, preprint 2000.
7032: \bibitem{Brieskorn} E. Brieskorn, Beispiele zur
7033: Differentialtopologie von Singularit\"aten, {\it Invent. Math.},
7034: {\bf 2}(1966), 1--14.
7035: \bibitem{Chekanov} Yu. Chekanov, Differential algebra of a Legendrian link.
7036: Preprint 1997.
7037: \bibitem{Conley-Zehnder} C. Conley and E. Zehnder,
7038:         {The {B}irkhoff--{L}ewis fixed point theorem and a conjecture of {V}.{I}.
7039: {A}rnold},
7040:     {\it Invent. Math.}, 73(1983), 33--49.
7041:     \bibitem{Donaldson-invariants} S.K. Donaldson,
7042:     Polynomial invariants for smooth four-manifolds,
7043: {\it Topology}, {\bf 29}(1990),  257--315.
7044: 
7045:      \bibitem{Donaldson} S.K. Donaldson, Symplectic
7046: submanifolds and almost-complex geometry, {\it J. Diff. Geom.},
7047: {\bf 44}(1996), 666-705.
7048: \bibitem{Eliash-ICM} Y. Eliashberg, Invariants in contact
7049: topology, {\it Proc. of ICM-98, Berlin}, Doc. Math., 1998,
7050: 327-338.
7051: 
7052: \bibitem{Eliash-Stein} Y. Eliashberg, Topological characterization of Stein manifolds
7053: of complex dimension $>2$, {\it Int. J. of Math}, {\bf 1}(1991),
7054: 29--46.
7055: 
7056:  \bibitem{Eliash-Gromov-convex} Y. Eliashberg and M. Gromov, Convex symplectic manifolds, {\it Proc.
7057: of Symp. in Pure Math.}, vol.52(1991), Part 2, 135--162.
7058:  \bibitem{EHS} Y. Eliashberg, H. Hofer and S. Salamon, Lagrangian intersections
7059: in contact geometry, {\it Geom. and Funct. Anal.}, 5(1995), 244--269.
7060: \bibitem{Etnyre-Sabloff} J. Etnyre and J. Sabloff,
7061: Coherent Orientations and Invariants of Legendrian Knots, preprint
7062: 2000.
7063:    \bibitem{EH-3ball} Y. Eliashberg, H. Hofer,  A Hamiltonian characterization
7064:    of the three-ball, {\it Differential Integral Equations},
7065:    {\bf 7}(1994),  1303--1324.
7066: 
7067: \bibitem{Floer} A. Floer,
7068: The unregularised gradient flow of the symplectic action,
7069:       {\it Comm. Pure Appl. Math.},
7070:      {\bf 41}(1988),
7071:   775-813.
7072: \bibitem{Floer-Hofer} A. Floer,  H. Hofer, Coherent
7073: orientations for periodic orbit
7074: problems in symplectic geometry, {\it Math. Z.}, {\bf  212}(1993), 13--38.
7075: \bibitem{Fukaya-Ono}
7076:  K. Fukaya and K. Ono, Arnold conjecture and Gromov-Witten invariants, preprint 1996.
7077:  \bibitem{Fukaya-Ono2} K. Fukaya,  K. Ono,  Arnold conjecture and Gromov-Witten invariant
7078: for general symplectic
7079: manifolds, in {\it The Arnoldfest (Toronto, ON, 1997)}, 173--190, Fields
7080: Inst. Commun., 24, Amer. Math. Soc., Providence, RI,
7081: 1999.
7082: 
7083:   \bibitem{FKO3}   K.Fukaya, K.Ono, Y.-G.Oh and H.Ohta, to appear.
7084:   \bibitem{Fulton-Pand} W. Fulton and R. Pandharipande, Notes on stable
7085:   maps and quantum cohomology, {\it Proc. of Symp. in Pure Math.},
7086:   {\bf 62}(1995), Part 2, 45--96
7087:   \bibitem{Gathmann}A. Gathmann,
7088:   Absolute and relative Gromov-Witten invariants of very ample
7089:   hypersurfaces, preprint 1999.
7090:   \bibitem{Getzler} E. Getzler, Topological recursion relations in genus $2$, in {\it Integrable systems
7091:    and algebraic geometry} (Kobe/Kyoto, 1997), 73--106, World Sci. Publishing  1998.
7092: \bibitem{Giroux} E. Giroux, Une structure de contact,  m\^eme tendue est plus ou
7093: moins tordue,
7094: {\it Ann. Scient. Ec. Norm. Sup.}, {\bf  27}(1994), 697--705.
7095: 
7096: \bibitem{Givental-toric} A. Givental,
7097: A symplectic fixed point theorem for
7098:  toric manifolds,
7099:  {\it The Floer memorial volume}, 445--481,
7100:  Progr. Math.,
7101:  133, Birkhäuser, Basel, 1995.
7102:  \bibitem{Givental-Maslov} A. Givental,
7103:  Nonlinear generalization of the Maslov index, {\it Adv. Soviet Math.}, {\bf 1}(1990),
7104:   71--103.
7105: \bibitem{Givental} A. Givental.
7106: Homological geometry and mirror symmetry. Proc. Int. Congress of
7107: Math., Z\"urich-1994. Birkh\"auser, 1995, v. 1, 472-480.
7108: 
7109: 
7110: \bibitem{Givental2} A. Givental.
7111: Homological geometry I: Projective hypersurfaces. Selecta Math.,
7112: New series 1 (2), 1995, 325-345.
7113: 
7114: \bibitem{Givental3}A. Givental.
7115:  Equivariant Gromov - Witten invariants. Intern. Math. Res. Notices, 1996,
7116: No.13, 613 - 663.
7117: \bibitem{Givental1} A. Givental and B. Kim.
7118: Quantum cohomology of flag manifolds and Toda lattices. Commun.
7119: Math. Phys. 168 (3), 1995, 609-641.
7120: 
7121: \bibitem{Graber-Pand} T. Graber and R. Pandharipande, Localization
7122: of virtual classes, {\it Invent. Math.},{\bf  135}(1999),    487--518.
7123: \bibitem{Gray} J.W. Gray, Some global properties of contact structures,
7124: {\it Annals of Math.}, {\bf 69}(1959), 421--450.
7125: \bibitem{Gromov-holomorphic} M. Gromov, Pseudo-holomorphic curves in symplectic manifolds,
7126: {\it Invent. Math.}, {\bf 82}(1985), 307--347.
7127: \bibitem{Gro-PDR} M. Gromov, Partial Differential Relations, Springer-Verlag,
7128: 1986.
7129: \bibitem{Hofer-Weinstein} H. Hofer, Pseudo-holomorphic curves and Weinstein conjecture
7130: in dimension three, {\it Invent. Math.}, {\bf 114}(1993), 515--563.
7131:  \bibitem{Hofer-Wysocki-Zehnder}    H. Hofer, K. Wysocki and E. Zehnder, Properties of
7132: pseudo-holomorphic curves in symplectisations. I. Asymptotics.
7133: {\it  Ann. Inst. H. Poincar\'e, Anal. Non Lin\'eaire}, {\bf
7134: 13}(1996), 337--379.
7135: 
7136: \bibitem{HoferWysockiZehnder:94d}
7137: H.~Hofer, K.~Wysocki, and E.~Zehnder,
7138:   The dynamics on a strictly convex energy surface in {${\bf R}^4$},
7139:   {\it Annals of Math.}, {\bf 148}(1998),  197-289.
7140: \bibitem{HoferWysockiZehnder:99}
7141: H. Hofer, K. Wysocki, and E. Zehnder,
7142:   Finite Energy Foliations of Tight Three-Spheres and Hamiltonian
7143: Dynamics, preprint 1999.
7144: \bibitem{Ionel} E.-N. Ionel, Topological recursive relations in
7145: $H^{2g}(M_{g,n})$, preprint 1999.
7146: \bibitem{Ionel-Parker2} E.-N. Ionel and T.H. Parker
7147:  Gromov-Witten invariants of symplectic sums.{\it  Math. Res. Lett.}
7148:  {\bf 5} (1998),  563--576.
7149:  \bibitem{Ionel-Parker} E.-N. Ionel and T.H. Parker,
7150:  Relative Gromov-Witten Invariants, preprint 1999.
7151: \bibitem{Kollar} J. Kollar,
7152: Rational curves on algebraic varieties, Springer-Verlag, 1996.
7153: 
7154: \bibitem{Kontsevich-quant} M. Kontsevich, Deformation quantization of Poisson manifolds, I,
7155: preprint 1997.
7156: \bibitem{Kontsevich-CP2} M. Kontsevich,  Enumeration of rational
7157: curves via torus action, in {\it The Moduli Space of Curves}, R.
7158: Dijgraaf, C. Faber and G. van der Geer eds., Birkhauser,1995,
7159: 335--368.
7160: \bibitem{Konts-Manin} M. Kontsevich, Yu. Manin,
7161:  Gromov-Witten classes, quantum cohomology, and enumerative
7162:  geometry, {\it Commun.Math.Phys.}, {\bf  164} (1994) 525-562.
7163: 
7164: \bibitem{Li-Tian} J. Li and G. Tian, Virtual moduli cycles and Gromov-Witten invariants of
7165: general symplectic
7166: manifolds, in {\it Topics in symplectic $4$-manifolds (Irvine, CA, 1996)}, 47--83,
7167: First Int. Press Lect. Ser., I, Internat. Press,
7168: Cambridge, MA, 1998.
7169: \bibitem{Liu-Tian} G. Liu and G. Tian, Floer homology and Arnold conjecture,
7170: {\it J. Diff. Geom.}, {\bf 49}(1998),  1--74.
7171: \bibitem{Lutz} R. Lutz,  Structures de contact sur les fibr\'es principaux en cercles de
7172: dimension $3$
7173: {\it  Ann. Inst. Fourier},   XXVII, {\bf 3}(1977), 1--15.
7174:               \bibitem{Martinet}J. Martinet, Formes de contact sur les vari\'et\'es de dimension $3$,
7175: {\it  Lecture Notes in Math.,} {\bf 209}(1971), 142--163.
7176: \bibitem{McDuff} D. McDuff, The virtual moduli cycle, in{\it  Northern California Symplectic
7177: Geometry Seminar}, 73--102,
7178: Amer. Math. Soc. Transl. Ser. 2, 196, Amer. Math. Soc., Providence, RI,
7179: 1999.
7180:     \bibitem{Mori-Miayoka} Y. Myayoka and S. Mori, A numerical
7181:     criterion for uniruleness, {\it Annals of Math}, {\bf
7182:     124}(1986), 65--69.
7183:     \bibitem{Morita} S. Morita, A topological classification of
7184:     complex structures on $S^1\times \Sigma^{2n-1}$,
7185:     {\it Topology}, {\bf 14}(1975), 13--22.
7186: \bibitem{Ng} L. Ng, On Invariants of Legendrian Knots, preprint
7187: 2000.
7188: \bibitem{Robbin-Salamon} J.  Robbin, D. Salamon,The Maslov index for
7189: paths, {\it Topology}, {\bf  32}(1993), 827--844.
7190: 
7191: \bibitem{Ruan} Y. Ruan,  Topological sigma model and
7192: Donaldson-type invariants in Gromov theory. {\it Duke Math. J.,},
7193: {\bf 83}(1996),  461--500.
7194:             \bibitem{Ruan-Li} Y. Ruan and A.-M.
7195:             Li,
7196:              Symplectic surgery and Gromov-Witten invariants of Calabi-Yau 3-folds
7197:              I, preprint 1999
7198:     \bibitem{Ruan-virtual} Y. Ruan,  Virtual
7199:     neighborhoods and pseudo-holomorphic curves, {\it Proceedings
7200:     of 6th G\"okova Geometry-Topology Conference, Turkish J. Math.}, {\bf  23} (1999),
7201:     161--231.
7202: \bibitem{Ruan-Tian} Y. Ruan and G. Tian, A mathematical theory of
7203: quantum cohomology, {\it J. Diff. Geom.}, {\bf 42}(1995),259--367.
7204: \bibitem{Siebert} B. Siebert,
7205: Gromov-Witten invariants of general symplectic manifolds, preprint 1997.
7206:    \bibitem{Salamon:Floer} D. Salamon. Lectures on Floer homology,
7207: in ``Symplectic Geometry and Topology'', IAS/Park City Mathematics Series, v. 7, AMS/IAS
7208: 1999, 144 -- 229.
7209:  \bibitem{Ustilovsky}  I. Ustilovsky, Infinitely many contact
7210:  structures on $S^{4m+1}$, {\it Int. Math. Res. Notices}, {\bf
7211:  14}(1999), 781--791.
7212: 
7213:  \bibitem{Vakil} R. Vakil, The enumerative geometry of rational
7214:  and elliptic curves in projective space, preprint 1997.
7215:  \bibitem{Viterbo} C. Viterbo, in preparation.
7216:  \bibitem{Weinstein}  A. Weinstein,
7217:  On the hypotheses of Rabinowitz's periodic orbits theorems
7218: {\it J. Diff. Eq. }, {\bf 33}(1979), 353--358.
7219:  \bibitem{Witten1} E. Witten,   Supersymmetry and Morse theory,
7220:  {\it J. Diff. Geom.}, {\bf 17}(1982), 661--692.
7221:  \bibitem{Witten2} E. Witten, Two-dimensional gravity and
7222:  intersection theory on moduli space, {\it Surveys in Diff.
7223:  Geom.}, {\bf 1}(1991), 243--310.
7224: \bibitem{MLYau} M.-L. Yau, Contact homology of subcritical Stein
7225: manifolds, thesis, Stanford University, 1999.
7226: 
7227: 
7228: 
7229:            \end{thebibliography}
7230: 
7231: \end{document}
7232: %%%% To be continued        1)
7233: 
7234: 2) Fukaya, Kenji; Ono, Kaoru Arnold conjecture and Gromov-Witten
7235: invariant.
7236: 
7237: 3) Li, Jun; Tian, Gang
7238: 
7239: 4)   dg-ga/9608005
7240: 
7241:     Notes: LaTeX2e with AMS-fonts, 73 pages, 3 figures, full revision
7242:            12/1997
7243: 
7244: 5)
7245: